+peters ekwere j. - petrophysics

1049

Upload: alfonsopzsz

Post on 13-Apr-2015

268 views

Category:

Documents


8 download

TRANSCRIPT

Page 1: +Peters Ekwere j. - Petrophysics
Page 2: +Peters Ekwere j. - Petrophysics

TABLE OF CONTENTS

Page 1 PETROLEUM RESERVOIR ROCKS ......................................... 1-1

1.1 PETROPHYSICS ............................................................................... 1-1 1.2 PETROLEUM RESERVOIR ROCKS ................................................1-2 1.3 MINERAL CONSTITUENTS OF ROCKS—A REVIEW ...................1-4

1.4 ROCKS............................................................................................... 1-5 1.4.1 Igneous Rocks.................................................................... 1-5 1.4.2 Metamorphic Rocks...........................................................1-6 1.4.3 Sedimentary Rocks............................................................1-6

1.5 CLASSIFICATION OF SEDIMENTARY ROCKS ............................. 1-7 1.5.1 Clastic Sedimentary Rocks ................................................ 1-7 1.5.2 Chemical Sedimentary Rocks............................................ 1-7 1.5.3 Organic Sedimentary Rocks ..............................................1-8

1.6 DISTRIBUTION OF SEDIMENTARY ROCK TYPES .................... 1-10

1.7 SANDSTONE RESERVOIRS (CLASTIC SEDIMENTARY ROCK) ............................................................................................. 1-10 1.7.1 Pore Space ....................................................................... 1-12 1.7.2 Compaction and Cementation......................................... 1-15 1.7.3 Classification ................................................................... 1-17

1.8 CARBONATE RESERVOIRS (LIMESTONES AND DOLOMITES) .................................................................................1-20 1.8.1 Classification ................................................................... 1-21 1.8.2 Pore Space .......................................................................1-22

1.9 FRACTURED RESERVOIRS ..........................................................1-28

1.10 RESEVOIR COLUMN.....................................................................1-29

REFRENCES...................................................................................1-32

2 POROSITY AND FLUID SATURATIONS..................................2-1

2.1 DEFINITION OF POROSITY ...........................................................2-1

2.2 FACTORS AFFECTING SANDSTONE POROSITY ........................ 2-2

2.3 FACTORS AFFECTING CARBONATE POROSITY ........................ 2-4

2.4 TYPICAL RESERVOIR POROSITY VALUES.................................. 2-5

2.5 LABORATORY MEASUREMENT OF POROSITY.......................... 2-6

Page 3: +Peters Ekwere j. - Petrophysics

2.5.1 Direct Porosity Measurement by Routine Core Analysis .................................................................... 2-6 2.5.2 Indirect Porosity Measurement by CT Imaging.............. 2-11

2.6 FLUID SATURATIONS ..................................................................2-16

2.7 INDIRECT POROSITY MEASUREMENTS FROM WELL LOGS................................................................................... 2-24 2.7.1 Introduction to Well Logging......................................... 2-24 2.7.2 Mud filtrate Invasion...................................................... 2-25 2.7.3 Porosity Logs .................................................................. 2-32 Density Log .................................................................. 2-32 Sonic Log (Acoustic Log) ............................................. 2-36 Neutron Log .................................................................2-41 Combination Porosity Logs ......................................... 2-45 2.7.4 Resistivity Log ................................................................ 2-46 Electric Log ................................................................. 2-54 Induction-Electric Log................................................ 2-56 Dual Induction Laterolog ........................................... 2-58 Focused Electric Log (Guard and Laterolog) ............. 2-62 Microresistivity Logs................................................... 2-65 2.7.5 Lithology Logs ................................................................ 2-68 Spontaneous Potential Log (SP)................................. 2-68 The Gamma Ray Log (GR)...........................................2-73 2.7.6 Nuclear Magnetic Resonance (NMR) Logs.................... 2-76 Nuclear Spins in a Magnetic Field.............................. 2-76 The Effect of Radiofrequency Pulses - Resonance Absorption ............................................... 2-79 Relaxation Processes...................................................2-80 Molecular Diffusion Effect.......................................... 2-84 NMR Signal and Corresponding T2 Spectrum ........... 2-84 Pore Size Distribution................................................. 2-89 Estimation of Permeability from NMR Relaxation Times............................................... 2-95 2.7.7 NMR Imaging of Laboratory Cores................................ 2-97 The Effect of Magnetic Field Gradients...................... 2-98 Slice-Selective Excitation............................................ 2-99 Frequency Encoding ................................................. 2-100 Phase Encoding..........................................................2-101 Image Reconstruction............................................... 2-102 Three-Dimensional NMR Imaging............................2-103 Signal-to-Noise Ratio and Image Contrast............... 2-104 Example NMR Images of Laboratory Cores..............2-105 2.7.8 A Comparison of Various Porosity Measurements for Shaly Sand....................................... 2-112 2.8 RESERVE ESTIMATION PROJECT ........................................... 2-113 2.8.1 Reserve Estimation........................................................ 2-114

Page 4: +Peters Ekwere j. - Petrophysics

2.8.2 Economic Evaluation..................................................... 2-115 2.8.3 Simulation Procedure.................................................... 2-116 2.8.4 Sampling Procedure ...................................................... 2-116 2.8.5 Simulation Output.........................................................2-124

2.9 PORE VOLUME COMPRESIBILITY............................................2-126

NOMENCLATURE .......................................................................2-135

REFRENCES AND SUGGESTED READINGS.............................2-138

3 PERMEABILITY .....................................................................3-1

3.1 DEFINITION ....................................................................................3-1

3.2 DIMENSIONS AND UNIT OF PERMEABILITY ............................ 3-6

3.3 LABORATORY DETERMINATION OF PERMEABILITY...............3-7

3.4 FIELD DETERMINATION OF PERMEABILITY ..........................3-14 3.4.1 Diffusivity Equation for Slightly Compressible Liquid........................................................3-15 3.4.2 Pressure Drawdown Equation ........................................3-19 3.4.3 Pressure Buildup Equation ............................................ 3-22 3.4.4 Diagnostic Plots.............................................................. 3-24 3.4.5 Skin Factor...................................................................... 3-30 3.4.6 Homogenous Reservoir Model with Wellbore Storage and Skin............................................. 3-33 3.4.7 Type Curve Matching ......................................................3-37 3.4.8 Radius of Investigation of a Well Test ........................... 3-40 3.4.9 Field Example of Well Test Analysis .............................. 3-40 3.4.10 Welltest Model for Dry Gas Reservoir .......................... 3-52

3.5 FACTORS AFFECTING PERMEABILITY..................................... 3-56 3.5.1 Compaction..................................................................... 3-56 3.5.2 Pore Size (Grain Size) ..................................................... 3-56 3.5.3 Sorting ............................................................................ 3-60 3.5.4 Cementation ................................................................... 3-60 3.5.5 Layering .......................................................................... 3-60 3.5.6 Clay Swelling....................................................................3-61

3.6 TYPICAL RESERVOIR PERMEABILITY VALUES .......................3-61

3.7 PERMEABILITY-POROSITY CORRELATIONS............................3-61

3.8 CAPILLARY TUBE MODELS OF POROUS MEDIA..................... 3-69 3.8.1 Carman-Kozeny Equation .............................................. 3-69 3.8.2 Tortuosity ........................................................................3-75 3.8.3 Calculation of Permeability from Pore

Page 5: +Peters Ekwere j. - Petrophysics

Size Distribution............................................................. 3-79 3.9 STEADY STATE FLOW THROUGH FRACTURES....................... 3-84

3.10 AVERAGING PERMEABILITY DATA .......................................... 3-85

3.11 DARCY’S LAW FOR INCLINED FLOW........................................ 3-88 3.12 VALIDITY OF DARCY’S LAW....................................................... 3-99

3.13 NON-DARCY FLOW..................................................................... 3-101

3.14 DARCY’S LAW FOR ANISOTROPIC POROUS MEDIA............. 3-106 3.14.1 Definition of Homogeneity and Anisotropy ................ 3-106 3.14.2 Darcy’s Law for Homogeneous and Anisotropic Medium.............................................3-107 3.14.3 Transformation of Permeability Tensor from One Coordinate system to Another .................... 3-114 3.14.4 Alternative Calculation of the Principal Values and the Principal Axes of the Permeability Anisotropy..............................................3-122 3.14.5 Directional Permeability...............................................3-124 3.14.6 Measurement of Transverse Permeability of a Cylindrical Core ....................................................3-137

3.15 EXAMPLE APPLICATIONS OF PERMEABILITY.......................3-140 3.15.1 Productivity of Horizontal Well ....................................3-140 Introduction............................................................... 3-141 Homogeneous and Isotropic Reservoirs ................... 3-141 Homogeneous and Anisotropic Reservoirs ...............3-145 3.15.2 Productivity of a Vertically Fractured Well..................3-152

NOMENCLATURE .......................................................................3-155

REFRENCES AND SUGGESTED READINGS.............................3-159

4 HETEROGENEITY ................................................................ 4-1

4.1 INTRODUCTION..............................................................................4-1

4.2 MEASURES OF CENTRAL TENDENCY AND VARIABILITY (HETEROGENEITY)............................................... 4-3 4.2.1 Measures of Central Tendency......................................... 4-3 Mean.............................................................................. 4-3 Geometric Mean............................................................ 4-3 Median .......................................................................... 4-3 Mode.............................................................................. 4-4 4.2.1 Measures of Variability (Heterogeneity or Spread) ............................................... 4-4 Variance ........................................................................ 4-4

Page 6: +Peters Ekwere j. - Petrophysics

Dykstra-Parsons Coefficient of Variation..................... 4-5 Lorenz Coefficient ......................................................... 4-8

4.3 MEASURES OF SPATIAL CONTINUITY ...................................... 4-11 4.3.1 Variogram........................................................................4-13 Definition .....................................................................4-13 How to Calculate the Variogram .................................4-16 Physical Meaning of the Variogram............................ 4-27 Variogram Models....................................................... 4-28 Fitting a Theoretical Variogram Model to an Experimental Variogram ....................................... 4-35 Variogram Anisotropy .................................................4-41 Example Experimental Variograms............................ 4-44 4.3.2 Covariance (Autocovariance) Function...........................4-51 Definition .....................................................................4-51 Physical Meaning of Covariance Function ................. 4-54 4.3.3 Correlation Coefficient Function (Autocorrelation Function) .............................................4-57

4.4 PROBABILITY DISTRIBUTIONS ................................................. 4-59 4.4.1 Normal (Gaussian) Distribution .................................... 4-60 4.4.2 Log Normal Distribution................................................ 4-72

4.5 ESTIMATION .................................................................................4-75 4.5.1 Introduction ....................................................................4-75 4.5.2 Ordinary Kriging Equations ........................................... 4-86 Derivation in Terms of the Covariance Function ................................................... 4-89 Derivation in Terms of the Variogram ....................... 4-94 Solution of the Kriging Equation in terms of the Covariance Function......................................... 4-98 Solution of the Kriging Equation in terms of Variogram ..............................................................4-103

4.6 CONDITIONAL SIMULATION....................................................4-132 4.6.1 Introduction ..................................................................4-132 4.6.2 Sequential Gaussian Simulation ...................................4-132 4.6.3 A Practical Application of Sequential Gaussian Simulation .....................................................4-136

NOMENCLATURE .......................................................................4-148

REFRENCES AND SUGGESTED READINGS.............................4-149

5 DISPERSION IN POROUS MEDIA ..........................................5-1

5.1 INTRODUCTION..............................................................................5-1

Page 7: +Peters Ekwere j. - Petrophysics

5.2 LABORATORY FIRST-CONTACT MISCIBLE DISPLACEMENTS........................................................................... 5-3

5.3 ORIGIN OF DISPERSION IN POROUS MEDIA.......................... 5-20 5.3.1 Molecular Diffusion.........................................................5-21 5.3.2 Mechanical Dispersion ....................................................5-21

5.4 CONVECTION-DISPERSION EQUATION................................... 5-23 5.4.1 Generalized Equation in Vector Notation...................... 5-23 5.4.2 One Dimensional Convection-Dispersion Equation ........................................................................ 5-25

5.4.2 Solution of the One-Dimensional Convection-Dispersion Equation ................................... 5-26

5.5 DISPERSION COEFFICENT AND DISPERSIVITY ..................... 5-42

5.6 MEASURMENT OF DISPERSION COEFFICENT AND DISPERSIVITY ......................................................................5-53

5.6.1 Traditional Laboratory Method with Breakthrough Curve .......................................................5-53 5.6.2 Laboratory Method of Peters et al. (1996) ..................... 5-56 5.6.3 Field Measurement of Dispersion Coefficient and Dispersivity ............................................ 5-71

5.7 FACTORS THAT COULD AFFECT DISPERSION COEFFICENT AND DISPERSIVITY ..............................................5-75

5.8 NUMERICAL MODELING OF FIRST-CONTACT MISCIBLE DISPLACEMENT .........................................................5-79 5.8.1 Introduction ....................................................................5-79 5.8.2 Mathematical Model of First-Contact Miscible Displacement ....................................................5-79 5.8.3 Numerical Modeling of Laboratory Experiments.......... 5-82 Experiment 1 ................................................................. 5-84 Experiment 2..................................................................5-91 Experiment 3................................................................. 5-99 Experiment 4................................................................5-106 Experiment 5................................................................ 5-116 Experiment 6................................................................ 5-121 NOMENCLATURE .......................................................................5-126

REFRENCES AND SUGGESTED READINGS.............................5-128 6 INTERFACIAL PHENOMENA AND WETTABILITY................ 6-1

6.1 INTRODUCTION..............................................................................6-1

6.2 SURFACE AND INTERFACIAL TENSIONS................................... 6-2 6.2.1 Surface Tension ................................................................ 6-2 6.2.2 Interfacial Tension .......................................................... 6-11

Page 8: +Peters Ekwere j. - Petrophysics

6.2.3 Measurement of Surface and Interfacial Tension......................................................... 6-20 Capillary Rise Experiment ............................................ 6-20 Sessile Drop Method ..................................................... 6-24 Pendant Drop Method .................................................. 6-26 Ring Method.................................................................. 6-27 Spinning Drop Method ................................................. 6-30

6.3 WETTABILITY................................................................................6-31 6.3.1 Definition.........................................................................6-31 6.3.2 Determination of Wettability ......................................... 6-36 Contact Angle Method ................................................. 6-37 Amott Wettability Test.................................................. 6-40 United State Bureau of Mines (USBM) Wettability Test............................................................. 6-42 6.3.3 Wettability of Petroleum Reservoirs.............................. 6-45 6.3.4 Effect of Wettability on Rock-Fluid Interactions........... 6-46 Microscopic Fluid Distribution at the Pore Scale ..................................................................... 6-47 Effect of Wettability on Irreducible Water Saturation .......................................................... 6-47 Effect of Wettability on Electrical Properties of Rocks ...................................................... 6-48 Effect of Wettability on the Efficiency of an Immiscible Displacement .............................................6-51

6.3 THERMODYMAMICS OF INTERFACES ..................................... 6-64 6.4.1 Characterization of Interfacial Tension as Specific Surface Energy.............................................. 6-64 6.4.2 Characterization of Microscopic Pore Level Fluid Displacements....................................................... 6-66 Case 1. Displacement of a Nonwetting Phase by a Wetting Phase ............................................ 6-67 Case 2. Displacement of a Wetting Phase by a Nonwetting Phase ................................................. 6-69

NOMENCLATURE .........................................................................6-71

REFRENCES AND SUGGESTED READINGS ..............................6-73 7 CAPILLARY PRESSURE .........................................................7-1

7.1 DEFINITION OF CAPILLARY PRESSURE ..................................... 7-1

7.2 CAPILLARY PRESSURE-SATURATION RELATIONSHIP FOR A POROUS MEDIUM.............................................................. 7-8 7.3 DRAINAGE CAPILLARY PRESSURE CURVE .............................. 7-17

Page 9: +Peters Ekwere j. - Petrophysics

7.4 CONVERSION OF LABORATORY CAPILLARY PRESSURE

DATA TO RESERVOIR CONDITIONS .......................................... 7-21 7.5 AVERAGING CAPILLARY PRESSURE DATA .............................. 7-21 7.6 DETERMINATION OF INITIAL STATIC RESERVOIR FLUID SATURATIONS BY USE OF DRAINAGE CAPILLARY PRESSURE CURVE.................................................. 7-28 7.7 CAPILLARY PRESSURE HYSTERESIS.........................................7-45 7.8 CAPILLARY IMBIBITION..............................................................7-54 7.9 CAPILLARY END EFFECT IN A LABORATORY CORE ...............7-57 7.9.1 Capillary End Effect.........................................................7-57 7.9.2 Mathematical Analysis of Capillary End Effect ..............7-59 7.9.3 Mathematical Model of Capillary End Effect During Steady State Relative Permeability Measurement.................................................................. 7-68 7.9.4 Experimental Evidence of Capillary End Effect...............7-70 7.10 CAPILLARY PRESSURE MEASUREMENTS ................................7-76 7.10.1 Restored State Method (Porous Plate Method)..............7-76 7.10.2 Mercury Injection Method .............................................7-77 7.10.3 Centrifuge Method..........................................................7-81 7.11 PORE SIZE DISTRIBUTION......................................................... 7-96 7.11.1 Introduction.................................................................... 7-96 7.11.2 Pore Volume Distribution .............................................. 7-98 7.11.3 Pore Size Distribution Based on Bundle of Capillary Tubes Model............................................7-103 7.11.4 Mercury Injection Porosimeter..................................... 7-115 7.12 CALCULATION OF PERMEABILITY FROM DRAINAGE CAPILLARY PRESSURE CURVE................................................. 7-118 7.12.1 Calculation of Absolute Permeability from Drainage Capillary Pressure Curve............................. 7-118 7.12.2 Calculation of Relative Permeabilities from Drainage Capillary Pressure Curve.............................7-132 7.13 EMPIRICAL CAPILLARY PRESSURE MODELS ........................ 7-133 7.13.1 Brooks-Corey Capillary Pressure Models ..................... 7-133 7.13.2 van Genuchten Capillary Pressure Model ....................7-143 7.14 CAPILLARY TRAPPING IN POROUS MEDIA........................... 7- 145 7.14.1 Pore Doublet Model of Capillary Trapping................... 7-145

Page 10: +Peters Ekwere j. - Petrophysics

7.14.2 Snap-Off Model of Capillary Trapping ......................... 7-152 7.14.3 Mobilization of Residual Non-Wetting Phase.............. 7-155 7.14.4 Oil Migration................................................................. 7-159

7.15 EFFECTS OF WETTABILITY AND INTERFACIAL TENSION ON CAPILLARY PRESSURE CURVES.......................7-162 NOMENCLATURE .......................................................................7-164

REFRENCES AND SUGGESTED READINGS ............................ 7-168 8 RELATIVE PERMEABILITY................................................... 8-1 8.1 DEFINITION OF RELATIVE PERMEABILITY...............................8-1

8.2 LABORATORY MEASUREMENT OF TWO-PHASE RELATIVE PERMEABILITY BY THE STEADY STATE METHOD ......................................................................................... 8-6 8.3 THEORY OF ONE DIMENSIONAL IMMISCIBLE DISPLACEMENT IN A POROUS MEDIUM..................................8-15 8.3.1 Mathematical Model of Two-Phase Immiscible Displacement................................................8-15 8.3.2 Buckley-Leverett Approximate Solution of the Immiscible Displacement Equation................................8-21 8.3.3 Waterflood Performance Calculations from Buckley-Leverett Theory .................................................8-31 Oil Recovery at any Time ...............................................8-31 Oil Recovery Before Water Breakthrough .....................8-31 Oil Recovery at Water Breakthrough............................ 8-32 Oil Recovery After Water Breakthrough ...................... 8-36 Water Production...........................................................8-41 8.4 LABORATORY MEASUREMENT OF TWO-PHASE RELATIVE PERMEABILITY BY THE UNSTEADY STATE METHOD ........................................................................................8-51 8.5 FACTORS AFFECTING RELATIVE PERMEABILITIES.............. 8-65 8.5.1 Fluid Saturation.............................................................. 8-65 8.5.2 Saturation History.......................................................... 8-66 8.5.3 Wettability ...................................................................... 8-67 8.5.4 Injection Rate ................................................................. 8-70 8.5.5 Viscosity Ratio ................................................................ 8-73 8.5.6 Interfacial Tension ......................................................... 8-74 8.5.7Pore Structure .................................................................. 8-75 8.5.8 Temperature ................................................................... 8-76 8.5.9 Heterogeneity ................................................................. 8-78

Page 11: +Peters Ekwere j. - Petrophysics

8.6 THREE-PHASE RELATIVE PERMEABILITIES .......................... 8-79 8.4 CALCULATION OF RELATIVE PERMEABILITIES FROM DRAINAGE CAPILLARY PRESSURE CURVE .............................8-82 NOMENCLATURE .........................................................................8-91

REFERENCES AND SUGGESTED READINGS ............................8-94 APPENDIX A: A Systematic Approach To Dimensional Analysis .. A-1

Summary.......................................................................................... A-1 Introduction..................................................................................... A-1 Algebraic Theory of Dimensional Analysis......................................A-2

Transformation of the Dimensionless π Groups .............................A-9 Example Problem.............................................................................A-9 Procedure ....................................................................................... A-10

Transformation of the Dimensionless π Groups for Example Problem........................................................................... A-21 Some Practical Considerations ......................................................A-28 Concluding Remarks...................................................................... A-31 Nomenclature ................................................................................ A-31 References ......................................................................................A-32

Page 12: +Peters Ekwere j. - Petrophysics

CHAPTER 1

INTRODUCTION

1.1 PETROPHYSICS

Petrophysics is the study of rock properties and their interactions

with fluids (gases, liquid hydrocarbons and aqueous solutions). Because

petroleum reservoir rocks must have porosity and permeability, we are

most interested in the properties of porous and permeable rocks. The

purpose of this text is to provide a basic understanding of the physical

properties of permeable geologic rocks and the interactions of the various

fluids with their interstitial surfaces. Particular emphasis is placed on

the transport properties of the rocks for single phase and multiphase

flow.

The petrophysical properties that are discussed in this text

include:

• Porosity

• Absolute permeability

• Effective and relative permeabilities

• Water saturation

1-1

Page 13: +Peters Ekwere j. - Petrophysics

• Irreducible water saturation

• Hydrocarbon saturation

• Residual oil saturation

• Capillary pressure

• Wettability

• Pore size

• Pore size distribution

• Pore structure

• Net pay thickness

• Isothermal coefficient of compressibility

• Mineralogy

• Specific pore surface area

• Dispersivity

1.2 PETROLEUM RESERVOIR ROCKS

A petroleum reservoir rock is a porous medium that is sufficiently

permeable to permit fluid flow through it. In the presence of

interconnected fluid phases of different density and viscosity, such as

water and hydrocarbons, the movement of the fluids is influenced by

gravity and capillary forces. The fluids separate, therefore, in order of

density when flow through a permeable stratum is arrested by a zone of

low permeability, and, in time, a petroleum reservoir is formed in such a

trap. Porosity and permeability are two fundamental petrophysical

properties of petroleum reservoir rocks.

1-2

Page 14: +Peters Ekwere j. - Petrophysics

Geologically, a petroleum reservoir is a complex of porous and

permeable rock that contains an accumulation of hydrocarbons under a

set of geological conditions that prevent escape by gravitational and

capillary forces. The initial fluid distribution in the reservoir rock, which

is determined by the balance of gravitational and capillary forces, is of

significant interest at the time of discovery.

A rock capable of producing oil, gas and water is called a reservoir

rock. In general, to be of commercial value, a reservoir rock must have

sufficient thickness, areal extent and pore space to contain a large

volume of hydrocarbons and must yield the contained fluids at a

satisfactory rate when the reservoir is penetrated by a well. Any buried

rock, be it sedimentary, igneous or metamorphic, that meets these

conditions may be used as a reservoir rock by migrating hydrocarbons.

However, most reservoir rocks are sedimentary rocks.

Sandstones and carbonates (limestones and dolomites) are the

most common reservoir rocks. They contain most of the world’s

petroleum reserves in about equal proportions even though carbonates

make up only about 25% of sedimentary rocks. The reservoir character

of a rock may be primary such as the intergranular porosity of a

sandstone, or secondary, resulting from chemical or physical changes

such as dolomitization, solution and fracturing. Shales frequently form

the impermeable cap rocks for petroleum traps.

The distribution of reservoirs and the trend of pore space are the

end product of numerous natural processes, some depositional and some

post-depositional. Their prediction, and the explanation and prediction of

their performance involve the recognition of the genesis of the ancient

sediments, the interpretation of which depends upon an understanding

of sedimentary and diagenetic processes. Diagenesis is the process of

1-3

Page 15: +Peters Ekwere j. - Petrophysics

physical and chemical changes in sediments after deposition that convert

them to consolidated rock such as compaction, cementation,

recrystallization and perhaps replacement as in the development of

dolomite.

1.3 MINERAL CONSTITUENTS OF ROCKS - A REVIEW

The physical and chemical properties of rocks are the consequence

of their mineral composition. A mineral is a naturally occurring

crystalline inorganic material that has specific physical and chemical

properties, which are either constant or vary within certain limits. Rock-

forming minerals of interest in petroleum engineering can be classified

into the following families: silicates, carbonates, oxides, sulfates

(sulphates), sulfides (sulphides) and chorides. These are summarized in

Table 1.1. Silicates are the most abundant rock-forming minerals in the

Earth’s crust.

Table 1.1 Rock - Forming Minerals

Name Chemical Formula Specific Gravity Silicates

Quartz Orthoclase Plagioclase Clay

SiO2

KAlSi2O8 NaAlSi3O8 CaAl2Si2O8

Al2Si2O5(OH)

and many more

2.65 2.57

2.62 - 2.76

2.5

Carbonates Calcite Dolomite

CaCO3

CaMg(CO3)2

2.72 2.85

Oxides Magnetite Hematite

Fe3O4 Fe2O3

5.18

4.9 - 5.3

1-4

Page 16: +Peters Ekwere j. - Petrophysics

Sulfates Anhydrite Gypsum Barite

CaSO4

CaSO4.2H2O BaSO4

2.89 - 2.98

2.32 4.5

Sulfide Pyrite

FeS2

5.02

Chloride Halite

NaCl

2.16

1.4 ROCKS

A rock is an aggregate of one or more minerals. There are three

classes of rocks: igneous, metamorphic and sedimentary rocks .

1.4.1 Igneous Rocks

These are rocks formed from molten material (called magma) that

solidified upon cooling either:

1. At the earth’s surface to form volcanic or extrusive rocks, e.g.,

basaltic lava flows, volcanic glass and volcanic ash.

or

2. Below the surface, usually at great depths, to form plutonic or

intrusive rocks, e.g., granites and gabbros.

Igneous rocks are the most abundant rocks on the earth’s crust,

making up about 64.7% of the Earth’s crust.

1-5

Page 17: +Peters Ekwere j. - Petrophysics

1.4.2 Metamorphic Rocks

These are rocks formed by transformation, generally in the solid

state, of pre-existing rocks beneath the surface by heat, pressure and

chemically active fluids, e.g., quartz is transformed to quartzite and

limestone plus quartz plus heat gives marble and carbon dioxide.

Metamorphic rocks are the second most abundant rocks on the

earth’s crust, making up 27.4% of the Earth’s crust.

1.4.3 Sedimentary Rocks

These are rocks formed at the surface of the earth either by

1. Accumulation and consolidation of minerals, rocks and/or

organisms and vegetation, e.g., sandstone and limestone.

or

2. Precipitation from solution such as sea water or surface water,

e.g., salt and limestone.

Sedimentary rocks are the source of petroleum and provide the

reservoir rock and trap to hold the petroleum in the earth’s crust.

Sedimentary rocks are the least abundant rocks on the earth’s crust,

making up about 7.9% of the earth’s crust. Because most reservoir

rocks are sedimentary rocks, they are of particular interest to us in the

study of petrophysics. Therefore, we will examine sedimentary rocks in

more detail than igneous and metamorphic rocks.

1-6

Page 18: +Peters Ekwere j. - Petrophysics

1.5 CLASSIFICATION OF SEDIMENTARY ROCKS

Sedimentary rocks may be classified by origin and composition as

clastic, chemical or organic. Tables 1.2 to 1.4 show the various rock

types for each class.

1.5.1 Clastic Sedimentary Rocks

These rocks are composed of fragments or minerals broken from

any type of pre-existing rock. Clastic sedimentary rocks are usually

classified by grain size as boulder, cobble, gravel, sand, silt and clay.

Figure 1.1 shows such a classification known as the Wentworth scale.

Table 1.2 Clastic Sedimentary Rocks

1.5.2 Chemical Sedimentary Rocks

These rocks are formed by chemical precipitation as carbonates,

e.g., limestone (CaCO3) and dolomite (CaMg(CO3)2) or as evaporties, e.g.,

gypsum (CaSO4.2H2O), anhydrite (CaSO4) and salt (NaCl).

1-7

Page 19: +Peters Ekwere j. - Petrophysics

1.5.3 Organic Sedimentary Rocks

These rocks are formed by biologic precipitation and by

accumulation of organic (plant and animal) material, e.g., peat, coal,

diatomite and limestone.

Figure 1.1 Classification of clastic rocks according to texture.

1-8

Page 20: +Peters Ekwere j. - Petrophysics

Table 1.3 Chemical Sedimentary Rocks (Precipitates)

Table 1.4 Organic Sedimentary Rocks

1-9

Page 21: +Peters Ekwere j. - Petrophysics

1.6 DISTRIBUTION OF SEDIMENTARY ROCK TYPES

Table 1.5 shows the approximate distribution of sedimentary rocks

in the earth’s crust. Shales make up over 50% of total sedimentary rock

volume in the earth’s crust.

