Vous êtes sur la page 1sur 112

EP00 - Introducing the E&P Business

Shell Learning Centre

CHAPTER 1 THE RESERVOIR


Contents
1.1 HYDROCARBON RESERVOIRS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Reservoir Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Seals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Sources of Hydrocarbons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Hydrocarbon Migration and Entrapment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 THE GEOLOGICAL TIMESCALE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 ROCK TYPES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 THE STRUCTURE OF THE EARTH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 PLATE TECTONICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 Divergent Plate Margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 Convergent Plate Margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 Conservative Plate Margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 SEDIMENTARY BASINS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 BASIN CLASSIFICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 HYDROCARBON ACCUMULATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 The Nature of Petroleum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 The Origin of Petroleum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 The Generation of Petroleum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 The Generation of Natural Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 The Generation of Solid Petroleum and Oil Shale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31 The Migration of Petroleum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 HYDROCARBON PLAYS AND PROVINCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 RESERVOIR STRUCTURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 THE THEORY OF STRESS AND STRAIN IN ROCKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 ELASTIC DEFORMATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 BRITTLE DEFORMATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .33 Fractures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .33 Faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .33 Stress and the Formation of Faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .33 Description of Faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .34 PLASTIC DEFORMATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .35 CLASSIFICATION OF STRUCTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .37 EXTENSIONAL STRUCTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .37 Faults in an Extensional Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .37 Sealing of Normal Faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .39 Folds in an Extensional Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .40 ROLLOVER ANTICLINE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .41 COMPRESSIONAL STRUCTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .42 Faults in a Compressional Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .42 Folds in a Compressional Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .42 WRENCH STRUCTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .43 Faults in a Wrench Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .44 Folds in a Wrench Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .45 DIAPIRISM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .46

1.2

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

13

EP00 - Introducing the E&P Business

Shell Learning Centre

1.3

RESERVOIR ARCHITECTURE . . . . . . . . . . . . . . . . . . . . . . . . GEOLOGICAL CYCLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Physical and Chemical Weathering of Rocks . . . . . . . . . . . . . . . . . . Transport of Weathered Material . . . . . . . . . . . . . . . . . . . . . . . . . . Transport of the Soluble Products of Chemical Weathering . . . . . . . . . SEDIMENTARY ENVIRONMENTS . . . . . . . . . . . . . . . . . . . . . . . . . SILICICLASTIC ENVIRONMENTS OF DEPOSITION . . . . . . . . . . . . . . Alluvial Fan Deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Braided River Deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Meandering River Deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Deltas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Syn-sedimentary Faulting in Deltas . . . . . . . . . . . . . . . . . . . . . . . . . CARBONATE ENVIRONMENTS OF DEPOSITION . . . . . . . . . . . . . . Deposition of Evaporites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . TEXTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Grain size distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Grain shape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Rock fabric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Facies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . DIAGENESIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Cementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Clay Minerals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dolomitisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . RESERVOIR PORE SYSTEM POROSITY . . . . . . . . . . . . . . Definition . . . . . . . . . . . . . . Factors Affecting Porosity . . . . Porosity in Carbonates . . . . . . Types of Porosity . . . . . . . . . . Typical Values of Porosity . . . . PERMEABILITY . . . . . . . . . . . Definition . . . . . . . . . . . . . . Factors Affecting Permeability . Typical Values of Permeability . Permeability Anisotropy . . . . . NET RESERVOIR . . . . . . . . . . Definition . . . . . . . . . . . . . . Practical Considerations . . . . . ...... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... ....... . . . . . . . . . . . . . . . ........ ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... ......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... ..........

......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... ......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... ......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... ..........

. . . . . . .49 . . . . . . . .49 . . . . . . . .49 . . . . . . . .51 . . . . . . . .52 . . . . . . . .53 . . . . . . . .53 . . . . . . . .53 . . . . . . . .54 . . . . . . . .55 . . . . . . . .57 . . . . . . . .60 . . . . . . . .61 . . . . . . . .63 . . . . . . . .63 . . . . . . . .63 . . . . . . . .63 . . . . . . . .64 . . . . . . . .64 . . . . . . . .65 . . . . . . . .65 . . . . . . . .66 . . . . . . . .66 . . . . . . . .67 . . . . . . . .67 . . . . . . .68 . . . . . . . .68 . . . . . . . .68 . . . . . . . .68 . . . . . . . .69 . . . . . . . .70 . . . . . . . .70 . . . . . . . .70 . . . . . . . .70 . . . . . . . .72 . . . . . . . .72 . . . . . . . .73 . . . . . . . .73 . . . . . . . .73 . . . . . . . .74 . . . . . . .75 . . . . . . . .75 . . . . . . . .75 . . . . . . . .77 . . . . . . . .77 . . . . . . . .78 . . . . . . . .79 . . . . . . . .82 . . . . . . . .82 . . . . . . . .83 . . . . . . . .84 . . . . . . . .85

1.4

1.5

RESERVOIR FLUIDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CHEMISTRY OF HYDROCARBONS . . . . . . . . . . . . . . . . . . . . . . . . Hydrocarbon Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Classification of Crude Oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CLASSIFICATION OF RESERVOIR FLUIDS . . . . . . . . . . . . . . . . . . . . THE BEHAVIOUR AND PHYSICAL PROPERTIES OF RESERVOIR FLUIDS Hydrocarbon Phase Behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . Dry Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wet Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Retrograde Gas Condensate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Volatile Oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Black Oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

14

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

PROPERTIES OF HYDROCARBON GASES . . . . . . . . . The Real Gas Equation of State and the "z-Factor" . . . . Amagat's law . . . . . . . . . . . . . . . . . . . . . . . . . . . . SUBSURFACE AND SURFACE VOLUME RELATIONSHIPS Oil Formation Volume Factor . . . . . . . . . . . . . . . . . . Gas Formation Volume Factor . . . . . . . . . . . . . . . . . . Solution Gas Oil Ratio . . . . . . . . . . . . . . . . . . . . . . . Producing Gas Oil Ratio . . . . . . . . . . . . . . . . . . . . . OTHER HYDROCARBON PROPERTIES . . . . . . . . . . . . Gas Compressibility . . . . . . . . . . . . . . . . . . . . . . . . Gas Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Gas Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Liquid Compressibility . . . . . . . . . . . . . . . . . . . . . . . Liquid Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . Liquid Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . RESERVOIR FLUID CORRELATIONS . . . . . . . . . . . . . . PROPERTIES OF FORMATION WATERS . . . . . . . . . . . Water Compressibility . . . . . . . . . . . . . . . . . . . . . . . Water Formation Volume Factor . . . . . . . . . . . . . . . . Water Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . CHARACTERISTICS OF FORMATION WATER . . . . . . . 1.6

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . .

. .87 . .88 . .90 . .92 . .94 . .95 . .96 . .96 . .97 . .97 . .97 . .97 . .98 . .98 . .99 . .99 .100 .100 .100 .101 .101

INITIAL RESERVOIR CONDITIONS . . . . . . . . . . . . . . . . . . . PRESSURE VERSUS DEPTH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NORMAL AND ABNORMAL PRESSURES . . . . . . . . . . . . . . . . . . . . Causes of Abnormal Pressures . . . . . . . . . . . . . . . . . . . . . . . . . . . The Impact of Abnormal Pressures on Drilling Operations . . . . . . . . . CAPILLARY PRESSURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surface Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Wettability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The Capillary Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . PRESSURE PROFILES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . SATURATION PROFILES AND CAPILLARY PRESSURE CURVES . . . . . . . DRAINAGE AND IMBIBITION . . . . . . . . . . . . . . . . . . . . . . . . . . . . RELATIVE PERMEABILITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . RESERVOIR TEMPERATURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... .......... ..........

. . . . . .103 . . . . . . .103 . . . . . . .104 . . . . . . .105 . . . . . . .106 . . . . . . .108 . . . . . . .108 . . . . . . .109 . . . . . . .109 . . . . . . .111 . . . . . . .112 . . . . . . .116 . . . . . . .118 . . . . . . .123

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

15

EP00 - Introducing the E&P Business

Shell Learning Centre

16

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

1.1 HYDROCARBON RESERVOIRS


INTRODUCTION
The purpose of this section is to introduce the important concepts of the formation of hydrocarbon accumulations and some of their key features.

Reservoir Units
Reservoirs are often made of sandstone, which consists of grains of the mineral quartz (pure silica, Si02). The hydrocarbons are situated in the spaces, or pores, between the grains. The porosity of the reservoir is the fraction of the total reservoir volume that consists of pores. Hydrocarbons cannot flow into or out of the reservoir unless the pores are connected together. The permeability of the reservoir is a measure of the connectivity of the pores.

Seals
Unless the reservoir is capped by an impermeable rock unit, or seal, there would be nothing to stop the relatively light hydrocarbons rising to the surface. Many oil and gas reservoirs are capped by seals of impermeable shales, which were muddy sediments when deposited. Some are sealed by impermeable units which were formed by the evaporation of seawater, such as anhydrite (CaS04) or halite (NaCI).

Traps
The reservoir must also have a three-dimensional shape which is capable of containing the accumulated hydrocarbons and preventing them from migrating elsewhere. This "container" is known as a trap, or structure. For example, the structure in Figure 1.1 is a dome which was formed by folding the layers of rock. There are three types of trap. Structural traps are formed as a result of the deformation of rock strata in the Earth's crust. StratigraphIc traps occur where there is a permeability barrier caused by variation in sedimentary rock types. Some traps have both structural and stratigraphic features and are known as combination traps. 78% of the crude oil which has been discovered globally to date is held in structural traps, 13% in stratigraphic traps and 9% in combination traps.

Sources of Hydrocarbons
Hydrocarbons are normally not generated in the reservoir - they migrate into the reservoir from a source rock. Source rocks must contain large amounts of organic material, from which oil and gas are created. Most source rocks are shales, as these rocks contain 90% of all the organic material which is found in
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

17

EP00 - Introducing the E&P Business

Shell Learning Centre

sediments. Source rocks themselves rarely form economic reservoirs because they have extremely low permeability. Source rocks need to be heated to over 10OC for a significant period of geological time in order to mature (yield oil or gas).

Hydrocarbon Migration and Entrapment


When mature, hydrocarbons are liberated from the source rock and float towards the surface, as they have a lower density than the water that saturates sedimentary rock. If the hydrocarbons encounter a reservoir which has a suitable seal and structure they will be trapped and form an accumulation. If a reservoir is not encountered, the hydrocarbons will simply continue migrating to the surface and escape. Over 95% of mature hydrocarbons have escaped in this manner. Clearly, the timing of events is crucial - the hydrocarbons must mature and migrate after the formation of the reservoir, seal and structure. These events take place over a very long timescale, an appreciation of which is necessary to understand the way in which hydrocarbon accumulations form.

THE GEOLOGICAL TIMESCALE


The Earth has been in existence for approximately 4.6 billion (4.6 x 109 ) years. During this time, numerous geological events and processes have taken place on the Earth's surface, leading to the formation of a series of rock layers. The age of these layers can be measured using absolute or relative dating techniques. Absolute ages of rocks can be determined using radioactive isotopes (e.g. isotopes of U, Th, Rb, Sr). Assuming the half-life of a particular isotope is accurately known, the formation age can be calculated from the present-day parent/daughter isotope ratio. However, sedimentary rocks contain little or no radioactive material. The formation age of these rocks is therefore measured using relative dating techniques. These techniques are based on simple concepts. Nicolauz Steno (1638-1687) proposed the law of superposition, which states that the older rocks in a normal sequence lie underneath younger ones. He also realized that most strata are deposited slowly and in a near-horizontal position, although later they may be folded or even overturned. Georges Cuvier (1769-1832) noticed that rock sequences often contain linear features or horizons which mark a change in orientation of the layers. These horizons, known as unconformities, represent breaks in

18

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

geological time during which the underlying beds were tilted and eroded before deposition of the next layer. The stages in the formation of an unconformity are shown in Figure 1.2. These methods for determining the ages of rocks allow geologists to determine a stratigraphic column, or sequence of rocks arranged in chronological order, for a particular locality. However, problems arise when it is necessary to correlate between localities that are separated by large distances or regions where no rocks are exposed. For these purposes, fossils are used to date rocks. Fossil evidence suggests that organisms first appeared on Earth c. 600 million years ago. Since then, evolutionary change has produced a large diversity of life forms. It is assumed that the evolution of a group of organisms proceeds from simpler to more complex organisms. Therefore, the fossil remains found in a series of rocks can be used to determine their relative ages. Shelly marine fossils are most commonly used as correlative tools, as the original organisms occurred in large numbers over a wide geographical region, and underwent rapid evolutionary change over short time intervals. For example most ammonite species lasted for ~ 2 million years before becoming extinct, which is very short compared to the entire geological timescale. By combining both absolute and relative dating methods, a geological timescale can be constructed (see Table 1.1).

When reading geological accounts of reservoirs or oilfields, you should be aware that three different sets of terminology may be used to describe the stratigraphy of a region; lithostratigraphy, biostratigraphy and chronostratigraphy. Lithostratigraphy (rock stratigraphy) describes the sequence of rock units in terms of their mineralogy, petrography, and internal structure. Biostratigraphy describes the rock succession in terms of fossil content
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

19

EP00 - Introducing the E&P Business

Shell Learning Centre

and the relative ages of the units. Chronostratigraphy uses an internationally agreed hierarchy of units to indicate the absolute age of the rock layers.

ROCK TYPES
All rocks fall into one of three broad categories : Igneous rocks contain minerals that have crystallized from magma. If the magma is situated deep in the Earth's crust, its rate of cooling will be slow, thus favouring the formation of coarse-grained rock, e.g. granite. Faster cooling at or near the Earth's surface results in finer grained rocks, such as basalt erupted from volcanoes. Metamorphic rocks are formed by the alteration of pre-existing rocks due to the application of heat and / or pressure. One of the best known types of metamorphic rock is marble, which is formed when limestone is subjected to high temperature and pressure. Sedimentary rocks. Clastic sedimentary rocks are formed by the weathering, transport and deposition of debris derived from pre-existing rocks. Carbonate rocks are formed by organic or inorganic precipitation of calcium carbonate. Evaporitic rocks, such as halite (NaCI), form due to partial or complete evaporation of seawater. When igneous and metamorphic rocks are exposed at the Earth's surface they are subjected to an environment which is totally different to that in which they were formed. The temperature and pressure are both much lower, and the environment is oxygen rich and often wet. As a result, many silicate minerals undergo changes that may lead to their total or partial breakdown and, subsequently, to the destruction of the rocks themselves. These breakdown processes are known collectively as weathering, and it is the products of weathering that eventually accumulate to form sedimentary rocks. Sedimentary rocks themselves may be subjected to renewed cycles of weathering. Nearly all hydrocarbon reservoir units are sedimentary rocks. After deposition, they were buried by later episodes of sedimentation. During burial, the porosity and permeability of the rocks often change as the grains become packed more tightly together and new minerals are precipitated in the pore spaces. Most hydrocarbons occur in the Earth's major sedimentary basins - regions of the crust which have subsided to accumulate thick sequences of sedimentary rocks. Oil and gas also occur in some mountain or orogenic belts. The development of sedimentary basins is linked to past movements of huge slabs of the Earth's crust, described by the theory of plate tectonics. Before describing sedimentary basins in more detail, it is necessary to first consider the structure of the earth and the theory of plate tectonics.

THE STRUCTURE OF THE EARTH


Almost 5 billion years ago, the Earth existed as a swirling mass of gases. With time, these gases cooled and condensed to form a solid body, spinning on a vertical axis. As a result, the most dense material became concentrated at the centre. This effect can be illustrated from calculations of the average density of the Earth (5500 kg/m2) from astronomic observations. Rocks at the surface have lower densities, ranging from 2500 to 3500 kg/m2, suggesting that, indeed, rocks towards the centre must have a higher than average density. The structure of the interior of the Earth has been investigated by refraction seismology. Shock waves from earthquakes and nuclear explosions can be detected at various points around the globe, and their arrival times used to determine the depths of major density changes within the Earth. These studies suggest that the Earth's interior can be divided into 3 main layers: the core, mantle and crust (Figure 1.3). The broad chemical composition of these layers is determined by comparison with rocky meteorites that have landed on the Earth's surface. The core is made of a complex iron-nickel alloy and consists of an inner and outer layer. The outer core is liquid and its movement is responsible for the Earth's magnetic field.

20

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The mantle is a very thick layer and occupies almost 70% of the Earth's volume. It is made of silicate (SO44-) minerals and, although solid, can undergo plastic deformation over very long time periods (so-called mantle convection currents). The Earth's crust varies in thickness from 6 to 90 km, with an average of 35 km. The crust is therefore very thin relative to the average radius of the planet (6370 km).

There are important differences between oceanic crust, which floors the ocean basins, and continental crust beneath the continents and continental shelves. The oceanic crust is thin (6-12 km) compared to the continental crust (30-90 km) (see Figure 1.4) and also has a more homogeneous composition. Seismic investigation of oceanic crust has shown that the deep ocean basins are floored by rocks which consist only of basalts and gabbros (coarse-grained equivalent of basalt). In contrast, continental crust has a more varied composition. The lower crust consists mainly of diorites (a type of igneous rock). The upper crust is more granitic with some metamorphic rocks, and is covered by a layer of sedimentary rocks which can be several kilometres thick.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

21

EP00 - Introducing the E&P Business

Shell Learning Centre

PLATE TECTONICS
In the 19th century, Alfred Wegener, a German geologist, recognised the geometric similarity of the coastlines of Africa and South America. He proposed the concept of continental drift -that continents, now far removed from one another, may have previously been joined. However, it is only since the development of modern geophysics that a feasible explanation has been provided for movements in the Earth's crust. In the 1960s, geophysicists noticed that major earthquakes and volcanoes are restricted to narrow zones on the Earth's surface (Figure 1.5). They concluded, therefore, that portions of the crust behave as rigid plates which deform at their edges. The term lithosphere is used to describe the crust and uppermost mantle that is mechanically rigid.

Further geophysical investigations into the Earth's interior showed that below the lithosphere there is a zone of ductile deformation known as the asthenosphere. It permits both lateral movement of the plates, and vertical crustal movements to compensate changes in crustal thickness during plate motions. Plate tectonic theory (Figure 1.6) therefore describes plates of lithosphere about 100 km thick overlying a layer of relatively low strength (the asthenosphere). The movement of these plates is driven by the continuous cooling of the Earth, which produces large scale convection cells in the mantle. In the oceans, oceanic lithosphere is created at divergent margins, and consumed at convergent margins. Along a third type of margin, oceanic plates slide past each other, forming conservative margins.

22

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Due to the structural and compositional heterogeneity of continental lithosphere, plate tectonics in the continental regions is considerably more complicated than in the oceans. Rather than occurring in narrow zones, deformation of the continents takes place over large regions. For example, the Himalayan fold belt, the result of continental collision between the Indian subcontinent and Eurasia, is over 2000 km wide. In the following sections, the typical structures formed at divergent, convergent, and conservative margins are described.

Divergent Plate Margins


Divergent plate margins in oceans generate oceanic crust by allowing molten mantle material to solidify at the surface. Oceanic crust covers about 70% of the Earth's surface and has a simple structure. The main features of divergent plate boundaries are shown in Figure 1.7.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

23

EP00 - Introducing the E&P Business

Shell Learning Centre

Ocean ridges form over the rising asthenosphere because this part of the crust is hotter, and therefore less dense, than the surrounding regions. Mantle material is shown in Figure 1.7 rising under the ridge, "filling the gap" as the two plates move away from each other. As the newly formed crustal material moves away from the spreading centre, it cools, and so its elevation decreases. Divergence in a continental plate causes thinning of the lithosphere by extension to form a sedimentary basin (Figure 1.8). As divergence continues, the two parts of the continental plate move apart, oceanic crust is formed at a new mid-ocean ridge and an ocean is born with continental or passive margins on either side.

Convergent Plate Margins


There are three types of convergent plate margins: ocean-ocean convergence, ocean-continent convergence, and continent-continent convergence. Convergence involving oceanic lithosphere results in consumption of lithospheric material at a subduction zone as it sinks into the underlying mantle (Figure 1.9). This material then heats up and melts to form a source of magma for volcanoes at the surface. A line of volcanoes, called a volcanic arc (or island arc if it is partly immersed), normally forms above a subduction zone. Continental lithosphere is too buoyant to be consumed at convergent margins, so the lithosphere becomes very thick to form mountain belts. The high pressures and temperatures associated with mountain formation, or orogenesis, cause metamorphism of the existing rocks. Most mountain belts have a thickened core of high pressure metamorphic rocks and some igneous material.

Conservative Plate Margins


The features formed at conservative margins can be considered as a combination of those from divergent and convergent margins, depending on the overall direction of plate movement and the shape of the plate boundary (Figure 1.10). If the plates are moving obliquely away from one another, the margin is said to be transtensional and small extensional basins are formed (Figure 1.10a). Conversely, if the relative motion of the plates is obliquely toward each other, small mountains and compressional basins form and the margin is termed transpressional (Figure 1.10b). Inevitably, real conservative margins do not have a straight line boundary, and both extensional and compressional features exist (Figure 1.10c). Tectonic activity at all of these margins causes the formation of sedimentary basins. Each margin is typified by certain kinds of basin. In the following section, the mechanisms are described which lead to the formation of sedimentary basins.

24

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

25

EP00 - Introducing the E&P Business

Shell Learning Centre

SEDIMENTARY BASINS
In simple terms, sedimentary basins are holes in the Earth's surface into which sediment is deposited. Many different types of sedimentary basins can be found at various localities around the world today. However, all of these basins can be explained by one of two simple basin-forming mechanisms; stretching or flexing of the lithosphere. Presumably, the same mechanisms have existed throughout geological time. As described in the previous section, when lithosphere is stretched it thins to form an extensional basin. The upper crust deforms by brittle failure, forming large faults and tilted blocks (Figure 1.11). The underlying mantle behaves in a ductile fashion and forms a 'necked' or thinned region, which is slightly wider than the brittle faulted zone. Subsidence in these basins occurs in two distinct phases; rift subsidence as a result of active stretching and thinning of the lithosphere, and thermal subsidence due to passive cooling and sinking of hot, upwelled mantle (thermal 'sagging').

During active stretching, subsidence is very rapid and large thicknesses of coarse-grained sediment are deposited. In the cooling phase, subsidence is slower and the basin is covered by a blanket of finer-grained material.

26

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

If stretching continues and oceanic crust is formed, the passive margins of the ocean continue subsiding and allow further sediment deposition. Passive margins, like those surrounding the North Atlantic, can accumulate several kilometres of sediment. This sediment load further amplifies the subsidence due to tectonic activity. Flexing occurs when a large load (e.g. a mountain belt or overriding plate) is imposed on the lithosphere. The resultant flexural basins subside very rapidly and fill up with sediment eroded from the nearby mountains or overriding plate (Figure 1.12). A further phenomenon of lithospheric flexure is the formation of a peripheral bulge which delineates the edge of the basin. Both basin-forming mechanisms can occur in oceanic or continental lithosphere. However, the largest accumulations of sediment occur in continental basins. Most hydrocarbon-bearing basins are of the extensional type.

BASIN CLASSIFICATION
In divergent regions, only extensional basins form. These can be intracontinental (within the continental plate, e.g. the North Sea Basin) or on the continental margin (e.g. the US North Atlantic Margin). In convergent zones, both flexural and extensional basins can form. During ocean-ocean or ocean-continent convergence, flexural basins form on the underriding plate (as trenches) and in front of the volcanic arc on the overriding plate (fore-arc basins - Figure 1.13). In addition, the action of the downgoing oceanic plate causes extension in the plate above, behind the volcanic arc, forming a back-arc basin (e.g. the Aegean Sea). During continent-continent convergence, flexural basins form on one or both sides of the mountain belt (see Figure 1.12). These are called foreland basins (e.g. the Po Valley and Molasse Basins in the Alps). Intracontinental, continental margin, and, to a lesser extent, foreland basins are of most importance in petroleum exploration.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

27

EP00 - Introducing the E&P Business

Shell Learning Centre

HYDROCARBON ACCUMULATIONS
The Nature of Petroleum
Petroleum is a general term given to naturally occurring substances which consist predominantly of hydrocarbons. Petroleum is found in gaseous (natural gas), liquid (crude oil) and solid (asphalt) states. Typical elemental compositions of petroleum in each of these states are given in Table 1.2.

Over 1200 different hydrocarbons compounds have been identified in crude oil. The number of carbon atoms in one molecule of each compound is used as the basis for subdividing the components of petroleum into seven main groups or fractions:C1-C3 Gases C4-C10 Gasoline C11-C13 Kerosene C14-C18 Diesel fuel C19-C25 Heavy gas oil C26-C40 Lubricating oil Over C40 Waxes

28

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The Origin of Petroleum


Most compounds that are found in petroleum are of organic origin. To be preserved for conversion into petroleum, the organic matter must be deposited and buried rapidly in an oxygen- and scavenger-free environment. These conditions are commonly found in continental shelf marine environments, where oxygen-free bottom waters exist together with a relatively high rate of sedimentation of inorganic material, such as clay or sand particles. While much of the organic material in marine sediments may be generated in situ, rivers may also transport land-derived material into shallow seas. Table 1.3 lists the annual production of organic carbon from each of the seven main types of environment found on Earth.

Land plants appeared 400 million years ago in the Devonian period, so the main source of organic material before this must have been marine. From the Precambrian until the Devonian, the largest producer of organic matter was phytoplankton, microscopic marine algae which live in the upper layers of the oceans and sink to the sea floor when they die. Here the water is oxygen-free and the algae are buried by deposition of mud. Muddy organic-rich deposits of this type have the potential to become source rocks for the generation of petroleum.

The Generation of Petroleum


On burial and exposure to heat and pressure, organic material degrades into an intermediate substance known as kerogen, from which petroleum is subsequently formed. Three kinds of kerogen have been identified. Differences in the elemental ratios of hydrogen:carbon and oxygen:carbon reflect the different source material for the kerogen types, each of which yields different products when mature:

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

29

EP00 - Introducing the E&P Business

Shell Learning Centre

The maturation of kerogen into crude oil and natural gas is achieved mainly by an increase in temperature. Therefore the formation of petroleum depends on the depth of burial and the geothermal gradient in the crust. Figure 1.13 shows the relative proportions of crude oil and gas formed at increasing depth below the surface, for a region where the geothermal gradient is about 3.5C/ 100m. Maximum crude oil production is about 2700 m and the peak for gas is at about 3500 m.

Petroleum is probably generated from kerogen in very small amounts at temperatures below 50C, with the peak production at about 100C. When the temperature rises above 150C, even for a short time, longer chain hydrocarbon molecules will begin to break down or "crack", yielding shorter chain molecules. The first gas to be produced contains some C4 to C10 compounds and is termed "wet gas", but as the temperature increases these break down to give C1 to C3 gaseous compounds and the product is then called "dry gas". Geothermal gradients vary from basin to basin. Source rocks that remain at very shallow depths will not normally generate petroleum. Exposure to excessive heat by very deep burial favours the production of gas only.

The Generation of Natural Gas


It is possible to trace a crude oil back to its source by comparing its geochemical analysis with that of hydrocarbon extracts from a particular source rock. Such an analysis is often not possible for a gas accumulation, particularly when the gas consists almost entirely of methane. There are different kinds of natural gas, with different compositions and modes of formation. Methane is produced by bacteria in the very early stages of accumulation and anaerobic decay - it occurs at low pressures and normally escapes into the atmosphere as "marsh gas", but it can be trapped at shallow depths. Coal is an important source of natural gas and it has been established that coal is able to generate large enough volumes of methane for commercial accumulations to form. Coal is formed from terrestrial organic matter and thus has a strong similarity to type III kerogen (Table 1.4), which tends to yield gas rather than
30
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

oil during maturation. It is thought that liquid hydrocarbons that are generated in small amounts during coal formation are largely absorbed by the coal and prevented from primary migration. Figure 1.13 and Table 1.4 indicate that during the maturation of any kerogen, natural gas and crude oil can form simultaneously. In the later stages of maturation the crude oil can be converted into wet gas and finally, with increased temperature, into dry gas.

The Generation of Solid Petroleum and Oil Shale


Crude oil that has migrated through a reservoir can sometimes reach the surface, where the volatile components escape and the rock becomes impregnated with a hydrocarbon residue. Such a rock is known as an oil sand or tar sand. The Athabasca oil sands of Alberta, Canada, are such a deposit, containing 600 x 109 barrels of asphaltic hydrocarbons. The oil is too viscous to flow and the rock has to be heated to extract the oil. The Athabasca oil sands are mined by open cast methods. Pitch lakes are large deposits of asphalt formed from oil that has seeped to the surface, accumulated in large depressions and become solid. Examples include the Bermudez Lake in Eastern Venezuela, which has estimated reserves of 6.1 million tonnes. Shales containing significant amounts of unaltered kerogen are called oil shales. "Synthetic" crude oil can be produced by heating oil shales to about 500C. If the kerogen forms less than 2.5% of the rock, its total calorific value is taken up heating the shale, which contributes to a typical economic limit for synthetic oil production from oil shales in the USA of 5% kerogen. This yields about 25 litres of oil per tonne of rock. Some of the largest deposits of oil shales are located in the American states of Utah, Wyoming and Colorado. It has been estimated that the oil shales could yield about 18000 x 109 barrels of oil, but because of the escalating costs, and a shortage of available water for processing, most of the projects involved with the production of synthetic crude oil were temporarily abandoned in the early 1980s. However, it is still being distilled in the Baltic region of the USSR and by the Chinese in Manchuria.

The Migration of Petroleum


Although the generation of petroleum in source rocks is fairly well understood, its migration into other rocks is not. The source rocks are mainly organic-rich shales, which are very fine grained and impermeable. So how can the generated petroleum move out of the source rocks? One current suggestion is that when kerogen is subjected to maturation temperatures, methane is formed which creates so much internal pressure within the source rocks that microfractures are formed which then permit the primary migration of the hydrocarbons. Having escaped out of the source rocks, the petroleum appears to move with greater freedom along joints, faults and bedding planes into a reservoir rock such as sandstone.

HYDROCARBON PLAYS AND PROVINCES


In this section so far we have described the set of geological circumstances necessary for the accumulation of oil and gas - the hydrocarbon play. As knowledge of a play is improved by exploration, quantitative estimates of hydrocarbons likely to have been trapped can be made in order to evaluate the likely success of exploration drilling. The North Sea is a large area of the Earth's crust in which circumstances favourable to hydrocarbon generation and entrapment have occurred, and in which many different plays occur. Such areas are termed hydrocarbon provinces. Let us now summarise the circumstances required for a hydrocarbon play to exist, using the Brent field from the North Sea province as an illustration: Source rock for oil: organic rich shales of the Kimmeridge Clay. Reservoir: delta and river sands of the Brent and Statfjord Formations. Seal: various shale units. Traps: due to faulting that preceded deposition of the seal. Timing of oil generation: 40 - 60 million years ago, after the formation of the trap.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

31

EP00 - Introducing the E&P Business

Shell Learning Centre

1.2 RESERVOIR STRUCTURE


THE THEORY OF STRESS AND STRAIN IN ROCKS
Rocks, like other materials, change shape when they are acted upon by stress, such as that generated by the motion of the Earth's plates. This induces strain in the rock (a change in shape or position). The relationship between stress and strain for rocks has been investigated using experiments on rock samples (Figure 1.15).

The relationship in Figure 1.15 is for a particular set of circumstances. The actual response of a rock to stress is affected by: Pressure and temperature: both increase with depth, which leads to more ductile behaviour. Rock competence (internal strength): well-cemented sandstones and carbonates can withstand the greatest effective stress and when they fail, they tend to do so in a brittle manner. These are examples of competent rocks. Shales, mudrocks and evaporite rocks tend to fail under smaller stresses and in a ductile manner - these are incompetent rocks. Reservoir rocks are usually competent. They are often faulted, and almost always contain many minor fractures, whereas cap rocks behave in a ductile manner and contain no fractures at all. Strain rate (rate of deformation): rocks deform over periods of millions of years. For example, the Viking Graben in the North Sea stretched by a factor of 50% over 20 million years, which represents a strain rate of around 10-13 to 10-14 per second. Strain rate has a large effect on the behaviour of a rock. Low strain rates normally cause permanent deformation of the rock, in either a brittle or ductile mode. Higher strain rates often lead to elastic behaviour of rocks (e.g. flexure of the lithosphere as a result of rapid loading). When considering the rheology or mechanical properties of a rock, it is therefore necessary to describe the conditions at the time of deformation. The types of deformation given in Figure 1.15, and described below, are for surface pressure and temperature conditions and moderate strain rates. It is assumed that behaviour changes are due to increased stress only.

32

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

ELASTIC DEFORMATION
Very low differential stress (less than 5 kbar) with very low confining pressure induces temporary or elastic deformation in rocks. The rock recovers to its original shape when the stress is removed. For example, when shock waves from earthquakes pass through the Earth. During this phase of deformation, there is a linear relationship between stress and strain. At a certain level of stress a rock will reach its elastic limit, after which this linear relationship breaks down. The maximum strength of the rock is defined by the yield point, which is followed either by a period of ductile deformation or brittle failure.

BRITTLE DEFORMATION
As the stress increases the linear relationship between stress and strain is lost. If the rock breaks or fractures there is a sudden drop in the effective stress. If the rock moves along the fracture or fault plane the stress continues to drop as the rock undergoes brittle deformation (Figure 1.15).

Fractures
Fractures are brittle failure planes which show no displacement of beds whatever from one side to the other. Fractures occur on surfaces that are exposed at the present day, where they are called joints, but they become less frequent at depth. They characteristically affect competent rocks rather than incompetent ones, and in sections of mixed competency, say limestones and shales, it is common to see vertical fracture planes through the limestones whilst the shales are apparently unfractured. Open fractures provide pathways for fluid movement both into and through reservoirs, whereas closed fractures inhibit movement of fluids, often forming an impermeable barrier within the reservoir. Fractures that start life as open sets may become closed by precipitation of minerals from solutions passing through them, or by a change of stress regime (e.g. changing them from extension-related to compression-related). Closed fracture sets may also open at a later stage in their geological history.

Faults
Faults are failure planes within a rock sequence along which some movement has occurred. Faults are one of the most important rock structures in petroleum geology as they are large fractures which control the distribution of rock units within fields, and provide both migration pathways and seals for fluids.

Stress and the Formation of Faults


A force acting on a surface can be resolved into two components (Figure 1.16) - one acting perpendicular to the surface (the normal force) and the other parallel to it (shear force - this may be further resolved into two orthogonal directions along the surface). Dividing the component forces by the area of the surface leads to the concept of normal stress, , and shear stress, . If the force is applied perpendicular to the surface, the normal stress will be at a maximum and the shear stress zero.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

33

EP00 - Introducing the E&P Business

Shell Learning Centre

If we now consider not a surface but a body of rock in the Earth's crust, it is subject to confining stress in every direction. Nevertheless, it is possible to resolve these stresses into three orthogonal directions such that the normal stresses acting on the rock are maximised and the shear stresses are zero. These directions are known as the principal axes of stress and are denoted by the labels 1, 2 and (3 for the greatest, intermediate and least normal stress respectively. Tensile stresses are negative. The Earth's surface is a plane of zero shear stress and therefore it follows that the principal axes of stress are orientated perpendicular and parallel to the Earth's surface. Precisely which principal axis is aligned with which orthogonal direction will be dictated by the relationship of the area to regional plate motions. The pattern of rock failure in experiments shows that rock fails along planes which lie at 45 or less to the axis of maximum stress ()). There will be two planes of failure, intersecting along the 2 axis and lying at equal angles from 3. This forms the basis for classifying the different types of faults which are observed in rocks according to the orientation of the principal axes of stress at the time of failure.

Description of Faults
Displacements vary from almost nothing up to many kilometres in major fault zones. Faults with a small displacement are usually of small areal extent while those with large displacements are large fault planes, often forming the boundary to several oil fields. The displacement of any fault can be determined by matching points on either side of the fault plane that were once in contact with each other. The slip is measured as the movement of the fault within the plane. The displacements seen in sections across the fault are the heave and the throw of the fault for horizontal and vertical movement respectively (Figure 1.17).

34

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The difference between the throw and the vertical separation is shown in Figure 1.18. Dip = tan [throw/heave]

When we describe the movement direction of faults it is almost always the relative displacement of the blocks that we are describing. The block which appears to have moved downwards is called the downthrown block. The block which appears to have moved upwards is called the upthrown block.

PLASTIC DEFORMATION
When a rock permanently deforms under stress without actually breaking this is known as ductile or plastic deformation (curve C-D in Figure 1.15). Folds in layers of rock are the most obvious evidence of plastic deformation. After plastic deformation the rock may also suffer brittle failure, but the shape of the fold is usually preserved in this case.

Description of Folds
Under natural conditions rocks fold into many different shapes, or fold styles, and so there are numerous terms used to describe folds and folding. Figure 1.19 shows a fold pair composed of an anticline or upfold of the rocks, together with its companion syncline or downfold. The curved beds in the crest and trough of the fold pair are joined by relatively planar dipping beds forming the fold limbs.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

35

EP00 - Introducing the E&P Business

Shell Learning Centre

If several layers of rock are folded together, as is usually the case, then lines of greatest curvature of beds within the folds, or hinge lines, can be defined which stack on top of each other. The stack of hinge lines defines a surface through the rock which is known as the axial plane or axial surface. Folds are commonly described with reference to the trend and dip of the axial surface, and the spacing between adjacent axial surfaces. In naturally deformed rocks, the hinges are not straight lines but are themselves curved. The fold style will often look rather like that shown in Figure 1.20, where the crests and troughs rise up and down. Upfolds in the axis of an anticline generate domes, downfolds in a synclinal axis form small basins. Domes are of fundamental importance in the petroleum industry as traps. For example, the major oilfields in the eastern Persian Gulf are large, open folds of this style, folding a fractured limestone reservoir with an evaporite seal.