Table 1.5 Distribution of Sedimentary Rocks Type % Earth’s Crust % Sedimentary Rock Shale 4.2 53

Sandstone 1.7 22 Limestone and

Dolomite 2.0 25

Total 7.9 100

1.7 SANDSTONE RESERVOIRS (CLASTIC SEDIMENTARY ROCK)

Sandstones are composed of fragmented materials, which have

been transported to the site of deposition by water currents and which

have been subjected to varying degrees of wave and current action

during transport and during deposition. Consequently, sandstone

reservoirs vary from clean, well sorted quartz sandstone with well

rounded grains (Figure 1.2a) through more angular feldspathic

sandstone containing varying amounts of clay (Figure 1.2b), to

argillaceous, very poorly sorted sandstone containing varying amounts of

rock fragments (Figure 1.2c) all of which may be affected by varying

degrees of compaction, cementation, solution and replacement.

1-10

Page 22: +Peters Ekwere j. - Petrophysics

Figure 1.2: Examples of sandstone reservoir rocks. (A) clean, well sorted sandstone, (B) angular, feldspathic sandstone and (C) argillacious, very

poorly sorted sandstone.

1-11

Page 23: +Peters Ekwere j. - Petrophysics

1.7.1 Pore Space

The basic framework of a sandstone reservoir is formed by the

sand grains between which the pore space may or may not contain

interstitial fine material and/or cement (Figure 1.3). The amount of this

intergranular pore space or porosity is controlled primarily by sorting of

the sediment and to a lesser extent by the packing of the grains. Sorting

is a measure of the spread of distribution of grain size on either side of

an average in a sediment. Theoretically, grain size has no effect on

porosity. This is true only for spherical grains of the same size.

However, the arrangement of such spheres has a large effect on the

porosity of the pack.

Figure 1.3: Framework of reservoir sand with interstitial clay and

cement.

1-12

Page 24: +Peters Ekwere j. - Petrophysics

Porosity is at its maximum for spherical grains but becomes

progressively less as the angularity of the grains increases because such

grains pack together more closely. Experimental data from artificial

sands confirm that the grain size essentially has no influence on porosity

for well sorted sand. However, porosities of wet packed sands show a

general decrease as sorting becomes poorer. This is because with a

mixture of sizes, the smaller grains partially fill the interstices between

the larger grains.

Permeability, being a measure of the ease with which a material

transmits fluids, depends primarily upon the size, shapes and extent of

the interconnections between individual pores rather than the size of the

pores themselves. However, since the interconnections are directly

related to the pore size which in turn is related to grain size, there are

relationships between these factors and permeability. Krumbein and

Monk (1942) have shown that permeability varies as the square of the

mean grain diameter and a complex function of sorting (other factors

being equal). Experimental data show a marked decrease in permeability

as grain size decreases and as sorting becomes poorer. Experience has

also shown that the permeability measured normal to the bedding is

usually less than permeability measured parallel to the bedding and that

large variations in permeability occur in different directions in the

bedding plane.

Clay in the pore space of a reservoir may affect the performance of

a reservoir very adversely. The amount and kind of clay, as well as

distribution throughout the reservoir rock, has an important bearing on

liquid permeability, whereas a small amount has little effect on porosity.

Figures 1.4 and 1.5 show how the dispersed clay morphology affects the

permeability and capillary pressure of sandstones.

1-13

Page 25: +Peters Ekwere j. - Petrophysics

Figure1.4: Three general types of dispersed clay in sandstone reservoir

rocks and their effects on permeability: (a) discrete particles of kaolinite; (b) pore lining by chlorite; (c ) pore bridging by illite (from Neasham,

1977).

1-14

Page 26: +Peters Ekwere j. - Petrophysics

Figure1.5: Three general types of dispersed clay in sandstone reservoir rocks and their effects on capillary pressure (from Neasham, 1977).

If fresh water, for example drilling fluid filtrate, invades a reservoir,

certain clays, such as montmorillonite, will swell and plug some of the

pore interconnections, drastically reducing the permeability, whereas

saline water may have no effect.

1.7.2 Compaction and Cementation

The pore space of the original unconsolidated sediment is reduced

in ancient rocks by many factors. Compaction and cementation are the

most important of these factors but they in turn are affected by

1-15

Page 27: +Peters Ekwere j. - Petrophysics

composition of the sediment, and the contained fluids and their

pressures. Compaction by the weight of the overburden commences as

soon as a sediment is deposited. It produces reduction of pore space as

a result of:

1. Closer packing of the grains which causes smaller pores and

connected passages.

2. Crushing and fracturing of grains, and dissolution at points of

contact sometimes accompanied by precipitation of silica in the

pore space nearby. Alkaline interstitial water seems to provide

conditions more conducive to dissolution than saline water.

3. Plastic deformation of the softer grains which tend to mold around

the harder grains thus destroying pore space. The softer grains

may be composed of limestone, shale, siltstone and other rock

fragments, and when the amount of such soft material exceeds 10

to 12%, the permeability may be largely destroyed even though

some porosity usually remains.

Early cementation of sand may produce a rigid framework which

will inhibit compaction until the depth of burial exceeds that at which

fracturing of the grains and cement is initiated. Abnormal fluid pressure

in sandstone reservoirs also inhibits compaction because the overburden

is partly supported by the fluid pressure.

Reduction of reservoir pressure by production of fluid can lead to

compaction of the producing zone by rearrangement of the sand grains.

This can produce serious and expensive subsidence of surface facilities

such as occurred in the Wilmington field in California, Bolivar District

fields in Venezuela and Ekofisk field in the North Sea. This form of

1-16

Page 28: +Peters Ekwere j. - Petrophysics

compaction leads to porosity and permeability reduction which is

irreversible and which may affect the producing characteristics of the

reservoir very adversely.

Cementation is the result of recrystallization from solution of silica,

carbonates and other soluble materials in the pores of clastic rocks. The

most common cementing materials in sandstone reservoirs are silica and

calcite but many others do occur. It is not uncommon to find both silica

and calcite present in the cement and in such cases, the silica in the

form of quartz has precipitated first and the calcite later. Silica cement

usually occurs in the form of quartz and grows in optical continuity with

the sand grains until finally the crystals interfere with one another and

an interlocking network results. Calcite cement is often patchy but may

completely fill the pore space.

Silica cement appears to have two origins: (1) early cementation

before the sands were compacted appreciably, and (2) deposition of silica

predissolved by pressure solution during compaction. In the Eocene

Wilcox sandstones of the Gulf Coast, the distribution of silica cement can

be related to both primary texture and depth of burial. The amount of

silica cement tends to increase in coarser and better sorted sands and, to

a lesser extent, with depth. Much of this cement appears to be early and

unrelated to the compaction process.

1.7.3 Classification

Sandstones may be classified by mineral composition (Figure 1.6).

The principal types are: (a) quartz sandstones consisting of over 95%

detrital quartz, (b) feldspathic sandstone consisting of 5 - 25% feldspar,

(c) arkosic sandstone consisting of over 25% feldspar, (d) sublithic

sandstone consisting of 5 - 25% rock fragments and (e) lithic sandstone

consisting of over 25% rock fragments.

1-17

Page 29: +Peters Ekwere j. - Petrophysics

Sandstone grain texture consists of five components: (a) size, (b)

sorting, (c) shape (sphericity), (d) roundness, and (e) packing. Figure 1.7

shows a grain size comparator where some of these qualitative terms are

presented visually. Porosity is independent of grain size for uniform

grains but decreases as sorting gets poorer. Close packing reduces

porosity and permeability. The effects of shape and roundness on

porosity are less definite. Permeability increases with increasing grain

size, but decreases with poorer sorting. Permeability generally increases

with angularity and decreasing sphericity.

Figure 1.6: Classification of sandstones by composition.

1-18

Page 30: +Peters Ekwere j. - Petrophysics

Figure 1.7: Grain size comparator chart (from Stow, 2005).

1-19

Page 31: +Peters Ekwere j. - Petrophysics

1.8 CARBONATE RESERVOIRS (LIMESTONES AND DOLOMITES)

Most carbonate rocks, like clastics, are composed of particles of

clay to gravel size that were generally deposited in a marine environment.

However, they differ from terrigenous clastics in that they are deposited

as lime particles which are produced locally, whereas, sandstones are

composed of particles transported from an outside source by water

currents. They differ even more importantly from sandstones by being

subject to more post-depositional diagenesis ranging from simple

cementation of the original particles to complete recrystallization or

replacement by dolomite or chert. In addition, they are susceptible to

solution at any stage in their diagenesis. They are usually more poorly

sorted than clastics.

Components of carbonate rocks are usually (1) grains of various

kinds, (2) lime mud, and (3) carbonate cement precipitated later. There

are several kinds of grains, of which four are the most important. These

are (1) shell fragments, called “bio”, (2) fragments of previously deposited

limestone called “intraclasts”, (3) small round pellets - the excreta of

worms, and (4) ooliths - spheres formed by rolling lime particles along

the bottom. Lime mud consists of clay-sized particles of lime.

The material between the grains may be primary lime mud

deposited at the same time as the grains which would be grain-supported

rock, or the grains may be “floating” in lime mud which would be mud-

supported rock.

1.8.1 Classification

Carbonates are usually classified according to depositional texture

as shown in Figure 1.8. The presence or absence of lime mud and the

1-20

Page 32: +Peters Ekwere j. - Petrophysics

type and abundance of grains form the basis of the classification. A

boundstone consists of original skeletal components bound together

during deposition (Figure 1.9). Grainstones consist of packed carbonate

grains with the texture being grain-supported and very little lime mud

(Figure 1.10). Packstones are grain-supported but contain very

substantial amounts of lime mud. Wackestones have a larger amount of

lime muds, such that the grains effectively “float” in the mud. Mudstones

consist of essentially lime muds only. The presence of lime mud may be

most important in the development of porosity in carbonates because

under the right conditions, lime mud may be preferentially dolomitized

and may also be more readily leached out than the grains.

Good porosity in carbonate reservoirs is usually due to

dolomitization. The largest volume of carbonate petroleum reserves

comes from dolomites. Dolomitization occurs from the substitution of

magnesium for calcium in half of the sites in a carbonate crystal. A

volume loss of 12 to 13% due to dolomitization results in a corresponding

increase in porosity. Due to their larger surface area, mud-size grains

are more easily dolomitized than sand-sized grains. Thus, the best

carbonate reservoirs may have the lowest primary porosity.

Dolomitization also creates planar crystal surfaces and harder crystal

structures. Thus, dolomites retain more of their porosity during

compaction than limestones.

1-21

Page 33: +Peters Ekwere j. - Petrophysics

Figure 1.8: Classification of carbonates by texture.

1.8.2 Pore Space

The porosity, permeability and pore space distribution in carbonate

reservoir rocks are related to both the depositional environment and the

diagenesis of the sediment. They are most commonly of secondary

(diagenetic) origin although residual primary pore space does occur.

Carbonates have a large range of pore structures due to the

complex nature of carbonate constituents and diagenetic features. The

pore structures have been classified by Choquette and Pray, 1970) as

shown in Figure 1.10.

1-22

Page 34: +Peters Ekwere j. - Petrophysics

Figure 1.9: Examples of boundstone.

1-23

Page 35: +Peters Ekwere j. - Petrophysics

Figure 1.9: Examples of grainstones.

1-24

Page 36: +Peters Ekwere j. - Petrophysics

Fabric-selective porosity includes:

•Interparticle porosity.

•Intercrystalline porosity - typical of dolomites.

•Fenestral porosity - by solution along bedding planes or joint

surfaces.

•Skeletal, framework, molding, or shelter porosity - selective solution of,

within, or around fossil material.

•Oomoldic porosity - selective solution of ooliths.

•Non fabric-selective porosity includes:

•Fracture porosity - by stress or shrinkage.

•Channel porosity - widening and coalescence of fractures.

•Vuggy or cavernous porosity.

•Bioturbation porosity - from boring and burrowing.

•Breccia porosity - in some cases, really high fracture porosity.

In carbonates, porosity and permeability are not well-correlated

with grain size or sorting. Porosity and permeability are controlled

largely by the amount of fines and by diagenesis. Correlation of

petrophysical properties with rock type is thus very difficult.

1-25

Page 37: +Peters Ekwere j. - Petrophysics

Figure 1.10: Classification of pore systems in carbonate rocks (Choquette and Pray, 1970).

1-26

Page 38: +Peters Ekwere j. - Petrophysics

Table 1.6 shows a comparison of the pore space characteristics of

clastic and carbonate rocks (Choquette and Pray, 1970).

Table 1.6. Comparison of Pore-space Properties in Clastic and Carbonate Rocks (Choquette and Pray, 1970).

1-27

Page 39: +Peters Ekwere j. - Petrophysics

1.9 FRACTURED RESERVOIRS

Natural reservoir fractures are caused by brittle failure, usually

due to such factors as (a) folding, (b) faulting, (c) fluid pressure, (d)

release of lithostatic pressure, (e) pressure solution, (f) dehydration, (g)

weathering, (h) cooling and (i) impact craters. Natural fractures can exist

in essentially any type of rock although they are particularly common in

carbonates.

Naturally fractured reservoirs are usually treated by a dual

porosity approach to deal with their properties. The matrix rock

(between fractures) usually has reasonable porosity and extremely low

permeability. Fractures range in size from hair-size to several

millimeters in aperture. Fractures that have not been filled with cement

have very high permeabilities, even though they may be fairly widely-

spaced. However, the fracture system generally contains only a small

fraction of the reservoir pore space. Thus, the matrix contains the bulk

of the reservoir pore volume while the fractures contain the bulk of the

reservoir flow capacity.

Figure 1.11 shows a naturally fractured rock together with its

idealized dual porosity approximation.

1-28

Page 40: +Peters Ekwere j. - Petrophysics

Figure 1.11: Idealization of naturally fractured reservoir (Warren and Root, 1963)

1.10 RESERVOIR COLUMN

Figure 1.12 shows a reservoir column penetrated by a well. The

total thickness of the reservoir as determined from the spontaneous

potential (SP) log, discussed in Chapter 2, is H. This reservoir contains a

hydrocarbon bearing zone and a water bearing zone at the bottom. The

gross pay thickness, which is the thickness of the hydrocarbon bearing

portion of the reservoir as determined from the resistivity log (see

Chapter 2), is . However, this thickness contains shale breaks which

will not be productive and must be discounted to determine the net pay

to be used in reserves estimation. The net pay for this example is given

by

0h

1-29

Page 41: +Peters Ekwere j. - Petrophysics

6

1Net pay

i

ii

h=

=

= = h∑ (1.1)

The net to gross (NTG) pay is defined as

6

1

0 0

Net to gross (NTG)

i

ii

hhh h

=

== =∑

(1.2)

The net to gross is a number that is less than or equal to 1 or if

expressed as a percentage, is a number that is less than or equal to

100%. Notice that in this example, there is a gas oil contact (GOC) and

an oil water contact (OWC) in the reservoir. The thickness of the gas zone

is gash and that of the oil bearing zone is . Of course, not all petroleum

reservoirs have a gas oil contact or an oil water contact. Net pay is used

along with other petrophysical properties of the reservoir to estimate the

hydrocarbon reserve as discussed in Chapter 2.

oilh

1-30

Page 42: +Peters Ekwere j. - Petrophysics

Figure 1.12. Reservoir column showing gross and net pay.

1-31

Page 43: +Peters Ekwere j. - Petrophysics

1-32

REFERENCES

Archie, G.E. :“Introduction to Petrophysics of Reservoir Rocks,” AAPG Bull., Vol. 34, No. 5 (May 1950) 943-961.

Beard, D.C. and Weyl, P.K. : “Influence of Texture on Porosity and Permeability of Unconsolidated Sand,” AAPG Bull. (Feb. 1973) 57, 349-369.

Choquette, P.W. and Pray, L.C. : “Geologic Nomenclature and Classification of Porosity in Sedimentary Carbonates,” AAPG Bull., Vol. 54, No. 2 (1970) 207-250.

Krumbein, W.C. and Monk, G.D. : “Permeability as a Function of the Size Parameters of Unconsolidated Sand,” Amer. Int. Mining and Met. Tech. Pub. 1492, 1942.

Levorsen, A.I. : Geology of Petroleum, W.H. Freeman and Company, San Francisco, 1967.

Neasham, J.W.: "The Morphology of Dispersed Clay in Sandstone Reservoirs and Its Effect on Sandstone Shaliness, Pore Space and Fluid Flow Properties," SPE 6858, Presented at the 52nd Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, Denver, Oct. 9-12, 1977.

Selley, R.C. : Elements of Petroleum Geology, W.H. Freeman, New York, 1985.

Stoneley, R. : An Introduction to Petroleum Exploration for Non-Geologists, Oxford University Press, New York, 1995.

Stow, D.A.V. : Sedimentary Rocks in the Field, Elsevier Academic Press, Burlington, 2005.

Page 44: +Peters Ekwere j. - Petrophysics

2-1

CHAPTER 2

POROSITY AND FLUID SATURATIONS

2.1 DEFINITION OF POROSITY

Porosity gives an indication of the rock’s ability to store fluids. It is

defined as the ratio of the pore volume to the bulk volume of the porous

medium as shown in the following equation:

p b s

b b

V V VV V

φ −= = (2.1)

Porosity may be classified as total or effective porosity. Total

porosity accounts for all the pores in the rock (interconnected and

isolated pores) whereas effective porosity only accounts for the

interconnected pores. Therefore, effective porosity will be less than or

equal to total porosity depending on the amount of isolated pores in the

rock. From a reservoir engineering standpoint, it is the effective porosity

that matters, not the total porosity.

Page 45: +Peters Ekwere j. - Petrophysics

2–2

Porosity may also be classified as primary or secondary. Primary

porosity is that which was formed at the time of deposition of the

sediments whereas secondary porosity was developed after deposition

and burial of the formation. Sandstone porosity is practically all primary

porosity whereas carbonate porosity tends to be secondary porosity.

2.2 FACTORS AFFECTING SANDSTONE POROSITY

Sandstone porosity is affected by packing, sorting and

cementation. Packing describes the arrangement of the sand grains

relative to one another. Figures 2.1 shows three idealized types of

packing for spherical sand grains and their theoretical porosities. The

cubic packing has a porosity of 47.6%; the hexagonal packing has a

porostiy of 39.5% and the rhombohedral packing has a porosity of

25.9%. As shown by the geometrical derivations in Figure 2.1, the

porosity of a pack of uniform spheres is independent of the grain size as

the grain diameter cancels out.

Well sorted sandstone consists of grains having approximately the

same size whereas poorly sorted sandstone consists of grains having a

wide range of different grain sizes. Poor sorting reduces the porosity of

the sandstone as may be seen in Figure 2.2 in which the small grains fit

into the pores created by the large grains, thereby reducing the porosity.

Page 46: +Peters Ekwere j. - Petrophysics

2-3

Figure 2.1. Effect of packing on porosity of uniform spheres.

Figure 2.2. Effect of sorting on porosity. (A) Irregular grains,

(B) Idealized spherical grains (from Tiab and Donaldson, 2004).

Page 47: +Peters Ekwere j. - Petrophysics

2–4

Table 2.1 shows experimentally measured porosities of various

artificial sandpacks. Note the general decrease of porosity with poor

sorting for all grain sizes and the approximately constant porosity of the

extremely well sorted sands for all grain sizes.

Table 2.1 Measured Porosities of Artificial Sandpacks (adapted from Beard and Weyl, 1973)

In consolidated rocks, the sand grains are cemented together

usually by quartz or carbonates. Cementation reduces the porosity of

the sand as shown in Figure 2.3.

Figure 2.3. Effect of cementation on porosity.

Page 48: +Peters Ekwere j. - Petrophysics

2-5

2.3 FACTORS AFFECTING CARBONATE POROSITY

In carbonates, secondary porosity is usually more important than

primary porosity. The major sources of secondary porosity are

fracturing, solution and chemical replacement.

Fractures are cracks in the rock. Figure 2.4 shows an idealized

fractured formation where the grains are bricks and the fractures

constitute the pore space. Although fracture porosity is generally small,

often 1-2%, the fractures are very useful in allowing fluids to flow more

easily through the rock. Therefore, they greatly enhance the flow

capacity of the rock.

Figure 2.4. Idealized fractured rock with low fracture porosity.

Solution is a chemical reaction in which water with dissolved

carbon dioxide reacts with calcium carbonate to form calcium

bicarbonate which is soluble. This reaction increases the porosity of the

limestone. The chemical reactions are

Page 49: +Peters Ekwere j. - Petrophysics

2–6

2 2 2 3CO H O H CO+ = (2.2)

( )2 3 3 3 2H CO CaCO Ca HCO+ = (2.3)

Chemical Replacement is a chemical reaction in which one type of

ion replaces another with a resulting shrinkage in the size of the new

compound. An example is dolomitization in which some of the calcium

ions in calcium carbonate are replaced by magnesium ions to form

calcium magnesium carbonate (dolomite). This replacement causes a

shrinkage of 12 to 13% in the grain volume, with a corresponding

increase in secondary porosity. The chemical reaction is

( )3 2 3 222CaCO MgCl CaMg CO CaCl+ = + (2.4)

2.4 TYPICAL RESERVOIR POROSITY VALUES

Sandstones have porosities that typically range from 8% to 38%,

with an average of 18%. About 95% of sandstone porosity is effective

porosity. Sandstone porosity is usually mostly intergranular porosity.

Carbonates have porosities that typically range from 3% to 15%, with an

average of about 8%. About 90% of carbonate porosity is effective

porosity. Carbonate porosities are much more difficult to characterize

and may consist of (1) intergranular, (2) intercrystalline, (3) fractures and

fissures, and (4) vugular porosities.

Page 50: +Peters Ekwere j. - Petrophysics

2-7

2.5 LABORATORY MEASUREMENT OF POROSITY

2.5.1 Direct Porosity Measurement by Routine Core Analysis

Direct measurement of porosity requires the measurements of two

of the three volumes Vb, Vs and Vp. In the laboratory, measurements

are usually performed on extracted cores, which have been cleaned and

dried.

Bulk volume can be determined by (1) caliper and (2) fluid

displacement. For well machined samples, the dimensions can be

measured with a caliper, from which the bulk volume can be calculated.

Two types of fluid displacements can be used to determine bulk

volume. In the first method, fluid that does not easily penetrate the

pores such as mercury is used. The apparatus, which is known as a

pycnometer, measures the volume of mercury displaced by the sample

(Figure 2.5a). Since mercury does not penetrate the pores at

atmospheric pressure, the volume of mercury displaced is equal to the

bulk volume of the sample.

In the second method, fluid which easily saturates the sample is

used. The sample is weighed in air, evacuated and then saturated with a

liquid (brine, kerosene, or toluene). The saturated sample is weighed in

air and then weighed fully immersed in the saturating liquid. The loss in

weight of the saturated sample when fully immersed in the saturating

liquid is proportional to the bulk volume of the sample (Archimedes

principle).

Grain volume can be determined by (1) fluid displacement and (2)

gas exapansion using Boyle’s law porosimeter. The loss in weight of the

Page 51: +Peters Ekwere j. - Petrophysics

2–8

dry sample and the sample fully immersed in a liquid is proportional to

the grain volume.

Figure 2.5b shows a schematic diagram of a Boyle’s law porosimeter for grain volume determination by gas expansion. The sample, which is confined in a vessel of known volume V1, is pressured

by gas (air, nitrogen or helium) to a pressure P1 (absolute units). The vessel of volume V1 is connected to a second vessel of known volume V2,

which is initially evacuated. The valve between the two vessels is opened

and the pressure in the two vessels is allowed to stabilize at P2 (absolute units). By Boyle’s law (PV = constant at a constant temperature),

( ) ( )1 1 1 2 2s sV V P V V V P− = − + (2.5)

Eq.(2.5) can be solved for the grain volume as

21 2

1 2s

PV V VP P

⎛ ⎞= − ⎜ ⎟−⎝ ⎠

(2.6)

The instrument can be calibrated with steel blanks of known volume.

Calibration consists of a plot of Vs versus P2/(P1-P2), which should be

linear with a slope, -V2, and an intercept, V1. At least three steel blanks

of different sizes should be used in the calibration to ensure reliability of

the calibration. The three data points should fall on the calibration line.

Also, the slope and the intercept should be checked against the known

volumes, V2 and V1. Once the calibration line has been established, it is

used to convert the measurements from core samples to grain volume.

Page 52: +Peters Ekwere j. - Petrophysics

2-9

Figure 2.5. Schematics of equipment for measurement of core plug porosity.(a) Bulk volume pycnometer; (b) Boyle’s law porosimeter; (c) Bulk and pore volume porosimeter.

Page 53: +Peters Ekwere j. - Petrophysics

2–10

Pore volume can be determined by (1) fluid saturation and (2)

mercury injection. The difference in the weight of the saturated sample

and the dry sample is proportional to the pore volume.

Mercury injection consists of forcing mercury under relatively high

pressure into the pores of the sample using a mercury porosimeter

(Figure 2.5c). Typically, the core is evacuated before mercury injection.

Because any air left in the pores is compressed to a negligibly small

value, the volume of mercury injected is essentially equal to the

connected pore volume of the sample. This is a destructive method

because after the test, the sample is no longer suitable for other

measurements. Mercury porosimetry is also used to determine capillary

pressure and pore size distribution of the sample (see Chapter 6).

The methods so far described determine the effective porosity of

the sample. To determine the total porosity, the sample is ground into a

fine powder after bulk volume measurement. The grain volume of the

ground sample can be determined either by liquid displacement or by

assuming an average grain density.

The measurement of porosity on consolidated samples in routine

core analysis might generally be expected to yield values of the true

fractional porosity plus or minus 0.005, i.e., a true value of 27% porosity

may be measured between 26.5% and 27.5% porosity.

Core porosities may differ from in-situ porosities for the following

reasons:

• The core may be altered during recovery.

Page 54: +Peters Ekwere j. - Petrophysics

2-11

• The core in the laboratory is no longer subjected to the overburden

and lateral stresses that it was subjected to in the reservoir.

• The porosities are measured on small plugs, which may not be

representative of the entire reservoir.

• The volume of the core analyzed is small and may not account for

the variability of the porosity in the reservoir.

Despite these limitations, core analysis provides the only direct

measurement of porosity. Frequently, the results of core analysis are

used to calibrate well logs.

Example 2.1

An experiment has been performed to determine the porosity of an

irregularly shaped core sample. The cleaned dry sample was weighed in

air. It was then evacuated and fully saturated with an oil with a density

of 0.85 gm/cc and then weighed again in air. Afterwards, the saturated

sample was weighed when it was fully immersed in the oil. Here are the

results of the experiment.

Weight of dry sample in air = 42.40 gm

Weight of the saturated sample in air = 45.49 gm

Weight of the saturated sample immersed in the oil = 28.80 gm

a. Calculate the porosity of the core.

Page 55: +Peters Ekwere j. - Petrophysics

2–12

b. Is there enough information from this experiment to determine the

mineralogy of the sample? If yes, what is it? Please justify your answer

with appropriate arguments.

Solution to Example 2.1

Wt of dry sample (Wdry) = 42.40 gm

Wt of saturated sample (Wsat) = 45.49 gm

Wt of sample immersed in oil (Wi) = 28.80 gm

Density of saturating oil (ρL) = 0.85 gm/cc

a. Required to calculate the porosity of the sample.

Pore volume (Vp) = (Wsat – Wdry)/ρL = (45.49–42.40)/0.85 = 3.64 cc

Bulk volume (Vb) = (Wsat – Wi)/ρL = (45.49–28.80)/0.85 = 19.64 cc

Porosity (φ) = Vp/Vb = 3.64/19.64 = 0.185 or 18.5%

b. Yes. There is enough information to determine the mineralogy of

the sample through the grain density.

Grain volume (Vs) = Vb–Vp =19.64–3.64 = 16.00 cc

Alternatively, Grain volume (Vs) = (Wdry–Wi)/ρL = (42.40–

28.80)/0.85 = 16.00 cc

Page 56: +Peters Ekwere j. - Petrophysics

2-13

Grain density (ρs) = Wdry/Vs = 42.40/16.00 = 2.65 gms/cc

Specific gravity of mineral (γm) = ρs /ρw = 2.65/1.00 = 2.65

Table 1.1 lists the specific gravities of common reservoir rock minerals.

From the table, quartz has a specific gravity of 2.65, which is the same

as the specific gravity of the sample matrix. Therefore, based on the

available information, the mineral of the sample is quartz.

2.5.2 Indirect Porosity Measurement by CT Imaging

With the availability of X-ray computed tomography (CT) imaging

systems in research laboratories, it is now possible to measure the

porosity distributions in core samples. Peters and Afzal (1992) have

made such measurements in an artificial sandpack and a Berea

sandstone approximately 60 cm long and 5 cm in diameter. CT imaging

gives rise to a very large data set, over 600,000 porosity values in some

cases. Therefore, it is convenient to present the results of the

measurements as images (Figures 2.7 and 2.8). It should be noted in

Figure 2.7 that a sandpack may not be as uniform as we always assume

it to be. The packing technique used in this test introduced significant

porosity variation into the pack. The packing history is clearly evident in

the image. The dominant feature in the porosity variation of the Berea

sandstone is layering which is clearly visible in Figure 2.8. The porosity

data also can be presented in histograms as shown Figures 2.9 and 2.10.