36

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

CLASSIFICATION OF STRUCTURES
Similar rock structures can be seen at many different scales. Explorationists are largely concerned with the large structures which define fields. For the production of a field, small-scale or minor structures are of great significance because they may displace the reservoir sufficiently to form a fluid barrier or produce separate pools of hydrocarbons which will require a carefully designed drilling programme to exploit fully. Since it is difficult to classify structures according to their size, we will look at different structures with respect to the type of deformation (brittle or plastic) and their relationship to plate tectonics. This gives 3 main groups: Extensional structures - related to divergent plate margins Compressional structures - related to convergent plate margins Wrench or strike-slip structures - related to conservative plate margins

EXTENSIONAL STRUCTURES
Extension is the process by which the crust is stretched when plates move apart from each other. The process is happening at the present-day in the North Atlantic, where America is moving away from Europe at about 3 cm per year (about as fast as your fingernails are growing).

Faults in an Extensional Setting


When the surface of the Earth stretches (3 horizontal), extensional faults will form at a high angle to the ground surface. Faults that extend the crust are called normal faults.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

37

EP00 - Introducing the E&P Business

Shell Learning Centre

Some tectonically active areas of the present day crust, like the East African Rift represent an early stage in the process of continental division prior to separation. The East African Rift is a failed attempt to split the Horn of Africa from Central Africa and form an ocean between them. It is a relatively young structure and is characterised by normal faulting on both sides of a prominent downthrown rift valley or graben. The Viking Graben in the Northern North Sea shows many of the typical features of a hydrocarbon bearing basin formed by extension (Figure 1.22).

In this example, the lower crust and mantle deform by ductile flow. The upper crust, however, deforms by brittle failure to form large, planar faults which separate tilled fault blocks. Many of the biggest oil and gas fields in the North Sea are trapped in such tilted fault blocks. The displacement on faults has an important effect in petroleum geology. A borehole that penetrates a normal fault will find some beds missing or cut-out near the fault plane compared to the full sequence recorded elsewhere. If unit 2 in Figure 1.23 is the reservoir sandstone, clearly it is missing in the borehole.

38

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

In soft, unconsolidated sediments, such as those seen in the Niger Delta, extensional faults form that are not planar, but curved both in plan and in section. These faults are called listric faults. Listric faults not only change their dip amount down the surface of the fault, but also their strike direction along the fault (Figure 1.24).

Sealing of Normal Faults


If a normal fault passes through a sedimentary sequence containing clays and shales, these ductile lithologies may be smeared along the fault plane due to the movement along the fault. This process especially takes place during movement of faults through unconsolidated sand/clay sequences. Clay smearing can lead to a dramatic reduction of permeability across the fault. The length over which clay is smeared is dependant on the amount of fault displacement, the nature and history of fault motion, the thickness of individual clay layers and the material properties of the clay. Figure 1.25 illustrates diagramatically the way in which a clay horizon may be smeared out along a normal fault. The clay smear may completely seal the permeable sandy units. Alternatively, the seal may be partial if the throw of the fault is large or if the clay units are thin.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

39

EP00 - Introducing the E&P Business

Shell Learning Centre

Computer models can be effective in correlating the smear potential of the various clay horizons with the fault throw to predict the sealing potential of the clay smear. However, seal quality is strongly dependant on differential pressures across the fault, which is independent of the clay smear potential of the sequence.

Folds in an Extensional Setting


When a body of rock breaks, gaps will not develop, because gravity would cause the overhanging portions of rock to collapse into the newly-formed hole. Therefore, the layers in the downthrown block above a fault must fold in order to keep the two blocks in continuous contact (Figure 1.26). These folds are distinctive since they are formed in response to the shape of the fault, and are known as rollovers. Rollovers are most common in rock sequences that were faulted before being completely consolidated.

40

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

ROLLOVER ANTICLINE

Anticlines formed in this way are obvious hydrocarbon traps and are important in both exploration and production. Figure 1.27 shows a seismic profile through a major fault of this type. The fault plane can be seen clearly and a broad, open anticline can be seen above the curved fault.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

41

EP00 - Introducing the E&P Business

Shell Learning Centre

COMPRESSIONAL STRUCTURES
In some parts of the world hydrocarbons are found in areas of compressional rather than extensional tectonics (e.g. Bolivia, Venezuela). Most of these areas were originally sedimentary basins in which the stress field changed and the basin started to shorten rather than to extend.

Faults in a Compressional Setting


When the surface of the Earth compresses (1 horizontal), faults that allow beds to overlap will form at a low angle to the ground surface. Faults that shorten the crust are called thrust or reverse faults.

These are generally low angle structures which dip at thirty degrees or less, and which stack parts of the rock sequence on top of each other: in Figure 1.29 the borehole penetrates the reservoir sandstone twice (along with some other beds).

Thrust zones or belts are therefore often areas of the earth's crust dominated by low angle faults which thicken the crust on a regional scale, resulting in mountainous areas of high relief and active erosion.

Folds in a Compressional Setting


The fold or anticlinal trap is the commonest form of petroleum accumulation, accounting for the largest volume of the world's resources. Anticlinal traps (Figure 1.30) are common in regions where the crust has been subjected to horizontal compression.

42

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The factors that determine the volume of petroleum on an anticlinal trap include : The size of the structure: anticlines can be hundreds of kilometres long and thousands of metres high. The volume of the source rocks that have matured since the structure was formed. The thickness and porosity of the reservoir rocks. The amount of closure on the structure is also important (Figure 1.31). Petroleum migrates from the source rocks into the reservoir until a stage is reached where it can escape from the trap at the spill-point. Such a trap is said to have only a limited closure.

WRENCH STRUCTURES
The lateral movement associated with conservative plate margins and zones of wrench faulting can produce extremely complex structures, with folds, normal faults and reverse faults. Figure 1.32 shows the direction of maximum compression and extension and the likely trajectories of different types of minor faults and folds associated with a zone of wrench faulting.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

43

EP00 - Introducing the E&P Business

Shell Learning Centre

Faults in a Wrench Setting


Many of the Earth's largest and most obvious faults, such as the San Andreas Fault of California, involve crustal blocks moving laterally past one another (2 vertical). This results in steep faults in diagonal arrays which are called wrench or strike-slip faults.

Wrench faults show no component of slip up or down the dip of the fault. The relative movement of the blocks is lateral, along the strike of the fault plane. The two sides of the fault are assigned a name by the relative direction of displacement of the two blocks. You imagine yourself standing on one block with the other moving past you. If the block you are looking at is moving to the left the fault is termed sinistral or left-lateral. Movement to the right is called dextral or right-lateral. Left-lateral displacement is shown in Figure 1.33 and you can see that this sense of displacement is true whichever block you are standing on. The deformation on large strike-slip faults, such as those in California (Figure 1.34), propagates down to the asthenosphere, as they involve the jostling of entire lithospheric plates. The blocks between these vast faults are very similar to the situation shown in Figure 1.32, with thrusts trending roughly ESE-WNW.

44

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Cross-sections through wrench zones often show a system of normal faults curving into the main wrench fault to produce what is known as a flower structure (Figure 1.35).

Folds in a Wrench Setting


Anticlines in wrench zones form normal to the direction of maximum compression and will therefore be aligned at angles of up to 45 to the shear direction. Figure 1.36 is a depth contour map in feet of the Newport-Inglewood dextral wrench zone in California. This fold zone is associated with the fault structures
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

45

EP00 - Introducing the E&P Business

Shell Learning Centre

shown in Figure 1.34. The en-echelon (= ladderlike) folds and the associated thrust and normal faults give rise to a complex structural geometry, but nevertheless a highly prospective oilfield trend. In general, the large number of trapping structures in strike-slip zones makes them exciting prospects, but they are difficult to interpret, and relatively poorly understood at present.

DIAPIRISM
There is one further geological process that creates both folds and faults, but which does not strictly fall into the category of tectonics. It is a process that happens both in areas that are tectonically active and inactive, and is almost exclusively associated either with evaporite salts or with overpressured clays. Salt beds have some unusual properties. They are both weak and have a low density, and despite being highly soluble in groundwaters, they have a very low natural porosity. Whilst other sediments compact with increased overburden by expelling water from their pore spaces, salt does not. With time, the contrast in density between the light salt horizons and the more dense sediments above and below becomes accentuated. Undercompacted clays will have a large amount of water trapped between their grains and are similarly very light and mobile. Both evaporites and overpressured clays are highly ductile even at the low temperatures that exist high in the crust. If a substantial thickness of evaporite or clay occurs within a sedimentary sequence it becomes unstable with increased depth of burial, being overlain by heavier sediment of "normal" density. The salt therefore tends to flow upward into any irregularities in the rocks above. Concentrations of the light lithology will then balloon upwards through the rock cover, piercing through each bed and eventually rising towards the surface. A body of rock that pierces through shallower rocks in the sequence like this is called a diapir, and the process is known as diapirism.
46
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Figure 1.37 shows some of the shapes of salt structures and the names given to them. Whilst these structures are collectively known as diapirs, strictly not all of them are, since not all of them pierce the overlying rock sequence. Salt domes, which can be up to 1 km in diameter, are impermeable to petroleum. In suitable areas, the upward moving plug will pierce the overlying strata and form traps in the upturned beds (Figure 1.38).

A characteristic radial pattern of faults often develops in the beds overlying salt domes, associated with the upward push of the salt (Figure 1.39). Such "accommodation structures" increase the permeability in the beds immediately overlying the dome, and in certain cases (e.g. North Sea Ekofisk) turn an otherwise impermeable rock into a reservoir of good permeability.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

47

EP00 - Introducing the E&P Business

Shell Learning Centre

48

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

1.3 RESERVOIR ARCHITECTURE


Both the geometry and internal characteristics of reservoir units are important in terms of estimating the reserves that are in place and determining how to extract them. Knowledge of the nature of the depositional environment of a reservoir unit helps to predict not only what its overall geometry might be, but also the extent to which its porosity and permeability might vary horizontally and vertically. In this section we consider the formation of sedimentary rocks and the typical geometries of sediment bodies formed in different depositional environments. We will then look at the internal features of sediment bodies, and at some of the changes that occur to sediments after they have been deposited.

GEOLOGICAL CYCLES
Geological cycles involve the manufacture and re-processing of the Earth's materials. The rock-cycle was first recognized by the pioneering Scottish geologist, James Hutton. In a book published in 1785, he showed how: existing rocks may be eroded to form sediments by weathering and erosion; the sediments may become compacted into rocks; and a later mountain-building event may expose these sedimentary rocks at the Earth's surface, here they may be eroded away to form a fresh generation of sediments. The parts of the cycle are illustrated in Figure 1.40. The burial and subsequent uplift of rocks is labelled "tectonic cycle" on this figure.

Physical and Chemical Weathering of Rocks


There are two types of weathering: physical weathering (mechanical disaggregation) and chemical weathering (or chemical decomposition). Both processes, either separately or together, lead to the formation of a wide range of sedimentary rocks. One of the most important forms of physical weathering, which causes the break up of large masses of rock into fragments, results from the repeated alternation of freezing and thawing of rainwater in cracks and fissures. Water expands on freezing, with an increase in volume of about 9%, and the resulting pressure increase on the sides of the fissures causes them to split further apart. Large blocks of rock fall under gravity from rock faces, where they are shattered into smaller angular fragments which accumulate as scree slopes. This process is called frost shattering and exposes a much larger surface area of rock to rainwater run-off. It is the dissolution of rock material by water that leads to chemical weathering. Three products result from chemical weathering: Water-soluble substances; these are carried away in solution by rainwater and rivers, eventually entering the sea. New minerals; these are formed when soluble substances are leached out of existing minerals and their atomic structures partially collapse. An example of this process is the formation of clay minerals through the weathering of feldspar and micas in granite. Preservation of resistant minerals, such as quartz, which do not undergo chemical attack.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

49

EP00 - Introducing the E&P Business

Shell Learning Centre

The new minerals and the resistant minerals are known as residual minerals. The most common resistant mineral is quartz but, if chemical weathering is not too prolonged, feldspar and micas will also resist chemical attack. Quartz is found in huge volumes as a residual mineral in the form of sand grains on beaches all over the world. Silica, in the form of SiO44-, from the silicate minerals, is also very slightly soluble. When the soluble cations are leached out of the more complex silicates (such as feldspar where the silicate framework forms large molecules), the collapsed atomic structure usually incorporates water to form a flaky, hydrated clay mineral. Climate plays a very important part in weathering. For chemical weathering to be most effective, annual rainfall should exceed 1 m and temperatures should be high to speed the chemical reaction. Frost shattering can still be effective at quite low annual rainfalls, but obviously it requires the temperature to
50
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

fluctuate above and below freezing. For these reasons chemical weathering predominates in the hot, humid low latitudes and physical weathering predominates in the colder high latitudes. Not all chemical weathering takes place at the site of a rock outcrop being weathered. As soon as rocks have become fragmented the fragments are moved, under gravity, by water, wind or glaciers. So both physical and chemical breakdown continue during transportation of the fragments.

Transport of Weathered Material


The soluble products of chemical weathering are removed in solution by rainwater and rivers. However, the residual products of chemical weathering, and the larger fragments produced by physical weathering, are also removed by various transporting media, such as water, wind and ice.

Water:
Rock fragments and mineral grains may be transported by water either by bouncing or rolling of the material along the bed of a river or the sea as bedload, or by transport of the material within the water itself as suspended load. Coarser material tends to move as bedload, and finer sediment is carried along within the water itself as suspended load. The speed (or energy) of the transporting medium and changes in speed are largely responsible for the selective uptake and deposition of sediment grains. High-energy environments result in the deposition of coarse-grained sediments, and low-energy environments lead to the deposition of fine-grained sediments. There can be a good deal of variation in the type of sedimentary rock within a small area. For example, coarse sand may be found on the outside of a river bend, where the flow is fastest, whereas finer sand may be deposited on the inside of the bend, where the flow is slowest. The sorting of sedimentary material into different sized grains is a progressive process. The further a mass of sedimentary material is transported, the greater is the chance that it will become separated out into fractions of different sizes. The mineral type is also important as it determines the shape and density of the sediment grains. Particles that have been picked up and transported are subjected to mechanical abrasion, and undergo changes in shape. The longer particles are in motion, the more eroded they will become as edges and comers are worn away. Large particles are more rapidly rounded than small ones. Therefore, rounded particles are much more common in coarse-grained sediments deposited in a high-energy environment than in fine-grained sediments deposited in a low-energy environment. For example, the pebbles on a beach are almost always smoothly rounded, but the sand grains are much more angular. Fine grains, fall out of suspension like confetti and come to rest on the bed beneath the water as very fine flat and even layers, or laminae, usually only a millimetre or so thick. Hence, fine-grained laminated sediments are characteristic of suspension deposits. Coarser grained sediments do not form such neat layers. Often they are built up by the motion of the water near the bed into regularly spaced mounds and hollows, such as ripple marks on beaches (Figure 1.41).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

51

EP00 - Introducing the E&P Business

Shell Learning Centre

Wind:
Air is a fluid, so that wind behaves in exactly the same way as water when transporting sediment. The only difference is that the density of air and its viscosity are very much lower than that of water, which means that air can move a much more restricted range of grain sizes than water. Most transport by air is as a kind of bedload sweeping across the ground. Only the very finest dusts are lifted up into suspension. This means that after re-deposition of the particles as wind speeds gradually fall, the resulting wind-blown sediments are usually very well sorted. Individual quartz grains are usually more rounded than water-borne sediments as the grains collide without the cushioning effect of water and so have their corners and edges worn down more rapidly. Wind-blown sediments can also accumulate to form the structures given in Figure 1.41. However, the low viscosity of the medium prevents upper stage flat beds and anti-dunes from forming.

Ice:
In contrast, moving ice in the form of glaciers or ice sheets is totally indiscriminate about what it transports. Anything from the fine clay-sized particles to vast boulders several metres in diameter can be embedded in the ice. In this state the fragments are not free to collide with each other, and most will not be deposited until the ice melts, often many miles away from where the fragments were picked up. When the ice melts, all the remaining fragments, from clay sized particles to boulders, are deposited at the same time to form a long ridge of sediment, known as a moraine. Consequently, true glacial sediments are very poorly sorted and the fragments are characteristically angular.

Transport of the Soluble Products of Chemical Weathering


The most abundant soluble cations released from silicate minerals by chemical weathering are Ca2+, Na+, K+ and Mg2+. These, along with some silica, in the form of Si044-, are carried away in rivers, eventually reaching the sea. Therefore, you might expect that as weathered material in solution has been added to the oceans over many millions of years, the sea should be becoming progressively more saline. This is not the case because the dissolved salts are removed at the same rate as they are added, which means that the composition of seawater has remained reasonably constant over at least the past 1000 million years. Two major ways in which dissolved salts are removed are by the action of marine organisms (fixing CaCO3 in their shells) and by direct chemical precipitation (to form evaporite minerals).

52

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

SEDIMENTARY ENVIRONMENTS
Variations of reservoir properties such as continuity, porosity and permeability exert a major control on the distribution and producibility of hydrocarbons. The geological factors which affect these properties are largely the result of depositional and/or diagenetic processes. We will first look at environments of deposition. The types of sediments that we see at the surface today fall into two broad groups: Sediments derived from weathering of existing rocks, which are subsequently transported and deposited in a new environment. This group of rocks are called the siliciclastics or clastics and comprise sands, silts and clays. Sediments resulting from chemical precipitation or the action of marine organisms. These usually occur in marine environments, where there is little input of clastic sediment. Two distinct groups of rock types are formed by these process; the carbonates, such as limestone, and the evaporites, such as halite or salt. Note that the detail and reliability of depositional interpretations varies widely from area to area. Reservoirs often contain composite bodies deposited in several environments.

SILICICLASTIC ENVIRONMENTS OF DEPOSITION


Siliciclastic sediments are deposited in many different environments, from rivers high in the mountains to the deep waters of the oceans. We will only discuss four clastic environments of deposition in this manual; alluvial fans, braided rivers, meandering rivers, and deltas. The first three of these are related to rivers and are called fluvial environments. For additional information, or for information on other clastic environments of deposition, we recommend that you read Sedimentary Environments and Facies, Edited by H.G. Reading (Blackwell Scientific Publications) or Sandstone Depositional Environments (AAPG Memoir 31).

Alluvial Fan Deposits


These are localized deposits in the form of cone-shaped aprons of sediment laid down where a river loaded with debris flows from the steep slope of a confined mountain valley onto a flat lowland plain. As the water flows out from the steep valley and onto the flat plain the sudden decrease in gradient results in a loss of velocity of the water current, and consequently extensive deposition of sediment. Fans form under a wide variety of climatic regimes from arid and semi-arid to humid and even glacial. A feature common to all alluvial fans is that the sediments are generally poorly sorted and become finer grained with increasing distance from the apex of the fan. The sediments of humid climate fans tend to pass downslope into normal flood-plain type sediments (Figure 1.42), whereas those of dry climate fans may terminate in lobe-shaped masses or grade into the sediments of temporary lakes.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

53

EP00 - Introducing the E&P Business

Shell Learning Centre

Braided River Deposits


Braided rivers develop where the slope over which the river flows is quite steep and when the discharge fluctuates widely. A high volume of coarse-grained, poorly sorted sediments are transported. A braided river is able to disperse water through a number of smaller channels and so the discharge is maintained by a network of small channels with relatively steep gradient. At periods of high discharge, new channels are cut rapidly through the sediments.

54

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Conditions that favour braiding are often found where heavily debris-laden rivers flow from a steeply dissected mountain range onto a lowland plain and where there are seasonally-high discharge periods. Braided river deposits will therefore consist mainly of coarse bed-load gravels and sands. Channel lag gravels are deposited when the discharge is highest, and the subsequent sandy deposits represent abating flow. As channel switching is frequent, a single sequence of deposits representative of one channel is thin (of the order of a couple of metres) and rather discontinuous laterally. Braiding occurs around bars (Figure 1.43) in a main channel. Longitudinal bars are formed as the river braids which prograde parallel to the channel. The vertical sequence of sediments from a braided river channel changes from gravel to sand, moving upwards through the sequence (Figure 1.43). This is known as a fining upwards sequence of sediments. Since the proportion of fine sediment preserved is usually very low, the channel sands are stacked on top of one another and so in reservoir terms are well connected. The absence of fine sediments also means that no reservoir seal units can be expected within braided river deposits.

Meandering River Deposits


Rivers flowing in broad, deep channels need only a shallow gradient to maintain their discharge. A river that meanders down a slope acquires effectively a more shallow gradient to its bed than one that flows straight down the slope. This is analogous to roads built in a zig-zag fashion on very steep hills for precisely the same reason. On the outside of a meander bend, water flows at a faster speed than on the inside. So while the outside bank of a meander is actively eroded, material is deposited on the inside bend of the next meander downstream. The sediment deposits are known as a point bar, which grows slowly sideways at the same time as the outside bend is eroded. The overall result is that lateral erosion is compensated by lateral deposition and the amplitude of the meander increases (Figure 1.44).

The main current of water does not flow straight down the river channel, but a helical flow is superimposed on the overall downstream movement. The point of maximum erosion therefore occurs slightly downstream of the mid-point thus extending the development of the meander further downstream (Figure 1.44).
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

55

EP00 - Introducing the E&P Business

Shell Learning Centre

The narrow neck of land between the meanders may be breached during periods of high discharge so that the river cuts a new channel (Figure 1.45). The entrances to old meanders soon become plugged with deposited sediment so that it is abandoned as an ox-bow lake, which gradually fills with flood sediments and decayed vegetation. The lateral growth and downstream migration of point bars associated with meanders produces a characteristic vertical sequence of fluvial sediments. At any one moment in time, there is a gradation of sediment grain sizes across a meander, which reflects the change in current velocity across the meander.

Only the coarsest gravels are deposited on the fast flowing outside of the bend, forming channel lag deposits. Finer sediment is deposited on the inside bend, where the velocity is lower, to form point bar deposits (Figure 1.46). The point bar deposition migrates downstream with the meander and covers the coarse lag gravels that were once part of the channel. Therefore, a typical sequence of meandering-stream deposits become finer grained moving up the sequence. This is known as a fining upwards sequence. During flooding water is not confined to the river channel and sediment-laden water spreads out across the flood plain (the area of flat land through which a meander migrates, Figure 1.46). The suspended load is deposited as the flow rate drops so that a thin layer of laminated silts and clays is spread across the top of the point bar sediments as overbank deposits. In humid climates the flood plain is extensively vegetated and so buried overbank deposits often contain layers rich in organic debris. Occasionally overbank deposits form levees (ridges) on either side of the river channel. They may rise up to 5m above the general level of the flood plain and consist of the coarsest fraction of the suspended load which is deposited as soon as the river floods. The migration of the meander belt and switching of a river channel by the breaching of a meander neck leads to a very complex and varied vertical sequence of sediments, as shown in Figure 1.46. These fining upwards sequences of sediments may persist laterally over several tens, or even hundreds, of metres reflecting the slow migration of the point bar. This produces long and relatively narrow sand bodies that are internally inhomogeneous. The degree to which individual sand bodies are connected with one another depends on the ratio of overbank to channel sediments.

56

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Deltas
The term delta, used to describe the mouths of large rivers such as the Nile, the Mississippi or the Ganges, comes from the similarity in shape between these sedimentary environments and the Greek capital letter, . Many modern deltas are growing out into the sea, or prograding. This happens because the continental margins are subsiding, and sediment is continuing to build up and outwards. By continued deposition and subsidence, the Mississippi delta has built a ten kilometre thickness of sediments that stretches several tens of kilometres into the North Atlantic basin. Deltaic settings contain a wide variety of sand bodies, all of which have different geometries and permeability distributions. As sediment-laden, fresh river-water enters the sea, there is an interaction between the two water bodies. The river is slowed, and the largest grains are deposited rapidly. The sediment forms a sandy, silty bar, which may project above sea-level near the channel mouth. The finer, silt and clay particles are carried further seawards and are deposited in the deeper water on the seawards dipping profile of the delta, known as the delta front and the prodelta (Figure 1.47).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

57

EP00 - Introducing the E&P Business

Shell Learning Centre

The uppermost delta bar sands are 'clean', pure sands rather than silty ones because the action of waves winnows out finer-grained sediments. Well-sorted, porous, delta-bar sandstones form ideal reservoir rocks. Continued seawards growth of the delta leads to bar type sediments being deposited over the top of the former prodelta sediments. The sediments therefore become coarser grained moving up through the sequence. This is known as a coarsening upwards sequence, which is a marked contrast to normal river sediments which fine upwards. Landwards of the delta front is the delta plain. The river that feeds the delta splits into a number of smaller distributary channels (Figure 1.48), so sediment is supplied to the delta front from several channels, rather than just one. In between distributaries the delta plain is covered by overbank muds. In humid climates, the interdistributary areas are richly vegetated by salt-marsh or mangrove swamps.

The sequences are further complicated because the distributary channels may switch direction, and previously abandoned channels may be reactivated. Over the past 6000 years, the active part of the Mississippi delta has changed its position fifteen times as the main river channel supplying the delta front has changed its position. Channel switching is initiated when a leve is breached by sediment-laden water. If the new channel forms a more satisfactory route, then a fresh pattern of distributaries is developed and the former delta is gradually abandoned. This process is called avulsion.

58

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Not all deltas show such prolific phases of abandonment as the Mississippi, where the rate of sediment supply is high and reworking by waves and tidal currents is low, so that progradation is rapid. This type of delta is dominated by the sediment supply from the river, and is known as a river-dominated delta, producing an elongate, "bird's-foot" shape (Figure 1.48). If the tidal range is high (2 to 4 m), tidal currents become important, and the pattern of sedimentation is disrupted. The "bird's foot" shape is lost and distributaries develop a braided pattern. Tidal sand bars form rather than mouth bars (Figure 1.49).

In many places around the North Sea, and in other parts of the world, tidal ranges are in excess of four metres (macrotidal coastlines). Macrotidal coastlines are subjected to intense battering by storm waves. Not only are these waves extremely erosive, generating new sediment, but they are also capable of transporting much coarser material than tidal currents. Inevitably, this process affects the pattern of sediment deposition. Generally, waves do not strike the coastline at right angles but approach it obliquely, causing longshore drift (Figure 1.50). This plays an important role in the transport, deposition and erosion of sediment along shorelines, and also in the formation, migration and destruction of such features as deltas and barrier beaches. When sediment is discharged into shallow water which is devoid of tidal currents but subject to moderate wave action, the sediment is reworked to form a virtually continuous sand-bar round the delta front (Figure 1.50). This is known as a wave dominated delta.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

59

EP00 - Introducing the E&P Business

Shell Learning Centre

Syn-sedimentary Faulting in Deltas


The pattern of sedimentation in deltas is often further complicated by faulting which occurs simultaneously with deposition of sediments. These are known as syn-sedimentary faults (Figure 1.51). Thick deltaic sequences, such as those of the Mississippi and the Niger, become unstable and tend to slip seaward along large, curved listric faults. Movement along the fault plane continues as sediments are accumulating, resulting in large differences in the thickness of sediments either side of the fault plane.

Such faults exert a major control on the formation of hydrocarbon traps within the Niger Delta, as shown in Figure 1.52.

60

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

CARBONATE ENVIRONMENTS OF DEPOSITION


The main difference between carbonate and clastic sediments is that the source for clastic sediments lies outside the area of deposition, whereas carbonates originate within the area of deposition. Carbonates are almost always made of fragments of calcareous skeletons built by organisms. This type of sediment is known as bioclastic. Most of the organisms which produce carbonate skeletons live in very shallow, clear sea water, where there is good penetration of light through the water (i.e. in the photic zone). The shells are built by extracting both Ca2+ and HCO-3 from seawater to form the mineral calcite (calcium carbonate, CaCO3). After the death of the organisms, their calcareous remains accumulate on the sea floor, where they become cemented by chemically precipitated calcite, in the form of a lime mud, to form limestone. Thus, most limestones are formed on the continental shelves in tropical and equatorial areas; where marine organisms thrive in the warm, sunlit waters and there is a low input of continental sediment. The best known example of carbonate rock formed in this way is coral reefs. Some limestones are precipitated directly from seawater. They require warm shallow waters (CaCO3 is less soluble in warm water than cold water) that are well agitated by waves or currents. Under these conditions, particles, such as sand grains or shell fragments, act as nuclei for the precipitated CaCO3 to grow around. The resulting calcareous grains are well-rounded and are termed ooids. Rocks formed from beds of ooids are called oolites and can form very good reservoirs, such as in the Gulf of Mexico and the Middle East.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

61

EP00 - Introducing the E&P Business

Shell Learning Centre

The Bahamas platform is a good example of a present-day carbonate environment of deposition (Figure 1.53). Many of the islands are rimmed by barrier reefs, which grow on the windward side of the islands on the edge of the platform, whilst muds are deposited on the sheltered leeward side. Banks of oolites are forming in the agitated waters around the Tongue of the Ocean and on the western edge of the platform. Some calcareous sediments form in the deeper ocean basins from the accumulation of calcareous microfossils. These unconsolidated sediments are called calcareous oozes. When compacted and cemented they form fine-grained limestones. Chalks are also fine-grained limestones. However, they form in warm, shallowish seas (a few hundreds of metres deep) from accumulations of small, planktonic organisms called coccolithophores. Chalks typically have high porosity (up to 40%) but very low permeability because of the small grain size, which can cause production problems. Some North Sea reservoirs, such as Ekofisk, Tor and Dan are formed of chalk. In the Dan Field the problem of low reservoir permeability is mitigated by vertical fractures in the rock. This field has been successfully developed by using horizontal wells which intersect the vertical fractures.

62

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Deposition of Evaporites
Sediments formed by direct precipitation are comparatively rare, but some are of great economic importance. The best known are salt deposits, or evaporites, which form in isolated inland seas or in narrow ocean basins where circulation with other sea-water does not take place. In these environments, water evaporation may exceed input from rivers, so that the water becomes super-saturated. Salts, such as halite (NaCI), gypsum (CaSO4.2H2O) and calcite (CaCO3), then crystallize out. Evaporites also form along arid shorelines like the present-day Persian Gulf. Here, they crystallize out within the sediments of the mud flats. High evaporative rates are common in hot, dry climates, so evaporites in rock sequences are a good indication of ancient arid conditions. Due to their extremely low porosity and permeability, evaporites can form excellent cap rock or seals. Furthermore, evaporite layers become mobile during burial and deform to form diapirs. These structures can form excellent traps for hydrocarbons. Evaporites are often found with carbonates as they are formed in similar environments. Many reservoirs in the North Sea and the north of Holland are found in the Zechstein Formation carbonate and evaporite play.

TEXTURES
The textures of sedimentary rocks (the shapes and range of sizes of the constituent grains) are very important, because they tell us about the depositional history of the sediment.

Grain size distributions


From the reservoir point of view, the most important parameter for grain size distribution is sorting, which describes the range of particle sizes present in a sedimentary rock. The terms used to describe sorting are illustrated in Figure 1.54. In well-sorted sediments most of the grains are close to the mean grain size, whereas in poorly sorted sediments there is a wide spread of grain sizes about the mean. Good sorting is achieved either when grains have undergone transport for long periods of time, or when the transporting medium is selective with respect to the grain size it can move, for example wind transport. Poorly sorted sediments are produced when a mass of material is transported and deposited fairly rapidly without being reworked, for example deposition by glaciers.

Sorting also depends on the source of the sediment. Erosion of a well-sorted sandstone cannot yield a poorly sorted sediment unless material is added from another source. Well sorted sediments have higher porosity and permeability than poorly sorted ones, so knowledge of the nature of the depositional environment helps predict the location of the best quality reservoir units.

Grain shape
The shape of the grains present in a sediment can also give an indication of the way the sediment was transported. The most useful feature is the roundness of the grains, which is a measure of how much the angular corners and edges have been worn away.
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

63

EP00 - Introducing the E&P Business

Shell Learning Centre

Grain roundness depends on the length of time sediment grains are transported. The longer the grains are kept moving, the more likely they are to collide with each other and with the bedrock. Wind transport is the most effective means of grain rounding. Sands that have been blown over long distances, as in deserts, are often very well rounded. The terms used to describe roundness (Figure 1.55) are purely descriptive, varying between very angular and well rounded.

Larger gravel fragments are rounded much more rapidly than finer particles. Thus pebble-sized fragments of rock may be quite well rounded after only a few kilometres of river transport, whereas fine quartz grains are still subangular after many kilometres. Sediments that have been transported far enough for the grains to become well rounded and efficiently separated so that they are well sorted are called texturaly mature. Conversely a sediment that is poorly sorted with angular fragments is said to be texturaly immature. Texturally mature deposits are the best reservoirs, providing their porosity is not altered during burial.

Rock fabric
One important feature of rock fabric is the relationship between the grains, which make up the coarser components of the sediment, and the finer grained matrix, which occupies the intergrain spaces. In the case of a set of marbles tipped into a box, the marbles are not only in contact with each other, but they support each other. In sediments, the grains form a self-supporting framework, and the fabric is described as grain supported. If the larger grains are separated from each other by the finer matrix with little or no intergrain contact, the rock fabric is said to show matrix support. Matrix supported rocks have lower porosity and permeability than grain supported ones.

Facies
The interpretation of ancient environments must be based on consideration of the features of an individual rock unit, its relationship with rock units above and below (where possible) and the way it changes laterally. In this context, the term facies is used widely by geologists, but often with rather different meanings. 'Facies' is used to: describe the features that can be observed in the field and laboratory. Lithofacies has this meaning, as does biofacies, the latter referring to the nature and distribution of the fossil flora and fauna present in a rock. imply the process that produced a rock body, such as deltaic facies. All studies, irrespective of their objectives or the time available for their completion must begin with a descriptive approach. However, it is best to avoid labelling facies according to their postulated origin, unless the evidence for doing so is convincing.

64

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

DIAGENESIS
Diagenesis encompasses the multitude of processes that may occur in sediments after they have been deposited. Many of these processes result in the reduction of the depositional (or primary) porosity and permeability of the sediment.

Compaction
The porosity of a rock is the fraction of the volume of the total sediment occupied by pores. In a freshly deposited sand, the porosity depends on a number of factors including how the grains are packed together (that is their fabric) and whether the sand contains fine clay and silt particles.

The sediment in Figure 1.56a has a higher porosity than that in Figure 1.56b, even though both are composed of grains of a uniform size. Those in (b) are packed more closely than those in (a). Compaction of a sand occurs in response to the pressure caused by deposition of further layers of sediment. During this process, the grains are packed more closely and there is a reduction in porosity. Porosity may be reduced from an initial value of 30-40% in a freshly deposited sand to 20-30% after compaction. The reduction is not very large, because sands are mostly composed of quartz grains which are hard and do not deform under pressure. Compaction can proceed still further if the grains begin to dissolve where they are in contact. This process of pressure dissolution can result in the elimination of porosity (Figure 1.57). Freshly deposited clay has a much higher initial porosity than sand and also contains a good deal of water (around 70-90% by volume of the sediment). This is rapidly expelled so that compaction occurs soon after burial. The water escapes up through the clay and any overlying sediment, where it may have an important role to play in the later cementation of rocks (see next section). The final compaction of a clay sediment to form a rock requires a high degree of burial and exposure to temperatures of 100C or more, during which chemical changes occur within the clay minerals themselves.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

65

EP00 - Introducing the E&P Business

Shell Learning Centre

Clay minerals are flaky, like micas, and so with increased compaction they will also become oriented parallel to each other. Eventually, the compacted rock may become fissile so that it splits easily along the planes of the flattened clay minerals. We call this rock a shale.

Cementation
Cementation involves the growth of mineral crystals, called cements, within the pore spaces of a sediment, and is a common process in sandstones and limestones. The cement is usually precipitated from the pore waters expelled from interbedded or underlying clay sediments during compaction. Some cementation may occur at a late stage in burial as a result of pressure solution between adjacent grains. This causes silica to be dissolved and reprecipitated elsewhere within the pore spaces, or in other sediments. Carbonates of calcium, magnesium and iron are common cements in sandstones. Calcareous cements may form in sediments even before burial, simply by the evaporation of groundwater or seawater rich in CaCO3. Cements may completely eliminate primary porosity.

Clay Minerals
Clay minerals are hydrous sheet silicates, similar to micas in structure. The fact that they contain water (i.e. they are hydrous) causes their presence in sand grade sediments to significantly modify electric and density log responses (see Chapter 2.3). Clay minerals are deposited from suspension, and so may filter down into the fabric of sands after the larger grains have been deposited. Clay minerals may also be precipitated in pore spaces during diagenesis, in which case they are known as authigenic clay minerals. The presence of clay minerals not only reduces the porosity of potential reservoir sands, but leads to the clogging of pore throats (see Figure 1.57). During production, the extraction of hydrocarbons and injection of acids or other fluids during secondary recovery operations can dislodge clay minerals within pores and clog pore throats still further.

66

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Dissolution
Once exposed at the Earth's surface, limestones are very susceptible to chemical weathering because they dissolve easily in rainwater that has been acidified by atmospheric carbon dioxide. The chemical reaction that takes place is: CaCO3 + Calcite in limestone H20 + rainwater C02 -> from the atmosphere 032 + 2HCO3 soluble ions in water

Dissolution of limestone can occur on all scales: from the dissolution of individual crystals in the rocks, to the dissolution of the whole rock forming cave systems. The weathering of limestones is a major source of both calcium and bicarbonate ions in river waters and groundwaters, which explains why water in limestone regions is "hard": calcium carbonate is re-deposited when the water is boiled.

Dolomitisation
Dolomitisation is an important diagenetic process which affects limestones. It involves replacement of calcite (CaCO3) crystals with dolomite (MgCa(CO3)2) crystals. The source of the magnesium is usually an evaporite layer. Dolomitisation is important as it can increase porosity. Dolomite crystals are smaller than those of calcite, therefore a crystal-by-crystal replacement results in an increase in porosity. If the process continues unabated however, the rock can become completely dolomitised, with even the original pore spaces filled with dolomite, resulting in a "tight" dolomite with little or no porosity. This process can best be illustrated by means of the graph below.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

67

EP00 - Introducing the E&P Business

Shell Learning Centre

1.4 RESERVOIR PORE SYSTEM


A physical description of the properties of the reservoir rock is required to quantify the amount of fluid present in a reservoir and to understand how fluids will flow through the reservoir. Three parameters describe the properties of the pore system: Porosity Permeability Net Reservoir fluid storage capacity of the rock flow of a single fluid thickness of reservoir rock

POROSITY
Rock consists of crystals or grains between which spaces, or pores, are left that are occupied by fluids. The pore space is of interest as it is potentially hydrocarbon-bearing. Porosity, is therefore the first parameter that needs to be quantified when calculating the volume of hydrocarbons in place. In this chapter a number of important aspects of porosity will be discussed.