The porosity in each voxel (volume element) was obtained by

scanning the sample dry and then scanning it fully saturated with a

wetting fluid such as brine. The x-ray attenuation equations for the two

scans are

Page 57: +Peters Ekwere j. - Petrophysics

2–14

( )1m air dryψ φ φψ ψ− + = (2.7)

( )1m brine wetψ φ φψ ψ− + = (2.8)

Eqs.(2.7) and (2.8) can be solved simultaneously to obtain the porosity in

each voxel as

wet dry

brine air

ψ ψφ

ψ ψ−

=−

(2.9)

The x-ray attenuation coefficient of the brine in Eq.(2.9) is obtained by

scanning a sample of the brine in a test tube and the attenuation for air

is assumed to be zero.

Figure 2.7. Porosity image of a sandpack from CT imaging. L = 54.2 cm,

Page 58: +Peters Ekwere j. - Petrophysics

2-15

d = 4.8 cm. (a) Cross-sectional slice. (b) Longitudinal vertical slice. (Peters and Afzal, 1992).

Figure 2.8. Porosity image of a Berea sandstone from CT imaging. L = 60.2 cm, d = 5.1 cm. (a) Cross-sectional slice. (b) Longitudinal vertical

slice. (Peters and Afzal, 1992).

Page 59: +Peters Ekwere j. - Petrophysics

2–16

Figure 2.9. Porosity histogram for a sandpack from CT imaging. L = 54.2 cm, d = 4.8 cm. Mean = 29.7%, Standard deviation = 2.5%.(Peters

and Afzal, 1992).

Page 60: +Peters Ekwere j. - Petrophysics

2-17

Figure 2.10. Porosity histogram for a Berea sandstone from CT imaging. L = 60.2 cm, d = 5.1 cm. Mean = 17.3%, Standard

deviation = 2.0%.(Peters and Afzal, 1992).

2.6 FLUID SATURATIONS

In a petroleum reservoir, there is always more than one fluid phase

occupying the pore space. In an oil reservoir, oil and water occupy the

pore space. In a gas reservoir, gas and water occupy the pore space. At a

certain point in the production of an oil reservoir, oil, water and gas

could occupy the pore space. There is a need to keep track of the

quantity of each type of fluid occupying the pore space. The

petrophysical property that describes the amount of each fluid type in

Page 61: +Peters Ekwere j. - Petrophysics

2–18

the pore space is the fluid saturation. It is defined as the fraction of the

pore space occupied by a fluid phase. Thus, in general,

Fluid VolumeFluid Saturation = Effective Rock Pore Volume

(2.10)

If Sw = water saturation, So = oil saturation and Sg = gas saturation,

then Sw = Vw/Vp, So = Vo/Vp, and Sg = Vg/Vp, where Vw, Vo, Vg and

Vp are the volumes of water, oil, gas and pore space, respectively. For an

oil reservoir without a free gas saturation, So + Sw = 1.0. For a gas

reservoir without a liquid hydrocarbon saturation, Sg + Sw = 1.0. For an

oil reservoir with a free gas saturation, So + Sw + Sg = 1.0. Fluid

saturation may also be expressed in %.

There are two methods of determining the in-situ fluid saturations

in a petroleum reservoir rock. The direct approach is to measure the

fluid saturations from a core cut from the reservoir. The indirect

approach is to measure some other physical property of the rock that can

be related to fluid saturation. The direct approach is discussed here.

The indirect approach such as using electric logs or capillary pressure

measurements to estimate water saturation will be discussed later.

One method of direct measurement of fluid saturations is the

retort method. In this method, a core sample is heated so as to vaporize

the water and oil, which are condensed and collected in a small,

graduated receiving vessel (Figure 2.11). The volumes of oil and water

divided by the pore volume of the core sample give the oil and water

saturations. The gas saturation is obtained indirectly by the requirement

that saturations must sum to one.

Page 62: +Peters Ekwere j. - Petrophysics

2-19

Figure 2.11. Retort distillation apparatus.

There are two disadvantages to the retort method of saturation

determination. In order to remove all the oil, it is necessary to heat the

core to temperatures in the range of 1000 to 1100 °F. At these

Page 63: +Peters Ekwere j. - Petrophysics

2–20

temperatures, the water of crystallization (hydration) of the rock is driven

off, resulting in an estimated water saturation that is higher than the

true interstitial (connate) water saturation. The second disadvantage is

that the oil when heated to high temperatures has a tendency to crack

and coke. This cracking and coking tend to reduce the oil volume

resulting in an oil saturation that is less than the true value.

Corrections can be made to the retort measurements to make them more

accurate.

Another method of direct saturation measurement is by extraction

with a solvent. This is accomplished in a Dean-Stark distillation

apparatus (Figure 2.12). The core is placed in the apparatus in such a

way that the vapor from a solvent (e.g., toluene) rises through the core

and is condensed back over the sample. This process leaches out the oil

and water from the sample. The water and solvent are condensed and

trapped in a graduated receiver. The water settles to the bottom of the

receiver while the solvent refluxes back into the main heating vessel. The

extraction is continued until no more water is collected in the receiving

vessel. The water saturation is calculated directly from the volume of

water expelled from the sample. The oil saturation is calculated

indirectly from the weight of the saturated sample before distillation, the

weight of the dry sample after distillation and the weight of the water

expelled from the sample. Again, the gas saturation is calculated

indirectly from the requirement that the saturations must sum to one.

To ensure that all the oil has been removed from the sample, the

sample may be transferred from the Dean-Stark apparatus to a Soxhlet

extractor for further extraction (Figure 2.13). The Soxhlet extractor is

similar to the Dean-Stark apparatus except that there is no provision for

trapping the extracted liquids.

Page 64: +Peters Ekwere j. - Petrophysics

2-21

The saturations determined by direct measurements on cores

should be treated with caution because they may not represent the in-

situ fluid saturations for several reasons. If the core was cut with a

water-based drilling mud, it would have been flushed by mud filtrate

resulting in a higher water saturation than the original, undisturbed

formation water saturation. The measured oil saturation in this case

would be the residual oil saturation after waterflooding, which is less

than the original in-situ oil saturation. If the core was cut with an oil-

based mud, the water saturation obtained by direct measurement will be

essentially the correct original water saturation, if it was at irreducible

level. If the original in-situ water saturation was not at the irreducible

level, then the oil mud filtrate could potentially displace some of the

water making the laboratory measured water saturation to be too low.

Figure 2.12. Dean-Stark apparatus.

Page 65: +Peters Ekwere j. - Petrophysics

2–22

Figure 2.13. Soxhlet extractor.

As the core is brought from the high pressure and temperature of

the reservoir to the low pressure and temperature of the laboratory,

changes occur in the fluid saturations which can make them

considerably different from the original in-situ saturations. The free gas,

if present, will expand, expelling water and oil in the process. Solution

gas will evolve from the oil, expand and further reduce the oil and water

volumes. The evolution of solution gas causes the oil to "shrink". These

changes cause the saturations determined by core analysis to be

different from the in-situ saturations. In particular, the changes cause

gas saturation to be excessive even when there was no free gas

saturation at the original in-situ conditions.

Page 66: +Peters Ekwere j. - Petrophysics

2-23

Although the saturations determined by direct measurements on

cores may not reflect the true in-situ saturations, they do provide useful

information about the reservoir. The saturation measurements can be

used to approximately locate the gas-oil and water-oil contacts in the

reservoir if they are present.

Fluid saturations, So, Sw and Sg, only tell us the proportion of

each fluid type in the pore space. They do not tell us how the fluids are

distributed in the rock. To determine the fluid distribution, we need to

consider the interfacial forces and phenomena that arise when

immiscible fluids are confined in reservoir pores of capillary dimensions.

The important interfacial forces and phenomena include surface tension,

interfacial tension, wettability, capillarity and capillary pressure (see

Chapters 6 and 7).

Table 2.2 shows the data obtained in an example core analysis

from a hydrocarbon bearing formation from a depth of 4805.5 to 4851.5

feet. The table shows the depth, core permeability, core porosity, oil

saturation, water saturation and gas saturation as determined in the

laboratory. Although the fluid saturations are not the true in-situ

saturations, nevertheless they provide useful information. Figure 2.14

shows the saturation distributions from the core data. One can easily see

a water bearing zone at the bottom where the measured water saturation

is very high, an oil bearing zone above it, and a gas cap on top of the oil

zone. A gas oil contact exists at 4828.5 ft, and a water oil contact exists

at 4848.5 ft. Note the misleading gas saturation below the gas oil

contact. There was no free gas saturation below the gas oil contact at in-

situ conditions.

Page 67: +Peters Ekwere j. - Petrophysics

2–24

Table 2.2: Core Analysis Data

Depth k φ So Sw Sg (ft) (md

) (%) (%) (%) (%)

4805.5 0 7.5 0.0 68.0

32.0

4806.5 0 12.3

0.0 78.0

22.0

4807.5 2.5 17.0

0.0 43.0

57.0

4808.5 59 20.7

0.0 29.0

71.0

4809.5 221 19.1

0.0 31.4

68.6

4810.5 211 20.4

0.0 38.7

61.3

4811.5 275 23.3

0.0 34.7

65.3

4812.5 384 24.0

0.0 26.2

73.8

4813.5 108 23.3

0.0 30.9

69.1

4814.5 147 16.1

0.0 29.2

70.8

4815.5 290 17.2

0.0 34.3

65.7

4816.5 170 15.3

0.0 24.2

75.8

4817.5 278 15.9

0.0 26.4

73.6

4818.5 238 18.6

0.0 39.8

60.2

4819.5 167 16.2

0.0 39.5

60.5

4820.5 304 20.0

0.0 38.0

62.0

4821.5 98 16.9

0.0 34.3

65.7

4822.5 191 18.1

0.0 34.8

65.2

4823.5 266 20.3

0.0 31.1

68.9

4824.5 40 15.3

0.0 22.9

77.1

4825.5 260 15.1

0.0 13.9

86.1

4826.5 179 14.0

0.0 21.4

78.6

4827.5 312 15.6

0.0 28.8

71.2

Page 68: +Peters Ekwere j. - Petrophysics

2-25

4828.5 272 15.5

0.0 34.8

65.2

4829.5 395 19.4

6.2 25.3

68.5

4830.5 405 17.5

13.1

25.7

61.2

4831.5 275 16.4

17.7

22.5

59.8

4832.5 852 17.2

19.8

19.2

61.0

4833.5 610 15.5

21.9

21.3

56.8

4834.5 406 20.2

16.3

22.3

61.4

4835.5 535 18.3

19.7

24.6

55.7

4836.5 663 19.6

19.4

16.3

64.3

4837.5 597 17.7

17.5

19.8

62.7

4838.5 434 20.0

14.0

27.5

58.5

4839.5 339 16.8

20.8

19.7

59.5

4840.5 216 13.3

18.1

23.3

58.6

4841.5 332 18.0

15.6

15.6

68.8

4842.5 295 16.1

19.3

15.5

65.2

4843.5 882 15.1

19.2

21.2

59.6

4844.5 600 18.0

20.6

22.2

57.2

4845.5 407 15.7

15.3

13.4

71.3

4847.5 479 17.

8 20.8

14.6

64.6

4848.5 0 9.2 14.1

8.7 77.2

4849.5 139 20.5

0.0 77.1

22.9

4850.5 135 8.4 0.0 57.2

42.8

4851.5 0 1.1 0.0 63.6

36.4

Page 69: +Peters Ekwere j. - Petrophysics

2–26

Figure 2.14. Saturation distributions from core analysis data.

Other useful observations can be made from the core

analysis data. The low residual oil saturation of about 20% indicates a

light oil reservoir in contrast to a heavy (more viscous) oil reservoir in

which the residual oil saturation would be much higher than 20%. All of

Page 70: +Peters Ekwere j. - Petrophysics

2-27

the measured properties vary with depth, which is an indication that the

reservoir is heterogeneous in nature. The porosity and permeability

distributions are shown in Figure 2.15. Variability of reservoir properties

is pervasive. Not only do the properties vary along the well depth, they

also vary laterally away from the well. The characterization of this

variability and the estimation of the properties at unmeasured locations

are the subjects addressed by geostatistics (see Chapter 4).

2.7 INDIRECT POROSITY MEASUREMENT FROM WELL LOGS

2.7.1 Introduction to Well Logging

In-situ porosity cannot be measured directly in the field as in the

laboratory. Therefore, only indirect measurements are made through well

logging. These measurements use either sonic energy or some form of

induced or applied radiation. Most log evaluation is concerned primarily

with determining in-situ porosity and water saturation. Neither in-situ

water saturation nor hydrocarbon saturation can be measured directly in

the wellbore. However, it is possible to infer the water saturation if the

porosity is known by measuring the resistivity of the formation.

Therefore, in this section, porosity and resistivity logs are discussed.

Care should always be taken in comparing core versus log-derived

porosities, particularly in rocks that have been highly affected by

diagenesis. Logs measure average porosities over a much larger volume

than conventional laboratory core analysis. Also, a laboratory core has

been relieved of the overburden and lateral stresses and because it is an

elastic medium, it will expand. Since the minerals have very low

coefficient of expansion, the increase in volume must be due almost

solely to the increase in porosity. Thus, the porosity measured in the

Page 71: +Peters Ekwere j. - Petrophysics

2–28

laboratory at ambient conditions may be expected to be higher than at in

situ conditions.

Figure 2.15. Porosity and permeability distributions from core analysis

data.

Page 72: +Peters Ekwere j. - Petrophysics

2-29

2.7.2 Mud Filtrate Invasion

Well log measurements are made in the borehole after the well has

been drilled. The drilling operation alters the formation characteristics

near the wellbore where the log measurements are made. In order to

interpret the logs, it is necessary to understand the changes that have

occurred in the formation caused by the drilling mud.

Drilling mud is a complex liquid usually composed of mainly water

(for water-based muds) and suspended solids and various chemicals that

control the mud properties (viscosity, fluid loss, pH and others). Clays

(bentonite) are added to give the mud viscosity and weighting material

(barite) is added to increase the mud density above that of water.

The mud is circulated during drilling to lift the cuttings out of the

borehole. Another important function of the mud is to exert a

backpressure on the formation to prevent the well from "kicking" during

the drilling operations. In general, during drilling, the pressure in the

mud column in the borehole is higher than the formation pressure.

If we take a mud sample and place it in a mud press as is typically

done in mud testing, we can separate the mud into its two main

components: mud filtrate and mudcake. Mud filtrate is a clear liquid

whose salinity varies according to the source of the water used to mix the

mud and the chemical nature of the additives. Usually, the filtrate

salinity is lower than the formation water salinity. Since the filtrate is

clear (no suspended solids), it can invade the formation if the pressure in

the wellbore is greater than the formation pressure, which is the case

during overbalance drilling. The mud filtrate can displace some of the

original formation fluids away from the wellbore into the formation.

Page 73: +Peters Ekwere j. - Petrophysics

2–30

The mudcake seals off the formation from further invasion by the

mud filtrate. The presence of the mudcake can be detected by the logging

tool and will cause the borehole diameter to be smaller than the bit

diameter. It is an indication of invasion and, indirectly, of permeability.

Since the mudcake is a solid, it will not normally invade the formation.

Drilling mud itself usually cannot invade the formation because it

contains a lot of suspended solids. However, whole mud can be lost into

the formation if the formation is inadvertently fractured. This is lost

circulation, which should not be confused with mud filtrate invasion. The

mudcake has a very low permeability and as a result, controls the

volume of filtrate invasion. The depth of invasion is determined by the

porosity of the formation. The depth of invasion will be greater in a low

porosity formation than in a high porosity formation everything else

being equal.

Figure 2.16 shows a schematic diagram of the undisturbed

formation and the altered formation after it has been penetrated by the

bit. The left figure shows the undisturbed formation before it was drilled.

Let us assume that this is a sandstone formation bounded above and

below by impermeable shales. The bottom of the sand has water

saturation, Sw, of 100% and the upper section of the sand is at an

irreducible water saturation, Swirr. Irreducible water saturation is that

saturation at which the water cannot be produced during normal

production operations. The water occupies the lower portion of the

formation because it is denser than oil which floats to the top of the

formation. There is a transition zone in which the water saturation

decreases from 100% at the bottom to 25% at the top over a finite length

of the formation. This transition zone is caused by capillarity (see

Page 74: +Peters Ekwere j. - Petrophysics

2-31

Chapter 7). It should be pointed out that not all hydrocarbon bearing

zones contain irreducible water. Some can contain mobile water.

Figure 2.16. Schematic diagram of mud filtrate invasion. Undisturbed formation is at left and invaded formation is at right.

On the right is the same formation after it has been drilled. Here,

invasion has occurred. Because of the higher pressure in the borehole

than the formation, the formation acts as a mud press and separates the

mud into mudcake which plasters the borehole wall, and mud filtrate,

which invades the formation. In order to invade the formation, the fluids

that were originally there must be displaced. The filtrate flushes or

Page 75: +Peters Ekwere j. - Petrophysics

2–32

displaces the fluids deeper into the formation and takes their place near

the wellbore.

In the bottom of the formation where Sw = 100%, the flushing is

nearly complete because the formation water is being displaced by water

which is different only in salinity and is miscible with it. The salinity of

any formation water left behind will soon reach equilibrium with the

filtrate because of diffusion.

In the upper part of the formation, we had an initially large oil

saturation (Soi = 1-Swirr). Although most of the formation water has been

displaced by mud filtrate, residual oil remains in the flushed zone with a

saturation, Sor, because an immiscible displacement can never be

complete.

The filtrate has flushed out all the original fluids that it can flush

out to a certain depth. This depth is called the flushed zone. The flushed

zone water saturation is designated Sxo and the flushed zone diameter is

designated dxo. The resistivity of the water in the flushed zone is Rmf, the

resitivity of the mud filtrate, and the resistivity of the flushed zone

formation is Rxo. If we go a little deeper into the formation, we will find a

transition zone, which contains a mixture of formation fluids and mud

filtrate. The zone, from the borehole wall to the end of the mud filtrate is

the invaded zone and includes the flushed zone. The diameter of the

invaded zone is designated as di, the water saturation is Si, and the

formation resisitivity Ri. Note that in the flushed zone, Si = Sxo and Ri =

Rxo. Since the water between dxo and di is a mixture of formation water

and filtrate, it is not possible to measure a single value for Ri in this zone.

Finally, as we pass the invaded zone, we return to the undisturbed or

uncontaminated formation. This is the virgin zone. The conditions in this

Page 76: +Peters Ekwere j. - Petrophysics

2-33

zone are the same as at the left side of Figure 2.16. The resistivity of this

undisturbed zone is Rt, the true formation resisitivity, which we ideally

would like to measure for estimating the undisturbed water saturation,

Sw.

Figure 2.17 gives the same information about the fluid

distributions but in a different format. On the left is a plot of Sw versus

depth for the undisturbed formation. At the bottom, Sw = 100% and is

constant for about 30 ft from the bottom. We then enter the transition

zone, where Sw changes with depth until it reaches an irreducible water

saturation of about 25%. Sw is constant for the last 20 ft or so at Sw =

Swirr. Note that the hydrocarbon saturation, Sh is equal to (1-Sw).

Figure 2.17. Variation of water saturation with distance from the borehole.

Page 77: +Peters Ekwere j. - Petrophysics

2–34

On the right are three sections drawn horizontally through the

formation to show how Sw varies with distance from the borehole at three

depths. At the bottom section, Sxo = Si = Sw = 100%. This is because there

was never any hydrocarbon in this section of the formation and as a

result, Sh = 0.

The depth in the middle section was chosen in the transition zone

where Sw was about 40%. We can see a change in the various water

saturations because oil is present in this zone and some of it has been

displaced by mud filtrate. Because of the residual oil, Sxo is less than

100% (Sxo = 1-Sor). Si will be lower than Sxo because some of the

hydrocarbon that was originally in the flushed zone has been displaced

into this zone, and Sw in the uncontaminated zone will be 40%.

In the uppermost section, Sw is at its irreducible saturation value

for the undisturbed formation. We see the maximum variation in the

various saturations after invasion. Sxo will be lower than the Sxo in the

transition zone, and (1-Sxo) will be close to the residual oil saturation, Sor.

The water saturation will vary throughout between dxo and di. Beyond di,

Sw = Swirr = 25%. In this uppermost section, we clearly see the flushed

zone diameter, where Sxo is constant, and the end of the invaded zone,

where Sw becomes constant.

Most of the difficulties in log evaluation and the proliferation of

many tool configurations are caused by the presence of mud invasion,

usually of unknown depth, in the logging environment. Figure 2.18

shows a schematic diagram of the borehole condition for logging

measurements.

Page 78: +Peters Ekwere j. - Petrophysics

2-35

Figure 2.18. Borehole conditions for well logging (Courtesy of Schlumberger).

Page 79: +Peters Ekwere j. - Petrophysics

2–36

2.7.3 Porosity Logs

Conventional logging techniques for measuring porosity are the

Density, Neutron and Sonic logs. All of these logs provide an indication

of total porosity.

Density Log

The Density log measures the electron density of the formation by

using a pad mounted chemical source of gamma radiation and two

shielded gamma detectors (Figure 2.19). The medium-energy gamma rays

emitted into the formation collide with electrons in the formation. At each

collision, a gamma ray loses some, but not all, of its energy to the

electron and then continues with reduced energy. This type of interaction

is known as Compton scattering. The scattered gamma rays reaching the

detector, at a fixed distance from the source, are counted as an

indication of the formation density.

Figure 2.19. Schematic of density logging tool.

Page 80: +Peters Ekwere j. - Petrophysics

2-37

The number of Compton scattering collisions is related directly to

the number of electrons in the formation. Therefore, the response of the

density tool is determined essentially by the electron density (the number

of electrons per cubic centimeter) of the formation. Electron density is

related to the true bulk density in gm/cc, which in turn depends on the

density of the rock matrix, the formation porosity and the density of the

pore fluids.

For a pure element, the electron density index, which is

proportional to the electron density is defined as

2e b

ZA

ρ ρ ⎛ ⎞= ⎜ ⎟⎝ ⎠

(2.11)

where ρe is the electron density index, ρb is the bulk density, Z is the

atomic number of the element and A is the atomic weight of the element.

For a molecule, the electron density index is given by

2 i

e b

ZM

ρ ρ⎛ ⎞

= ⎜ ⎟⎜ ⎟⎝ ⎠

∑ (2.12)

where M is the molecular weight and iZ∑ is the sum of the atomic

numbers of the atoms making up the molecule, which is equal to the

number of electrons per molecule. For most materials encountered in the

formation, the quantities in brackets in Eqs.(2.11) and (2.12) are

approximately equal to unity as shown in Tables 2.3 and 2.4. The density

tool is calibrated in a fresh water filled limestone formation of high purity

to give an apparent density that is related to the electron density index

by

Page 81: +Peters Ekwere j. - Petrophysics

2–38

1.0704 0.1883a eρ ρ= − (2.13)

For liquid filled sandstones, limestones and dolomites, the apparent

density read by the tool is practically equal to bulk density of the

formation.

The bulk density of a clean formation is given by

( )1b f mρ φρ ρ φ= + − (2.14)

Eq.(2.14) can be solved for the porosity as

m b

m f

ρ ρφρ ρ

−=

− (2.15)

To calculate porosity from Eq.(2.15), the matrix and fluid densities must

be known or assumed. The depth of investigation of the density log is

relatively shallow. Therefore, in most permeable formations, the pore

fluid is the drilling mud filtrate, along with any residual hydrocarbons.

Usually, the fluid density is assumed to be 1.0 gm/cc. When residual

hydrocarbon saturations are fairly high, this can cause the calculated

porosity values to be greater than the true porosity, and should be

corrected for this effect. Table 2.4 gives the densities of various rock

matrices.

Figure 2.20 shows a typical presentation of a density log. Track 1

shows the Gamma Ray log, which measures the natural gamma

radiation of the formation. Radioactive elements such as uranium,

potassium and thorium tend to occur more in shales than in sands. As a

result, the Gamma Ray log is a lithology log that identifies shales from

sands. The caliper in the same track measures the borehole diameter.

Page 82: +Peters Ekwere j. - Petrophysics

2-39

The formation density and the porosity derived from it are presented in

Tracks 2 and 3. Also shown is the correction or compensation applied to

account for mud cake effect and small borehole irregularities.

Table 2.3: Atomic Properties of Common Elements in the Formation

Elemen

t

A

Z 2 Z

A⎛ ⎞⎜ ⎟⎝ ⎠

H 1.008 1 1.984

1

C 12.01

1

6 0.999

1

O 16.00

0

8 1.000

0

Na 22.99 11 0.956

9

Mg 24.32 12 0.986

8

Al 26.98 13 0.963

7

Si 28.09 14 0.996

8

S 32.07 16 0.997

8

Cl 35.46 17 0.958

8

K 39.10 19 0.971

9

Ca 40.08 20 0.998

0

Page 83: +Peters Ekwere j. - Petrophysics

2–40

Table 2.4: Densities of Rock Formations and Fluids

Compound

Formula

ρb (gm/cc)

2 iZM

⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠

∑ ρe

(gm/cc) ρa

(gm/cc)

Quartz SiO2 2.654 0.9985 2.650 2.648

Calcite CaCO3 2.710 0.9991 2.708 2.710

Dolomite CaMg(CO3)2 2.870 0.9977 2.863 2.876

Anhydrite CaSO4 2.960 0.9990 2.957 2.977

Sylvite KCl 1.984 0.9657 1.916 1.863

Halite NaCl 2.165 0.9581 2.074 2.032

Gypsum CaSO4.H2O 2.320 1.0222 2.372 2.351

Fresh Water H2O 1.000 1.1101 1.110 1.00

Salt Water 200,000

ppm

1.146 1.0797 1.237 1.135

“Oil” N(CH2) 0.850 1.1407 0.970 0.850

Methane CH4 ρmeth 1.247 1.247ρmet

h

1.335ρmeth-

0.188

“Gas” C1.1H4.2 ρgas 1.238 1.238ρgas 1.325ρgas -0.188

Page 84: +Peters Ekwere j. - Petrophysics

2-41

Figure 2.20. Presentation of density log.

Page 85: +Peters Ekwere j. - Petrophysics

2–42

Sonic Log (Acoustic Log)

The Sonic log measures the time, Δt, required for compressional

sound wave to traverse one foot of formation. Known as the interval

transit time, Δt is the reciprocal of the velocity of the compressional

sound wave. To avoid fractions, the interval transit time is scaled by 106

and reported in micro-seconds per ft (μsec/ft). Thus,

610t

vΔ = (2.16)

where Δt is the interval transit time in μsec/ft and v is the compressional

wave velocity in ft/s.

The sonic tool contains a transmitter and two receivers (Figure

2.21). When the transmitter is energized, the sound wave enters the

formation from the mud column, travels through the formation and back

to the receivers through the mud column. The difference between the

arrival times at the two receivers divided by the distance between the

receivers gives the interval transit time. The speed of sound in the tool

and the drilling mud is less than that in the formation. Accordingly, the

first arrival of sound energy at the receivers corresponds to sound travel

paths in the formation near the borehole. The logging tool has circuits to

compensate for hole size changes or any tilting of the tool in the hole.

The interval transit time in a formation depends upon lithology and

porosity. In general, the more dense or consolidated a formation, the

lower the interval transit time. An increase in travel time indicates an

increase in porosity. Based on laboratory measurements, Wyllie (1956)

concluded that in clean and consolidated formations with uniformly

Page 86: +Peters Ekwere j. - Petrophysics

2-43

distributed small pores, there is a linear relationship between porosity

and interval transit time as follows:

( )1f mt t tφ φΔ = Δ + − Δ (2.17)

where Δt is the interval transit time measured by the log, Δtf is the

interval transit time in the pore fluid, Δtm is the interval transit time in

the rock matrix, and φ is the formation porosity.

Figure 2.21. Schematic of sonic logging tool.

Page 87: +Peters Ekwere j. - Petrophysics

2–44

Eq.(2.17) can be solved for porosity as

m

f m

t tt t

φ Δ − Δ=

Δ − Δ (2.18)

To calculate porosity from Eq.(2.18), the transit times for the rock matrix

and the pore fluid must be known or assumed. Table 2.5 gives the sonic

speeds and interval transit times for common rock matrices. The depth of

investigation of the sonic log is relatively shallow. Thus, the pore fluid is

usually assumed to be mud filtrate with an interval transit time of 189

μsec/ft, corresponding to a fluid velocity of 5300 ft/sec.

If any shale laminae exist in the sandstone, the apparent sonic

porosity is increased by an amount proportional to the bulk volume

fraction of such laminae. The interval transit time is increased because

transit time for shale generally exceeds that of the matrix.

In carbonates having intergranular porosity, Wyllie’s average

formula still applies. But sometimes, the pore structure and pore size

distribution are significantly different from sandstones. Also, there is

often some secondary porosity such as vugs and fractures with much

larger dimensions than the pores of the primary porosity. In vuggy

formations, according to Wyllie, the velocity of sound depends mostly on

the primary porosity, and the porosity derived from the sonic reading

through the time average formula will tend to be too low by an amount

approaching the secondary porosity.

Direct application of the Wyllie formula to unconsolidated and

insufficiently compacted sands gives porosity values that are too high.

When shale transit time exceeds 100 μsec/ft, which is an indication of

Page 88: +Peters Ekwere j. - Petrophysics

2-45

undercompaction, then the compaction correction should be made to

obtain more accurate porosity values. This is accomplished by applying

an empirical correction factor as shown in Eq.(2.19):

100m

f m sh

t tt t t

φ⎛ ⎞⎛ ⎞Δ − Δ

= ⎜ ⎟⎜ ⎟⎜ ⎟Δ − Δ Δ⎝ ⎠⎝ ⎠ (2.19)

where Δtsh is the interval transit time in the adjacent shale. Figure 2.22

shows a typical presentation of the sonic log.