Definition
Porosity is defined as the ratio of the amount of pore space present in a volume of rock to the total volume (pore space plus rock matrix) of that rock. Porosity is thus a fraction in the range 0 (no pore space present) to 1 (no rock matrix present). Porosity is often expressed as a percentage. There is essentially nothing wrong with this practice as long as it is remembered that in all calculations (e.g. of hydrocarbon in place) porosity values should be entered as fractions.

Factors Affecting Porosity


The actual value of the porosity in a particular sample of rock is determined by a number of factors. For the following calculations, it is assumed that the rock is sandstone, in which the matrix consists of spherical quartz grains, and that all spheres are identical and are cubically packed (Figure 1.59).

68

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The porosity of this system can be calculated as follows:

Note that the porosity is independent of grain size.

In a sandstone the grains will not be perfect spheres and they can be packed together in different ways (Figure 1.56). If not all grains are of the same size, the small grains will fit in between the large ones and cause a reduction of the porosity. The porosity can also be reduced by compaction or by the presence of some other material in the pore space, like calcite cement or clay minerals. Porosity is thus dependent on: grain shape grain packing grain sorting overburden & compaction cementation clay content

Porosity in Carbonates
In crystalline carbonates (limestone or dolomite) the rock structure also consists of a packing, in this case of crystals. The size of the crystals is usually less than that of sand grains (150 urn vs 500 urn). In carbonates dissolution can lead to porosity also being present inside crystals. This is known as intraparticle porosity. Large dissolution features may also be present (e.g. vugs) and extensive dissolution can even lead to the creation of caverns.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

69

EP00 - Introducing the E&P Business

Shell Learning Centre

Carbonates are more brittle than sandstones and will tend to give way and break under stress. This accounts for the common occurrence of fractures in carbonate reservoirs. Some of the porosity types that can be encountered in carbonates are illustrated in Figure 1.60.

Types of Porosity
The intergranular porosity of sandstones and the interparticle porosity in carbonates is often referred to as primary porosity. The macroscopic porosity features often encountered in carbonates are called secondary porosity. Together these two porosity types make up the total porosity of the rock. A porosity concept that is of relevance when reservoir fluids are being produced is the interconnected porosity. Obviously, pore space has to be interconnected to make the flow of fluid possible. If only part of the pore space is interconnected, the term effective porosity is used to describe the porosity which is interconnected.

Typical Values of Porosity


Actual porosity values can be as high as 45% in unconsolidated sands, and as low as 1% in still prolific fractured carbonates. Table 1.5 gives common ranges of porosity:

Some non-productive rocks also have high porosities. Examples are shales and clays; they can contain substantial amounts of water which is not movable, as they have a low permeability.

PERMEABILITY
The permeability is a measure of the ease with which a fluid can flow through a rock. In this section absolute permeability will be discussed, which is a property of the rock itself and which is independent of the fluids that flow. Later, the effects of the fluid properties and the relative fluid saturations on the flow behaviour will be discussed.

Definition
To make fluid flow through a porous rock sample, a pressure differential AP has to be applied across it (Figure 1.60). The resulting flow rate through the sample depends on the sample dimensions, the fluid viscosity and the internal pore geometry of the sample. The latter is described by the empirical rock parameter, k, the permeability.

70

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Permeability has the dimensions of area (m2), but is commonly expressed in Darcies (D) or millidarcies (mD). One Darcy is equivalent to 1.01325 x 10'^ m^.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

71

EP00 - Introducing the E&P Business

Shell Learning Centre

Factors Affecting Permeability


Experiments using capillary tube models of the pore space in rocks easily show that the flow rate is higher in wide capillaries than in several narrow ones having the same total flow area permeability is dependent on the radius of the capillaries or pore throats. An increase in the radius of the capillaries will also lead to an increase in the porosity, suggesting that permeability is dependent on the same factors as porosity (i.e. pore throat size, grain size variation), but to a differing degree.

Typical Values of Permeability


Although porosity and permeability are related, there are also some important differences between the two properties. Typical relations between porosity and permeability for a number of rock types are outlined in Figure 1.62. The logarithmic scale for the permeability axis indicates that permeability is much more strongly dependent on pore (throat) radius than is the porosity. Typical permeability values range from 0.1 mD to several Darcies. Figure 1.62 indicates that low porosity material containing few large pores (e.g. reefs) can have higher permeability than high porosity material containing many small pores (e.g. chalks). Sometimes rock possesses a dual porosity system, for example a low porosity and permeability carbonate that is fractured. The fractures themselves (secondary porosity) contribute only very little to the total porosity, but provide an excellent flow path (very high permeability) for reservoir fluids. In such a system the low permeability matrix can produce into the fracture system, which then allows rapid flow of the fluid to the wellbore, thus resulting in a prolific producer of hydrocarbons.

72

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Permeability Anisotropy
In most rocks the permeability depends to some extent on the direction of fluid flow. The permeability is usually lowest in the direction perpendicular to the bedding planes due to the following effects: Most rock grains are elongated along one axis, however slightly, and will tend to settle with their long axis approximately horizontal. The overburden pressure compacting a sediment acts vertically, and will also tend to make the grains lie with their longest axis horizontal. Layers of less permeable sediments, e.g. muds, will be deposited horizontally, and will tend to hinder flow perpendicular to their final orientation in the reservoir. On a reservoir scale, the average permeability observed in a particular direction depends less on the microscopic rock properties (as might be measured on core plugs) and more on sedimentary features, for example grading or cross bedding on a small scale, or palaeo-coastline cut-offs on a large scale. For this reason it is very difficult to directly measure permeabilities appropriate to the behaviour we want to describe. Average permeabilities must be inferred by observing the behaviour of flowing wells. It is usual to distinguish between the horizontal and vertical permeabilities in a reservoir, the difference being expressed as the ratio between the vertical and horizontal permeabilities (ky/k^). In a clean, well-developed sand the vertical permeability may be as high as 70-80% of the horizontal permeability. However, a figure of 10% or less is more typical in many reservoirs due to the presence of small amounts of widely distributed shale. Variations in permeability in different directions in the horizontal plane have also been observed, but are rarely accounted for in reservoir models.

NET RESERVOIR
"Net reservoir" thickness is an important descriptive parameter of the reservoir, as it quantifies the vertical dimension when calculating the volume of hydrocarbons in place.

Definition
Net reservoir is defined as the total thickness of rock in a reservoir that contains movable hydrocarbons. Shale intervals, with effectively zero porosity, would thus be excluded from 'net reservoir'. Net reservoir is also frequently defined using minimum, or 'cut-off', values for porosity, permeability, and hydrocarbon saturation. However, these should be used with care, as they may lead to under-valuation of a prospect by the exclusion of low quality reservoir units. These units do not contribute to reservoir behaviour in the short term, but may do significantly in the long term. For this reason, it is recommended that the volume of hydrocarbons initially in place should always be calculated with no cut-offs. Any contribution to reservoir performance from low quality intervals should be incorporated via a reduced recovery factor. Net reservoir may also be defined using a Net to Gross Ratio (N/G), which is the ratio of the net reservoir thickness to the gross formation or interval thickness.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

73

EP00 - Introducing the E&P Business

Shell Learning Centre

Practical Considerations
Information about net reservoir, like information about porosity and permeability, can only be obtained from wells. This means that only "spot readings" are available and the volume of data depends on the number of wells. Usually the physical characteristics of the reservoir are not constant over the field, requiring interpolation between wells and extrapolation outside areas of well control to be done. The geological model of the field plays an important role in this process - assumptions about the lateral continuity of sand bodies must be based on such a model (Figure 1.63).

74

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

1.5 RESERVOIR FLUIDS


Petroleum is a general term applied to naturally occurring mixtures of organic compounds, primarily hydrocabons, that are found in the earth. Beyond this shared definition, specimens of petroleum collected from various locations might appear to have very little in common. They can vary from clear liquids that look much like water, to thicker, greenish or reddish-brown mixtures, or even highly viscous, semi-solid black substances. Their odour can range from a sweet, aromatic bouquet to the distinctly disagreeable smell of rotten eggs. The bulk of the chemical compounds that comprise crude oil are hydrocarbons and the ability of carbon atoms to form long branching and cyclic chains allows an almost limitless diversity in the molecular composition of petroleum accumulations. Different specialists involved in petroleum exploration and production have different reasons for wishing to examine and characterize the hydrocarbon fluids and water found together in petroleum reservoirs. For example, a chemical engineer may be interested in a crude oil's composition only as it relates to the amount of commercial products that the oil will yield after refining. The petroleum engineer, on the other hand, is primarily concerned with the analysis of hydrocarbons in order to determine their behaviour under the varying conditions of pressure and temperature that occur in the reservoir and production systems. In this section, we shall concentrate on the sampling and analysis of reservoir fluids for the latter purpose.

CHEMISTRY OF HYDROCARBONS
Hydrocarbon Series
If different samples of petroleum acquired from a variety of locations are analysed for their elemental composition it is found that the proportion of the total mixture which is made up of carbon, hydrogen, sulphur, nitrogen, and oxygen atoms is remarkably uniform, despite the wide diversity in the physical appearance and properties of the samples (refer back to Table 1.2, Section 2.1 HYDROCARBON ACCUMULATIONS). For example, the proportion of carbon in crude oil is virtually always in the narrow range 82 - 87 % by weight. This consistency is brought about by the bonding characteristics of carbon atoms, resulting in the observation that most hydrocarbon molecules exhibit similar ratios of carbon to hydrogen atoms present. A number of hydrocarbon series have been identified. Each series may contain many thousands of different hydrocarbon molecule types, each of which exhibits the same C:H atomic ratio. Furthermore, the variation in the ratio of C:H atoms between different series is also small. The differences in the properties of crudes may be explained in part by the relative amounts of each series present. The classification of hydrocarbon molecules into series is based on the arrangement of the carbon atoms (straight chains, branched chains, or cyclical) and the number of bonds between the carbon atoms (single, double, or triple). Examples of the more common series found in petroleum are:Alkanes (also called paraffins). These exhibit the general formula CnH2n+2, the first member of the series being methane (CH4), the second ethane (C2H6) and so on. Alkanes contain only single carbon - carbon bonds and, as this frees up the maximum number of valence electrons for bonding with hydrogen, alkanes are said to be saturated (with hydrogen). The first four members of this series (methane, ethane, propane and butane) exist as gases under standard conditions of pressure and temperature. Those from C5H12 (pentane) to about C17H36 are liquids while C18H38 and higher are wax-like solids. All the straight-chain alkanes up to C40H82 (tetracontane) have been identified in crude oil, typically amounting to 15% to 20% of the oil. Alkanes are characterised by their chemical inertness, which probably accounts for their stability over geological time.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

75

EP00 - Introducing the E&P Business

Shell Learning Centre

Cycloalkanes (or "naphthenes"). These have the general formula CnH2n and are also saturated but, unlike the chain-like structure of alkanes, form closed rings of carbon atoms. They are relatively stable, with chemical properties similar to those of alkanes. Arenes (also called aromatics). These are derivatives of benzene (C6H6), some examples of which are illustrated in Figure 1.64. The structure of benzene makes it relatively stable and unreactive and arenes are either liquids or solids under standard conditions. Arenes are referred to as unsaturated (with hydrogen) as the molecules contain multiple carbon - carbon bonds which reduce the proportion of valence electrons which are available for bonding with hydrogen. Alkenes and cycloalkenes (also known as olefins), which contain double carbon carbon bonds, and acetylenes, which include compounds having triple carbon-carbon bonds. Olefins are very uncommon in crude oils and the acetylenes are virtually absent due to their high degree of reactivity. Of the eighteen different named hydrocarbon series, alkanes, cycloalkanes, and arenes are the most common hydrocarbon constituents of crude oils.

When more than three carbon atoms are present in a hydrocarbon molecule there will be more than one way of arranging the atoms within the molecule. Different structural arrangements of molecules with the same atomic composition are called isomers and possess different physical properties (Table 1.6). The isomer count for alkanes ranges from two for butane to 6.2x1013 for tetracontane.

76

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Classification of Crude Oils


Crudes are normally characterized according to their base (the predominant series type in the crude), which fundamentally dictates its nature. A three-category system is used - paraffinic, naphthenic and intermediate. The term 'naphthenic' is something of a misnomer, for it relates not to the hydrocarbon series but to the fact that on distillation the crude leaves a residue of asphalt rather than wax. The Universal Oil Products (UOP) Company introduced the "UOP factor", which was an attempt to establish a single index relating to the base of a given crude oil and giving an indication of the commercial value of refined products obtained from it. The factor is based on the average boiling point and specific gravity of a sample. Values between 12.5 and 13.0 are found for paraffinic hydrocarbons; between 11.0 and 12.0 for naphthenic hydrocarbons; and between 9.0 and 12.0 for aromatics. More modern crude oil classification is based on dividing the hydrocarbons into alkanes, naphthenes, and aromatics, in combination with nitrogen, sulphur, and oxygen compounds (called resins and asphaltenes). Few oils have more than 50% naphthenic content. Paraffinic-naphthenic, paraffinic, and aromaticintermediate oils are most common. Natural gas is composed primarily of the lighter (low carbon) members of the paraffin series. Methane and ethane frequently comprise 80% to 90% by volume of a natural gas and if significant amounts of higher carbon constituents are present the gas may also yield valuable liquid products at surface (e.g. natural gasoline).

Non-Hydrocarbon Components
All hydrocarbon mixtures contain at least a few non-hydrocarbon substances. Crude oils typically contain organic compounds which incorporate certain amounts of sulphur, nitrogen, and oxygen atoms. Some organic compounds also contain metallic elements, such as vanadium and nickel. While these components are distributed throughout the whole distillation range of a crude oil, they tend to be concentrated in the heavier components. The colour and odour of crude oil stems mainly from the nitrogen, sulphur, and oxygen compounds concentrated in the C26-C40+ fractions. Sulphur is the most common nonhydrocarbon constituent, typically comprising 0.65% of the crude by weight. Carbon dioxide, nitrogen, hydrogen sulphide, helium, and hydrogen are the more common impurities found in natural gas. These are important because they can cause deviations from normal hydrocarbon behaviour if present in sufficient quantities. In addition, impurities such as hydrogen sulphide and carbon dioxide can cause severe corrosion and handling problems for a production engineer. Sulphur contamination can also lead to serious refining problems.

CLASSIFICATION OF RESERVOIR FLUIDS


The precise properties of reservoir fluids will usually be measured or predicted through sampling and laboratory analysis of the chemical composition. However, it is also convenient to describe them in terms of a loose category system. This is based on the volume ratio of the gaseous to liquid phases produced at surface (the gas:oil ratio, or GOR) and the relative density (also referred to as specific gravity or just 'gravity') of each produced phase. The volumes related by the GOR are always measured at standard conditions of temperature and pressure. Specific gravity is the ratio of the density of a substance to the density of some reference substance. For gases, the reference is dry air at the same temperature and pressure as the gas in question. It therefore follows that the specific gravity of a gas will be equal to the ratio of the molecular weights of the gas and air (which is 28.97). For liquids, the reference is pure water at 289 K and one atmosphere (60F and 101 kPa or 14.7 psia).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

77

EP00 - Introducing the E&P Business

Shell Learning Centre

For hydrocarbon liquids, the American Petroleum Institute (API) defined the API gravity scale, which was derived to allow hydrometers to be calibrated easily with a linear scale. The API gravity of a crude bears the following relationship to its specific gravity:-

Note that the scale is inverse with respect to specific gravity - dense crudes register low on the API scale (water has a value of 10). Table 1.7 shows the fluid classifications in use and typical compositions of light end paraffinic compounds for each. The boundaries between the categories are not rigourously defined. In general, the proportion of light paraffinic hydrocarbons present is almost 100% for a dry gas and decreases as you move down the table.

THE BEHAVIOUR AND PHYSICAL PROPERTIES OF RESERVOIR FLUIDS


The pressure and temperature conditions acting on a substance (and the specific volume that it occupies) determine whether it exists as a solid, liquid, gas, or a combination of these phases. Changes in pressure and/or temperature may alter the phase state of the substance. For example, methane, which is a gas at standard temperature and pressure, will liquefy if the pressure is increased or the temperature reduced sufficiently. Phase behaviour is the term applied to the description of the manner in which these phase changes occur. We can relate the phase behaviour of a hydrocarbon mixture in terms of the changes we observe in the phase state as the pressure, temperature and volume of the system are varied. However, it is the forces of
78
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

molecular attraction and repulsion which ultimately dictate the phase behaviour of a mixture and thus any attempt to describe the physics of the system must include a description of its molecular composition. Over the range of temperatures and pressures encountered in normal petroleum operations the main concern is with the gas and liquid phases of hydrocarbons, although solid hydrocarbons may occur and give rise to production problems (e.g. paraffin wax in production equipment, asphaltic deposits in the reservoir rock, hydrocarbon-water compounds, or hydrates, that cause problems in gas production).

Hydrocarbon Phase Behaviour


Let us start by examining the phase behaviour of a single pure component, then extend this to a mixture of only two pure components and finally discuss the more complicated phase behaviour of the reservoir hydrocarbon mixtures listed in Table 1.7. It is customary to represent the phase behaviour of hydrocarbon fluids on a plot of either pressure versus volume or pressure versus temperature. To consider the behaviour of a pure hydrocarbon substance as temperature and pressure are varied, imagine that a pressured container is filled with liquid ethane at 289 K (60F) and 6895 kPa (1000 psia). By injecting or withdrawing mercury from the container (or cell), the volume available for the ethane can be varied (Figure 1.65). Initial conditions in the cell correspond to point (1) on the p-V diagram of Figure 1.65. As mercury is withdrawn from the cell the pressure falls sharply, in response to the increase in the volume of the ethane, until the first bubble of gas appears (point (2)). This is called the bubble point and is followed by a period during which the volume of ethane can be increased without any change in pressure. At any point along this constant pressure line (points (2) - (3)) the amounts of ethane liquid and ethane gas are in equilibrium and the work done to expand the ethane in this period is expended in the vaporization of the liquid. The volume of gas in the cell thus increases at this constant pressure until the point is reached where all the liquid is vaporized - the so-called dew point (3). As the cell volume is increased further at constant temperature, the ethane gas expands and the pressure drops further.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

79

EP00 - Introducing the E&P Business

Shell Learning Centre

Figure 1.66 shows how the temperature axis can be added to the graph to describe the phase behaviour using all three variables. Liquids are much less compressible than gases and hence the marked difference in the slope of the lines in the P-V plane above the bubble point (steep) and below the dew point (gradual). The locus of bubble points obtained at various temperatures projected onto the pressure-temperature plane defines a line called the vapour pressure curve. At pressures and temperatures above this line ethane exists in the liquid phase - beneath the line it exists as a gas. The vapour pressure curve for a single-component system such as this one ends at the critical point. As the critical point is approached, the properties of the gas and liquid phases become similar, and they are identical at the critical point. Above and to the right of the critical point the fluid is neither gaseous nor liquid in terms of its physical appearance - it is a super-critical fluid, or plasma. For a single pure component, the bubble point and dew point at each particular temperature correspond to the same pressure. Between the bubble point and dew point, however, the relative amounts of liquid and gas coexisting in equilibrium will depend on the specific volume of the gas and liquid. The average density of the two-phase material changes according to the relative amounts of each phase present, but the density of the liquid and gas phases remain constant at any given pressure and temperature. The densities of the liquid and gas phases converge as conditions approach the critical point.

Reservoir fluids are not pure hydrocarbons. To understand the added complexity of a multicomponent system we will begin by observing a binary system - ethane and normal heptane (Figure 1.67). The bubble point line and dew point line of the pressure - volume plot no longer coincide to form the vapour pressure curve when plotted on the pressure -temperature plane. A two-phase region is now apparent on the pressure - temperature plot, as a two-phase envelope. Depending on the amount of ethane and heptane present, the two-phase envelope will adopt a different position on the plot, moving to the right as the proportion of heptane increases.

80

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

As the percentage of n-heptane in the mixture increases, the critical point of the system moves up to a maximum critical pressure and then down and to the right to a maximum critical temperature, that of n-heptane. When the composition is evenly distributed between components, the two-phase region increases in size, whereas when one component predominates, the two-phase envelope shrinks in size and approaches the vapour pressure curve of that component. Inside the two-phase region of any particular envelope, amounts of liquid and gas phases coexist in equilibrium. Just as in the case of a single-component system, the pressure and temperature conditions below the dewpoint curve dictate that the mixture exists as a gas, and the conditions above the bubblepoint dictate a liquid state. An important difference, however, is that the compositions of the liquid and gas phases within the two-phase envelope are not constant. The liquid and gas phases present in the twophase region will contain more or less of each of the two components, depending on the temperature, the pressure, and the overall composition of the system. As pressure is decreased at a constant temperature through the two-phase region, the composition of the gas phase and liquid phase changes. The initial gas phase is rich in the lighter component, but gradually acquires more and more of the heavier component as the pressure is dropped and more gas evolves. Concurrently, the liquid phase loses its portion of lighter component and the amount of liquid decreases.

This change in phase composition is, of course, quite different from the single-component case, where the liquid and gas phases always consist of 100% pure component. As the complexity of the hydrocarbon mixture increases, determining the composition of the liquid and gas phases becomes increasingly difficult. However, knowing the composition can be very important in describing the behaviour of the fluid in the reservoir and predicting the nature of the products at surface, particularly in the case of gas condensates and volatile oils. We will now go on to discuss the phase behaviour of each of the five main categories of petroleum fluid of interest in the oil industry: dry gas, wet gas, retrograde gas condensate, volatile oil and black oil.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

81

EP00 - Introducing the E&P Business

Shell Learning Centre

Dry Gas
The phase diagram for a dry gas will be similar to that shown in Figure 1.68. Point A represents the initial pressure and temperature conditions in the reservoir. It can be seen that this point lies to the right of the maximum temperature of the two-phase region (the so-called cricondentherni), which means that this mixture exists as a gas in the reservoir. Over the production life of a field the pressure in the reservoir will usually decline as material is extracted. However the temperature in the reservoir usually remains constant (any loss of heat energy being replaced from the surrounding rocks). Isothermal expansion therefore exists in the reservoir and reservoir conditions will gradually progress from point A to the abandonment point on Figure 1.68, a vertical straight line. From this we see that the reservoir fluid remains as a gas throughout the life of a field.

Separator conditions are also shown on the plot. This point is still to the right of the cricondentherm, which means that the hydrocarbons remain entirely in the vapour phase throughout their journey from reservoir to surface. It is this feature of dry gas fluids which gives them their name - the term 'dry' refers to the fact that no liquid hydrocarbon phase is produced at surface. It has nothing to do with the water content. Most dry gases will yield some water through condensation and it is possible for water to be produced from the underlying aquifer along with the gas. However, the reservoir fluid is still referred to as 'dry'.

Wet Gas
By increasing the percentage of heavier hydrocarbons in our mixture, the phase envelope shifts and the critical point moves upward and to the right (Figure 1.69). Wet gases still exist entirely in the vapour phase in the reservoir during depletion, but differ from dry gases in that separator conditions are within the two-phase region. There will therefore be both a gas and liquid product at surface, the liquid being referred to as condensate. The precise amount of liquid recovered with a given volume of gas production will depend on the separator
82
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

conditions and the position of the lines of constant liquid volume inside the two-phase region. Their position depends on the composition of the mixture.

Retrograde Gas Condensate


In the case of a dry gas or a wet gas, the reservoir temperature is higher than the cricondentherm. As the amount of heavy hydrocarbon components in the mixture is increased, the phase envelope is pushed further to the right and the system behaves as what is known as a retrograde gas condensate (Figure 1.70). The reservoir temperature is now situated between the critical temperature and the cricondentherm. Normally one would expect vaporization to occur with decreasing pressure, but in the case of a retrograde gas condensate the position of the phase envelope with respect to the reservoir temperature causes liquefaction to occur as the pressure drops below the dew point curve into the two-phase region. Hence the term retrograde, implying a process which is the reverse of the norm. In this manner a liquid phase is formed in the reservoir. In principle, retrograde condensation will continue until a certain maximum volume fraction of liquid phase is reached. Thereafter the volume fraction (or liquid quality) begins to reduce and 'normal' vaporization of the liquid in the two-phase region occurs. However, in practice this re-vaporization is inhibited because production of the gas phase continues while the retrograde condensation is going on. This causes the overall composition of the reservoir fluid to change. By selectively leaving heavy hydrocarbons in the reservoir, while simultaneously withdrawing the lighter hydrocarbons as gas, the composition of the remaining reservoir fluid is shifted toward a higher percentage of heavy hydrocarbons. This causes the overall phase envelope to shift further to the right and this promotes retention of hydrocarbons in the liquid phase as progressively lower pressures are required for vaporization to take place.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

83

EP00 - Introducing the E&P Business

Shell Learning Centre

Retention of liquid hydrocarbons in the reservoir is further promoted by their interaction with the reservoir rock itself. In most porous systems, a fluid needs to build up a certain 'critical' saturation before it can begin to flow. Such is the case with retrograde condensate, which may need to build up to a saturation of 10 - 20% before it becomes mobile. Therefore the condensate often becomes trapped as an immobile liquid phase within the reservoir pores. The development of retrograde gas condensate reservoirs is characterized by a number of problems that are unique to this type of fluid. To prevent the loss of high-value liquid hydrocarbons in the reservoir, attempts may be made to artificially maintain the reservoir pressure at a high level - above or close to the dew point. This may be achieved by reinjecting the produced gas after the more valuable heavy (liquid) hydrocarbons have been stripped out at surface - such a scheme is called gas cycling. The reservoir geology must be favourable for this to be effective - the presence of high permeability streaks may lead to the early breakthrough of the lean re-injected gas at the production sites, or alternatively if there are significant permeability barriers in the reservoir (such as sealing faults) this may prevent pressure communication between the injectors and producers. Even if the average reservoir pressure can be kept high, it is still possible that the pressure in the immediate vicinity of the producing wells will dip below the dew point (there must be a pressure gradient towards the wells for flow to occur). This causes the deposition of immobile condensate banks around the producers which can greatly reduce the productivity of the gas.

Volatile Oil
If the original hydrocarbon mixture has enough heavy components to move the critical point of the twophase envelope to the right of reservoir conditions, the mixture exists as liquid in the reservoir. Figure 1.71 shows an example of the two-phase envelope for a volatile or high-shrinkage oil. This terminology simply means that the oil contains relatively large amounts of lighter and intermediate hydrocarbons that vaporize easily. Such a composition is reflected in the spacing of the percent liquid lines within the two-phase region. With a fairly small drop in pressure, the relative amount of gas to liquid volume increases rapidly.
84
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Volatile oils are in general those for which initial reservoir temperature lies between the critical point and the point of maximum pressure in the two-phase region (the so-called cricondenbar).

If reservoir conditions are such that the mixture exists as a liquid (point A) and a significant drop in pressure is required to reach the bubble point, the oil is said to be undersaturated (with gas). This term implies that in principle it would be possible to add more gaseous light components to the hydrocarbon mixture, but the mixture would still exist as 100% liquid at initial reservoir conditions. When the bubble point is reached, or if initial conditions are coincident with a point on the bubble point curve, the oil is said to be saturated (with gas). The bubble point pressure is therefore also known as the saturation pressure for the crude oil system. Gas which evolves from the liquid phase, either within the reservoir or in the surface facilities, is referred to as solution gas. This name is something of a misnomer and is derived from the observation that the the gas appears to have been dissolved in the oil. In truth the gas is generated simply by a change in the pressure and temperature state of the fluid from a point outside the two-phase region to a point within it.

Black Oil
Figure 1.72 is a phase diagram for the final common hydrocarbon category, a low shrinkage, or "black" oil. Low-shrinkage oils require a much greater drop in pressure before a significant amount of gas phase is formed in the two-phase region. The percent liquid lines ("quality lines") are more closely grouped toward the dewpoint line of the phase diagram, which results in relatively small amounts of gas phase production at separator conditions compared with more volatile oils. Initial reservoir conditions for black oils typically lie to the left of the cricondenbar (the maximum pressure at which two phases can co-exist in equilibrium) on the p-T diagram. "Oil shrinkage" is used to quantify the volume reduction experienced by the reservoir liquid during the production process. It is defined as the ratio of the stock tank oil volume to that of the original reservoir liquid. This ratio is virtually always less than 1, for any expansion of the liquid during the production
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

85

EP00 - Introducing the E&P Business

Shell Learning Centre

process is more than offset by the volume reduction which occurs as solution gas is liberated. Oil shrinkage is the reciprocal of the "oil formation volume factor", to be discussed later.

If reservoir conditions are such that the initial pressure-temperature point is within the two-phase region, there must by definition be both a liquid and gas phase present in the reservoir, which will thus contain an oil accumulation with a gas cap. The gas in the gas-cap at initial conditions is termed "free gas" so as to maintain the distinction between it and "solution gas" (that gas which is dissolved in the liquid phase at initial conditions). The free gas phase is usually in thermodynamic equilibrium with the liquid oil phase, so that the gas is at its dew point and the oil at its bubble point. The phase diagrams for each individual phase, and for the total mixture, can therefore be represented as in Figure.1.73.

86

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

A reservoir hydrocarbon mixture may therefore exist as a liquid, a gas or an equilibrium mixture of both. Each of these has properties that vary with pressure, temperature, volume and composition.

PROPERTIES OF HYDROCARBON GASES


The need to know values for the physical properties of gases is evident in practically any calculation dealing with reservoir or production engineering. For example, we need to be able to describe the degree of volume change that the gas will undergo on its journey from the reservoir to standard conditions in order to define the total amount of gas that is available to be produced and to predict the volumes of gas that will be available for sales. Any attempt to describe the flow properties of the gas will require an understanding of fluid viscosity. The interpretation of the pressure data acquired in well production tests also requires knowledge of gas compressibility. Gas density and viscosity values are necessary for the design of surface separation and metering facilities. Before describing each of these key gas properties we will look at the real gas equation of state, which provides the basis for our ability to predict gas properties over a wide range of pressure and temperature conditions.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

87

EP00 - Introducing the E&P Business

Shell Learning Centre

The Real Gas Equation of State and the "z-Factor"


The physical properties of hydrocarbon gases can be described by the ideal gas equation of state. This equation is grounded in several basic laws: Boyle's law: states that at a constant temperature the pressure of an ideal gas is inversely proportional to its volume Charles's or Gay-Lussac's law: states that at a constant pressure the volume of an ideal gas varies directly with the temperature, and that at a constant volume the pressure varies directly with the temperature Avogadro's law: states that all ideal gases at a given pressure and temperature have the same number of molecules for a given volume The ideal gas equation of state that is derived from these laws is expressed as: pV = nRT where: p V n T absolute pressure volume number of moles of gas absolute temperature

The constant R takes different numerical values depending on the units system being used in the equation. However, it is independent of the type of gas and is therefore called the universal gas constant. The equation is only valid for ideal gases in which: The gas molecules themselves occupy a volume that is insignificant compared to the volume occupied by the gas. There are no molecular forces that cause the molecules to either attract or repel one another. The gas molecules do not lose any energy when they collide with one another. At very low pressures, many gases exhibit ideal behaviour and the equation is valid. However, at higher pressures, such as those found in oil and gas reservoirs, the equation does not hold. The petroleum industry has found that a simple correction factor for the ideal gas equation of state is useful for describing the behaviour of hydrocarbon gas mixtures under oilfield conditions. This real gas deviation factor is defined as the ratio of the actual volume of a gas to the volume that would be occupied if the gas behaved "ideally" under the same conditions. It is commonly referred to as the "z factor" and also sometimes as the "compressibility factor". The z factor is a function of pressure, temperature and composition and is usually determined experimentally for individual gas samples. The typical variation of the z factor with pressure and temperature for a given composition is shown in Figure 1.74. At very low pressures the z factor for all temperatures converges towards unity, implying a tendency towards ideal gas behaviour at low pressure.

88

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

89

EP00 - Introducing the E&P Business

Shell Learning Centre

By incorporating the z factor we can write down the ideal gas equation of state as follows: -pV = znRT where: z dimensionless compressibility factor or gas deviation factor

The simplicity of this expression disguises the complexity of actually determining the appropriate value of the z factor to use, particularly with gases for which experimental measurements have not been made. However, the observation that the z factor curves for different gas samples exhibit similar (though not identical) trends lead to the development of the law of corresponding states. This law expresses the fact that gases have the same z factor at the same conditions of reduced pressure and reduced temperature. Reduced pressure is simply the ratio of the pressure and the critical pressure of the gas, while reduced temperature is the ratio of the temperature and the critical temperature of the gas. A generalized chart of z factor as a function of reduced pressure and temperature can thus be generated and used for any gas. This avoids the need for a complex equation of state, but requires that the critical temperature and pressure be known. However, determining the critical pressure and temperature for a gas mixture is also rather difficult. To determine the reduced pressure and temperature of a gas mixture another basic law, Amagat's law, is observed.

Amagat's law
states that the total volume occupied by a gas mixture at certain conditions of temperature and pressure is equal to the sum of the volumes that the individual pure components would occupy at the same conditions. From this law, it can be shown that the volume fraction of a pure gas in a mixture is equal to the mole fraction of that gas in the mixture. Furthermore, we can calculate pseudo-critical properties for the gas by generating the mole fraction-weighted average critical pressure and temperature of each of the pure components present in the gas as determined by compositional analysis. The pseudocritical properties of gas mixtures are used in the same manner as actual critical properties of pure gases when determining z factors. However, pseudocritical constants are not the actual critical constants of the gas mixture. Figure 1.75 shows the relationship between z factor, pseudo-reduced pressure and pseudo-reduced temperature. The methodology for using this chart is as follows:-

When a compositional analysis giving the mole fraction of each component in a gas mixture is not available, use an empirical correlation between gas gravity and pseudo-critical properties. Figure 1.76 gives a commonly used correlation that requires only the gas gravity as measured in the field or laboratory. The gas gravity can also be calculated from a compositional analysis by using Amagat's law to determine the apparent molecular weight of the mixture - simply multiply the mole fraction of each component by 'its molecular weight and sum them. This value, divided by the molecular weight of air (28.9), is the gas
90
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

gravity. However, when a compositional analysis is available, it should be used to determine the pseudocritical constants directly. The correlations in Figure 1.76 are less accurate and should be used only when a simple gas gravity measurement is the only available data. The compressibility factor chart in Figure 1.75 is only accurate for mixtures of hydrocarbon gases. If a natural gas accumulation has significant amounts of non-hydrocarbon components, the compressibility factor calculated using reduced pressure and temperature will be in error.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

91

EP00 - Introducing the E&P Business

Shell Learning Centre

SUBSURFACE AND SURFACE VOLUME RELATIONSHIPS


We can now use what we know about the phase behaviour and physical properties of hydrocarbon mixtures to describe the way in which the volume of a liquid or gas in the reservoir will change on its journey to stock tank or standard conditions. Table 1.8 summarises the units which oil, water and gas are measured in at reservoir and standard conditions. Standard conditions are conventionally taken to be 15C (60F) and 1 atmosphere (14.7 psia), although you may encounter slightly different local definitions. At surface conditions liquids are measured with respect to stock tank volumes, whilst gases are measured with respect to standard conditions.

Figure 1.77 schematically shows the phase changes and volume changes that a reservoir liquid might undergo during production. These are the result of pressure and temperature changes. Figure 1.78 introduces some of the terminology used to describe these volume changes - in this diagram the system has been expanded to include the presence of free gas and water in the reservoir. The major volume changes are associated with the expansion of free gas from a primary gas-cap and the liberation of gas originally dissolved in oil. The PVT parameters which are used to relate the surface production to underground withdrawal are the oil and gas formation volume factors, Bo and Bg, respectively, and the dissolved gas-oil ratio, Rs. We will now discuss each of these parameters in more detail.
92
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

93

EP00 - Introducing the E&P Business

Shell Learning Centre

Oil Formation Volume Factor


The relationship between a volume of liquid in the reservoir (i.e. oil, which may contain a considerable amount of 'dissolved' gas) and the resulting volume of liquid that is recovered from the stock tank on surface is defined by BO, the oil formation volume factor (rv/sv):

Figure 1.77 shows a relatively large volume of oil at bubble point conditions in the reservoir. The volume of this oil we shall call 'B', as indicated on the diagram. If this oil is produced to surface, through the separator and eventually to the stock tank, we will end up with a certain volume of oil in the stock tank which we shall call 'F'. The formation volume factor of the oil at the bubble point is thus B / F. Volume 'B' is usually greater than volume 'F' because any expansion of the liquid under the reduction of pressure is more than offset by a volume reduction when the dissolved gas is liberated. The oil formation volume factor is therefore generally greater than one. After a certain amount of production the reservoir pressure may have declined to a value p1. This will have resulted in a volume reduction of the oil phase from B to C as some gas will have been released from the oil in the reservoir. This means that the formation volume factor of the oil at pressure 'p1' will be smaller than at the bubble point pressure. Figure 1.79 shows the overall relationship between Bo and pressure. As the pressure, P, decreases below the initial reservoir pressure (P1), but remains above the bubble point pressure (Pbp) for the oil, Bo increases as the oil expands and gas remains in solution. Below the bubble point, Bo is observed to decrease as gas begins to come out of solution, this loss of gas from the oil being greater than the expansion of the remaining oil. This oil shrinkage of the oil volume from bubble point to stock tank pressure and temperature can be significant for volatile oils, which may have initial formation volume factors (Boi) as high as 2 m3/stm3.

94

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Gas Formation Volume Factor


The factor used to relate reservoir gas volumes to standard conditions is termed the gas formation volume factor, Bg, and is defined as follows:-

This relationship can be calculated by determining the z factor at the desired conditions, using experimental laboratory data, a gas analysis or a gas gravity. This method for calculating the gas formation volume factor is only applicable when the reservoir and surface gas have the same composition. If a mixture's phase behaviour is such that significant volumes of liquid condensate are produced in going from reservoir to surface conditions, the composition of the gas phase in the reservoir may be significantly different from that at the surface. In this case another approach must be taken to mathematically combine the surface fluid volumes and relate them to a reservoir volume. From this expression for the gas formation volume factor it can be seen that Bg is approximately inversely proportional to the gas pressure, p, which confirms that gas has high compressibility. The number of standard surface volumes compressed into a reservoir volume is much greater at high pressures that at low pressures. However, Bg changes more rapidly with pressure at low pressures than at higher pressures (Figure 1.80). An alternative way of representing the free gas volume change is the gas expansion factor, E, (sv/rv), which is simply the reciprocal of Bg:

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

95

EP00 - Introducing the E&P Business

Shell Learning Centre

Solution Gas Oil Ratio


The volume of dissolved gas per stock tank volume of oil is referred to as the solution gas-oil ratio, RS (sv gas/sv oil) and its relationship with pressure is shown in Figure 1.81. Above the bubble point, Rs is a constant because the total amount of gas dissolved in the oil, by definition, is constant. Below the bubble point, gas starts to be liberated from the reservoir liquid. As there is less gas dissolved in the reservoir liquid, the solution gas oil ratio of the liquid will be lower than at the bubble point.