Table 2.5: Sonic Speed and Interval Transit Time for Rock Formations

vm (ft/sec)

Δtm (μsec/ft)

Δtm (μsec/ft)

(commonly used)

Sandstones 18,000 – 19,000 55.5 – 51.0 55.5 or 51.0

Limestones 21,000 – 23,000 47.6 – 43.5 47.5

Dolomites 23,000 43.5 43.5

Anhydrite 20,000 50.0 50.0

Salt 15,000 66.7 67.0

Casing (iron) 17,500 57.0 57.0

Page 89: +Peters Ekwere j. - Petrophysics

2–46

Figure 2.22. Presentation of sonic log.

Neutron Log

The Neutron log measures induced formation radiation produced

by bombarding the formation with fast moving neutrons (Figure 2.23).

Page 90: +Peters Ekwere j. - Petrophysics

2-47

The tool responds primarily to the hydrogen present in the formation.

Thus, in clean formations, whose pores are filled with water or oil, the

neutron log reflects the amount of liquid-filled porosity.

Neutrons are electrically neutral particles with a mass almost

identical to the mass of a hydrogen atom. High-energy (fast) neutrons are

continuously emitted from a radioactive source mounted in the logging

tool. These neutrons collide with the nuclei of the formation materials.

With each collision, a neutron loses some of its energy. The

Figure 2.23. Schematic of neutron logging tool.

amount of energy lost per collision depends on the relative mass of the

nucleus with which the neutron collides. The greatest energy loss occurs

when the neutron collides with a nucleus of practically equal mass, i.e.,

Page 91: +Peters Ekwere j. - Petrophysics

2–48

hydrogen. Collisions with heavy nuclei do not slow the neutrons down

very much. Thus, the slowing down of neutrons depends largely on the

amount of hydrogen in the formation.

Within a few microseconds, the neutrons have been slowed down

by successive collisions to thermal velocities, corresponding to energies

of around 0.025 electron volt (eV). They then diffuse randomly, without

losing any more energy, until they are captured by the nuclei of atoms

such as chlorine, hydrogen, silicon and others. The capturing nucleus

becomes intensely excited and emits a high-energy gamma ray of

capture. Depending on the type of Neutron logging tool, either these

capture gamma rays or the slow neutrons themselves are counted by a

detector in the tool.

When the hydrogen concentration of the material surrounding the

neutron source is large, most of the neutrons are slowed down and

captured within a short distance from the source. However, if the

hydrogen concentration is small, the neutrons travel farther from the

source before they are captured. Accordingly, the counting rate at the

detector increases for decreased hydrogen concentration and decreases

for increased hydrogen concentration. The porosity based on the neutron

count is given by

logN a b φ= − (2.20)

where N is the slow neutrons counted, a and b are empirical constants

determined by appropriate calibration and φ is the porosity.

Since there is very little difference in the concentration of hydrogen

in oil or water, neutron logs measure the liquid filled porosity. A high

Page 92: +Peters Ekwere j. - Petrophysics

2-49

neutron counting rate indicates low porosity and a low neutron counting

rate indicates high porosity. Figure 2.24 shows a typical presentation of

the Neutron log. The neutron count is presented in API (American

Petroleum Institute) units. The porosity is in neutron porosity units

based on calibration with limestone or sandstone.

Two additional factors should be considered in the interpretation of

neutron logs. First, shales and zones containing a significant amount of

shale, will indicate a high neutron porosity due to the bound water

associated with the shale. Secondly, because of the lower concentration

of hydrogen in gas than in oil or water, a zone containing gas will

indicate a neutron porosity that is lower than it should be. These

features are really an advantage since a comparison of the neutron

porosity to cores and other porosity logs provides a convenient method

for determining shale volumes and for distinguishing gas zones from oil

or water zones. In a gas zone, the fluid density is very much lower than

the 1.0 gm/cc used in Eq.(2.15) to calculate the density porosity. As a

result, the density porosity in a gas zone is higher than it should be.

Thus, in a gas zone, the neutron porosity is too low and the density

porosity is too high. When the two porosity logs are superimposed, the

two curves will agree is shales and in liquid zones and will cross over in

gas zones. This cross over of the two logs can be used to identify gas

bearing zones as shown in Figure 2.25.

Page 93: +Peters Ekwere j. - Petrophysics

2–50

Figure 2.24. Presentation of Neutron log.

Page 94: +Peters Ekwere j. - Petrophysics

2-51

Figure 2.25. A comparison of neutron and density porosities. Shaded areas indicate gas zones.

Page 95: +Peters Ekwere j. - Petrophysics

2–52

Combination Porosity Logs

In many areas, it is common practice to record more than one

porosity log on a well. Common combinations are Density-Neutron,

Density-Sonic and Sonic-Neutron. Sometimes, all three logs are run in

the same well. These logs are usually recorded along with a Gamma Ray

curve and a Caliper.

Combination porosity logs are used to (1) differentiate oil or water

from gas zones, (2) calculate quantitative values for lithology, and (3)

determine volume of shale in the rock matrix. Figure 2.26 shows a

section of the three porosity logs run in the same well.

2.7.4 Resistivity Log

Resistivity is one of the most useful physical properties measured

in the borehole. Formation resistivity measurements, in conjunction with

porosity and water resistivity, are used to obtain values of water

saturation and consequently, hydrocarbon saturation. They are also

used in conjunction with lithology logs to identify hydrocarbon bearing

intervals and to estimate the net pay thickness.

Resistivity is the degree to which a substance “resists” or impedes

the flow of electrical current. It is a physical property of the material,

independent of size and shape. In well logging, both resistivity and

conductivity are used frequently. One is the reciprocal of the other. Thus,

1ResistivityConductivity

= (2.21)

Page 96: +Peters Ekwere j. - Petrophysics

2-53

Figure 2.26. A comparison of the three porosity logs in the same formation.

Page 97: +Peters Ekwere j. - Petrophysics

2–54

Low resistivity corresponds to high conductivity and high resistivity

corresponds to low conductivity. The resistivity unit used in well logging

is ohm-meter2/meter, which is usually shortened to ohm-meter.

Electrical conductivity is expressed in mhos per meter. In order to avoid

decimal fractions, in electrical logging, it is expressed in millimhos per

meter.

In reservoir rocks, the sedimentary minerals that make up the

formation matrix are non-conductors. Also, hydrocarbons such as gas

and oil are non-conductors. Therefore, current flow in sedimentary rocks

is associated with the water in the pore space. Most of the waters

encountered in well logging contain some sodium chloride (NaCl) in

solution. The current then is carried by the ions of the salt, which is

dissolved in the water. Therefore, conductivity is proportional to the salt

concentration (salinity) of the water. Although each ion is capable of

carrying only a definite quantity of charge, as the formation temperature

is increased, these ions are capable of moving faster. This results in

increased conductivity. Figure 2.27 shows the variation of water

resistivity with temperature at various salinities.

The amount of water contained in the formation is directly related

to the porosity and, also, affects the formation resistivity. As the volume

of water increases, the capacity for ions increases and the conductivity

increases. Thus, the formation resistivity is affected by (1) salt

concentration in the water (salinity), (2) reservoir temperature, (3) water

volume (porosity) and (4) hydrocarbon content. Thus, although we

cannot directly measure the amount of hydrocarbon in a formation, we

can infer the hydrocarbon content from resistivity measurements.

Page 98: +Peters Ekwere j. - Petrophysics

2-55

Figure 2.27. Variation of water resistivity with temperature and salinity.

Let Rw be the resistivity of the formation water, Ro be the

resisitivity of the formation saturated 100% by the formation water of

resistivity Rw and Rt be the true resistivity of the formation partially

saturated with water of resistivity Rw and hydrocarbon. Based on

laboratory measurements, Archie (1953) found that Ro was directly

proportional to Rw for clean, consolidated sandstone cores for a fixed

porosity. Thus,

or oo w

w

RR FR FR

= = (2.22)

where F is a constant of proportionality at a given porosity known as the

formation resistivity factor. He further found that F could be related to

the porosity of the core by an equation of the form

Page 99: +Peters Ekwere j. - Petrophysics

2–56

m

aFφ

= (2.23)

where a is an empirical constant and m is a cementation factor that

varies from 1.3 for unconsolidated sands to 2.5 for consolidated

sandstones. Thus, for clean sandstone cores of varied porosity,

ln ln lnF a m φ= − (2.24)

Eq.(2.24) suggests that a graph of lnF versus lnφ should be linear. Figure

2.28 shows such a graph for clean sandstone cores. For these data, a =

1.10 and m = 1.73.

y = 1.1016x-1.7332

R2 = 0.8648

1

10

100

1000

0.01 0.1 1

Porosity

Form

atio

n R

esis

tivity

Fac

tor

Figure 2.28. Log-Log graph of formation resistivity factor versus porosity for various water resistivities.

Page 100: +Peters Ekwere j. - Petrophysics

2-57

For the cores used in his measurements, Archie found a to be

approximately 1.0. Therefore, he proposed the following relationship

between formation resistivity factor and porosity

1mF

φ= (2.25)

Others performed similar measurements using their own core samples. A

group in Humble Oil (Now ExxonMobil) performed similar measurements

and found that their data were best fitted by an equation of the form

2.15

0.62o

w

RFR φ

= = (2.26)

Eq. (2.26) is known as the Humble formula and is still widely used in the

petroleum industry.

Archie also conducted resistivity measurements in partially

saturated cores to measure Rt at various water saturations, Sw. He

defined formation resistivity index as

t

o

RIR

= (2.27)

He found that the formation resistivity index, I, was related to the water

saturation by an equation of the form

1t

no w

RIR S

= = (2.28)

Page 101: +Peters Ekwere j. - Petrophysics

2–58

where n is the water saturation exponent. Eq.(2.28) can be solved for the

water saturation as

1 o wn n nw

t t

R FRSI R R

= = = (2.29)

Figure 2.29 shows lnI versus lnSw for a Berea sandstone core. For this

sample, n = 2.27. Archie found n to be approximately 2 for the core

samples in his study. The hydrocarbon saturation is given by

1h wS S= − (2.30)

Figure 2.30 shows an example invasion profile for resistivity

measurements. This profile is analogous to the water saturation profile of

Figure 2.17.

Page 102: +Peters Ekwere j. - Petrophysics

2-59

Figure 2.29. Resistivity index for Berea sandstone core

Page 103: +Peters Ekwere j. - Petrophysics

2–60

Figure 2.30. Example invasion profile for resistivity logs.

Three types of logging tools are used to measure formation

resistivity: Induction tools, focused resistivity tools and unfocused

resistivity tools. These tools can be further subdivided into those that

measure a very small volume of the formation (microresistivity logs) and

those that measure a relatively large volume of the formation. Table 2.6

presents a summary of the various resistivity tools and their limitations.

Page 104: +Peters Ekwere j. - Petrophysics

2-61

Table 2.6: Resistivity Tools

Electric Log

The Electric log was the basic and most frequently used log until

the mid 1950's. This log was invented and developed by two French

brothers, Conrad and Marcel Schlumberger. Figure 2.31 shows the

presentation of an electric log. It consists of a Spontaneous Potential (SP)

curve in Track 1 and a combination of resistivity curves designated as

normal and lateral depending on the electrode arrangements.

The normal curve is produced by two effective electrodes downhole,

a current electrode and a pickup electrode as shown in Figure 2.32.

Resistivity values are measured by recording the voltage drop across

these electrodes. A short normal, with electrode spacing of 18 inches, is

used for correlation to define bed boundaries, and to measure the

Page 105: +Peters Ekwere j. - Petrophysics

2–62

resistivity near the wellbore. Normal curves have a radius of investigation

of approximately twice the electrode spacing.

Figure 2.31. Presentation of an electric log.

Page 106: +Peters Ekwere j. - Petrophysics

2-63

The lateral curve is produced by three effective electrodes, one

current and two pickup electrodes (Figure 2.32). The radius of

investigation is approximately equal to the electrode spacing, which is

the distance from the current electrode to the midpoint between the two

pickup electrodes. The spacing is usually in the range of 16 to 19 feet.

Lateral curves are nonsymmetrical and highly distorted by adjacent beds

and thin beds, but are effective in measuring true resistivity in thick

homogeneous formations.

Figure 2.32. Schematic of electric logging tool: (A) Normal curve, (B)

Lateral curve.

Page 107: +Peters Ekwere j. - Petrophysics

2–64

Induction-Electric Log

The Induction-Electric log is a combination of electric log curves

with induction curves. The induction tool was developed to provide a

means of logging wells drilled with oil-based (nonconductive) muds. All

the original electric logging tools used the mud column to conduct the

current into the formation, so they could not be run in nonconductive

muds or air-drilled holes. Although the induction tool was developed to

meet the need for a resistivity tool that could operate in a nonconductive

mud, it soon became recognized that the tool worked better than the

original electric log in freshwater muds. The induction curve was easier

to read than the electric log, and it read close to true formation resistivity

in formations where the resistivity was not over 200 ohm-meter and Rmf

was greater than Rw.

The induction tool works by the principle of electromagnetic

induction. A high-frequency alternating current flows through a

transmitter coil mounted on the logging tool (Figure 2.33). This current

sets up a high-frequency magnetic field around the tool, which extends

into the formation. The alternating magnetic field causes currents to flow

through the formation concentric with the axis of the induction tool. The

currents, called ground loops, are proportional to the conductivity of the

formation. They alternate at the same frequency as the magnetic field

and the transmitter current flowing through the induction coil. The

ground loop currents set up a magnetic field of their own. This secondary

magnetic field causes a current to flow in the receiver coil located in the

logging tool. The amount of current flowing in the receiver coil is

proportional to the ground loop currents and therefore to the

conductivity of the formation. The signal in the receiver coil is detected,

Page 108: +Peters Ekwere j. - Petrophysics

2-65

processed and recorded on the log as either a conductivity measurement

or a resistivity measurement.

The tool illustrated in Figure 2.33 is a simple two-coil device. In

practice, bucking coils are used to help focus the effects of the main

transmitter and receiver coils and to remove unwanted signals from the

borehole. One popular induction tool used today has six different coils.

The depth of investigation (the depth from which most of the

measurement is obtained) for a typical deep induction tool is about 10

feet. The vertical resolution (the thinnest bed that the tool will detect) is

40 inches. Both the depth of investigation and the vertical resolution are

affected by the spacing between the main transmitter and receiver coils

as well as by the placement of the focusing coils. By judicious selection of

these parameters, different depths of investigation can be designed into a

tool. Thus, one can measure the resistivity profile through the invaded

zone and correct the deep induction reading to move it close to the true

formation resistivity, Rt.

Figure 2.33. Schematic of induction logging tool.

Page 109: +Peters Ekwere j. - Petrophysics

2–66

Figure 2.34 shows a typical presentation of the induction electric

log (IEL). It includes an SP and/or Gamma Ray curve, 18" normal and

the induction curve on both the resistivity and conductivity scales. An

amplified 18" normal curve is often recorded in areas where low

resistivities are encountered.

For many years, the induction electric log was the most popular

induction tool in high-porosity formations such as in California, along

the US Gulf Coast and in other high-porosity, moderate-resisitivity

formations. A single induction curve with a vertical resolution of about 3

feet and a depth of investigation of about 10 feet was combined with

either a short normal curve or a shallow laterolog curve. Since mud

filtrate invasion is seldom deep in high-porosity formations, these two

curves, corrected for borehole and bed boundary effects, could be used to

determine Rt.

Dual Induction Laterolog

The Dual Induction Laterolog was developed for those areas that

had low porosities and deep invasion. The tool has two induction curves

(ILd and ILm) with a vertical resolution of about 40 inches. However, one

induction curve, the ILd, reads very deeply into the formation, while the

medium induction curve, ILm, reads only half as deep. A shallow-reading

laterolog combined with the two induction curves gives a good

description of the resistivity profile.

Page 110: +Peters Ekwere j. - Petrophysics

2-67

Figure 2.34. Presentation of induction-electric log.

Page 111: +Peters Ekwere j. - Petrophysics

2–68

Figure 2.35 shows a typical presentation of a dual induction

laterolog. An SP and/or Gamma Ray curve and three resistivity curves

having different depths of investigation are recorded. The shallow

laterolog measures the resistivity of the flushed zone, Rxo. The medium

induction curve, ILm, measures the combined resistivity of the flushed

and invaded zones, Ri. The deep induction curve responds primarily to

the resistivity of the uncontaminated zone, Rt. The resistivity curves may

be recorded on logarithmic or linear scales. The logarithmic presentation

permits a greater dynamic range for resistivities and is convenient for

determining ratios since the difference of two logarithms is equal to their

ratio.

The ratios of shallow to deep curves and medium to deep curves

are used to determine the diameter of invasion, di, the resistivity of the

flushed zone, Rxo, and the true formation resistivity, Rt. Figure 2.36

shows a typical "tornado chart" (so called because of its distinctive shape)

used to correct the dual induction log to obtain Rt.

Page 112: +Peters Ekwere j. - Petrophysics

2-69

Figure 2.35. Presentation of dual induction laterolog.

Page 113: +Peters Ekwere j. - Petrophysics

2–70

Figure 2.36. Tornado chart used to correct deep induction resistivity to true resistivity.

Page 114: +Peters Ekwere j. - Petrophysics

2-71

Focused Electric Log (Guard and Laterolog)

In boreholes which contain extremely saline drilling muds or very

high resistivity formations, the current that is emitted from a normal or

lateral electrode is almost entirely confined within the borehole and flows

up and down within the mud column. Very little, if any, of the current

penetrates surrounding resistive material. Under similar borehole

conditions, the induction logging tool is also adversely affected because

too much of the receiver voltage is derived from high conductivity of the

invaded zone. Focused-current logging tools have been designed to

overcome these problems in part.

There are two different focused-current logging systems, referred to

as the guard and laterolog, in use today (Figure 2.37). In the guard

system, guard electrodes are placed above and below a current electrode

and kept at the same potential to focus the formation current into a thin

disc, which flows perpendicularly to the borehole. The radius of

investigation is approximately three times the length of the guard

electrode.

The guard log defines bed boundaries very well and is affected very

little by adjacent bed resistivities. Shallow guard systems, utilizing short

guard electrodes (approximately 30 inches), are used with tools like the

Dual Induction for measuring the flushed zone resistivity, Rxo, or the

invaded zone resistivity, Ri. The longer guard (5 feet in length) systems

are used for measuring the true resistivity of the uncontaminated

formation, Rt. Figure 2.38 shows an example guard log presentation.

The laterolog electrode arrangement consists of a center current

electrode placed symmetrically between three short-circuited pairs of

electrodes. A controlled current is emitted from the short-circuited outer

Page 115: +Peters Ekwere j. - Petrophysics

2–72

pair of electrodes in such a manner that the voltage difference between

the two inner short-circuited pairs of electrodes is essentially zero. As in

the guard system, these electrode arrangements focus the formation

current into a thin disc, which flows perpendicularly to the borehole.

Figure 2.37. Focused-current electrode arrangements.

Various laterolog tools have been developed over the years. Among

the most commonly used tools, the dual laterolog is common. This tool,

similar to the dual induction tool, has both deep and shallow measuring

laterologs. It is often run in conjunction with a very shallow reading

laterolog tool, which is mounted on a pad pressed against the borehole.

Page 116: +Peters Ekwere j. - Petrophysics

2-73

This shallow reading curve, called the micro-spherically focused log,

measures the flushed zone resistivity (Rxo). This combination of

measurements can define the resistivity profile from the borehole,

through the invaded zone to the undisturbed formation. Since the

current path for these logs is through the mud to the borehole wall,

through the invaded zone, and then to the uncontaminated zone, the

resistivity readings are a combination of these different zones. However,

mud and the invaded zones affect the tool's resistivity measurement

much less than unfocused tools, a feature which minimizes corrections.

Microresistivity Logs

Microresistivity tools are designed to measure the resistivity of the

flushed zone (Rxo). Since the flushed zone could be only 3 or 4 inches

deep, Rxo tools have very shallow readings, with depths of investigation

approximately 1 to 4 inches. The electrodes are mounted on flexible pads

pressed against the borehole wall, thereby eliminating most of the effects

of the mud on the measurement (Figure 2.39). Microresistivity logs

include the microlog, microlaterologs and microspherically focused logs.

Collectively, these logs can be used to estimate

• Depth of invasion

• Flushed zone water saturation (Sxo)

• Moveable hydrocarbon saturation (Sxo-Sw)

• Corrections for deep induction and laterologs

• Permeability

• Hole diameter

• Pay zone thickness

Page 117: +Peters Ekwere j. - Petrophysics

2–74

• Porosity.

Figure 2.38. Presentation of a guard log.

Page 118: +Peters Ekwere j. - Petrophysics

2-75

Figure 2.39. Microresitivity logging tools.

Page 119: +Peters Ekwere j. - Petrophysics

2–76

2.7.5 Lithology Logs

Two lithology logs are commonly used in formation evaluation, the

Spontaneous Potential (SP) log and the Gamma Ray (GR) log. Both are

recordings of naturally occurring phenomena in the formation.

Spontaneous Potential Log (SP)

The SP curve records the electrical potential produced by the

interaction of formation water, conductive drilling mud, and certain ion-

selective rocks such as shale. It is a recording versus depth of the

difference between the electrical potential of a moveable electrode in the

borehole and the electrical potential of a fixed surface electrode. Opposite

shales, the SP curve usually defines a more or less straight line on the

log, called the shale baseline. Opposite permeable formations, the curve

shows deflections from the shale baseline. In thick beds, these

deflections tend to reach an essentially constant deflection defining a

sand line. The deflection may be to the left (negative) or to the right

(positive), depending primarily on the salinities of the formation water

and of the mud filtrate. If the formation water is more saline than the

mud filtrate, the deflection is to the left. If it is less saline than the mud

filtrate, the deflection is to the right.

The position of the shale baseline on the SP log is arbitrary as it is

set by the logging engineer so that the curve deflections remain in the SP

track of the log. The SP is measured in millivolts (mV).

An SP curve cannot be recorded in boreholes filled with

nonconductive muds, such as oil muds or air, because such muds do not

provide electrical continuity between the SP electrode and the formation.

Also, if the resistivities of the mud filtrate and formation water are about

Page 120: +Peters Ekwere j. - Petrophysics

2-77

equal, the SP deflections will be small and the curve will be rather

featureless and useless.

The deflections on the SP curve result from electric currents

flowing in the mud in the borehole. These currents are caused by

electromotive forces in the formation, which are of electrochemical and

electrokinetic origins. Consider a permeable formation with thick shales

above and below it. Assume that the two electrolytes present, formation

water and mud filtrate, contain sodium chloride (NaCl) only. Because of

the layered clay structure and the charges on the layers, shales are

permeable to the Na+ cations but impermeable to the Cl- anions. Only the

Na+ cations are able to move through the shale from the more saline to

the less saline NaCl solution. This movement of charged ions constitutes

an electric current, and the force causing them to move constitutes a

potential across the shale.

The curve arrow in the upper section of Figure 2.40 shows the

direction of the current corresponding to the flow of Na+ ions through the

adjacent shale from the more saline formation water to the less saline

drilling mud in the borehole. Since shales pass only the cations, shales

resemble ion-selective membranes, and the potential across the shale is

called the membrane potential.

A second component of the electrochemical potential is produced

at the edge of the invaded zone where the mud filtrate and formation

water (the electrolytes) are in direct contact. Here Na+ and Cl- can diffuse

from one electrolyte to the other. Since Cl- are more mobile than Na+

ions, the net result of the diffusion is the flow of negative Cl- ions from

the more saline to the less saline electrolyte. This is equivalent to a

conventional current flow in the opposite direction as shown by the

Page 121: +Peters Ekwere j. - Petrophysics

2–78

arrow A in the upper half of Figure 2.40. The current flowing across the

junction between solutions of different salinity is produced by an

electromotive force called liquid-junction potential. The magnitude of the

liquid-junction potential is much smaller than the membrane potential.

If the permeable formation is clean (not shaly), the total

electrochemical emf, Ec, corresponding to these two phenomena is given

by

log wc

mf

aE Ka

= − (2.31)

where aw and awf are the chemical activities of the two solutions at

formation temperature, K is a coefficient proportional to temperature,

and for NaCl formation water and mud filtrate is 71 at 25 ºC (77 ºF). The

chemical activity of a solution is roughly proportional to salinity and

hence to its conductivity. If the solutions contain substantial amounts of

salts other than NaCl, the value of K at 77 ºF may differ from 71. If the

permeable formation is shaly, or contains dispersed clay, the total

electrochemical emf will be reduced since the clay produces an

electrochemical membrane of opposite polarity to that of the adjacent

shale bed.

Page 122: +Peters Ekwere j. - Petrophysics

2-79

Figure 2.40. Schematic representation of potential and current distribution in and around a permeable bed.

Page 123: +Peters Ekwere j. - Petrophysics

2–80

An electrokinetic potential, Ek (also known as streaming potential

or electrofiltration potential), is produced when an electrolyte flows

through a permeable, nonmetallic, porous medium. The magnitude of the

electrokinetic potential is determined by several factors, among which are

the differential pressure producing the flow and the resistivity of the

electrolyte. In the borehole, an electrokinetic emf, Ekmc, is produced by

the flow of the mud filtrate through the mud cake. An electrokinetic emf,

Eksh, may also be produced across the shale, since it may have sufficient

permeability to permit a tiny amount of filtrate flow from the mud. Each

of these electrokinetic emfs contributes to a more negative SP reading

opposite the permeable bed and opposite the shale. The net contribution

to the SP deflection is, therefore, the difference between Ekmc and Eksh,

which is generally small and negligible.

The movement of ions, which causes the SP phenomenon, is

possible only in formations that have a certain minimum permeability.

However, there is no direct relationship between the magnitude of the SP

deflection and permeability, nor does the SP deflection have any direct

relationship with the porosity.

The lower portion of Figure 2.40 shows how the SP currents flow in

the borehole and formations. The current direction shown corresponds to

the more usual case where the salinity of the formation water is greater

than that of the mud filtrate. Thus, the potential observed over the

permeable bed is negative with respect to the potential opposite the

shale. This negative variation corresponds to an SP curve deflection to

the left of the SP log as shown in the figure.

As shown in Figure 2.40, the SP currents flow through four

different media: the borehole (mud), the invaded zone, the noninvaded

Page 124: +Peters Ekwere j. - Petrophysics

2-81

part of the permeable formation, and the surrounding shale. In each

medium, the potential along a line of current flow drops in proportion to

the resistance encountered. The total potential along a line of current

flow is equal to the total emf. The deflections on the SP curve are,

however, a measurement of only the potential drop in the borehole (mud)

resulting from the SP currents. This potential drop represents only a

fraction (although usually the major fraction) of the total emf. If the

currents could be prevented from flowing by means such as the

insulating plugs as shown in the upper part of Figure 2.40, the potential

differences observed in the mud would equal the total emf. The SP curve

recorded in such an idealized condition is called the static SP curve and

is shown in the lower part of Figure 2.40.

Figure 2.41 shows the presentation of an SP curve. In general,

shale-free permeable beds of moderate to low resistivity are sharply

defined by the SP curve. High resistivity beds distort the SP currents,

causing a change in the shape of the SP curve at bed boundaries and

thus poor boundary definitions. Also, the SP curve is depressed in

permeable zones that contain shale or hydrocarbon. The shape of the SP

curve is influenced by (1) the thickness (h) and resistivity (Rt) of the

permeable bed, (2) the resistivity (Ri) and the diameter (di) of the invaded

zone, (3) the resistivity (Rs) of the surrounding formation, and (4) the

resistivity of the mud (Rm) and the diameter (d) of the borehole.

Page 125: +Peters Ekwere j. - Petrophysics

2–82

Figure 2.41. Presentation of an SP curve in a sand-shale sequence.

The Gamma Ray Log (GR)

The GR log is a measurement of the natural radioactivity of the

formation. In sedimentary formations, the log normally reflects the shale

content of the formation because the radioactive elements tend to

concentrate in clays and shales. Clean formations usually have a very

low level of radioactivity, unless a radioactive contaminant such as

volcanic ash or granite wash is present or formation waters contain

dissolved radioactive salts.

The GR log can be recorded in cased holes, which makes it very

useful as a correlation curve in completion and workover operations. It is

Page 126: +Peters Ekwere j. - Petrophysics

2-83

frequently used to complement the SP log and as a substitute for the SP

curve in wells drilled with salt mud, air or oil-based muds. In each case,

it is useful for the location of shales and nonshaly beds and, most

importantly, for general correlation.

Gamma rays are bursts of high-energy electromagnetic waves that

are emitted spontaneously by some radioactive elements. Nearly all

gamma radiation encountered in the earth is emitted by the radioactive

potassium isotope of atomic weight 40 (K40) and by the radioactive

elements of the uranium and thorium series. These elements emit

gamma rays, the number and energies of which are characteristic of each

element. Figure 2.42 shows the energies of the emitted gamma rays.

Potassium emits gamma rays of a single energy at 1.46 MeV, whereas the

uranium and thorium series emit gamma rays of various energies.

In passing through matter, gamma rays experience successive

Compton-scattering collisions with atoms of the formation material,

losing energy with each collision. After the gamma ray has lost enough

energy, it is absorbed by means of photoelectric effect, by an atom of the

formation. Thus, natural gamma rays are gradually absorbed and their

energies reduced as they pass through the formation. Two formations

having the same amount of radioactive material per unit volume, but

having different densities, will show different levels of radioactivity. The

less dense formations will appear to be slightly more radioactive than the

more dense formations. The GR log response, after appropriate

corrections, is proportional to the weight concentrations of the

radioactive material in the formation:

i i i

b

V AGR

ρρ

= ∑ (2.32)

Page 127: +Peters Ekwere j. - Petrophysics

2–84

where ρi are the densities of the radioactive minerals, Vi are the bulk

volume factors of the minerals, Ai are proportionality factors

corresponding to the radioactivity of the mineral, and ρb is the bulk

density of the formation. In sedimentary formations, the depth of

investigation of the GR log is about 1 foot.

Figure 2.42. Gamma ray emission spectra of radioactive minerals.

The gamma ray logging tool contains a detector to measure the

gamma radiation originating in the volume of formation near the tool.

Page 128: +Peters Ekwere j. - Petrophysics

2-85

Scintillation counters are now generally used for this measurement. They

are much more efficient than the Geiger-Mueller counters used in the

past. Events resulting in gamma rays are random. For this reason,

gamma ray logs have a ragged appearance (see Figure 2.20).

The primary calibration for the gamma ray tools is in the API

(American Petroleum Institute) test facility in Houston. A field calibration

standard is used to normalize each tool to the API standard and the logs

are calibrated in API units. The radioactivities in sedimentary formations

generally range from a few API units in anhydrite or salt to 200 or more

units in shales.

The GR log is particularly useful for defining shale beds when the

SP is distorted (in very resistive formations), when the SP is featureless

(in fresh water-bearing formations or in salty muds, when Rmf = Rw), or

when the SP cannot be recorded (in nonconductive mud, empty or air-

drilled hole, cased holes). The bed boundary is picked at the point

midway between the maximum and minimum deflection of the anomaly.

The gamma ray reflects the proportion of shale and, in many

regions, can be used to quantitatively as a shale indicator. It is also used

for the detection and evaluation of radioactive minerals, such as potash

and uranium ore. The GR log is part of most logging programs in both

open hole and cased hole. It is readily combined with most other logging

tools and permits the accurate correlation of logs made on one trip into

the borehole with those that were made on another trip.

2.7.6 Nuclear Magnetic Resonance (NMR) Logs

NMR measurements of fluids in porous media can be used under

favorable conditions to estimate porosity, irreducible water saturation,

Page 129: +Peters Ekwere j. - Petrophysics

2–86

moveable fluid saturation, fluid viscosity, pore size distribution, surface

to volume ratios and permeability. In this section, the principles of NMR

and its use in petrophysical measurements are presented.

Nuclear Spins in a Magnetic Field

Most atomic nuclei possess a quantum mechanical property called

spin angular momentum. This means that the nucleus spins around an

axis. Since atomic nuclei are charged particles, the spinning motion

causes a magnetic moment that is co-linear with the direction of the spin

axis. The strength of this magnetic moment is a property of the type of

nucleus. 1H nuclei (protons) possess a strong magnetic moment (second

only to radioactive 3H), which, together with the high natural abundance

of hydrogen and prevalence in most fluids, makes it the ideal nucleus for

NMR logging and NMR imaging in radiology.

Consider a collection of 1H nuclei as in Figure 2.43. In the absence

of an externally applied magnetic field, the individual magnetic moments

have no preferred orientation and the net magnetization of the collection

of spins is zero. However, if an externally supplied magnetic field

(denoted as B0) is imposed, there is a tendency for the magnetic moments

to align with the external field. 1H nuclei with a quantum number of I = 12

in this situation may adopt one of two possible orientations: alignment

parallel or anti-parallel to B0 as shown in Figure 2.44. Thus, depending

on their orientation, we can define two groups or populations of spins.

Alignment parallel to B0 is the lower energy orientation and is thus

preferred, while the anti-parallel alignment is the higher energy state.

However, the energy difference between the two states is very small.

Thermal energy alone causes the two states to be almost equally

populated. The remaining population difference results in a net bulk

Page 130: +Peters Ekwere j. - Petrophysics

2-87

magnetization aligned parallel to B0. It is only this net magnetization

arising from a small population difference that is detectable by NMR

techniques.

The individual spins do not align exactly parallel or anti-parallel to

B0, but at an angle to B0. This is analogous to the case of a spinning top:

Figure 2.43. Random orientations of the nuclear magnetic moments in the absence of an externally applied magnetic field.

Bo

parallel

antiparrallel

Figure 2.44. Two orientations of the nuclear magnetic moments in the

Page 131: +Peters Ekwere j. - Petrophysics

2–88

presence of an externally applied magnetic field (Bo).

the top precesses about the axis defined by the pull of gravity. This

precession defines the surface of a cone. Figure 2.45 shows a model for a

large collection of spins at any given instant. Here, each of the vectors

(arrows) represents an individual spin. The vector M represents the bulk

net magnetization that results from the vector sum of the contributions

from each of the spins.

Bo M

parallel

antiparallel

Figure 2.45. The bulk net magnetization M

The individual magnetic moments precess at a certain frequency,

known as the Larmor frequency, which is determined by the strength of

the magnetic field and the type of nucleus. The Larmor equation gives

this frequency as

0 0Bω γ= (2.33)

where ω0 is the Larmor frequency (in radians per second), Bo is the

magnetic field strength, and γ is a constant for each nucleus known as

Page 132: +Peters Ekwere j. - Petrophysics

2-89

the gyromagnetic ratio. Eq.(2.33) is the fundamental equation for all

NMR methods. Table 2.7 gives γ values for some NMR active nuclei.

From Table 2.7, we can see that 1H nucleus has the highest

relative sensitivity and a high natural abundance. So 1H nuclei (protons)

possess a strong magnetic moment and are most commonly used in NMR

logging and NMR imaging.

Table 2.7: Magnetic Resonance Properties of Some Important Nuclei

Nucleus

Natural

Abundance %

γ

Hz / Gauss

Relative

Sensitivity

1H 13C 19F

23Na 31P

99.98

1.10

100.0

100.0

100.0

4257

1071

4005

1126

1723

1.0

0.016

0.830

0.093

0.066

The Effect of Radiofrequency Pulses - Resonance Absorption

In order to detect a signal, a condition of resonance needs to be

established. The term “resonance” implies alternating absorption and

dissipation of energy. Energy absorption is caused by radiofrequency (RF)

perturbation, and energy dissipation is mediated by relaxation processes.

RF radiation, like all electromagnetic radiation, possesses electric

and magnetic field components. We may consider the RF as another

magnetic field of strength B1 perpendicular to B0 as shown in Figure

2.46. At equilibrium, M (=Mo) is stationary and difficult to observe. By

Page 133: +Peters Ekwere j. - Petrophysics

2–90

applying an RF field B1 perpendicular to B0, M can be rotated so that it

has a component transverse to B0. A maximum transverse component is

obtained by applying B1 with an amplitude and duration that rotates M

by 90 degree and into the plane transverse to B0. This is called a 90-

degree pulse. Once in the transverse plane, M precesses (rotates) around

the B0 axis at the Larmor frequency. This rotating magnetization can

induce an alternating current (AC) in a receiver coil, and that current can

be used to record the action of the magnetization in the transverse plane.

The signal that is induced in the receiver coil decays over time. The

signal decay is due to a process known as relaxation.

Relaxation Processes

In resonance absorption, RF energy is absorbed by the nuclei when

it is broadcast at the Larmor frequency. Relaxation is the process by

which the nuclei release this energy and return to their original

configuration. There are two relaxation processes involved: transverse

relaxation and longitudinal relaxation.

Bo M

B1

Figure 2.46. A second magnetic field B1 generated by Radiofrequency (RF)

Page 134: +Peters Ekwere j. - Petrophysics

2-91

Given that at equilibrium, the net magnetization is longitudinal (B0

direction), the equilibrium magnetization in the transverse plane is zero.

So the signal that is induced in the receiver coil will decay over time and

reach zero some time after the 90º RF pulse is turned off. This decay

process is exponential and the decay envelop can be expressed by the

equation

( )0 *2exp /transverse transverseM M t T= − (2.34)

where 0transverseM is the initial transverse magnetization immediately after a

90º RF pulse, transverseM is the transverse magnetization at any given time t

after a 90º RF pulse, *2T is the apparent NMR transverse relaxation time

or spin-spin relaxation time, which characterizes the rate of signal decay.

The decay mechanism is that different components of the magnetization

may precess at slightly different rates, a process known as “dephasing”

in the transverse plane. Since the signal recorded is the vector sum of all

the transverse components, sufficient dephasing will lead to complete

cancellation of the signal.

One of the major causes of this dephasing is B0 inhomogeneity

(ΔB0) effects. Spins at different locations are not exposed to exactly the

same B0 field, which in turn yields a range of Larmor frequencies. ΔB0

effects are largely suppressed by correcting the magnetic field uniformity

as much as possible, and by employing NMR “spin-echo” techniques.

In spin-echo NMR (Hahn, 1950), the dephasing ΔB0 effects are

largely reversed by following a 90-degree RF pulse with a short delay τ

and a 180-degree “refocusing” pulse. After a second delay τ, the

transverse magnetization transverseM is refocused as an NMR spin-echo. ΔB0

Page 135: +Peters Ekwere j. - Petrophysics

2–92

effects are canceled at the center of the echo, where the dephasing from

ΔB0 is momentarily rephased.

If the induced signal is measured at time TE after a 90º RF pulse,

then equation (2.34) can be rewritten as

( )02exp /transverse transverse EM M T T= − (2.35)

where TE=n2τ and n is the number of 180-degree pulses. T2 is the true

NMR transverse relaxation time or spin-spin relaxation time. In Eq.(2.35), *

2T is substituted with T2 upon the correction of ΔB0 effects.

T2 can be measured by repeating the spin-echo experiment with

different TE, then fitting the resulting transverseM versus TE curve to

Eq.(2.35). Another more accurate method, and one that is used in NMR

logging, is to employ a multiple-spin-echo sequence, where the NMR

signal is refocused multiple times as shown in Figure 2.47. This method

removes unwanted molecular diffusion effects. In NMR logging, 400 or

more echos are used in the measurements. The exponential curve, which

measures the amplitudes of the decaying echos is the fundamental well

log measurement and is used to compute the T2 spectrum.

Page 136: +Peters Ekwere j. - Petrophysics

2-93

Figure 2.47. Carr-Purcell-Meiboom-Gill spin echo pulse sequence.

Longitudinal relaxation is a process that restores the longitudinal

magnetization to its equilibrium state after an RF pulse is turned off.

Immediately after a 90º pulse, the net magnetization vector lies in the

transverse plane. Therefore, longitudinal magnetization is zero. However,

as time passes, the longitudinal magnetization will approach the

equilibrium value. This buildup of the longitudinal magnetization is also

exponential in time and can be expressed as

( ) ( )0 11 exp /zM t M t T= − −⎡ ⎤⎣ ⎦ (2.36)

where M0 is the equilibrium value of the longitudinal magnetization, Mz(t)

is the longitudinal magnetization at any given time t after a 90º RF

pulse, and T1 is the NMR longitudinal relaxation time or spin-lattice

relaxation time that characterizes the rate of buildup. For any given

system, T1 is greater than or equal to T2. (T1 and T2 of 1H nuclei in water

Page 137: +Peters Ekwere j. - Petrophysics

2–94

in a test tube is about 3 seconds. In a porous medium, the T1 and T2 are

significantly lower than in a test tube and are of the order of

milliseconds). The longitudinal direction or direction of B0, is

conventionally assigned the z direction, i.e., B0=Bz.

In an NMR experiment, sufficient time must elapse between

successive 90º RF pulses to allow M to achieve equilibrium. This time,

known as the repetition time TR, depends on the sample’s T1. In general,

TR should be greater than 5T1. Substituting TR into Eq.(2.36) gives

( ) ( )0 11 exp /z R RM T M T T= − −⎡ ⎤⎣ ⎦ (2.37)

where Mz(TR) is the longitudinal magnetization at the repetition time TR

after a 90º RF pulse is turned off. Since another 90º RF pulse is applied

immediately at time TR, the 0transverseM for this subsequent 90º RF pulse

should be equal to Mz(TR)) of Eq.(2.37). Thus,

( )00 11 exp /transverse RM M T T= − −⎡ ⎤⎣ ⎦ (2.38)

Substituting Eq.(2.38) into Eq.(2.35) gives the transverse magnetization

at the time of measurement as

( ) ( )0 1 21 exp / exp /transverse R EM M T T T T= − − −⎡ ⎤⎣ ⎦ (2.39)

As Eq.(2.38) implies, T1 can be measured by repeating an NMR

experiment with several different TR values, then fitting the resulting

transverseM versus TR to Eq.(2.38).

Molecular Diffusion Effect

Page 138: +Peters Ekwere j. - Petrophysics

2-95

Random thermal motion will affect the NMR signal as spins move

from one part of the sample to another and experience a different

magnetic field strength because of static field inhomogeneities. Variation

in the frequency and phase of these mobile spins introduces a phase

incoherence that causes a reduction in the signal amplitude. This effect

can be expressed by Eq.(2.40) (Hahn, 1950):

32 2

0 2exp exp3 2

E Etransverse transverse

R

T TD GM MT

γ⎡ ⎤⎛ ⎞ ⎛ ⎞= − −⎢ ⎥⎜ ⎟ ⎜ ⎟⎝ ⎠⎢ ⎥⎝ ⎠ ⎣ ⎦

(2.40)

where D is the diffusion coefficient and G (∝ΔB0) is the magnetic field

gradient.

NMR Signal and Corresponding T2 Spectrum

The exponential decay curve shown in Figure 2.47 suggests that all

the spins relax at the same T2 relaxation time. This would be applicable

to the spins in a bulk fluid such as the protons in water in a test tube.

However, when the fluids are confined in the pore space of a porous

medium, the protons near the pore wall will relax faster (shorter T2) than

those in the center of the pores. Further, the protons in small pores will

relax faster than those in large pores. As a result, the NMR decay signal

will contain a spectrum of T2 relaxation times. The signal can be

decomposed into its T2 spectrum by using a multi-exponential model as

( ) ( )210

exp /j m

j jj

M tf t T

M

=

=

= −∑ (2.41)

Page 139: +Peters Ekwere j. - Petrophysics

2–96

where fj is proportional to the population of protons which have a

relaxation time of T2j. The inversion problem is to find a set if T2

amplitudes, fj, from a set of measurements of echos, gi. Application of

Eq.(2.41) to each echo gives the following system of linear equations in fj

( ) ( )210

exp / for 1,2,...,j m

ii j i j i

j

M tg f t T i n

=

=

= = − + =∑ (2.42)

where gi is the amplitude of the echo measured at time ti, εi is the error

caused by noise in the data and n is the total number of echos. A least

square fit is used to determine the fjs that minimize the sum

( )2

22

1 1 1

exp /n m m

j i j i ji j j

f t T g fα= = =

⎛ ⎞− − +⎜ ⎟

⎝ ⎠∑ ∑ ∑ (2.43)

where αi is a regularization constant or smoothing parameter.

Figure 2.48 shows an example log measurement for n = 500, TE =

1.2 milliseconds, with 12 measurements averaged together. Note the

noisy data. The magnitude of the NMR signal at t =0 (M0) is proportional

to the number of hydrogen nuclei in the measurement volume. This

number can be calibrated into the total NMR porosity. Figure 2.49 shows

the corresponding T2 spectrum obtained from the data of Figure 2.48.

Page 140: +Peters Ekwere j. - Petrophysics

2-97

Time (ms)

Figure 2.48. Carr-Purcell-Meiboom-Gill spin echo measurements.

T2 (ms)

Figure 2.49. T2 spectrum of the data of Figure 2.49.

Page 141: +Peters Ekwere j. - Petrophysics

2–98

Based on laboratory measurements in sandstone cores, the NMR

spectrum can be divided into several segments to represent the fluid

distribution in the pore space as shown in Figure 2.50. In this figure, the

area under the spectrum is equal to the total NMR porosity.

Figure 2.50. T2 spectrum and fluid distribution in the pore space.

The nomenclature for Figure 2.50 is as follows. The Free Fluid

Index (FFI) is the percent of the bulk volume occupied by movable fluids

(water + hydrocarbon). In Figure 2.50, this is shown as the fraction of the

Page 142: +Peters Ekwere j. - Petrophysics

2-99

bulk volume for T2 greater than 30 milliseconds. The Free Fluid Index is

also known as Bulk Volume Movable (BVM). The Bulk Volume

Irreducible (BVI) is the percent of the bulk volume containing irreducible

water. The irreducible water consists of two components: (1) Clay Bound

Water (CBW) (T2 from 0.3 to 3 milliseconds), which is the water bound to

the clay minerals and (2) Bulk Volume Capillary (BVC) (T2 from 3 to 30

milliseconds), which is the water trapped by capillary forces. The Bulk

Volume Water (BVW) (T2 from 0.3 to 300 milliseconds) is the percent of

the bulk volume occupied by water (moveable, capillary bound water and

clay bound water). These bulk volumes are related to porosity and

saturations as follows:

ptotal

b

VV

φ = (2.44)

ii

p

VSV

= (2.45)

p CBW BVC BVMV V V V= + + (2.46)

CBW BCV BVMtotal

b b

V V VV V

φ += + (2.47)

CBW BCV

b

V VBVIV+

= (2.48)

BVM

b

VBVMV

= (2.49)

total BVI BVMφ = + (2.50)

Page 143: +Peters Ekwere j. - Petrophysics

2–100

total MPHIφ = (2.51)

totalBVM BVI MPHI BVI FFIφ= − = − = (2.52)

w pww total

b b

S VVBVW SV V

φ= = = (2.53)

BVC BVMe

p

V VV

φ += (2.54)

BVC BVMe

total b

V VV

φφ

+= (2.55)

etotal

BVC BMVφφ+

= (2.56)

CBW BVC bwirr

p b total

V V V BVISV V φ

⎛ ⎞⎛ ⎞+= =⎜ ⎟⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(2.57)

total wirrBVI Sφ= (2.58)

( )1total wirrFFI Sφ= − (2.59)

Figure 2.51 shows an NMR log in which the pore fluids in the

second track have been divided into free fluid (above 30 ms), capillary

bound fluid (3 to 30 ms), small pore bound fluid (0.28 to 3 ms), and very

small pore bound fluid (0.2 to 0.28 ms). The two small pore bound fluids

(0.2 to 3 ms) correspond to clay bound fluid. Also shown in the third

track is the T2 spectrum, with a T2 cutoff of 30 ms for free fluid. Note the

two streaks with free fluid at depths of X163 m and X189 m, where the

Page 144: +Peters Ekwere j. - Petrophysics

2-101

T2 is greater than 30 ms. This example is from a predominantly shaly

formation where the differences in the bound fluids are more apparent.

Pore Size Distribution

NMR relaxation measurements have been shown to be a sensitive

probe in the study of the microscopic structure of porous media. The

connection between NMR measurements and pore size is based on the

strong effect that the rock surface has on promoting proton relaxation.

The NMR T1 relaxation behavior of a fluid confined within a pore is

sensitive to both the pore geometry and size, and thus yields much

useful information when related to the pore-size distribution via an

appropriate mathematical model. The simplest and most common model

is the “two-fraction fast-exchange” model (Senturia and Robinson, 1970;

Howard et al., 1990), which assumes that there are two magnetically

distinct phases within the pore: a bulk phase with relaxation

characteristic of the bulk fluid and a surface phase with much faster

relaxation. Assuming that diffusion of the fluid is much faster than the

relaxation process, the observed relaxation rate is given by a single

average T1 value as

1 1 1

1 1

bulk surface

ST T V T

λ= + (2.60)

where S/V is the pore surface area to volume ratio and λ is the thickness

of the surface monolayer. The T1 technique basically determines the

surface/volume ratio as a characteristic pore-size parameter. Usually,

T1bulk>>T1surface, so Eq.(2.60) can be simplified as

Page 145: +Peters Ekwere j. - Petrophysics

2–102

Figure 2.51. Example NMR log showing fluid distribution in the pore

space.

Page 146: +Peters Ekwere j. - Petrophysics

2-103

11

1 ST V

ρ ⎛ ⎞= ⎜ ⎟⎝ ⎠

(2.61)

where ρ1 = λ/T1surface is the NMR surface relaxivity, which is a measure of

the ability of the surface to cause relaxation of proton magnetization. The

surface relaxivity ρ1 has dimensions of length/time. Eq.(2.61) shows that

T1 responds to pore size. Small pores (large S/V ratio) exhibit small

values for T1, and the converse is true for large pores.

Since transverse relaxation time T2 is closely related to T1, it is

expected that a similar relationship exists between the distribution of T2

and the pore-size distribution. However, in this case, apart from the bulk

relaxation process, which can often be neglected, the T2 relaxation is also

controlled by a surface relaxation mechanism as well as the diffusion

effect because of the magnetic field gradient. The relaxation rate equation

is of the form

( )2

2 2 2

1 1 13bulk surface

S B DT T V T

λ τγ= + + ∇ (2.62)

where the first two terms on the right side correspond to similar

expressions in Eq.(2.60), while the last term accounts for spin dephasing

because of restricted diffusion in a magnetic field gradient ∇B. τ is the

pulse spacing (pulse-to-echo delay), γ is the gyromagnetic ratio for

protons, and D is the diffusion coefficient. Usually, the first and the last

Page 147: +Peters Ekwere j. - Petrophysics

2–104

terms on the right side of Eq.(2.62)are small and can be neglected. Thus,

22

1 ST V

ρ ⎛ ⎞= ⎜ ⎟⎝ ⎠

(2.63)

where ρ2 = λ/T2surface is the NMR T2 surface relaxivity. Eq.(2.63) shows

that T2 also responds to pore-size. Thus, in a water bearing zone without

the complications of hydrocarbons, the T2 spectrum such as shown in

Figure 2.49 could be viewed as a pore-size distribution.

It should be emphasized that what is measured by NMR relaxation

is the distribution of volume-to-surface ratio, (V/S). Since (V/S) has the

dimension of length, then its distribution could be viewed as a “pore-size

distribution.” The ratio (V/S) is not a pore size or a pore diameter. Its

magnitude is markedly affected by the pore shape. The ratio will be a

maximum for a spherical pore (a sphere has the smallest surface area for

a given volume) and will decrease for other pore shapes. If all pores are

geometrically similar, then the T2 spectrum could be viewed qualitatively

as a pore size distribution.

Figure 2.52 shows a clastic sequence with shales overlying a

sandstone in a water zone. The volumetric calculations in the first track

indicate an upward decreasing clay in the sandstone interval. Geological

analysis suggests that this corresponds to a coarsening-upward

sequence. While this assertion is based on inference, the T2 spectrum in

the third track shows an upward increase in the relaxation times, a trend

Page 148: +Peters Ekwere j. - Petrophysics

2-105

which can be explained by an increase in the pore size and, therefore, an

increase in the grain size.

Figure 2.52. Example NMR log showing pore size distribution in a water-bearing sandstone section.

Page 149: +Peters Ekwere j. - Petrophysics

2–106

Carbonates often have a variety of pore types such as moldic,

intercrystalline and interparticle pores. The relationship between

relaxation times and pore size distribution is therefore more complex in

carbonates than in sandstones. Further, the surface relaxivity of

carbonates is less than that of sandstones, typically 1.7μm/s for

carbonates compared to 5μm/s for sandstones. As a result, the protons

align and relax faster in sandstones than in carbonates. Therefore, the

cut off T2 for free fluid in carbonates is higher than in sandstones,

typically on the order of 100 ms.

Estimation of Permeability from NMR Relaxation Times

The fact that T1 and T2 NMR relaxation times can be used to

estimate permeability stems from the fact that these relaxation times can

be correlated with pore size distribution in water-bearing zones.

Permeability is proportional to the square of some characteristic length of

the porous medium and as such would be proportional to 21T or 2

2T .

Various empirical equations have been proposed for estimating the

absolute permeability of a porous medium. Wyllie and Rose (1950)

suggested an empirical permeability equation of the form

x ywirrk C Sφ −= (2.64)

where C is an empirical constant, φ is the porosity, Swirr is the irreducible

water saturation and x and y are numerical exponents. Timur developed

a permeability equation in the spirit of Eq.(2.64) of the form

4.5

4210wirr

kSφ

= (2.65)

Page 150: +Peters Ekwere j. - Petrophysics

2-107

where the porosity and irreducible water saturation are in fractions and

the permeability is in millidarcy. One version of the Timur/Coates

equation that is widely used to estimate permeability from NMR logs is

given by

4 2

NMR FFIkC BVI

φ⎛ ⎞ ⎛ ⎞= ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(2.66)

where C=10 or can be determined from laboratory measurements on

cores. In Eq.(2.66) the porosity, FFI and BVI are in porosity units (p.u.)

and the permeability is in millidarcy. BVI and FFI are related to the

irreducible water saturation by Eqs.(2.58) and (2.59).

Another permeability equation based on laboratory measurements

at Schlumber Doll Research (SDR) is given by

4 22gmk a Tφ= (2.67)

where T2gm is the logarithmic (geometric) mean of the T2 spectrum

defined as

2

12

1

logm

j jj

gm m

jj

f TT

f

=

=

=∑

∑ (2.68)

Other equations proposed by Kenyon et al. are

4 21 1NMRk C Tφ= (2.69)

and

Page 151: +Peters Ekwere j. - Petrophysics

2–108

4 22 2NMRk C Tφ= (2.70)

Figure 2.53 shows a comparison of the permeability from NMR

measurements using Eq.(2.66) with those from core analysis. The

agreement between the two sets of data is good. Also shown is the

comparison of the NMR porosity measurements with those from core

analysis. The agreement also is good.

Figure 2.53. A comparison of NMR-derived permeability and porosity with measurements from core analysis (from Dunn et al., 2002).

Page 152: +Peters Ekwere j. - Petrophysics

2-109

2.7.7 NMR Imaging of Laboratory Cores

Nuclear magnetic resonance imaging (NMRI) can be used to map

the spatial distribution of NMR observables in core analysis. Thus,

porosity distribution and fluid saturation distribution or solvent

concentration distributions can be imaged in the laboratory.

In order to create an image, the NMR experiments must be

modified to spatially encode the NMR signals. Figure 3.54 is a typical

2DFT (2-dimensional Fourier Transform) spin-echo NMR imaging

sequence showing (a) radiofrequency pulses, (b) slice-selection gradient,

(c) frequency-encoding gradient, (d) phase-encoding gradient, and (e)

NMR signal (echo). The experiment is repeated with multiple phase-

encoding gradient amplitudes.

The Effect of Magnetic Field Gradients

For most NMR imaging applications, the B0 field must be made to

vary in a linear fashion with distance. A magnetic field gradient or simply

gradient refers to the spatial variation of the strength of the B0 field.

Magnetic field gradient is a key factor in NMR imaging. A magnetic field

gradient causes the transverse magnetization to precess at a frequency

that is proportional to position along the gradient axis as follows:

( )0r rB rGω γ= + (2.71)

where ωr is the Larmor frequency at position r, Gr is the gradient, and r

is the position along the gradient axis. Since we can measure ωr and we

know B0, Gr and γ, the position of the resonating nucleus can be

determined. Gr can be applied concurrently with slice selection,

frequency encoding or independently of other events (phase encoding).

Page 153: +Peters Ekwere j. - Petrophysics

2–110

The magnitude of the gradient, its direction (i.e., along what axis), and

timing need to be controlled.

180o

90o

a. RF

b. ssG

c. ROG

d. PEG

Timing

RT p1 pw tro

ET

Echo

e. NMR

tpe

Figure 2.54. Simplified timing diagram for a 2DFT spin-echo NMR imaging sequence

Page 154: +Peters Ekwere j. - Petrophysics

2-111

Slice-Selective Excitation

In most NMR imaging applications, it is desirable to generate a

single slice or multislice images. The initial step in generating such an

image is the localization of the RF excitation to a region of space. This is

accomplished by the use of frequency-selective excitation in conjunction

with a slice-selection gradient, Gss, on an axis perpendicular to the

chosen slice plane. A frequency-selective RF pulse has two parts

associated with it, a central frequency and a bandwidth of frequencies

Δωss defined by the shape of the pulse envelope. When such a pulse is

broadcast in the presence of the slice-selection gradient, a narrow region

of the object will achieve the resonance condition and will absorb the RF

energy. The central frequency of the pulse determines the particular

location that is excited by the pulse when the slice-selection gradient is

present. Different slice positions are achieved by changing the central

frequency. The slice thickness is determined by the bandwidth of

frequencies Δωss incorporated into the RF pulse and is given by

1ss ss sG dω γΔ = (2.72)

where ds1 is the slice thickness. Typically, Δωss is fixed so that the slice

thickness is changed by changing the amplitude of Gss. Thinner slices

require larger Gss. Once Gss is determined, the central frequency ωr is

calculated using Eq.(2.71) to bring the desired location into resonance.

Multislice imaging uses the same Gss for each slice but a unique RF

pulse during excitation. Each RF pulse has the same bandwidth but a

different central frequency, thereby exciting a different region of the

object.

Page 155: +Peters Ekwere j. - Petrophysics

2–112

Frequency Encoding

The next task in the imaging process is encoding the image

information within the excited slice. The image information is actually

the amplitude of the NMR signal arising from the various locations in the

slice. Two distinct processes are used for encoding the two dimensions:

frequency encoding and phase encoding.

The frequency encoding provides one of the two visual dimensions

of the image. The NMR signal is always detected in the presence of a

frequency encoding gradient in an imaging pulse sequence. After slice-

selective excitation, the frequency encoding gradient, also known as the

readout gradient, GRO, is applied perpendicularly to the slice direction.