Producing Gas Oil Ratio


RS must not be confused with Rp, the produced-gas oil ratio. Rp is the total gas produced (solution gas + free gas) per stock tank volume of oil, and it is apparent that Rp - Rs represents the free gas produced at surface from the gas cap, per stock tank volume of oil. If the reservoir pressure is higher than the liquid bubble point, the reservoir is defined as undersaturated and only one phase (oil) exists. When a gas cap is initially present and two phases (oil and gas) exist in the reservoir, at the gas-oil interface the oil is described as saturated and must be at its bubble point pressure.

96

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

OTHER HYDROCARBON PROPERTIES


Gas Compressibility
Compressibility is an important property of reservoir fluids in terms of describing the performance of the production process. In this context it is the isothermal compressibility (i.e. at a constant temperature) that is important and this is defined as follows:-

This should not be confused with the z-factor, which is also sometimes referred to as the "compressibility factor" (see above).

Gas Density
The gas equation of state can be used to calculate the density of a gas under any condition.

At a constant temperature the density of a gas increases with pressure. However, the z factor influences the relationship, and this factor also changes with pressure. The density of a gas is the most commonly measured property. It is easily obtained experimentally by measuring the specific gravity of the gas, which is the density of the gas relative to that of air at the same conditions of pressure and temperature.

Gas Viscosity
Whenever we need to describe the flow properties of reservoir fluids we need some idea of the fluid viscosity, which is a measure of it's internal resistance to flow and is often thought of as a sort of internal "friction". The effects of temperature, pressure, and molecular size on gas viscosity can be explained in terms of molecular kinetic energy. As temperature is increased, the kinetic energy of the gas molecules increases, resulting in more frequent molecular collisions and a greater internal friction or viscosity. This increase in gas viscosity with temperature is the direct opposite of the decrease that is normally observed in liquids when the temperature is increased. If the temperature is held constant, an increase in gas pressure causes the gas molecules to move closer together, again increasing the number of collisions and thus the viscosity. The size of the molecules also affects the viscosity. At a given temperature, heavier molecules have lower velocities and, therefore, are responsible for fewer molecular collisions. However, as pressure is increased, the molecules are confined together and the heavier molecules have greater attractive forces.
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

97

EP00 - Introducing the E&P Business

Shell Learning Centre

This contributes to the number of collisions and thus to an increase in viscosity. As a result, gases with higher densities tend to exhibit higher viscosities at reservoir conditions. Viscosity is defined as the force per unit area required to maintain a unit velocity gradient in a fluid:

The S.I. unit of viscosity is pascal.second (Pa.s), corresponding to a force of one newton acting on one m2 which maintains a fluid velocity of 1 m/s over a distance of 1 m. Another commonly encountered unit of viscosity is the centipoise (cp), 1 cp being equivalent to 0.001 Pa.s. Natural gases exhibit viscosities of 1.0.10-5 to 5.0.10-5 Pa.s (0.01 - 0.05 cp) under typical oilfield conditions. Measuring gas viscosities at reservoir temperatures and pressures is difficult. Correlations based on gas gravity are commonly used in place of actual laboratory measurements.

Liquid Compressibility
Liquids are far less compressible than gases. This is evident from the slope of the isotherm on the pressurevolume phase diagram. It takes a much larger increase in pressure to reduce the volume of the liquid phase than to reduce the volume of the gas phase by a similar amount. The greater the pressure, the smaller the effect of changes in pressure on the compressibility of the liquid. As temperature increases, the effect of pressure variations increases. The coefficient of isothermal compressibility for liquid is defined in the same way as for a gas:-

When we produce from an undersaturated oil reservoir (pressure above the bubble point - no free gas in the reservoir), the ability of the liquid remaining in the reservoir to expand is important in determining the performance of the reservoir. Description of this process therefore requires a clear understanding of the liquid compressibility. There are correlations available to estimate values of compressibility for hydrocarbon liquids.

Liquid Density
The density, or mass per unit volume, of hydrocarbon mixtures in the liquid state is easily determined at the surface using hydrometers that measure the API gravity of an oil. Determining the density of an oil at reservoir conditions can be done in the laboratory or from liquid analyses at both reservoir and surface conditions. Knowing the density of the reservoir liquid can be useful, for example, in determining a pressure gradient within the oil column. Many liquid mixtures follow the additive rule: the mass of a component in one gramme-mole of a mixture is equal to the product of its molecular weight and the mole fraction of that component in the mixture. Also the
98
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

volume of the component is the product of the mass of that component in the mixture and its specific volume at the given conditions. The density of the mixture is then the ratio of the sums of the masses and volumes of each component. However, reservoir liquids often contain large quantities of lighter paraffins, the specific volumes of which are influenced by the attractive forces of the larger hydrocarbon molecules present in the mixture. As a result, the additive volume method of determining reservoir fluid density must be modified to account for the percentage of methane and ethane in the mixture and the molecular weight of the rest of the components. The thermal expansion of hydrocarbon liquids must also be considered when determining their density, particularly when correcting the volume and density measurements of stock tank oil to standard conditions. For example, we may sample a crude oil directly from the flowline and find its API gravity to be 52'API at 85C (185F), indicating a density of 769 kg/m3. The same oil sample taken from the stock tank at 15C would have a density of 817 kg/m3, or 42API. Heavier crudes with lower reservoir temperatures will be affected less by temperature.

Liquid Viscosity
Viscosity is another important property of hydrocarbon liquids and is defined in exactly the same way as for gas (see above). In contrast to gases, liquid viscosity decreases with increasing temperature. This is attributed to the thermal expansion of liquids, which moves the individual molecules away from each other and decreases internal "friction". Liquid viscosities are higher than gas viscosities in general, with most oils being in the range of 0.0002 - 0.03 Pa.s (0.2 - 30 cp) or higher at reservoir conditions. Liquid viscosity is dependent on the composition of the liquid, with the presence of lighter hydrocarbons in the mixture promoting lower overall viscosity. Liquid viscosity increases as pressure is increased, just as with a gas, although the magnitude of the change is not nearly as great. A typical oil viscosity may increase by about 15% per 100 bar, while a gas viscosity may increase by four times as much. The relationship between pressure and viscosity is reversed when the pressure falls below the bubble point and gas is released from the oil. Now the viscosity increases with decreasing pressure because of the loss from the liquid of the lighter hydrocarbons which contributed to the lower viscosities. Understanding the viscosity of hydrocarbon liquids is important for analysing the performance of reservoirs, for evaluating enhanced oil recovery projects (where the viscosity of the oil is often an important factor), and for the effective design of surface separation and treating facilities.

RESERVOIR FLUID CORRELATIONS


Fluid properties may be determined in several ways: Direct measurement on a sample of the fluid in the laboratory Calculation from a compositional analysis and knowledge of the properties of the components Empirical correlations with other physical parameters Correlations exist for many properties and new ones have been developed. Most correlations are available as graphs, nomographs, tables, or equations. They are useful for making estimates for designing the correct sampling procedure required for a given well and as a check against results obtained in the laboratory. If sampling is impossible or uneconomic, correlation data may have to suffice for engineering calculations. In most cases several correlations exist for any given parameter. The values calculated are usually not radically different from one another. Each correlation fits the data upon which it was based within a certain percentage error, generally less than 5%. If the oil or gas is representative of the fluids used in the correlation's development, a more accurate estimate of the correlated parameter can be expected. Additionally, some efforts have been made to determine which correlations are most representative of other producing areas. If one or several PVT analyses are available within a producing area, calculations using each correlation can be compared to the actual laboratory data, and the most representative correlation may be chosen.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

99

EP00 - Introducing the E&P Business

Shell Learning Centre

PROPERTIES OF FORMATION WATERS


Formation water properties are required to analyse and predict reservoir performance, particularly when large volumes of water are produced or when large volumes of water enter the reservoir from adjacent aquifers during the life of the field. The properties of interest to the engineer or geologist include compressibility, gas solubility, formation volume factor, density, viscosity, and salinity.

Water Compressibility
The compressibility of water is dependent upon the temperature, pressure, and amount of gas in solution. The coefficient of compressibility is defined as for other fluids:

Water compressibility is smaller than oil compressibility, and much smaller than gas compressibility (see Table.1.9). Compressibility decreases with increasing temperature at low temperatures, and increases with increasing temperature at higher temperatures. Compressibility decreases with increasing pressure. Compressibility increases with increasing amounts of gas in solution. The solubility of natural gas in water is itself a function of pressure, temperature, and salinity. Normally this value is about 1 - 5 m3/m3 (5 - 25 ft3/bbI ). Knowing the pressure, temperature, and water salinity allows the determination of gas solubility. Knowing gas solubility, and using the same temperature and pressure, we can determine water compressibility.

Water Formation Volume Factor


The definition of formation volume factor of formation water, Bw, is similar to that of oil in that it represents the change in volume undergone by an amount of water in moving from reservoir to surface conditions. As with oil, the release of gas from solution decreases the water volume, the expansion of the water with reduced pressure increases its volume, and the contraction of the water with reduced temperature decreases its volume. The expansion and contraction due to pressure and temperature reductions are relatively small and offset each other. The solubility of gas in water is much less than in oil, typically one hundred times less. The net result is that below the bubble point the volume reduction due to the liberation of gas is often not sufficient to offset the liquid expansion and so Bw increases with decreasing pressure (the opposite to normal oil behaviour). In practice, the value of Bw is almost always close to unity and values in the range 1.00-1.02 are typical. Correlations also exist which enable the formation volume factor to be estimated at any given pressure and temperature depending on the salinity of the water.

100

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Water Viscosity
Formation water viscosity decreases as temperature increases, at a constant pressure. Pressure and salinity changes do not affect water viscosity in a regular manner, nor to any great degree. Water viscosity can be important in evaluating improved oil recovery techniques such as waterfloods, steamfloods or any process that uses water as the drive fluid.

CHARACTERISTICS OF FORMATION WATER


In addition to gas and / or oil, petroleum reservoirs always contain a third fluid, water. Formation water is termed as: Connate water if it is believed to be a remnant of the original water in which the sediment was deposited. Meteoric water if it originated as rainfall and was carried into the ground via outcrops, fractures or permeable sediments. Interstitial water is the term used for the formation water that shares the pore space of the hydrocarbon reservoir with oil and gas, regardless of origin. Interstitial water saturations in petroleum reservoirs usually range from 10% to 50% of the pore space, and water saturations can vary throughout a reservoir, depending on the pore structure of the rock. Formation waters are most commonly distinguished by there varying degrees of salinity, that is the amount of dissolved ions present in the water, which is usually expressed in milligrams per litre (mg/1) or parts per million (ppm). Two methods are commonly used to represent diagrammatically the amounts of each dissolved ion found in formation water samples (Figure 1.82). In the Tickell method the reaction value of each ion is plotted, this being the mg/1 concentration times the valence of the ion, divided by its atomic weight. Thus 420 mg/1 calcium is equal to a reaction value of 420 x 2 / 40.08 = 20.09. Reaction values are often expressed as a percentage of the total reaction value of all the dissolved ions in a solution, or alternatively as equivalents per million, or milliequivalents per litre (meq/l).

The Stiff method also uses meq/l, but the scale is horizontal, resulting in a "butterfly" pattern. Usually the sodium and chloride scales are graduated in scale units of 100 meq/1 and the other ionic scales in units of 10, although the scales can be varied for any group of diagrams under comparison. Salinity is the most important characteristic that can be measured for a formation water sample as this will
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

101

EP00 - Introducing the E&P Business

Shell Learning Centre

dictate the ways in which formation water and hydrocarbon mixtures change with pressure and temperature. For example, the solubility of natural gas in a formation water with 150,000 ppm total dissolved solids is only about half the solubility in pure water. However, even in pure water the solubility of natural gas is not great; it is only about 2 to 5 m3/m3 (10 to 30 scf/bbI) under most reservoir conditions. Determination of the formation water salinity is more important for correct interpretation of electric well logs. The resistivity of a formation water is related to its salinity, and a value of formation water resistivity is necessary for the quantitative evaluation of resistivity logs. Although this value can be estimated from log data and correlations, the best results are obtained when a sample of formation water is retrieved and analyzed.

102

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

1.6 INITIAL RESERVOIR CONDITIONS


PRESSURE VERSUS DEPTH
The stratigraphic column in a sedimentary basin consists of a solid phase and a fluid phase. The solid phase, or matrix is composed of rock. The fluid phase is composed of the water, oil or gas which fill the pore spaces within the rock. At any depth, the total weight of the overlying porous rock and the fluid it contains causes an overburden stress on the formation. This is usually denoted by the symbol SV. The overburden stress is supported partly by grain-to-grain contacts, and partly by the pressure of the fluid in the pores. The loading carried by the grain-to-grain contacts is called the effective or matrix stress, and is denoted by the symbol SV. The fluid pressure is denoted by the

This relationship can be expressed in the form of a graph of the overburden stress and the fluid pressure versus depth (Figure 1.83). The overburden stress SV, and the effective stress V act vertically, whilst the fluid pressure P acts equally in all directions. Horizontal confining stresses Sh1 and Sh2, and horizontal effective stresses h1 and (h2, can be defined, and are similarly related to the fluid pressure. The overburden gradient increases with depth because compaction reduces the porosity and increases the average density of the formation. The gradient varies locally according to the lithology of the formation, because of the different densities of the various minerals forming the grains of rock. This variation is usually smoothed out by averaging the gradient over large depth intervals. Typical values range from 14 kPa/m to 23 kPa/m. In areas where active sedimentation takes place, such as the Gulf coast, NW Borneo and Nigeria, 21 kPa/m is an acceptable average.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

103

EP00 - Introducing the E&P Business

Shell Learning Centre

The fluid pressure results from the weight of the column of water extending from surface to the depth of interest. The variation of pressure with depth is described by the hydrostatic pressure equation:-

The fluid pressure gradient ranges from 10 kPa/m for fresh water up to about 11.5 kPa/m for saturated brine. The gradient will also change slightly with depth because water is slightly compressible, and expands with increasing temperature at greater depths. The effective stress is the difference between the overburden stress and the fluid pressure at any depth.

NORMAL AND ABNORMAL PRESSURES


A reservoir is termed normally, or hydrostatically pressured if there is a continuous column of water extending from surface down to the reservoir, through which the water pressure is being transmitted. One way of checking whether or not a reservoir is normally pressured is to extrapolate the water pressure gradient observed immediately below the oil water contact of a reservoir back up to surface. If there is a continuous column of water the extrapolated pressure will be the atmospheric pressure acting on the top of the water column.

In carrying out such an extrapolation, small allowances should be made for changes in water salinity, and for the small compressibility and the thermal expansion of the water. Together these effects should appear as a small change in the average water pressure gradient. The "surface" to which the extrapolation is carried should either be mean sea level when dealing with an offshore accumulation, or the ground water level onshore, which may lie several tens of metres below the surface of the ground in some areas.
104
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

If the column of water between the reservoir and the surface is broken by a sealing layer of some kind an abnormally pressured system may result (i.e. the system will exhibit either underpressures or overpressures). The extrapolated water pressure at surface would be either higher or lower than atmospheric pressure (Figure 1.84). The true pressure distribution might look similar to the situation shown in Figure 1.85. Above the sealing layer there is a column of normally pressured water extending down from the surface. A sharp transition occurs across the seal, with the pressure gradient continuing below the seal but at an offset from the normal pressure line.

Causes of Abnormal Pressures


Compaction, where permeable rock is isolated by a substantial thickness of clay or other impermeable sediments, and then either buried or uplifted so that the overburden stress is altered. Normally, in the case of an increased overburden stress due to burial, the rock would compact, its pore space would reduce in volume and water would be expelled from the pores into the surrounding rocks. However, if the surrounding sediments are too impermeable to allow the expelled water to dissipate the water will stay in the pore space and an overpressure situation will result. Underpressure situations can develop in an analogous way if the reservoir rock is uplifted rather than buried. Overpressures are created in some relatively impermeable rocks, because of the difficulty with which liquids can escape from the pores as the rock is compacted. These "depopressures" occur particularly in thick clays and shales which undergo rapid burial. Thermal effects where water expansion due to changes in temperature is prevented because water cannot escape through the surrounding impermeable sediments. Phase changes such as the dewatering of clay minerals and conversion of gypsum (CaSO4.2H2O) to anhydrite (CaSO4), which contribute additional free water to the pore space. Production from a well sealed reservoir may cause underpressures due to "depletion". Removing fluid from such a closed system causes the fluid pressure to be reduced, because no liquid can enter the
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

105

EP00 - Introducing the E&P Business

Shell Learning Centre

reservoir to replace the produced volume. The same may apply to a relatively impermeable reservoir, in which any influx of fluid to replace the produced volume takes place very slowly. The reservoir is underpressured during its producing life, but will repressure over a longer time period. "Inflation" overpressures occur when a normally pressured reservoir becomes connected to an already overpressured zone. This effect may occur over "geological" time, e.g. by leakage of fluid from a deeper overpressured interval into a reservoir, or rapidly during production activities, e.g. by rupturing a fault seal between a normally pressured and an overpressured fault block as the normally pressured block is depleted. Inflation overpressures may also occur as a result of an internal blowout in a well, where fluid flows from a high pressure zone, through the borehole into lower pressure zones. Changes in horizontal stress caused by tectonic activity have a similar effect to changes in the vertical overburden stress. Osmosis where clays act as semi-permeable membranes separating formation waters with different ionic concentrations, act to pump water molecules into or out of a sealed system.

The Impact of Abnormal Pressures on Drilling Operations


During normal drilling operations there is a column of drilling mud in the borehole. One of the functions of the mud is to generate a fluid pressure in the borehole which is greater than the pressure of the fluids in the formation. This overbalance prevents the formation fluids from flowing into the borehole (Figure 1.86a). The overbalance is kept small to minimize the loss of drilling mud from the borehole into the more permeable parts of the surrounding formation. If overpressures are encountered during drilling the mud overbalance may be lost, and fluids may flow from the overpressured formation, if it is permeable, into the borehole causing a "kick" (Figure 1.86b, mud gradient 1). The density of the fluid influx from the formation will be less than the density of the mud in the borehole, causing a reduction in the weight of the fluid column to the surface. The pressure in the borehole at the point of influx will be reduced, and the underbalance will increase. This condition is unstable and may develop into a blowout if it is not quickly controlled by re-establishing an overbalance at the point of influx. If the mud weight is increased to achieve an overbalance against the overpressured formation (Figure 1.86b, mud gradient 2), the large overbalance which results against the shallower, normally pressured formation may cause unacceptable losses of drilling fluid from the borehole. The shallower formations might therefore have to be sealed off behind casing before drilling into the overpressures. If underpressures are encountered during drilling, the sudden large overbalance may cause large losses of drilling fluid to the underpressured formation (Figure 1.86c, mud gradient 1). If these losses cannot be replaced by adding drilling fluid to the wellbore at the surface, the level of the mud in the borehole may drop (Figure 1.86c, mud gradient 2). The overbalance against the shallower, normally pressured formations may be lost, allowing an influx into the borehole - a kick. One way to guard against this, is to set casing before drilling into the underpressured formation, removing the need to maintain an overbalance against the shallower formation.