Under the influence of this new gradient field, the nuclei within the slice

will precess at different frequencies depending on their positions in the

readout gradient’s direction, in accordance with Eq.(2.71). Each of these

frequencies will be superimposed in the echo. The echo signal is detected

in the presence of GRO and digitized at a chosen sampling interval for

later Fourier transformation. The magnitude of GRO and the frequencies

that are detected enable the positions of the nuclei to be determined.

Two user-selectable parameters determine the image resolution in

the frequency encoding, or the readout, direction: the field of view (FOV)

in the readout direction and the number of readout data points in the

matrix, NRO. The pixel size in the readout direction is given by

( )RO

RORO

FOVN

Δ = (2.73)

and

Page 156: +Peters Ekwere j. - Petrophysics

2-113

( )2

RORO

RO RO ro

NBWFOVG G tγ γ

= = (2.74)

where BW is the receiver bandwidth, γ is the gyromagnetic ratio and tro is

the duration of the frequency encoding gradient.

Phase Encoding

In order to produce a two-dimensional image of the slice, one can

cause a systematic variation in phase that would encode the spatial

information along the one remaining axis of the image plane. This is

accomplished by the use of a phase encoding gradient, GPE. GPE is

perpendicular to both Gss and GRO and is the only gradient that changes

amplitude during the data acquisition loop of a standard two-

dimensional imaging sequence. The NMR imaging information is

obtained by repeating the slice excitation and signal detection many

times (typically 128 or 256 times), each with a different amplitude of GPE

applied before detection. The resulting signals are stored separately for

subsequent processing.

Separate Fourier transformation of each of these data sets yields a

set of projections onto the readout axis. Specifically, the second Fourier

transformation converts signal amplitude at each readout frequency from

a function of GPE to a function of frequency.

The image resolution in the phase encoding direction depends on

two user-selectable parameters, the FOV in the phase encoding direction

and the number of phase encoding steps in the matrix, NPE. The pixel

size in the phase encoding direction is given by

( )PE

PEPE

FOVN

Δ = (2.75)

Page 157: +Peters Ekwere j. - Petrophysics

2–114

and

( ) PEPE

PE pe

NFOVG tπ

γ= (2.76)

where γ is the gyromagnetic ratio and tpe is the duration of the phase

encoding gradient.

Image Reconstruction

Two types of matrices are used in NMR imaging: raw data and

image data. The raw data matrix is a grid of complex points with the

frequency encoding direction displayed in the horizontal direction and

the phase encoding direction displayed in the vertical direction. All image

information is contained within the raw data matrix.

The image data matrix is obtained by a two-dimensional Fourier

transformation of the raw data matrix. The first Fourier transformation of

each row of the raw data matrix yields a set of modulated projections of

the slice onto the frequency encoding axis. The second Fourier

transformation of each column of the temporary data matrix converts the

signal magnitude at each readout frequency from a function of GPE to a

function of frequency, resulting in the image. The image matrix is a

spatial map of the nuclei signal intensity. While the Fourier

transformation contains information regarding both the magnitude and

the phase of the measured signals, the phase information is often

discarded so that the normal image matrix contains only the magnitude

information. The image matrix is usually displayed as a rectangular

image with readout in one direction and phase encoding in the other

direction.

Page 158: +Peters Ekwere j. - Petrophysics

2-115

Figure 2.55 shows a graphical representation of how an image is

spatially encoded during 2DFT imaging. In this figure, the frequency

encoding direction is horizontal and the phase encoding direction is

vertical. A set of NMR signals is acquired using the same frequency

encoding gradient but different values of the phase encoding gradient.

Each of these NMR signals is Fourier transformed to provide a frequency

spectrum of each phase encoding step, which constitutes a one-

dimensional projection in the frequency encoding direction. Each column

of the data from the first Fourier Transform projection images is Fourier

transformed again to determine the spatial projection in the vertical

image plane.

Three-Dimensional NMR Imaging

Three-dimensional volume imaging technique is, in essence, a

double phase encoding technique. The slice-selective excitation is

replaced with another phase encoding process along that axis. Each RF

pulse excites the entire imaging volume instead of just one slice. The

second phase encoding is applied to partition or subdivide the volume

into individual layers. The number of layers is determined by the number

of phase encode steps. For example, if the number of the second phase

encoding gradient steps is changed from 32 to 128, 3D-FFT of the 3D

data set then yields 32 to 128 image layers in that direction.

Page 159: +Peters Ekwere j. - Petrophysics

2–116

Figure 2.55. Graphical demonstration of 2DFT NMR image data acquisition and reconstruction.

The advantage of the 3D-imaging technique is that very thin

contiguous layer images can be obtained with minimal interslice

crosstalk. Also, the signal-to-noise ratio is greater than for a comparable

2D-sequential imaging method.

The disadvantage of the technique is the time required. The total

scan time for 3D volume imaging is much longer than for 2D slice or

multislice imaging. In practice, 3D volume imaging is not widely used.

Instead, 2D-multislice imaging is often used as a replacement for 3D

volume imaging.

Signal-to-Noise Ratio and Image Contrast

Page 160: +Peters Ekwere j. - Petrophysics

2-117

Voxel size is a very important factor in increasing the signal-to-

noise ratio (SNR). Voxel size is defined by

( ) ( )

1RO PE

sRO PE

FOV FOVv d

N N⎛ ⎞⎛ ⎞

Δ = ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(2.77)

If the voxel volume is large, then there are more spins in each voxel to

contribute to the signal, thus increasing the SNR. But large voxels imply

a low spatial resolution. The converse is also true. Small voxels imply a

low SNR but a high spatial resolution. Therefore, the user-controlled

parameters, slice thickness ds1, FOV, NRO and NPE, are important

parameters that affect the signal-to-noise ratio and the image resolution.

The echo time TE, repetition time TR, longitudinal relaxation time T1

and transverse relaxation time T2 are the important factors that

determine image contrast as defined in Eq.(2.39). Remembering that T2

relaxation describes the rate of decay of the NMR signal in the transverse

plane, a long TE would yield different signal intensity from objects

possessing different T2 values. The long T2 object will contribute more

signal, causing it to appear hyperintense in the NMR image relative to

the short T2 object. This is termed a “T2-weighted image.” Similarly, since

T1 relaxation describes the rate of recovery of the longitudinal

magnetization, a short TR would yield different signal intensity from

objects possessing different T1 values. The object exhibiting a short T1

value will contribute more signal, causing it to appear hyperintense in

the NMR image relative to the long T1 object. This is termed a “T1-

weighted image.”

Page 161: +Peters Ekwere j. - Petrophysics

2–118

In summary, TE controls T2-weighting and TR controls T1-

weighting of an image. Short T2 objects are dark on T2-weighted images,

but short T1 objects are bright on T1-weighted images.

Example NMR Images of Laboratory Cores

Figures 2.56 and 2.57 show NMR-derived porosity images of a

brine saturated sandstone core. Figure 2.56 shows the images at six

cross sections along the core whereas Figure 2.57 shows the images of

four, longitudinal vertical slices. The layered nature of the core is clearly

apparent from these images. Although the images are not calibrated with

numerical porosity values, nevertheless, the first three slices in Figure

2.56 indicate that the core appears to be less porous at the top than

elsewhere. The core is 10 cm long, 5 cm in diameter, and has a

permeability of 97 md and an average porosity of 15.9 %.

Figure 2.58 shows NMR-derived images of the solvent

concentration for a first-contact miscible displacement conducted in the

core of Figures 2.56 and 2.57. In this experiment, a more viscous

deuterium oxide (D2O) was used to displace the less viscous brine from

Page 162: +Peters Ekwere j. - Petrophysics

2-119

Figure 2.56. NMR images of transverse slices of a brine saturated sandstone core.

the core at a favorable mobility ratio of 0.84. D2O is practically devoid of

protons, so the injection of D2O reduces the NMR signal in the voxels

Page 163: +Peters Ekwere j. - Petrophysics

2–120

invaded by the D2O. The reduced NMR signals can be used to image the

progress of the displacement as shown in Figure 2.58. Because the

mobility ratio is favorable, the displacement pattern is controlled entirely

by the permeability variations in the core. Note the chanelling of the

injected solvent due to permeability variation in the core. Clearly, the top

of the core is less permeable than the rest of the core. This correlates well

with the low porosity of the top of the core shown in Figure 2.56. Figure

2.59 shows one-dimensional solvent concentration profiles for the same

displacement. The information is quantitative and can be used to

calibrate a numerical model of the displacement.

Figure 2.57. NMR images of longitudinal, vertical slices of a brine saturated sandstone core.

Page 164: +Peters Ekwere j. - Petrophysics

2-121

Figure 2.58. Solvent concentration images of a first-contact miscible displacement in the sandstone core of Figures 2.55 and 2.56.

Page 165: +Peters Ekwere j. - Petrophysics

2–122

Figure 2.59. Solvent concentration profiles of the first-contact miscible displacement of Figure 2.57.

Figure 2.60 shows NMR-derived permeability images of another

layered sandstone core. The permeability for each voxel of the image was

calculated from T1 distributions using Eq.(2.69). Figure 2.61 shows a

comparison of NMR-derived permeability with flow-derived permeability

for five core samples. The agreement between the two sets of permeability

values is reasonable for four of the five measurements.

Page 166: +Peters Ekwere j. - Petrophysics

2-123

Figure 2.60. Permeability images of a layered sandstone core.

Page 167: +Peters Ekwere j. - Petrophysics

2–124

Figure 2.61. A comparison of NMR-derived permeability with flow-derived

permeability of core samples.

2.7.8 A Comparison of Various Porosity Measurements for Shaly Sand

Figure 2.62 shows the porosities measured by the various logging

tools along with the porosity from core analysis. This figure shows that

the total porosity measured by a neutron tool is larger than those

measured by the density tool, the NMR tool or the sonic tool. The total

porosity from the sonic tool is less than that of the density tool because

the density tool measures the porosity in isolated pore whereas the sonic

Page 168: +Peters Ekwere j. - Petrophysics

2-125

tool does not. If isolated pores are present, the effective porosity

measured in core analysis will be less than the effective permeability

measured by NMR. Thus, Figure 2.62 could be helpful in reconciling

differences in the porosity measurements from the various methods.

Figure 2.62. A comparison of porosity measurements by various methods for a shaly sand.

2.8 RESERVE ESTIMATION PROJECT

The objective of this project is to compute the recoverable oil

reserve and the anticipated undiscounted net cash flow for a new oilfield

discovery using Monte Carlo Simulation. The petrophysical parameters

that go into the reserve estimation are uncertain and as such should be

Page 169: +Peters Ekwere j. - Petrophysics

2–126

treated as random variables with distinct probability distributions. The

outcome of the simulation will be a range of reserve and net cash flow

estimates with their associated probabilities and uncertainties. This

project gives an example practical application of the petrophysical

properties of porosity and water saturation, which are the main subjects

of this chapter. The project also introduces the important subject of risk

analysis of petroleum development, which should be of interest to all

petroleum engineers.

2.8.1 Reserve Estimation

Based on volumetric considerations (Figure 2.63), the recoverable

oil reserve is given by

( )7758 1 wr f

o

Ah SN R

Bφ −

= (2.78)

where

Nr = recoverable oil reserve (stock tank barrels, STB)

A = area of the reservoir (acres)

h = net pay thickness (feet)

φ = porosity (fraction)

Sw = average water saturation (fraction)

Rf = recovery factor (fraction)

Bo = oil formation volume factor (reservoir barrels/stock tank barrel, RB/STB)

7758 = conversion constant (barrels/acre-foot)

Page 170: +Peters Ekwere j. - Petrophysics

2-127

Figure 2.63. Reservoir volume.

To account for the uncertainties in the variables on the right side

of Eq.(2.78) various experts within the company have been requested to

provide their best estimates of these variables based on their professional

judgments. These estimates are shown in Table 2.8 along with their

assumed probability distributions.

Table 2.8 Reservoir parameter estimates from experts.

Property Minimum (x1) Most Likely (x2) Maximum (x3) Probability

Distributio

n

A (acres) 2500 6000 9000 Triangular

h (ft) 200 300 500 Triangular

φ 0.15 0.25 0.35 Triangular

Bo (RB/STB) 1.20 1.30 1.35 Triangular

RF 0.20 0.40 Uniform

Oil Price ($/STB) 25 30 40 Triangular

Page 171: +Peters Ekwere j. - Petrophysics

2–128

Based on preliminary evaluation, the water saturation has been

determined to correlate with porosity as shown in Eq.(2.79).

0.325 0.500wS φ= − (2.79)

2.8.2 Economic Evaluation

In order to assess the profitability of the proposed development, it

will be necessary to perform a detailed year-by-year discounted cash flow

projection for the field. However, for the purpose of this project, a

preliminary undiscounted net cash flow analysis will be sufficient. It is

assumed that the petroleum fiscal regime applicable to this field is a

royalty/tax system.

The undiscounted net cash flow is given by

Net Cash Flow (NCF) = (Gross Revenue - Royalty - Costs)(1 - Tax Rate) (2.80)

Gross Revenue = Reserve x Price (2.81)

For this project, assume the following:

Royalty = 12.5% of Gross Revenue

Costs (CAPEX + OPEX) = 38% of Gross Revenue

(CAPEX = capital expenditures, OPEX = operating expenses)

Tax Rate = 40%

Page 172: +Peters Ekwere j. - Petrophysics

2-129

2.8.3 Simulation Procedure

You should draw a random sample for each variable of interest

from its probability distribution and compute the recoverable reserve and

the net cash flow for each iteration of the simulation. Each variable can

be sampled using a random number generator. A different random

number should be used to sample each variable because the variables

are assumed to be uncorrelated, except for the water saturation, which is

correlated to porosity. Normally, you should perform enough simulation

iterations so that the means and standard deviations of the reserve and

the net cash flow are approximately constant (i.e., they are no longer

sensitive to the number of iterations). This could require several

thousand iterations. For this exercise, you should perform at least 5,000

iterations.

2.8.4 Sampling Procedure

Presented herein is the procedure for sampling from a triangular

distribution and a uniform distribution using a random number

generator. Figure 2.64 shows the probability density function (pdf) for a

triangular distribution. The first step is to compute the cumulative

distribution function (F) as a function of x. Two cases are examined.

Case 1. x1≤ x ≤x2

For this case, the probability density function is given by

( )( )( )

11

3 1 2 1

2( )

x xf x

x x x x−

=− −

(2.82)

Page 173: +Peters Ekwere j. - Petrophysics

2–130

Figure 2.64. Probability density function for a triangular distribution.

The cumulative distribution function (F), which is the probability that x

is less than or equal to a prescribed value, is obtained by integrating

Eq.(2.82) to obtain

( )

( )( )1

21

13 1 2 1

( )x

x

x xF f x dx

x x x x−

= =− −∫ (2.83)

Solving Eq.(2.83) for x gives

( )( )1 3 1 2 1x x F x x x x= + − − (2.84)

It turns out that F is uniformly distributed between 0 and 1, just like the

random number generator in spreadsheets or other computer software.

Therefore, to sample from the first part of a triangular distribution,

generate a uniformly distributed random number (Rn) between 0 and 1

and substitute it for F in Eq.(2.84) to obtain the sample value for x as

( )( )1 3 1 2 1x x Rn x x x x= + − − (2.85)

Page 174: +Peters Ekwere j. - Petrophysics

2-131

Case 2. x2≤ x ≤x3

For this case, the probability density function is given by

( )

( )( )3

23 1 3 2

2( )

x xf x

x x x x−

=− −

(2.86)

The cumulative distribution function (F) is obtained by integrating

Eqs.(2.82) and (2.86) to obtain

( )( )( )

2

1 2

23

1 23 1 3 2

( ) ( ) 1x x

x x

x xF f x dx f x dx

x x x x−

= + = −− −∫ ∫ (2.87)

Solving Eq.(2.87) for x gives

( )( )( )3 3 1 3 21x x F x x x x= − − − − (2.88)

To sample from the second part of the triangular distribution, substitute

the random number for F in Eq.(2.88) to obtain

( )( )( )3 3 1 3 21x x Rn x x x x= − − − − (2.89)

For each iteration, it is necessary to test the random number to

determine if Eq.(2.85) or Eq.(2.89) should be used to calculate x. Such a

test is straightforward. For x = x2, Eq. (2.85) gives the critical value of Rn

as

( )( )

2 1

3 1

x xRn

x x−

=−

(2.90)

Page 175: +Peters Ekwere j. - Petrophysics

2–132

Eq.(2.85) should be used to calculate x if

( )( )

2 1

3 1

x xRn

x x−

≤−

(2.91)

Otherwise, Eq.(2.89) should be used to calculate x.

Figure 2.65 shows a graphical demonstration of Monte Carlo

sampling using the triangular distribution for net pay thickness of Table

2.8. The upper part of the figure shows the probability density function

whereas the lower part shows the cumulative distribution function. Also

shown in the lower part of the figure is the sampled value of 380 ft of net

pay for a random number of 0.760. The data used to construct Figure

2.65 are shown in Table 2.9.

Figure 2.66 shows the probability density function for a uniform

distribution. For this case, the probability density function is given by

( )3 1

1( )f xx x

=−

(2.92)

The cumulative distribution function (F) is obtained by integrating

Eq.(2.92) to obtain

( )( )1

1

3 1

( )x

x

x xF f x dx

x x−

= =−∫ (2.93)

Solving Eq.(2.93) for x gives

( )1 3 1x x F x x= + − (2.94)

Page 176: +Peters Ekwere j. - Petrophysics

2-133

0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0 100 200 300 400 500 600

x (ft)

f(x)

F

0.0000.1000.2000.3000.4000.5000.6000.7000.8000.9001.000

0 100 200 300 400 500 600

x (ft)

F

Figure 2.65. Graphical demonstration of Monte Carlo sampling

Table 2.9: Data for Monte Carlo Sampling of Net Pay Thickness

x F f(x)

200 0.000 0.0000

0

210 0.003 0.0006

7

220 0.013 0.0013

Page 177: +Peters Ekwere j. - Petrophysics

2–134

3

230 0.030 0.0020

0

240 0.053 0.0026

7

250 0.083 0.0033

3

260 0.120 0.0040

0

270 0.163 0.0046

7

280 0.213 0.0053

3

290 0.270 0.0060

0

300 0.333 0.0066

7

310 0.398 0.0063

3

320 0.460 0.0060

0

330 0.518 0.0056

7

340 0.573 0.0053

3

350 0.625 0.0050

0

360 0.673 0.0046

7

370 0.718 0.0043

3

380 0.760 0.0040

0

390 0.798 0.0036

Page 178: +Peters Ekwere j. - Petrophysics

2-135

7

400 0.833 0.0033

3

410 0.865 0.0030

0

420 0.893 0.0026

7

430 0.918 0.0023

3

440 0.940 0.0020

0

450 0.958 0.0016

7

460 0.973 0.0013

3

470 0.985 0.0010

0

480 0.993 0.0006

7

490 0.998 0.0003

3

500 1.000 0.0000

0

Page 179: +Peters Ekwere j. - Petrophysics

2–136

Figure 2.66. Probability density function for a uniform distribution

To sample from a uniform distribution, substitute the random number

for F in Eq.(2.94) to obtain

( )1 3 1x x Rn x x= + − (2.95)

The sampling procedure outlined above is the basic Monte Carlo

Sampling procedure. It is not a very efficient sampling technique because

there is no guarantee that all parts of the distribution will be sampled

equally. A more efficient and sophisticated stratified sampling procedure,

known as the Latin Hypercube Sampling, is available. This sampling

procedure ensures that all parts of the distribution are sampled equally

and will result in a faster convergence of the simulation to the final

results than the traditional Monte Carlo Sampling Method.

A brief description of the Latin Hypercube Sampling is as follows.

The cumulative distribution function, 0<F<1, for each variable is divided

into n subintervals of equal probability and numbered 0, 1, 2, 3, …, n-1.

Page 180: +Peters Ekwere j. - Petrophysics

2-137

The objective is to sample from each subinterval during the first n

iterations. However, the subinterval to be sampled is selected randomly.

For example, if n = 100, the subintervals will be numbered 0, 1, 2, 3, …,

99. Each subinterval can be selected randomly by generating a random

integer from 0 to 99. The subintervals could be picked at random in

advance and stored in a one-dimensional array of 100 elements. Let’s

look at the array for the Area, A(100), as an example. The contents of this

array after the random selections might look like this:

A(1) = 45

A(2) = 2

A(3) = 7

A(4) = 92

A(5) = 23

A(6) = 0

A(100) = 8

Make sure every subinterval is represented in the array. A similar array

is generated with a new set of random numbers for each of the other

variables. During the first simulation iteration, the area will be sampled

from subinterval 45. To do so, generate a random number (r) between 0

and 1, say 0.263. The random number used to sample subinterval 45 is

computed as

(1) 45 0.263 0.45263100 100 100 100A rRn = + = + = (2.96)

Page 181: +Peters Ekwere j. - Petrophysics

2–138

In the second iteration, the area will be sampled from subinterval 2 by

generating another random number (r) between 0 and 1, say 0.950 and

computing

(2) 2 0.950 0.02950100 100 100 100A rRn = + = + = (2.97)

At the end of the first 100 iterations, each of the subintervals has been

sampled once. A similar sampling procedure is applied to the other

variables.

To perform the next 100 iterations of the simulation, the sequence

in which the subintervals are sampled is picked randomly. For example,

for the second 100 iterations, the array A may look like this:

A(1) = 3

A(2) = 86

A(3) = 50

A(4) = 14

A(5) = 6

A(6) = 99

A(100) = 72

The sampling for the second 100 iterations then proceeds in the same

manner as for the first 100 iterations. Thus, the entire simulation is

performed in increments of 100 iterations until the planned total number

of iterations is achieved. Figure 2.67 shows qualitatively the effect of

Page 182: +Peters Ekwere j. - Petrophysics

2-139

stratification in Latin Hypercube sampling from a normal distribution for

n = 20. The bands get progressively wider towards the tails of the

distribution as the probability density drops off.

Figure 2.67. Effect of stratification in Latin Hypercube sampling.

2.8.5 Simulation Output

Based on your simulation results, calculate the following for the

Reserve (Nr) and Net Cash Flow (NCF) estimates:

1. Minimum 2. Maximum 3. Mean 4. Standard deviation 5. Skewness 6. Kurtosis

Page 183: +Peters Ekwere j. - Petrophysics

2–140

7. Mode (most likely value) 8. P90 value (There is a 90% probability that the variable of interest is

equal to or greater than the P90 value – Low Value) 9. P50 value (median) 10. P10 value (There is a 10% probability that the variable of interest is equal to or greater than the P10 value – High Value)

Note: Based on the above definitions, P10 > P90.

Plot the following graphs:

1. Reserve histograms (Frequency vs N, and Frequency vs natural log of N)

2. Net Cash Flow (NCF) histograms (Frequency vs NCF, and Frequency vs natural log of NCF)

3. Cumulative Distribution Function (F) for the reserve (F vs N) 4. Expectation Curve for the reserve ( (1– F) vs N on the same graph

as F vs N) 5. Cumulative Distribution Function (F) for the net cash flow (F vs

NCF) 6. Expectation Curve for the net cash flow ( (1– F) vs NCF on the

same graph as F vs NCF) 7. Histograms for A, h, φ, Sw, Bo, RF and Oil Price based on your

sampling to see how close they are to their theoretical input probability density functions .

Suggestion:

You can easily perform this simulation in Excel with VBA or by

writing a high level computer program in Fortran, C++ or Matlab. The

choice is yours. Please do not perform the simulation with a commercial

risk analysis software such as @Risk or Crystal Ball because I want you

to experience the simulation first hand rather than using the commercial

software as a black box.

Note that a similar simulation approach can be used to compute

the recoverable gas reserve in the case of a gas reservoir. In this case, the

appropriate equations are as follows:

Page 184: +Peters Ekwere j. - Petrophysics

2-141

( )43560 1 wr

g

Ah SG RF

Bφ −

= (2.98)

0.02827gZTBP

= (2.99)

where

Gr = recoverable gas reserve (standard cubic feet, scf)

Bg = gas formation volume factor (reservoir cubic feet/scf)

Z = gas compressibility factor (dimensionless)

T = absolute reservoir temperature (ºRankine)

P = reservoir pressure (psia)

43560 = conversion constant (ft2/acre)

2.9 PORE VOLUME COMPRESSIBILITY

Reservoir rock at in situ conditions is subjected to overburden

stress, whereas at the surface, recovered core has been relieved of the

overburden stress. It is not usual to perform routine porosity

measurements under stress approaching reservoir conditions. Because

of this, laboratory-measured porosities are generally expected to be

higher than in situ values. Pore volume compressibility can be used to

correct laboratory-measured porosity to an in situ value and for other

reservoir engineering calculations.

Page 185: +Peters Ekwere j. - Petrophysics

2–142

In general, the isothermal coefficient of compressibility of a

material is defined as

1

T

VcV P

∂⎛ ⎞= − ⎜ ⎟∂⎝ ⎠ (2.100)

where the negative sign convention is used to ensure that c is a positive

number because when a material is compressed, its volume decreases

thereby making T

VP

∂⎛ ⎞⎜ ⎟∂⎝ ⎠

a negative quantity. Pore volume compressibility is

defined without the negative sign as

1 1pf

Tp T

Vc

V P Pφ

φ∂⎛ ⎞ ∂⎛ ⎞= =⎜ ⎟ ⎜ ⎟∂ ∂⎝ ⎠⎝ ⎠

(2.101)

where P is the pore pressure (internal pressure) rather than the

overburden pressure (external pressure). When the pore pressure

declines during production, the overburden pressure causes the reservoir

to compact. Thus, a decrease in pore pressure causes a decrease in the

pore volume or porosity thereby making T

VP

∂⎛ ⎞⎜ ⎟∂⎝ ⎠

or TP

φ∂⎛ ⎞⎜ ⎟∂⎝ ⎠

positive.

For many years, the petroleum industry has relied on Hall’s (1953)

correlation, shown in Figure 2.68, for estimating pore volume

compressibility. This correlation was developed from measurements on

seven consolidated limestone and five consolidated sandstone samples, a

rather small dataset for a universal correlation.

Compressibilities, however, are highly affected by reservoir type

and overburden conditions. Newman (1973) has presented a more

comprehensive pore volume compressibility data based on 256 samples.

Page 186: +Peters Ekwere j. - Petrophysics

2-143

Of the 256 samples, 197 were sandstones from 29 sandstone reservoirs,

and 59 were limestones from 11 limestone reservoirs. Figure 2.69 shows

a correlation of the pore volume compressibilities with initial sample

porosities for all 256 samples. Also shown is Hall's correlation. Viewed

collectively, the data show little or no correlation between pore volume

compressibility and initial sample porosity. The data show considerable

disagreement with Hall's correlation.

Figure 2.68. Pore volume compressibility versus porosity (Hall, 1953).

Page 187: +Peters Ekwere j. - Petrophysics

2–144

Figure 2.69. Pore volume compressibility at 75% lithostatic pressure versus initial sample porosity for sandstone and limestone samples

(Newman, 1973).

Newman divided his data into four qualitative classifications in an

effort to improve the correlation between pore volume compressibility and

initial porosity. He classified the samples as (1) consolidated limestone,

(2) consolidated sandstone, (3) poorly consolidated (friable) sandstone,

and (4) unconsolidated sandstone. Poorly consolidated or friable

sandstone could be broken by hand whereas consolidated sandstone

Page 188: +Peters Ekwere j. - Petrophysics

2-145

could not. Unconsolidated sandstone would collapse under its own

weight if it is not protected in a sleeve.

Figure 2.70 shows the correlation for consolidated limestones

along with the correlations from Hall and van der Knaap (1959).

Qualitatively, all three dataset show a general decrease in pore volume

compressibility with an increase in porosity. However, there is

considerable scatter in the data.

Figure 2.70. Pore volume compressibility at 75% lithostatic pressure versus initial sample porosity for limestone samples (Newman, 1973).

Page 189: +Peters Ekwere j. - Petrophysics

2–146

Figure 2.71 shows the correlation for consolidated sandstones

along with Hall's correlation. Qualitatively, the data show a general

decrease in pore volume compressibility with an increase in porosity with

some scatter. Newman's data decrease more rapidly with an increase in

porosity than Hall's data.

Figure 2.71. Pore volume compressibility at 75% lithostatic pressure

versus initial sample porosity for consolidated sandstone samples (Newman, 1973).

Page 190: +Peters Ekwere j. - Petrophysics

2-147

Figure 2.72 shows the correlation for poorly consolidated or friable

sandstones along with Hall's correlation. There appears to be no

correlation between pore volume compressibility and porosity for this

class of sandstones.

Figure 2.72. Pore volume compressibility at 75% lithostatic pressure

versus initial sample porosity for poorly consolidated sandstone samples (Newman, 1973).

Page 191: +Peters Ekwere j. - Petrophysics

2–148

Figure 2.73 shows the correlation for unconsolidated sandstones

along with Hall's correlation. For this class of sandstones, the pore

volume compressibility increases with an increase in porosity, in contrast

to the trend for consolidated samples.

Figure 2.73. Pore volume compressibility at 75% lithostatic pressure versus initial sample porosity for unconsolidated sandstone samples

(Newman, 1973).

Page 192: +Peters Ekwere j. - Petrophysics

2-149

Newman's data clearly show that there is no universal correlation

for pore volume compressibility and porosity. For detailed analysis, pore

volume compressibility should be measured in the laboratory on core

samples of interest. Figure 2.74 shows a schematic diagram of a typical

device for measuring pore volume compressibility. These measurements

are fairly difficult to make, and must include calibrations for the

compressibilities of the fluids and the equipment.

Figure 2.74. Typical apparatus for measuring pore volume

compressibility (Hall, 1953).

Pore volume compressibility is not always constant. For example,

in abnormally pressured reservoirs, it is common to find that pore

volume compressibility at high reservoir pressures is much larger than

Page 193: +Peters Ekwere j. - Petrophysics

2–150

the pore volume compressibility at low pressures. This is due to

compaction effects as the pore pressure declines.

Pore volume compaction can lead to undesirable operational

problems. A famous case of this is the subsidence of Ekofisk field in the

Norwegian sector of North Sea. The seafloor subsided about 3 meters

from 1971 to 1984 due to reservoir depletion. Continuous monitoring

indicated subsidence of 33 to 38 cm per year. By 1997, the cumulative

subsidence was 7.8 meters. In 1998, the operator undertook expensive

measures to jack up their platforms by 6 meters. On the other hand,

pore volume compaction provides valuable reservoir drive energy to

increase production from the field.

Total system compressibility is a parameter of importance in

pressure transient testing and reservoir material balance analysis,

especially above the bubble point pressure of the oil. Total system

compressibility is defined as the combined compressibility of the pore

volume and all the fluids saturating the medium:

t o o w w g g fc S c S c S c c= + + + (2.102)

NOMENCLATURE

amf = activity coefficient of mud filtrate

aw = activity coefficient of formation water

A = atomic weight

Bg = gas formation volume factor

Bo = oil formation volume factor

B0 = magnetic field strength

Page 194: +Peters Ekwere j. - Petrophysics

2-151

BVI = bound fluid index

c = isothermal coefficient of compressibility

cf = isothermal coefficient of compressibility of pore volume

cg = isothermal coefficient of compressibility of gas

co = isothermal coefficient of compressibility of oil

cw = isothermal coefficient of compressibility of water

dh = borehole diameter

di = diameter of invaded zone

ds1 = slice thickness

dxo = diameter of flushed zone

D = diffusion coefficient

Ec = electrochemical potential

Ek = electrokinetic potential

Ekmc = electrokinetic potential of mud cake

F = formation resistivity factor

FFI = free fluid index

FOV = field of view

G = gradient

GPE = phase encode gradient

GRO = readout gradient

Gss = slice selection gradient

Gr = recoverable gas reserve

h = net pay thickness

I = formation resistivity index

k = permeability

m = cementation factor

M = net magnetization

Mz = net magnetization in the z direction

n = saturation exponent

Nr = recoverable oil reserve

MNR = nuclear magnetic resonance

Page 195: +Peters Ekwere j. - Petrophysics

2–152

P = pressure

RF = radiofrequency

Rf = recovery factor

Ri = resistivity of invaded zone

Rmf = resistivity of mud filtrate

Rn = random number

Ro = resistivity of water saturated zone

Rs = resistivity of surrounding formation

Rt = true formation resistivity

Rw = resistivity of formation water

Rxo = resistivity of flushed zone

S = surface area of pore

Sh = hydrocarbon saturation

Sg = gas saturation

So = oil saturation

Soi = initial oil saturation

Sor = residual oil saturation

Sxo = water saturation of flushed zone

Sw = water saturation

Swirr = irreducible water saturation

t = time

tpe = duration of phase encoding gradient pulse

tro = duration of frequency encoding gradient pulse

T = temperature

T1 = longitudinal relaxation time or spin-lattice relaxation time

T2 = transverse relaxation time or spin-spin relaxation time

T2gm = geometric mean time *

2T = apparent transverse relaxation time spin-spin relaxation time

T2bulk = transverse relaxation time of bulk fluid

T2surface = transverse relaxation time of surface fluid

TE = echo time

Page 196: +Peters Ekwere j. - Petrophysics

2-153

TR = repetition time

v = compressional wave velocity

V = volume of pore

Vp = pore volume

Vb = bulk volume

Vs = grain volume

Z = atomic number

φ = porosity

φe = effective porosity

φtotal = total porosity

ρ1 = NMR T1 surface relaxivity

ρ2 = NMR T2 surface relaxivity

ρa = apparent density

ρb = bulk density

ρe = electron density

ρf = fluid density

ρm = rock matrix density

ΔB0 = magnetic field inhomogeneity

ΔPE = pixel size in the phase encoding direction

ΔRO = pixel size in the readout direction

Δv = voxel size

Δt = interval transit time

Δtf = fluid interval transit time

Δtm = rock matrix interval transit time

Δtsh = shale interval transit time

∇B = magnetic field gradient

α = regularization constant or smoothing parameter

λ = thickness of surface monolayer

ε(tj) = noise

τ = interpulse spacing time

Page 197: +Peters Ekwere j. - Petrophysics

2–154

ωo = Lamor frequency

ωr = Lamor frequency at position r

γ = gyromagnetic ratio

Ψair = x-ray attenuation coefficient of air (= 0)

Ψbrine = x-ray attenuation coefficient of brine

Ψdry = x-ray attenuation coefficient of dry sample

Ψm = x-ray attenuation coefficient of rock matrix

Ψwet = x-ray attenuation coefficient of brine saturated sample

REFERENCES AND SUGGESTED READINGS

Amyx, J.W., Bass, D.M., Jr. and Whiting, R.L. : Petroleum Reservoir Engineering, McGraw-Hill Book Company, New York, 1960.

Anderson, G. : Coring and Core Analysis Handbook, Petroleum Publishing Company, Tulsa, Oklahoma, 1975.

Archer, J.S. and Wall, C.G. : Petroleum Engineering, Graham & Trotman, London, England, 1986.

Archie, G.E. :“The Electrical Resistivity Log as an Aid in Determining Some Reservoir Characteristics,” Trans. AIME (1942) 146, 54-62.

Baker Atlas Publication: Introduction to Wireline Log Analysis, 1992.

Bassiouni, Z. : Theory, Measurement, and Interpretation of Well Logs, SPE Textbook Series, Vol. 4, Society of Petroleum Engineers, Richardson, TX, 1994.

Beard, D.C. and Weyl, P.K. : “Influence of Texture on Porosity and Permeability of Unconsolidated Sand,” AAPG Bull. (Feb. 1973) 57, 349-369.

Bradley, H.B. (Editor-in-Chief) : Petroleum Engineering Handbook, SPE, Richardson, Texas, 1987.

Calhoun, J.C. : Fundamentals of Reservoir Engineering, University of Oklahoma Press, Norman, Oklahoma, 1982.

Page 198: +Peters Ekwere j. - Petrophysics

2-155

Carman, P.C. : "Fluid Flow Through A Granular Bed,” Trans. Inst. Chem. Eng. London (1937) 15, 150-156.

Carman, P.C. : "Determination of the Specific Surface of Powders,” J. Soc. Chem. Indus (1938) 57, 225-234.

Carpenter, C.B. and Spencer, G.B. : “Measurements of the Compressibility of Consolidated Oil-Bearing Sandstones,” RI 3540, USBM (Oct. 1940).

Coats, G.R., Xiao, L. and Prammer, M.G. : NMR Logging: Principles and Applications, Halliburton Energy Services, Houston, 1999.

Collins, R.E. : Flow of Fluids Through Porous Materials, Research & Engineering Consultants Inc., 1990.

Core Laboratories Inc. : Special Core Analysis, 1976.

Corey, A.T. : Mechanics of Heterogeneous Fluids in Porous Media, Water Resources Publications, Fort Collins, Colorado, 1977.

Cosse, R. : Basics of Reservoir Engineering, Editions Technip, Paris, 1993.

Dobrynin, V.M. : "The Effect of Overburden Pressure on Some Properties of Sandstones,” Soc. Pet. Eng. J. (Dec. 1962) 360-366.

Dunn, K.J., Bergman, D.J. and Latorraca, G.A.: Nuclear Magnetic Resonance Petrophysical and Logging Applications, Handbook of Geophysical Exploration, Seismic Exploration Vol. 32, Pergamon, New York, 2002.

Fatt, I. : “Pore Volume Compressibilities of Sandstone Reservoir Rocks,” Trans., AIME (1958) 213, 362-364.

Fraser, H.J. and Graton, L.C. : "Systematic Packing of Spheres - With Particular Relation to Porosity and Permeability,” J. Geol., Vol. 43, No. 8, Nov. - Dec., 1935, 789-909.

Geertsma, J. : “The Effect of Fluid Pressure Decline on Volumetric Changes of a Porous Medium,” Trans. AIME (1957) 210, 331-340.

Hall, H.N. : “Compressibility of Reservoir Rocks,” Trans., AIME (1953) 198, 309-311.

Haughey, D.P. and Beveridge, G.S.G. :”Structural Properties of Packed Beds - A Review,” Can. J. Chem. Eng., Vol. 47 (April 1969) 130-149.

Jennings, H.J. : “How to Handle and Process Soft and Unconsolidated Cores,” World Oil (June 1965) 116-119.

Jorden, J.R. and Campbell, F.L. : Well Logging I - Rock Properties, Borehole Environment, Mud and Temperature Logging, Monograph

Page 199: +Peters Ekwere j. - Petrophysics

2–156

Vol. 9, Society of Petroleum Engineers of AIME , Richardson, Texas, 1984, Ch. 2.

Keelan, D.K. : "A Critical Review of Core Analysis Techniques” The Jour. Can. Pet. Tech. (April-June 1972) 42-55.

Kleinberg, I. : “Probing Oil Wells with NMR,” The Industrial Physicist, (June/July 1996) 18-21.

Li, Ping: Nuclear Magnetic Resonance Imaging of Fluid Displacements in Porous Media, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1997.

Mann, R.L. and Fatt, I. : “Effect of Pore Fluids on the Elastic Properties of Sandstones,” Geophysics (1960) 25, 433-444.

Mayer-Gurr, A. : Petroleum Engineering, John Wiley & Sons, New York, 1976.

Monicard, R.P. : Properties of Reservoir Rocks, Gulf Publishing Company, Houston, TX, 1980.

Neasham, J.W.: "The Morphology of Dispersed Clay in Sandstone Reservoirs and Its Effect on Sandstone Shaliness, Pore Space and Fluid Flow Properties," SPE 6858, Presented at the 52nd Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, Denver, Oct. 9-12, 1977.

Newman, G.H. : "Pore-Volume Compressibility of Consolidated, Friable, and Unconsolidated Reservoir Rocks Under Hydrostatic Loading,” J. Pet. Tech. (Feb. 1973) 129-134.

Peters, E.J. and Afzal, N. : “Characterization of Heterogeneities in Permeable Media with Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 7, No. 3/4, (May 1992) 283-296.

Peters, E.J. and Hardham, W.D. : “Visualization of Fluid Displacements in Porous Media Using Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 4, No. 2, (May 1990) 155-168.

Pirson, S.J. : Oil Reservoir Engineering, McGraw-Hill Book Company, Inc., New York, 1958.

Ruth, D. and Pohjoisrinne, T. : "The Precision of Grain Volume Porosimeters” The Log Analyst (Nov.-Dec. 1993) 29-36.

Schlumberger Publication: Log Interpretation Principles/Applications, Schlumberger Educational Series, 1987.

Page 200: +Peters Ekwere j. - Petrophysics

2-157

Tamari, S. : "Optimum Design of the Constant-Volume Gas Pycnometer for Determining the Volume of Solid Particles," Meas. Sci. Tech.,Vol.15, 2004, 549-558.

Tiab, D. and Donaldson, E.C. : Petrophysics, Second Edition, Elsevier, New York, 2004.

van der Knaap, W. : "Nonlinear Behavior of Elastic Porous Media,” Trans., AIME (1959) 216, 179-187.

Vose, D. : Quantitative Risk Analysis: A Guide to Monte Carlo Simulation Modeling, John Wiley and Sons, New York, 1998.

Winsauer, W.O., Shearin, H.M., Jr., Masson, P.H. and Williams, M. : "Resistivity of Brine-Saturated Sands in Relation to Pore Geometry," AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 253-277.

Wyllie, M.R.J. and Rose, W.D. : "Some Theoretical Considerations Related to the Quantitative Evaluation of the Physical Characteristics of Reservoir Rock from Electrical Log Data,” Trans., AIME (1950) 189, 105-118.

Wyllie, M.R.J. and Spangler, M.B. : "Application of Electrical Resistivity Measurements to Problems of Fluid Flow in Porous Media,” AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 359-403.

Wyllie, M.R.J., Gregory, A.R. and Gardner, G.H.F. : "Elastic Wave Velocities in Heterogeneous and Porous Media” Geophysics, Vol. 21, No. 1 (1956) 41-70.

Page 201: +Peters Ekwere j. - Petrophysics

3-1

CHAPTER 3

PERMEABILITY

3.1 DEFINITION

Permeability gives an indication of the porous medium’s ability to

transmit fluids (i.e., permit fluid flow). It is defined through Darcy’s law. For a

horizontal system, Darcy’s law for single phase flow in differential form is

kA dPqdxμ

= − (3.1)

For steady-state linear flow of a single phase liquid in a horizontal medium as

shown in Figure 3.1, Darcy’s Law may be integrated to give

( )1 2P PkA kA PqL Lμ μ− Δ

= = (3.2)

Figure 3.1. Linear flow geometry.

Page 202: +Peters Ekwere j. - Petrophysics

3-2

Eq(3.2) can be solved for the absolute permeability of the porous medium as

q LkA P

μ=

Δ (3.3)

The absolute permeability is the permeability of the medium when it is

saturated 100% by a single phase, non-reactive liquid. It is a property of the

porous medium and is independent of the fluid properties.

Eq.(3.2) can be written as

LP qkAμ⎛ ⎞Δ = ⎜ ⎟

⎝ ⎠ (3.4)

Eq.(3.4) is useful for processing laboratory data to determine the absolute

permeability of a sample using a non-reactive liquid. If a steady state liquid

flow experiment is performed at several rates, Eq.(3.4) shows that a graph of

ΔP versus q should be linear with a slope, m, given by

LmkAμ

= (3.5)

from which the absolute permeability can be calculated as

LkmAμ

= (3.6)

Table 3.1 shows the data from such an experiment performed on an

unconsolidated sandpack using Dow Corning mineral oil. The fluid and

sandpack properties were as follows:

μ = 105.363 cp

L = 115.6 cm

d = 4.961 cm

φ = 37.80%

Page 203: +Peters Ekwere j. - Petrophysics

3-3

Table 3.1: Pressure Drop and Flow Rate Data for Steady State Flow Experiment in a Sandpack (from Peters, 1979).

q

(cc/s)

ΔP

(atm)

0 0

0.0014 0.0476

0.0556 1.9284

0.0889 3.0573

0.1333 4.5439

0.2222 7.5303

0.3111 10.465

Figure 3.2 shows the graph of ΔP versus q from the experiment. The data plot

as a straight line through the origin, thereby verifying the validity of Darcy's

law for the experiment. The slope of the line is m = 33.658. The permeability

is calculated with Eq.(3.6) as

( )( )

( )2

105.363 115.618.72

4.96133.6582

= =⎡ ⎤⎛ ⎞⎢ ⎥⎜ ⎟

⎝ ⎠⎢ ⎥⎣ ⎦

darcys.

If the graph of Figure 3.2 were nonlinear, then Darcy's law would not be

valid for the experiment. This situation would occur if the liquid reacted with

the porous medium. An example of such a reaction would occur if the

experiment were performed on a sandstone containing a significant amount of

montmorrillonite-type clays using fresh water. Can you sketch how ΔP would

vary with q in such an experiment?

Page 204: +Peters Ekwere j. - Petrophysics

3-4

Figure 3.2. Graph of ΔP verus q for steady state flow of a non-reactive liquid

through an unconsolidated sandpack.

Table 3.2 compares Darcy units with oilfield units. Using the

information in the table, Eq.(3.2) can be written in oilfield units as

0.001127kA PqB Lμ

Δ= (3.7)

or

1887.2

kA PqB Lμ

Δ= (3.8)

For radial flow into a wellbore as shown in Figure 3.3, Darcy’s law may

be expressed in radial coordinates in oilfield units as

Page 205: +Peters Ekwere j. - Petrophysics

3-5

( )0.00708

ln

e w

e

w

kh P Pq

rBr

μ

−=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.9)

or

( )1141.2

ln

e w

e

w

kh P Pq

rBr

μ

−=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.10)

Table 3.2 Comparison of Darcy and Oilfield Units

Variable Darcy Unit Oilfield Unit Pressure atm psia Time second day Flow rate cm3/s STB/D Wellbore radius cm ft Well drainage radius cm ft Porosity fraction fraction Permeability darcy millidarcy Pay thickness cm ft Fluid viscosity cp cp Compressibility atm-1 psi-1

Figure 3.3. Radial flow into a wellbore.

Page 206: +Peters Ekwere j. - Petrophysics

3-6

3.2 DIMENSIONS AND UNITS OF PERMEABILITY

Eq.(3.2) can be rearranged to determine the dimensions of permeability

as follows:

[ ] [ ][ ][ ][ ][ ]q L

kA p

μ=

Δ (3.11)

where the square brackets refer to the dimensions of the enclosed variable.

Using mass (M), length (L) and time (T) as the fundamental dimensions, the

dimensions of the variables on the right side of Eq.(3.11) are

[ ]3Lq

T=

[ ] MLT

μ =

[ ]L L=

[ ] 2A L=

[ ] 2

MpLT

Δ =

Thus,

[ ]( )

( )

3

22

L M LT LT

kML

LT

⎛ ⎞⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠=⎛ ⎞⎜ ⎟⎝ ⎠

or

[ ] 2k L= (3.12)

Eq.(3.12) shows that the permeability of a porous medium has the dimensions

of length squared. This means that permeability is proportional to the square

of some characteristic dimension of the porous medium.

Page 207: +Peters Ekwere j. - Petrophysics

3-7

Based on Eq.(3.12), a rational unit for permeability would be foot

squared or centimeter squared. Both units were found to be too large for use

with porous media. Therefore, the petroleum industry adopted the darcy as

the unit of permeability. It can be shown that

9 2 13 2 11 21 9.869 10 9.869 10 1.062 10darcy x cm x m x ft− − −= = = (3.13)

3.3 LABORATORY DETERMINATION OF PERMEABILITY

Permeability is an empirical parameter, which must be determined by

measuring all the parameters in Darcy’s law. Core permeabilities are usually

measured in the laboratory using dry gas (air, nitrogen or helium) as the

flowing fluid to minimize rock-fluid reaction. In this case, another form of

Darcy’s law for steady-state flow of an ideal gas must be used, recognizing

that the volumetric flow rate of a gas varies with pressure. For an ideal gas at

a constant temperature, Boyle's law can be written for a fixed mass of gas at

two conditions as

sc scqP q P= (3.14)

where qsc is the gas volumetric flow rate at a reference pressure Psc.

Substituting Eq.(3.14) into (3.1) and rearranging gives

21

2g

scsc

k A dPqP dxμ

⎛ ⎞= − ⎜ ⎟

⎝ ⎠ (3.15)

where kg is the permeability to gas of the medium. Eq.(3.15) can be integrated

to give the working equation for a gas permeameter as

2 2

1 2

2g

scsc

k A P PqP Lμ

⎛ ⎞−= ⎜ ⎟

⎝ ⎠ (3.16)

which can be used to calculate the permeability to gas as

( )2 2

1 2

2 sc scg

q LPkA P P

μ=

− (3.17)

Page 208: +Peters Ekwere j. - Petrophysics

3-8

The permeability to gas determined from Eq.(3.17) at the low pressures

typically used in the laboratory measurement is usually higher than the

absolute permeability of the porous medium because of an electro-kinetic

phenomenon known as the Klinkenberg (1941) effect. Klinkenberg effect is

due to the fact that at low mean pressures, the mean free path of the gas

molecules is about the same size as the pores in the rock. This gives rise to

gas slippage at the wall of the pores. As a result, the permeability to gas at

low pressure is higher than the absolute permeability of the porous medium.

Klinkenberg found that the permeability to gas, kg, is related to the absolute

permeability of the medium, kL, by the equation

1g Lbk kP

⎛ ⎞= +⎜ ⎟⎝ ⎠

(3.18)

where P is the mean gas pressure given by

1 2

2P PP +

= (3.19)

and b is a constant, which depends on the gas used in the measurement. The

absolute permeability (liquid permeability) of the medium can be determined

in the laboratory by measuring permeabilities to gas, kg, at different average

core pressures. A graph of kg versus 1P

yields a straight line with an

intercept equal to kL and a slope equal to kLb as shown in Figure 3.4.

Figue 3.5 shows the Klinkenberg correction plots for a core for gas

permeabilities measured with hydrogen, nitrogen and carbon dioxide. It

should be noted that the degree of gas slippage varies with the nature of the

gas. Hydrogen with the smallest molecules has the most gas slippage and

carbon dioxide with the largest molecules has the least gas slippage. All three

straight lines extrapolate to the absolute permeability of the core at infinite

mean pressure. The Klinkenberg effect is a laboratory scale phenomenon. It

is usually not important at the high mean pressures of a petroleum reservoir.

Page 209: +Peters Ekwere j. - Petrophysics

3-9

Figure 3.4. Klinkenberg permeability correction.

Figure 3.5. Permeability of a core sample to hydrogen, nitrogen and carbon dioxide. Absolute permeability of the core to isooctane = 2.55 md (from

Klinkenberg, 1941).

Page 210: +Peters Ekwere j. - Petrophysics

3-10

Figure 3.6 shows a schematic diagram of a gas permeameter. In this

figure, the gas volumetric flow rate, qsc, is measured at the outlet pressure, P2.

Therefore, in this case, P2 is the reference pressure, Psc.

Figure 3.6. Schematic diagram of a gas permeameter.

In general, permeability is an anisotropic property of a porous medium,

that is, it is directional. Routine core analyses are usually made on core

plugs drilled horizontally from a core. It is sometimes possible to specify that

plugs be cut along bedding planes. Plugs are sometimes cut vertically if it is

desired to obtain vertical as well as horizontal permeabilities. These plugs are

shown in Figure 3.7.

Page 211: +Peters Ekwere j. - Petrophysics

3-11

Figure 3.7. Possible core plug samples.

Because of interactions of fluids with reservoir rocks, absolute

permeabilities to different fluids may not always be the same. The

permeability to brine, for example, is often somewhat less than the

Klinkenberg-corrected gas permeability. For this reason, it is usually

beneficial to obtain at least a few brine permeability measurements, especially

if waterflooding is anticipated. These measurements are time-consuming, and

Page 212: +Peters Ekwere j. - Petrophysics

3-12

thus are more costly. Another permeability that is sometimes referenced is

the permeability to oil at irreducible water saturation (kor). This permeability

is known as the effective permeability to oil at irreducible water saturation. Its

value is typically around 80% to 98% of the absolute permeability depending

on the quality of the porous medium. Table 3.3 presents typical values for

clean, unconsolidated sandpacks.

Table 3.3: Petrophysical Data of Sandpacks (from Peters, 1979)

Sample L d φ k kor kor/k Swirr

No (cm) (cm) (%) (darcy) (darcy) (%)

1 22.9 4.81 38.98 16.43 15.82 0.96 11.20

2 23.6 4.84 36.37 18.36 17.99 0.98 8.66

3 112.8 4.96 38.38 21.89 18.05 0.82 9.18

4 23.7 4.81 38.94 18.93 18.33 0.97 11.34

5 110.5 4.83 38.91 20.52 18.50 0.90 10.83

6 110.0 4.81 37.61 21.91 19.40 0.89 10.48

7 23.6 4.84 37.49 14.19 11.03 0.78 15.26

8 116.1 4.96 34.66 18.28 15.57 0.85 8.32

9 113.0 4.96 39.62 22.99 19.90 0.87 8.61

10 115.9 4.96 35.86 19.22 15.49 0.81 12.53

11 22.8 4.84 37.52 16.23 15.19 0.94 11.20

12 110.4 4.81 37.46 20.85 18.48 0.89 10.00

13 115.9 4.97 37.80 18.54 15.53 0.84 9.29

14 110.0 4.81 38.11 22.50 18.62 0.83 8.91

15 112.8 4.97 35.48 22.72 20.90 0.92 9.92

16 23.4 4.84 34.61 14.51 9.58 0.66 11.03

17 23.6 4.84 38.60 20.87 18.49 0.89 10.15

18 23.7 4.81 35.44 12.11 11.94 0.99 12.02

Page 213: +Peters Ekwere j. - Petrophysics

3-13

Example 3.1

The permeability of a clean, dry core plug is to be determined. The core is

cylindrical with a diameter of 24 mm and a length of 32 mm. The core was

installed in a gas permeameter and air was flowed through it at an average

rate of 100 cm3 in 2 minutes 20 seconds, measured at atmospheric pressure

(Figure 3.8). The pressure differential across the sample was kept constant at

12 cm of mercury. The upstream gauge pressure (at the inlet of the core) was

76 cm of mercury. The gas viscosity at the test temperature was 0.01808 cp.

The barometric pressure was 76 cm of mercury.

Figure 3.8. Gas permeameter for Example 3.1.

1. Calculate the permeability to air at the test conditions.

2. Does the permeability calculated in part 1 represent the true absolute permeability of this core? If yes, why? If no, why not?

Page 214: +Peters Ekwere j. - Petrophysics

3-14

Solution to Example 3.1

d = 24 mm = 2.4 cm

L = 32 mm = 3.2 cm

qsc = 100cm3/(2 minutes 20 seconds) = (100/140) cm3/s

P1 = 76cm of Hg gauge = 76+76 = 152 cm of Hg absolute

= 152/76 = 2 atm absolute

P2 = (P1 - 12) = 152 - 12 = 140 cm of Hg absolute

= 140/76 = 1.842 atm absolute

Psc = 1 atm

μg = 0.0808 cp

Substituting into Eq.(3.17) gives

( )( )( )( )( )( ) ( )2 2 2

2 100 /140 0.01808 1 3.20.0301 darcy = 30.1 md

2.4 / 2 2 1.842gkπ

= =−

No. The calculated permeability to gas of 30.1 md is not the true absolute

permeability of the core because of Klinkenberg effect. It is larger than the

absolute permeability of the core.

3.4 FIELD DETERMINATION OF PERMEABILITY

Permeability can be determined in the field by use of pressure transient

tests, essentially measuring the permeability of the in-situ reservoir rock on a

large scale. A transient pressure test consists of changing the flow rate of a

well and then recording the bottomhole pressure response as a function of

time. The pressure data can be analyzed to obtain formation permeability and

other reservoir and well parameters. Oil permeability at irreducible water

saturation is normally used for comparison with the permeability from most

pressure transient tests that involve only oil flow.

Page 215: +Peters Ekwere j. - Petrophysics

3-15

3.4.1 Diffusivity Equation for Slightly Compressible Liquid

The partial differential equation that describes the transient pressure

response of a reservoir undergoing single phase flow is obtained by combining

the law of mass conservation, Darcy’s law and an equation of state for the

fluid. The mass conservation equation, also known as the continuity equation,

is

( ) ( ). 0vt

φρρ

∂∇ + =

∂ (3.20)

Darcy's law is given by

kv Pμ

= − ∇ (3.21)

The equation of state for a slightly compressible liquid such as oil or water is

given by

( )oc P Poeρ ρ −= (3.22)

where ρo is the density at a reference pressure Po. Eqs.(3.20) through (3.22)

can be combined to give

2 1tc P PPk t t

φμα

∂ ∂∇ = =

∂ ∂ (3.23)

where α is a constant and is given by

t

kc

αφμ

= (3.24)

Eq. 3.23 is known as the diffusivity equation in the petroleum engineering

community. Others may recognize it as the diffusion equation or the heat

conduction equation. Thus, Eq.(3.23) is a standard partial differential

equation of mathematical physics. It is a second order, parabolic, linear

partial differential equation. The solutions of this equation for specific initial

Page 216: +Peters Ekwere j. - Petrophysics

3-16

and boundary conditions are used for transient pressure analysis and for

natural water influx calculations.

The constant, α, is known as the diffusivity constant and can be written

as

t t

khk Tc hc S

μα

φμ φ

⎛ ⎞⎜ ⎟⎝ ⎠= = = (3.25)

where T is the transmissibility of the reservoir given by

khTμ

= (3.26)

and S is the storativity of the reservoir given by

tS hcφ= (3.27)

Thus, the diffusivity constant is the ratio of the transmissibility to the

storativity (storage capacity) of the reservoir.

The mathematical model (initial-boundary value problem) for transient

pressure analysis is as follows:

2

2

1 tcP P Pr r r k t

φμ∂ ∂ ∂+ =

∂ ∂ ∂ (3.28)

( )P r,0 iP= (3.29)

0

lim2

sf

r

qPrr kh

μπ→

∂⎛ ⎞ =⎜ ⎟∂⎝ ⎠ (3.30)

( )lim , irP r t P

→∞= (3.31)

Eq.(3.28) is Eq.(3.23) for one-dimensional radial flow. Eq.(3.29) is the initial

condition, which specifies that before production commenced, the pressure in

the reservoir was constant and equal to the initial reservoir pressure, Pi.

Page 217: +Peters Ekwere j. - Petrophysics

3-17

Eq.(3.30) is the internal boundary condition at the wellbore, which specifies a

constant instantaneous sandface rate upon commencement of production.

This instantaneous constant sandface boundary condition is unrealistic

because we expect that when a well is put on production, the sandface rate

should increase from zero to the final constant rate over a finite time.