106

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

107

EP00 - Introducing the E&P Business

Shell Learning Centre

CAPILLARY PRESSURE
When two immiscible fluids of different density are poured into a tank, the lighter fluid will float on top of the heavier fluid and a sharp interface will exist between the two fluids. If a narrow "capillary" tube (such as a thin glass tube) is inserted through the interface, the fluid contact inside the tube will establish itself at a different level than that outside the tube (Figure 1.87).

This capillary effect is caused by interactions between the molecules of the two fluids and between the fluids and the solid (in this case glass). Two phenomena play a role here: surface tension and wettability. We need to understand the capillary effect because it plays an important role in the way fluids distribute themselves in the reservoir, where the glass tube in the above experiment is replaced by countless narrow pores and pore throats.

Surface Tension
Consider a drop of water surrounded by air. The water molecules at the surface of the drop experience a relatively strong inward attraction to the water molecules inside the drop and a relatively weak attraction to the air molecules outside it. This imbalance in forces acting on the molecules at the surface causes the drop to contract to a geometry in which the surface area of the drop is minimised - a sphere (Figure 1.88). Any attempt to change the shape of the drop will meet resistance, because this will increase the surface area of the drop and hence increase the imbalance of forces acting on the surface molecules. This resistance is described by the surface tension () of the two fluids. For any two fluids there will be a specific value of the surface tension, depending on the forces between the molecules of each of the fluids and their mutual interaction. To prevent the drop imploding, the surface tension will be counterbalanced by a pressure differential from inside the drop to the outside, the pressure in the water being higher than that in the air. The magnitude of the pressure difference depends on the magnitude of the net inward force at the surface of the drop and is therefore proportional to the surface tension of the two fluid system. The pressure difference also depends on the radius of the drop. The smaller the drop, the larger is the surface area in relation to the total volume and so the larger is the pressure differential. The net effect can be expressed as follows:

= 2

108

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

Wettability
Wettability describes the behaviour at the contact between two fluids and a solid, e.g. air, water and glass (Figure 1.89). A contact angle, , will form at any point where the two fluids and the solid meet. This contact angle is a property of the system and depends on the molecular interactions within and between the two fluids and the solid. We can select either of the two fluids and measure the contact angle through that fluid. If the contact angle is less than 90 degrees then the fluid through which the angle is measured is defined as the wetting phase. The other fluid is the non-wetting phase. A fluid that is non-wetting in one system, e.g. air in an air-water-solid system, may be wetting in another system, e.g. air in an air-mercury-solid system.

The Capillary Effect


The capillary effect is a combined result of the influence of surface tension and wettability. In a glass tube passing through the interface in an oil-water system, we can measure the contact angle between the inner wall of the glass tube and the oil-water interface (Figure 1.90).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

109

EP00 - Introducing the E&P Business

Shell Learning Centre

The oil-water interface in the glass tube will be curved, with the radius of curvature given by:

The narrower the tube, or the smaller the contact angle, the smaller the radius of curvature of the interface. Surface tension causes a pressure differential across the curved interface between the two fluids. This is called the capillary pressure, and denoted by the symbol Pc.

To balance the surface tension effect, the fluid pressure on the concave side of the interface will be higher than that on the convex side. Returning to the oil-water example (Figure 1.87), the pressure in the water at any depth is described by the hydrostatic pressure equation, as is the pressure in the oil at any depth. The pressure in the oil and the water at the flat interface outside the glass tube must be equal, but the pressure in the water at the curved interface inside the capillary tube will be lower than the pressure in the oil at the same point. Water is drawn up inside the capillary tube so that the increase in the pressure drop through the water in the capillary tube, compared with the pressure drop through the oil it replaced, balances the capillary pressure drop across the curved interface. We can therefore relate the hydrostatic pressure differential to the capillary pressure as follows:-

110

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

This expression indicates that for a given fluid-fluid-solid system, i.e. for a given and , the height of the capillary rise is inversely proportional to the radius of the capillary tube and inversely proportional to the density difference between the fluids.

PRESSURE PROFILES
We are now in a position to compare the pressure profiles that will exist in a two - fluid system with and without the influence of capillary effects. Figure 1.91 a shows a tank containing oil and water. Three capillary tubes of different sizes pass through the oil-water interface. Consider first the pressure distribution outside the three capillary tubes. Starting at the oil water interface, the pressure will increase as we go down into the water as described by the hydrostatic pressure equation:-

Going up from the oil-water interface, the same expression applies for the pressure in the oil at any given point, with the density being o instead of w. This pressure distribution (Figure 1.91b) prevails in the absence of any capillary forces. The interface of the oil and the water is called the Free Water Level.

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

111

EP00 - Introducing the E&P Business

Shell Learning Centre

Turning now to the Inside of the tubes, the pressure in the oil is the same as the oil pressure outside the capillaries at the same depth, because the two regions of the oil are connected via the top end of the capillary. Similarly, the pressure in the water inside the capillary tubes is the same as the pressure in the water outside the capillaries at the same depth below the free water level. In the previous section we saw that the surface tension at the oil water interface inside the tubes together with the wettability characteristics of the oil-water-glass system will generate a capillary pressure differential at the fluid contact inside the tubes. This can only be supported in an equilibrium state by a rise of the contacts in the tubes to a level which is above the free water level outside the tubes. The magnitude of the capillary pressure, and hence the height of the capillary rise, is inversely proportional to the radius of the tube. The greater the capillary pressure difference the greater the capillary rise. In general, if fluids are confined in narrow spaces where capillary effects will play a role, the oil water contact will be higher than the free water level. This is also true for a reservoir, where the same effects take place in a more complicated configuration of pores and narrow connections. The oil and water pressure gradients in a reservoir can be determined by measuring pressures at a series of depths. The intersection of these gradients will be at the free water level of the reservoir, exactly as in the case of the capillary tubes (Figure 1.91 c). The description of the pressure distribution in a gas-water system is the same as that discussed for an oilwater system. The pressure gradient in the gas will be much smaller than that in the oil due to the lower gas density 9. The surface tension between gas and water is slightly higher than between oil and water, resulting in higher capillary pressure differences for a given pore size.

SATURATION PROFILES AND CAPILLARY PRESSURE CURVES


Most sediments are saturated with water when they are deposited, or become water saturated shortly after. Traps are formed by impermeable cap rock and deformation of the sediments by movements in the Earth's crust (Figure 1.92a). Hydrocarbons are usually formed in source rocks somewhat remote from the eventual reservoir rock, and subsequently migrate into the traps to form accumulations. The migrating hydrocarbons will move through the upper part of the reservoir rock, just below the cap rock, and fill the trap from the top down (Figure 1.92b). The pressure distribution will be similar to that in the capillaries of the previous section (Figure 1.91 c). The oil water contact is at different elevations in capillaries of different sizes. Note that although the position of the free water level can be defined, it is not physically present in the reservoir as an interface between oil and water.

112

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

This process is called drainage - the non-wetting phase, oil, is pushing or draining the wetting phase, water, out of the system. The opposite process, which may occur when water from the aquifer re-enters the reservoir during oil production, is called imbibition. Drainage (and imbibition) can be simulated in the laboratory on a sample of the reservoir rock obtained from cores. In the experiment the sample is filled with water and surrounded by oil. This is illustrated in Figure 1.93a, where the rock sample is represented by a bundle of different sized glass capillary tubes. Pressure is applied to the oil, which will try to enter the rock pores. Due to capillary action, the initial increase in pressure is 'expended' purely in developing a curved contact at the oil-water interface - we find that a certain threshold pressure differential has to be built up before the capillary effect can be overcome and oil can enter the largest pore throat (or glass tube, Figure 1.93b). The capillary effect in the smaller pores (tubes) is stronger and therefore a progressively greater pressure differential must be applied in order to make oil enter the smaller pores (Figure l.93c).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

113

EP00 - Introducing the E&P Business

Shell Learning Centre

As the drainage experiment continues, we can easily measure the amount of water that is displaced from the core sample as a function of the differential pressure that is applied via the oil phase. If we also know the porosity of the sample we can then determine the capillary pressure profile as a function of the water saturation of the rock. This function is known as the capillary pressure curve for the rock, an example of which is shown in Figure 1.94c.

114

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

We are now in a position to examine the true fluid distribution in the reservoir due to the influence of capillary effects. The glass capillary tubes in the laboratory experiment are replaced by the porous reservoir rock (Figure 1.94a). An effectively infinite range of different 'radius' pores and pore throats will be present in the reservoir, the range being limited by a finite minimum and a finite maximum size. In principle, the largest pore throat will determine the minimum height of capillary rise of water above the free water level. It is therefore the largest pore throat that determines the height of the Oil-Water Contact (OWC, Figure 1.94). Everywhere in the reservoir below the OWC will contain 100% water saturation, regardless of whether it is above or below the free water level. The smallest pore throat will cause the greatest capillary effect and therefore result in the largest capillary rise. Everywhere in the reservoir above this point will contain water at its minimum, or connate, saturation. This water is immobile and is present in the form of a thin layer smeared over the surface of the rock within the pores (assuming the rock to be water wet). In between the OWC and the point at which connate water saturation is reached, the water saturation will reduce progressively from the OWC upwards as the height of the capillary rise in progressively smaller pores is reached and the transition from the water to oil phase occurs (Figure 1.94b). This interval of variable saturation is known as the capillary transition zone. The height of the transition zone and the shape of the saturation distribution versus height within it are functions of the range and distribution of pore sizes within the rock. We can convert the height of the capillary rise into a capillary pressure scale by applying the hydrostatic pressure equation:-

This equation also provides a link between the capillary pressure curve measured on a rock sample in the laboratory, and the variation of the water saturation (and hence oil saturation) with height above the free water level observed in the reservoir (Figure 1.94a and b). Therefore, measurements of the capillary pressure curve in the laboratory allows us to describe and predict the distribution of fluids within the reservoir. In reality reservoirs do not consist of a single type of rock. Each type of rock will have its own characteristic distribution of pore sizes, and its own characteristic capillary pressure curve. A full description of the fluid distribution in a reservoir requires capillary pressure curves determined for a series of samples of the various rock types in the reservoir (Figure 1.95). The oil saturation at any particular depth will depend not only on the pressure difference between the oil and water, but also on the rock type. The smaller the average pore size of a rock type, the lower the average oil saturation at a particular depth. The permeability of a rock is also related to the average pore size and, in general, high capillary pressures are associated with low permeabilities. The same types of models also apply to gas - water situations.
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

115

EP00 - Introducing the E&P Business

Shell Learning Centre

In the laboratory the capillary pressure curves are often measured with a mercury-air system rather than an oil-water or air-water system. The capillary pressure curves have exactly the same shape, but a different pressure scale due to differences in surface tension, contact angle and density between the different fluid systems. Mercury-air capillary pressures can be translated into oil-water or gas-water capillary pressures, or into a height above a free water level by a simple linear scaling. The different pressure scale results from a much greater surface tension between mercury and air, and a different contact angle compared with those of oil-water or gas-water systems.

DRAINAGE AND IMBIBITION


The process of oil displacing water in a water wet system, as in the case of a reservoir filling up with oil, is called drainage.The capillary pressure curves as discussed in the preceding section represent this drainage process. The reverse process, namely that of water displacing oil may also take place in a reservoir, e.g. when oil is being driven out of the reservoir by water from an expanding aquifer below the reservoir. This
116
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

process is called imbibition, and it can be studied in the laboratory by reversing a drainage experiment and lowering the pressure applied to the oil surrounding the rock sample. The sequence of events taking place in the pores during a complete experiment is shown schematically in Figure 1.96.

The experiment begins with a water saturated sample (point 1), which undergoes a drainage process as the surrounding oil pressure is increased. No oil enters the sample until a large enough pressure difference has been created between the oil and water (point 2), whereupon the oil enters the larger pores (point 3). Continuing to increase the oil pressure progressively pushes the oil into smaller and smaller pores (points 4 and 5). If at any time during drainage the process is reversed by lowering the oil pressure again, an imbibition process takes place. Due to capillary effects, not all of the oil that was pushed into the sample in the drainage process will be produced back again. Oil drops will be 'snapped off' at pore throats and remain behind in the sample (points 3A, 4A and 4B). When the pressure difference between the oil and water has been reduced to zero again (points 3A and 4A) the average oil saturation left behind in the sample is termed the residual oil saturation. From the diagram it is clear that the residual oil saturation depends on the point at which the imbibition process starts, i.e. on the initial oil saturation. High initial oil saturation (point 4) results in high residual oil saturation (point 4A). When the drainage process is carried on to high capillary pressures, corresponding to high elevations above the free water level in the reservoir, the capillary pressure curve becomes asymptotic to a minimum water saturation (point 5). This is water which is trapped in the very small pores and as a film of water coating the grains of rock. The irreducible water saturation is also referred to as connate water. Not all reservoir systems are water wet. There are examples of oil wet reservoirs in which the oil preferentially coats the grains of rock as a result of chemical interactions between some of the molecules in the oil and the rock grains. An oil wet system can be described by the same models used for a water wet system, but the pressure difference across any oil-water interface will be such that the water pressure is
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

117

EP00 - Introducing the E&P Business

Shell Learning Centre

greater than the oil pressure. Capillary effects will draw oil down below the free water level, and the oil water contact in a reservoir would be observed below the depth of the free water level determined from pressure measurements. Some reservoirs may have an intermediate wettability, and show no preference for oil or water wetting the rock. In these systems the oil-water interfaces need not be curved, i.e. the contact angle is close to 90C, and capillary effects will be negligible. Many reservoirs may have a mixed wettability, in which the smallest pores have remained full of water and so are water wet, whilst the larger pores have been completely swept of their original water and now contain only oil and so are oil wet. There are no simple models to describe such a system, which would be assumed to be either water or oil wet depending on the more dominant wetting phase.

RELATIVE PERMEABILITY
When there is more than one fluid present in a pore space, the fluids will interfere with each other's ability to flow. The flow rate of a single fluid through a rock depends on the viscosity of the fluid, the permeability of the rock, and the imposed pressure gradient. The permeability of the rock is independent of the fluid that is flowing through the rock. However, when more than one fluid is present, each fluid will occupy part of the pore space and reduce the pore space available for the other fluids to flow through. The permeability "seen" by a particular fluid depends on what fraction of the pore space it has available to flow in. The higher the saturation of a fluid in the pore space, the greater will be its effective permeability. The ratio between the effective permeability experienced by a fluid at any given saturation level to the absolute permeability of the rock is known as the relative permeability of the fluid. Under the initial conditions found in many reservoirs, the pore space contains a high oil saturation, typically 80% to 90%, with only a small irreducible or connate water saturation present (Figure 1.97a). The oil can flow freely through the rock and is hardly affected by the presence of the water, which in a water wet system is distributed around the edges of the pores, thus presenting only a minor impediment to the flow of the oil. The absolute permeability of the rock is defined as the permeability when only a single fluid is present in the pore space (i.e. at 100% fluid saturation) and is independent of the fluid type. If there would be no water at all present, the permeability of the rock to the oil would be therefore be equal to the absolute permeability and the relative permeability to oil would be one. With the small water saturation present, the relative permeability of the rock to oil is less than one, typically in the range 0.7 to 0.9. The relative permeability to the water is zero, because the water is trapped by capillary effects and cannot flow at this low saturation. After an oil zone has been swept by water, for example as a result of a water injection project, the pore space will contain a low, residual oil saturation and a high water saturation (Figure 1.97b). The oil is no longer able to flow at these low saturations (typically some 30%) because it has become trapped by capillary forces during the imbibition process. The relative permeability of the rock to oil is now zero. The water is able to flow, but is hindered by the presence of the residual oil. Note that the oil droplets occupy positions in the middle of the pore. They therefore present a much greater impediment to water flow than did the connate water to oil flow in Figure 1.97a. The relative permeability of the rock to water in this situation is very much less than one, typically 0.2 to 0.4.

118

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

These two extremes of saturations are termed the endpoint conditions. The oil endpoint occurs at high oil saturation and connate water saturation. The water endpoint occurs at residual oil saturation and high water saturations.

At connate water saturation, Swc, the oil is at its endpoint relative permeability, kro'. As the water saturation increases, the oil relative permeability decreases until residual oil saturation is reached, Sorw, at which point the oil has zero relative permeability because it is immobile. Similarly the water relative permeability increases from zero at connate water saturation to the water endpoint relative permeability, krw, at residual oil saturation (Figure 1.98).
2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

119

EP00 - Introducing the E&P Business

Shell Learning Centre

The full relative permeability curves can be measured on rock samples in the laboratory, but often only the endpoints are required as the conditions in most of a reservoir approximate to one or other of the endpoints throughout the production lifetime. The region where both oil and water are flowing at intermediate saturations is often negligibly small compared with the size of the reservoir (Figure 1.99).

Water displacing oil in a water wet a system is an imbibition process and modelling such a process requires the imbibition relative permeability curves. It is rare to encounter a drainage process in water-oil systems during the production life of a field. Drainage processes would be active if oil were displacing water, or if water were displacing oil in an oil wet system.

120

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

The behaviour of gas-oil systems, usually in the presence of connate water, can be modelled in a similar way to oil-water systems, but close attention should be paid to whether the process being modelled is a drainage or an imbibition process. Strictly speaking such a system is water wet, but from the point of view of the gas any liquid, be it oil or water, is the wetting phase. Gas-oil relative permeability curves are plotted as a function of liquid saturation. If oil or water is moving into a gas cap, the wetting phase is increasing and an imbibition process is taking place. If a gas cap is expanding and displacing oil downwards, the wetting phase is decreasing and a drainage process is taking place. The relative permeability curves for the two processes have different endpoints and follow completely different paths (Figure 1.100).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

121

EP00 - Introducing the E&P Business

Shell Learning Centre

122

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

EP00 - Introducing the E&P Business

Shell Learning Centre

RESERVOIR TEMPERATURE
When specifying the initial conditions of a reservoir, the formation temperature is important for the prediction of reservoir fluid properties. Usually, a simple model of a linear temperature variation with depth adequately matches measured formation temperatures. The slope of the line is termed the geothermal gradient, and typically lies in the range 0.02 - 0.06C/m. The precise geothermal gradient at a particular depth will depend on the thermal conductivity of the surrounding rock, and on the presence of any nearby sources of heat, such as concentrations of radioactive minerals or magma intrusions from below the crust. Temperatures are recorded by many logging tools to correct the tool response for thermal effects. In many cases this data will be available as a log. Such data should be checked by comparing all the temperature logs available for a well, and selecting those measurements which show the least influence of the cooling effect of the circulating drilling mud. The later a log was run after stopping circulation, the greater the opportunity for the formation to return to its normal temperature (Figure 1.101).

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

123

EP00 - Introducing the E&P Business

Shell Learning Centre

124

2 0 0 1 S h e l l I n t e r n a t i o n a l E x p l o r a t i o n a n d P r o d u c t i o n B . V.

Vous aimerez peut-être aussi