Eq.(3.29) also specifies a vanishingly small wellbore radius thereby treating

the well as a line source. These two specifications, though unrealistic, have

been made for mathematical expediency because with them, we can obtain a

simple approximate solution to the partial differential equation. Eq.(3.31) is

the external boundary condition, which specifies that far away from the

wellbore, the reservoir is undisturbed and as such, the reservoir pressure will

remain at the initial pressure, Pi. This boundary condition is the infinite

acting reservoir boundary condition.

Eqs.(3.28) through (3.31) constitute the initial-boundary value problem

that describes the transient pressure response of a reservoir that has been

put on instantaneous constant sandface rate production. It is useful to put

these equations in dimensionless forms. This will reduce the number of

variables and also will facilitate the use of existing solutions from other fields

for pressure transient analysis. For example, there is a wealth of solutions of

linear, second order, parabolic partial differential equation in the heat

conduction literature. Casting our model in dimensionless variables allows

the immediate use of these existing solutions from other disciplines for our

purpose.

In dimensionless forms, Eqs.(3.28) through (3.31) become

2

2

1D D D

D D D D

P P Pr r r t

∂ ∂ ∂+ =

∂ ∂ ∂ (3.32)

( )P r ,0 0D D = (3.33)

Page 218: +Peters Ekwere j. - Petrophysics

3-18

0

lim 1D

DDr

D

Prr→

⎛ ⎞∂= −⎜ ⎟∂⎝ ⎠

(3.34)

( )lim , 0D

D D DrP r t

→∞= (3.35)

where the dimensionless variables are defined as

( ),

2

iD

sf

P P r tP

qkhμ

π

−=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.36)

Dw

rrr

= (3.37)

and

2Dr w

kttc rφμ

= (3.38)

In oilfield units, the dimensionless pressure and dimensionless time are given

by

( ),

141.2

iD

P P r tP

q Bkhμ

−=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.39)

and

2

0.0002637D

r w

kttc rφμ

= (3.40)

In Eq.(3.39), q is surface rate and qB is the sandface rate.

Eqs.(3.32) through (3.35) can be solved, for example, by Laplace

transformation. It can be shown that the solution is of the form

( ) 1 ( )2DP x Ei x= − − (3.41)

Page 219: +Peters Ekwere j. - Petrophysics

3-19

where

2

4D

D

rxt

= (3.42)

and

( )y

x

eEi x dyy

−∞− = −∫ (3.43)

Ei(-x) is the exponential integral function, which is tabulated in mathematical

handbooks. Table 3.4 gives the Ei function for various values of the argument,

x. The solution given in Eq.(3.41) can be written as

( )21,

2 4D

D D DD

rP r t Eit

⎛ ⎞= − −⎜ ⎟

⎝ ⎠ (3.44)

at any dimensionless radius and time, and

( ) 1 11,2 4D D

D

P t Eit

⎛ ⎞= − −⎜ ⎟

⎝ ⎠ (3.45)

at the wellbore, where rD = 1. In terms of actual pressure, radius and time in

oilfield units, Eq.(3.44) and (3.45) become

( )2948141.2 1,

2t

ic rq BP r t P Ei

kh ktφμμ ⎡ ⎤⎛ ⎞

= − − −⎢ ⎥⎜ ⎟⎝ ⎠⎣ ⎦

(3.46)

and

( ) ( )2948141.2 1,

2t w

wf w ic rq BP t P r t P Ei

kh ktφμμ ⎡ ⎤⎛ ⎞

= = − − −⎢ ⎥⎜ ⎟⎝ ⎠⎣ ⎦

(3.47)

3.4.2 Pressure Drawdown Equation

For 0.01x ≤ , the exponential integral function can be accurately

approximated as

Page 220: +Peters Ekwere j. - Petrophysics

3-20

( ) ln 0.5772Ei x x− + (3.48)

This approximation is usually valid at the wellbore in all well testing

situations. Using the approximation, the dimensionless solution at the

wellbore, Eq.(3.45), can be written as

( ) ( )11, ln 0.80907 for 252D D D DP t t t= + ≥ (3.49)

Eq.(3.49) can be written in terms of log to base 10 as

( ) 21, 1.1513 log log 3.23D Dt w

kP t tc rφμ

⎡ ⎤⎛ ⎞= + −⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.50)

Eq.(3.50) gives the flowing bottomhole pressure as

( ) ( ) 2

162.6, log log 3.23wf w it w

q B kP t P r t P tkh c r

μφμ

⎡ ⎤⎛ ⎞= = − + −⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.51)

Eq.(3.51) provides the theoretical basis for the semilog analysis of drawdown

pressure data. It shows that if the well is produced instantaneously at a

constant sandface rate, then a graph of Pwf versus logt will be linear with a

negative slope given by

162.6q Bmkh

μ= − (3.52)

and a pressure intercept at logt = 0 given by

int 2log 3.23it w

kP P mc rφμ

⎡ ⎤⎛ ⎞= + −⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.53)

The formation permeability can be calculated from the slope of the semilog

line as

162.6q Bkmh

μ= − (3.54)

Page 221: +Peters Ekwere j. - Petrophysics

3-21

Table 3.4: Exponential Integral Function

Page 222: +Peters Ekwere j. - Petrophysics

3-22

and the initial pressure can be calculated from the pressure intercept as

int 2log 3.23it w

kP P mc rφμ

⎡ ⎤⎛ ⎞= − −⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.55)

3.4.3 Pressure Buildup Equation

We wish to solve the diffusivity equation for the case of a well, which

has been producing at a constant instantaneous rate q for a time and then

shut in at time tp as shown in Figure 3.9C. Since the diffusivity equation is a

linear partial differential equation, superposition principle can be used to

construct this solution from the constant rate solution that we already have.

Figure 3.9 shows a graphical construction of how two constant rates can be

used to generate the rate schedule for a pressure buildup test. If the constant

production rate A is added to the constant injection rate B, the outcome will

be the rate in C, which simulates the rate for a pressure buildup test for a

well that had produced for time tp before it was shut in for the buildup test.

Figure 3.9. Rate schedule for pressure buildup test.

Page 223: +Peters Ekwere j. - Petrophysics

3-23

Using Eq.(3.49), the dimensionless pressure for the buildup test can easily be

derived by applying superposition principle as follows:

DC DA DBP P P= + (3.56)

( )1 ln 0.809072DA p D

P t t⎡ ⎤= + Δ +⎣ ⎦ (3.57)

[ ]1 ln 0.809072DB DP t= − Δ + (3.58)

Substituting Eqs.(3.57) and (3.58) into Eq.(3.56) gives the dimensionless

pressure for the buildup test as

1 ln2

pDBU DC

t tP P

t+ Δ⎛ ⎞

= = ⎜ ⎟Δ⎝ ⎠ (3.59)

Eq.(3.59) gives the following pressure buildup equation

( ) 162.6 log pws i

t tq BP t Pkh t

μ + Δ⎛ ⎞Δ = − ⎜ ⎟Δ⎝ ⎠

(3.60)

Eq.(3.60) is known as the Horner pressure buildup equation for an infinite

acting reservoir. It shows that if the sandface rate (qB) can be reduced to zero

instantaneously as shown in Figure 3.8C, then a graph of the buildup

bottomhole pressure, Pws(Δt), versus Horner time, pt tt

+ Δ⎛ ⎞⎜ ⎟Δ⎝ ⎠

, will be linear with

a negative slope given by Eq.(3.54) and a pressure intercept equal to Pi. The

formation permeability can be calculated from the slope of the Horner line and

the initial reservoir pressure can be determined by extrapolating the Horner

line to an infinite shut in time. For a bounded reservoir (in contrast to an

infinite acting reservoir), Horner pressure buildup equation takes the form

( ) * 162.6 log pws

t tq BP t Pkh t

μ + Δ⎛ ⎞Δ = − ⎜ ⎟Δ⎝ ⎠

(3.61)

Page 224: +Peters Ekwere j. - Petrophysics

3-24

where the extrapolated pressure, P*, is less than the initial reservoir pressure,

Pi.

3.4.4 Diagnostic Plots

The drawdown equation, Eq.(3.51), was derived using an internal

boundary condition in which the sandface rate increased instantaneously

from zero to a constant value. The buildup equation, Eq.(3.60), was derived

using an internal boundary condition in which the sandface rate decreased

from a constant value to zero instantaneously. Given that the wellbore

contains compressible fluids, when the well is opened for production, the

surface rate may be constant but the sandface rate will increase from zero to

a constant value over a definite period of time as shown in Figure 3.10.

Similarly, when the well is shut in at the wellhead for a pressure buildup test,

the surface rate is zero but the sandface rate will decline from its constant

value to zero over a finite period of time as shown in Figure 3.9. This

phenomenon, whereby the sandface rate lags behind the surface rate, is

known as wellbore unloading during drawdown and afterflow during buildup.

The term wellbore storage is normally used to describe the phenomenon either

during drawdown or during buildup.

Figure 3.10. Idealized and actual sandface rates during drawdown and buildup.

Page 225: +Peters Ekwere j. - Petrophysics

3-25

The consequence of neglecting the effect of wellbore storage in our

mathematical model is that the very early time pressure data will deviate from

the semilog line of Eq.(3.51) for the drawdown test or from the Horner semilog

line of Eq.(3.60) for the buildup test as shown in Figures 3.11 and 3.12. It is

therefore necessary to perform preliminary diagnostic plots of the pressure

data to identify those data that should be fitted to the semilog lines if any. The

function normally used to diagnose the presence of wellbore storage in the

pressure data is the welltest derivative function defined as

'

lnD D

D DD D

dP dPP td t dt

= = (3.62)

Figure 3.11. Semilog plot for pressure drawdown test showing deviation from semilog line due to wellbore storage effect.

Page 226: +Peters Ekwere j. - Petrophysics

3-26

Figure 3.12. Horner plot for pressure buildiup test showing deviation from

semilog line due to wellbore storage effect.

The welltest derivative function is defined in the manner shown in Eq.(3.62) to

take advantage of the fact that the data that fit the welltest model for a

drawdown test fall on the semilog line of Eq.(3.42). Along this semilog line, the

dimensionless welltest derivative function will be a constant as shown below:

' 1 a constantln 2

D DD D

D D

dP dPP td t dt

= = = = (3.63)

In actual variables, the welltest derivative function becomes

' 70.6 a constantln

d P d P q BP td t dt kh

μΔ Δ ⎛ ⎞Δ = = = =⎜ ⎟⎝ ⎠

(3.64)

Thus, after the effect of wellbore storage has subsided, the welltest derivative

function will become constant along the semilog line. It is only the pressure

data that have a constant welltest derivative (making allowance for

fluctuations due to noise) that should be fitted to the drawdown semilog line.

Page 227: +Peters Ekwere j. - Petrophysics

3-27

The diagnostic plots consist of log-log plots of ΔP versus t and ΔP' versus t

superimposed on the same graph. Figure 3.13 shows typical diagnostic plots

for a pressure drawdown test affected by wellbore storage.

Figure 3.13. Diagnostic plots for pressure drawdown data affected by wellbore storage.

The diagnostic plots can be divided into three time segments. In the first

time segment, ΔP and ΔP' are equal and have a unit slope on the log-log scale.

The data in this time segment are dominated 100% by wellbore storage. They

cannot be used to estimate reservoir properties. They can only be used to

estimate wellbore storage coefficient. In the second time segment, ΔP and ΔP'

separate from each other. The derivative function has a characteristic hump

indicative of wellbore storage. The magnitude of the hump is a measure of the

severity of the wellbore storage problem (and skin damage). The pressure data

Page 228: +Peters Ekwere j. - Petrophysics

3-28

in this time segment are still affected by wellbore storage but to a lesser

degree than in the first time segment. The data in this segment can be

analyzed by type curve matching to estimate approximate formation

properties. In the third time segment, the welltest derivative becomes constant

in accordance with Eq.(3.64). Data in this time segment are no longer affected

by wellbore storage and will plot on the correct semilog line of Eq.(3.51). Note

that the constant value of the welltest derivative function in this time segment

is equal to 70.6 q Bkhμ⎛ ⎞

⎜ ⎟⎝ ⎠

as shown in Eq.(3.64). This fact can be used to estimate

the formation permeability from the derivative function. If the test ended

during the second time segment, which happens in many tests, such a test

cannot be analyzed with a semilog plot.

Figure 3.14 shows typical diagnostic plots for a pressure buildup test

affected by wellbore storage. In general, the diagnostic plots are similar to

those of a drawdown test. The only difference is that in the third time

segment, when the pressure data are no longer affected by wellbore storage,

the welltest derivative function for a buildup may not be constant but may be

distorted as shown in Figure 3.14. This distortion is caused by the fact that

an approximate ΔP is used to calculate the derivative function for a buildup

test instead of the true ΔP, which is not accessible.

The welltest derivative function can be calculated with the following

central difference approximation:

( ) ( )

( ) ( )

1 11 1

1 1'

1 1i

i i i ii i i i

i i i ii i

t i i i i

P P P Pt t t tt t t td PP t t

dt t t t t

− ++ −

− +

− +

⎡ ⎤⎛ ⎞ ⎛ ⎞Δ − Δ Δ − Δ− + −⎢ ⎥⎜ ⎟ ⎜ ⎟− −Δ⎛ ⎞ ⎝ ⎠ ⎝ ⎠⎢ ⎥Δ = =⎜ ⎟ ⎢ ⎥− + −⎝ ⎠

⎢ ⎥⎢ ⎥⎣ ⎦

(3.65)

Page 229: +Peters Ekwere j. - Petrophysics

3-29

Figure 3.14. Diagnostic plots for pressure buildup data affected by wellbore storage.

Figure 3.15 shows the pressures used to define ΔP for drawdown and buildup

tests. For drawdown test, ΔP is defined as

( )i wfP P P tΔ = − (3.66)

For buildup test, ΔP should be defined as

( ) ( )ws wf pP P t P t tΔ = Δ − + Δ (3.67)

However, because Pwf(tp+Δt) is not available, ΔP for buildup test is

approximated by

Page 230: +Peters Ekwere j. - Petrophysics

3-30

( ) ( )ws wf pP P t P tΔ = Δ − (3.68)

This approximation could distort the welltest derivative function after the

effect of wellbore storage has subsided. The derivative function for a pressure

buildup test can be calculated with Eq.(3.65) using actual shut in time, Δt, or

Horner time, pt tt

+ Δ⎛ ⎞⎜ ⎟Δ⎝ ⎠

, or effective shut in time, Δte, defined as

( )( )p

ep

t tt

t tΔ

Δ =Δ +

(3.69)

Figure 3.15. Pressures used to define ΔP for drawdown and buildup.

3.4.5 Skin Factor

During drilling operations, the near wellbore permeability could be

reduced by formation damage caused by mud filtrate invasion. Also, if the well

is treated by an acid job, the near wellbore permeability could be enhanced

and the well is said to be stimulated. Thus, for a damaged or stimulated well,

Page 231: +Peters Ekwere j. - Petrophysics

3-31

there is a region of altered permeability near the wellbore as shown in Figure

3.16.

The additional pressure change near the wellbore caused by the region

of altered permeability can be incorporated into the welltest model through

the concept of a skin factor. The skin factor, S, is a dimensionless pressure

change at the wellbore given by

2

skinPSq

khμ

π

Δ=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.70)

in Darcy units or

141.2

skinPSq Bkhμ

Δ=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.71)

in oilfield units where ΔPskin is defined as

skin wfundamaged wfdamagedP P PΔ = − (3.72)

for a damaged well or

skin wfundamaged wfstimulatedP P PΔ = − (3.73)

for a stimulated well. It can be seen from Figure 3.16 that ΔPskin and hence, S,

is positive for a damaged well and negative for a stimulated well. It can be

shown that the permeability of the altered zone, ks, is related to the true

formation permeability, k, by the equation

1 ln s

s w

rkSk r

⎛ ⎞ ⎛ ⎞= −⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ (3.74)

Eq.(3.74) contains two unknowns, ks and rs, an as such is not as useful for

estimating the skin factor as it appears to be.

Page 232: +Peters Ekwere j. - Petrophysics

3-32

Figure 3.16. Schematic diagram showing skin.

Skin factor can be incorporated into the welltest model as

( ) ( )11, ln 0.80907 +S for 252D D D DP t t t= + ≥ (3.75)

Eq.(3.75) can be written in terms of log to base 10 as

Page 233: +Peters Ekwere j. - Petrophysics

3-33

( ) 21, 1.1513 log log 3.23 0.87D Dt w

kP t t Sc rφμ

⎡ ⎤⎛ ⎞= + − +⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.76)

Eq.(3.76) gives the flowing bottomhole pressure in the presence of a skin as

( ) ( ) 2

162.6, log log 3.23 0.87wf w it w

q B kP t P r t P t Skh c r

μφμ

⎡ ⎤⎛ ⎞= = − + − +⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.77)

Eq.(3.77) can be solved for the skin factor from the drawdown semilog line as

( )

2

11.1513 log 3.23

162.6wf i

t w

P hr P kSq B c r

khμ φμ

⎡ ⎤⎢ ⎥− ⎛ ⎞⎢ ⎥= − +⎜ ⎟⎛ ⎞⎢ ⎥⎝ ⎠−⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.78)

where Pwf(1hr) is read on the semilog straight line or its extrapolation as

shown in Figure 3.10. For a buildup test, the skin factor is given by

( ) ( )

2

11.1513 log 3.23

162.6wf p ws

t w

P t P hr kSq B c r

khμ φμ

⎡ ⎤⎢ ⎥− ⎛ ⎞⎢ ⎥= − +⎜ ⎟⎛ ⎞⎢ ⎥⎝ ⎠−⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.79)

where Pws(1hr) is read on the Horner straight line or its extrapolation as

shown in Figure 3.12. Thus, the skin factor, S, can be determined from a

drawdown or a buildup test. Note that the presence of a skin factor does not

affect the shape of the welltest derivative function and as such, it can still be

used to perform its diagnostic function.

3.4.6 Homogeneous Reservoir Model with Wellbore Storage and Skin

The welltest model can be reformulated to include wellbore storage and

skin. The resulting initial-boundary value problem to be solved consists of the

following equations in Darcy units

Page 234: +Peters Ekwere j. - Petrophysics

3-34

2

2

1 tcP P Pr r r k t

φμ∂ ∂ ∂+ =

∂ ∂ ∂ (3.28)

( )P r,0 iP= (3.29)

( )lim , irP r t P

→∞= (3.31)

2

w

wf

r r

dPkh Pr C qr dt

πμ =

∂⎛ ⎞ − =⎜ ⎟∂⎝ ⎠ (3.80)

( ),w

wfr r

PP P r t S rr =

⎡ ∂Δ ⎤⎛ ⎞Δ = Δ − ⎜ ⎟⎢ ⎥∂⎝ ⎠⎣ ⎦ (3.81)

The previous single internal boundary condition, Eq.(3.30), has now

been replaced by two internal boundary conditions, Eqs.(3.80) and (3.81).

Eq.(3.80) states that the constant surface rate, q, consists of the sum of the

sandface rate, given by the first term on the left side of the equation, and the

wellbore storage rate, given by the second term on the left side of the

equation. The constant, C, is the wellbore storage coefficient defined as

( )( )

i

i wf

V V tVCP P P t

−Δ= =

Δ − (3.82)

where

C = wellbore storage coefficient, reservoir cm3/atm

ΔV = change in wellbore fluid volume, reservoir cm3

ΔP = change in bottomhole pressure, atm

Vi = initial wellbore fluid volume before unloading, reservoir cm3

V(t) = wellbore fluid volume during unloading, reservoir cm3

Pi = initial wellbore pressure before unloading, atm

Pwf(t) = wellbore pressure during unloading, atm

Page 235: +Peters Ekwere j. - Petrophysics

3-35

The wellbore storage coefficient, C, is a measure of the severity of the wellbore

storage problem and can be determined from the unit slope line of the first

time segment on the diagnostic plots. The unit slope line in oilfield units is

given by

24qBP t

C⎛ ⎞Δ = ⎜ ⎟⎝ ⎠

(3.83)

from which C can be calculated in RB/psi as

unit slope line24

qB tCP

⎛ ⎞⎛ ⎞= ⎜ ⎟⎜ ⎟Δ⎝ ⎠⎝ ⎠ (3.84)

Eq.(3.81) states that the pressure change at the wellbore, (Pi-Pwf),

consists of the sum of the pressure change for the undamaged or

unstimulated well, given by the first term on the right side of the equation,

and ΔPskin, given by the second term on the right side of the equation.

The initial-boundary value problem can be put in dimensionless form as

2

2

1D D D

D D D D

P P Pr r r t

∂ ∂ ∂+ =

∂ ∂ ∂ (3.32)

( ),0 0D DP r = (3.33)

( )lim , 0D

D D DrP r t

→∞= (3.35)

1

1D

wD DD D

D D r

dP PC rdt r

=

⎛ ⎞∂− =⎜ ⎟∂⎝ ⎠

(3.85)

1D

DwD D D

D r

PP P S rr

=

⎡ ⎤⎛ ⎞∂= −⎢ ⎥⎜ ⎟∂⎝ ⎠⎣ ⎦

(3.86)

where the dimensionless wellbore storage coefficient is defined as

Page 236: +Peters Ekwere j. - Petrophysics

3-36

22Dt w

CCc hrπφ

= (3.87)

in Darcy units or

2

5.6152D

t w

CCc hrπφ

= (3.88)

in oilfield units. The larger the dimensionless wellbore storage coefficient, the

more severe is the wellbore storage problem.

The initial-boundary value problem represented by Eqs.(3.32), (3.33),

(3.35), (3.85) and (3.86) can be solved by Laplace transformation. It can be

shown that the dimensionless pressure at the wellbore in Laplace space is

given by

( )( ) ( )

( ) ( ) ( )1

3/ 2 21 1

1,o

wD

D o

K z S zK zP z

z K z z C K z S zK z

+=

⎡ ⎤+ +⎣ ⎦

(3.89)

where z is the Laplace parameter and Ko and K1 are Bessel functions of the

second kind of order 0 and 1. Eq.(3.89) can easily be inverted numerically, by

say Stehfest algorithm, to obtain the PwD function in the time domain. To

obtain the welltest derivative function, the Calculus derivative with respect to

tD in Laplace space is first obtained, inverted and multiplied by tD. The

Calculus derivative in Laplace space is given by

( ) ( )( ) ( )

( ) ( ) ( )1'

3/ 2 21 1

1, 1,o

wD wD

D o

z K z S zK zP z zP z

z K z z C K z S zK z

⎡ ⎤+⎣ ⎦= =⎡ ⎤+ +⎣ ⎦

(3.90)

Figure 3.17 shows the solutions and welltest derivative functions for

various values of CDe2S plotted on log-log scales. This family of solutions

constitutes Bourdet et al.'s wellbore storage type curves, which can be used to

analyze transient pressure tests that fit a homogeneous reservoir model with

wellbore storage and skin by type curve matching.

Page 237: +Peters Ekwere j. - Petrophysics

3-37

3.4.7 Type Curve Matching

A type curve is a graph of dimensionless pressure together with the

corresponding welltest derivative function versus dimensionless time, often on

a log-log scale. Such a graph is a convenient way to present the solution to

the diffusivity equation obtained by numerical computations. Figure 3.17 is

an example of type curves obtained by solving the diffusivity equation in the

presence of wellbore storage and skin. From the definitions of PwD, tD and CD,

BOURDET TYPE CURVES

0.1

1

10

100

0.1 1 10 100 1000 10000

tD/CD

PwD

- Pw

D'

CDe2S

1060

1050

1040

1030

1020

1015

1010

108

106

104

103

102

Figure 3.17. Bourdet et al type curves for a homogeneous reservoir with wellbore storage and skin for CDe2S from 102 to 1060.

Page 238: +Peters Ekwere j. - Petrophysics

3-38

log log log141.2wD

khP Pq Bμ

⎛ ⎞= Δ + ⎜ ⎟

⎝ ⎠ (3.91)

42.951 10log log logD

D

t x khtC Cμ

−⎛ ⎞ ⎛ ⎞= +⎜ ⎟ ⎜ ⎟

⎝ ⎠⎝ ⎠ (3.92)

Eqs.(3.91) and (3.92) show that the graph of log wDP versus log D

D

tC

⎛ ⎞⎜ ⎟⎝ ⎠

is the

same as the graph of log PΔ versus log t with the axes shifted by the constants

log141.2

khq Bμ

⎛ ⎞⎜ ⎟⎝ ⎠

and 42.951 10log x kh

−⎛ ⎞⎜ ⎟⎝ ⎠

. Thus, the ΔP versus t obtained from a

transient pressure test can be matched graphically with the appropriate type

curve to estimate the reservoir and well parameters.

The procedure for using the Bourdet et al.'s type curves to analyze a

transient pressure test is as follows:

1. Prepare a log-log plot of ΔP versus t on tracing paper using the same

scale as the type curves. Use only major grid lines on this plot. This is

the field plot.

2. Slide the field plot over the type curves both vertically and horizontally

to obtain the best match of the PD function and the corresponding

welltest derivative function. Trace the matched type curve on the field

plot to obtain a visual record of the best match.

3. Choose a convenient match point and record DMP , D

D M

tC

⎛ ⎞⎜ ⎟⎝ ⎠

, Mt and

( )2SD M

C e .

4. Calculate the formation permeability from the pressure match as

( )( )( )( )

141.2 DM

M

q B Pk

h Pμ

(3.93)

Page 239: +Peters Ekwere j. - Petrophysics

3-39

5. Calculate the wellbore storage coefficient from the time match as

( )0.0002951

M

D

D M

kh tC

tC

μ⎛ ⎞

Δ⎜ ⎟⎝ ⎠=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.94)

6. Calculate the dimensionless wellbore storage coefficient.

7. Calculate the skin factor as

( )2

1 ln2

SD M

D

C eS

C

⎡ ⎤⎢ ⎥=⎢ ⎥⎣ ⎦

(3.95)

Figure 3.18 shows qualitatively a type match using the Bourdet et al.'s type

curves.

Figure 3.18. Example type curve match.

Page 240: +Peters Ekwere j. - Petrophysics

3-40

3.4.8 Radius of Investigation of a WellTest

An approximate equation for estimating the radius investigated by a

welltest is given by

0.03248invt

ktrcφμ

= (3.96)

where the variables are in oilfield units.

3.4.9 Field Example of a Well Test Analysis

A pressure buildup test was conducted on a new well in a new reservoir with

the results shown in Table 3.5.

Table 3.5: Pressure Buildup Data for Field Example

t tΔ ( )wsP tΔ

(hrs) (hrs) (psia) 15.33000 0.00000 3086.33 15.33417 0.00417 3090.57 15.33833 0.00833 3093.81 15.34250 0.01250 3096.55 15.34667 0.01667 3100.03 15.35083 0.02083 3103.27 15.35500 0.02500 3106.77 15.35917 0.02917 3110.01 15.36333 0.03333 3113.25 15.36750 0.03750 3116.49 15.37583 0.04583 3119.48 15.38000 0.05000 3122.48 15.38833 0.05833 3128.96 15.39667 0.06667 3135.92 15.40500 0.07500 3141.17 15.41333 0.08333 3147.64 15.42583 0.09583 3161.95 15.43833 0.10833 3170.68 15.45083 0.12083 3178.39 15.46333 0.13333 3187.12 15.47583 0.14583 3194.24 15.49250 0.16250 3205.96 15.50917 0.17917 3216.68 15.52583 0.19583 3227.89 15.54250 0.21250 3238.37 15.55917 0.22917 3249.07 15.58000 0.25000 3261.79

Page 241: +Peters Ekwere j. - Petrophysics

3-41

15.62167 0.29167 3287.21 15.66333 0.33333 3310.15 15.70500 0.37500 3334.34 15.74667 0.41667 3356.27 15.78833 0.45833 3374.98 15.83000 0.50000 3394.44 15.87167 0.54167 3413.90 15.91333 0.58333 3433.83 15.95500 0.62500 3448.05 15.99667 0.66667 3466.26 16.03833 0.70833 3481.97 16.08000 0.75000 3493.69 16.14250 0.81250 3518.63 16.20500 0.87500 3537.34 16.26750 0.93750 3553.55 16.33000 1.00000 3571.75 16.39250 1.06250 3586.23 16.45500 1.12500 3602.95 16.51750 1.18750 3617.41 16.58000 1.25000 3631.15 16.64250 1.31250 3640.86 16.70500 1.37500 3652.85 16.76750 1.43750 3664.32 16.83000 1.50000 3673.81 16.95500 1.62500 3692.27 17.08000 1.75000 3705.52 17.20500 1.87500 3719.26 17.33000 2.00000 3732.23 17.58000 2.25000 3749.71 17.70500 2.37500 3757.19 17.83000 2.50000 3763.44 18.08000 2.75000 3774.65 18.33000 3.00000 3785.11 18.58000 3.25000 3794.06 18.83000 3.50000 3799.80 19.08000 3.75000 3809.50 19.33000 4.00000 3815.97 19.58000 4.25000 3820.20 19.83000 4.50000 3821.95 20.08000 4.75000 3823.70 20.33000 5.00000 3826.45 20.58000 5.25000 3829.69 20.83000 5.50000 3832.64 21.08000 5.75000 3834.70 21.33000 6.00000 3837.19 21.58000 6.25000 3838.94 22.08000 6.75000 3838.02 22.58000 7.25000 3840.78 23.08000 7.75000 3843.01 23.58000 8.25000 3844.52

Page 242: +Peters Ekwere j. - Petrophysics

3-42

24.08000 8.75000 3846.27 24.58000 9.25000 3847.51 25.08000 9.75000 3848.52 25.58000 10.25000 3850.01 26.08000 10.75000 3850.75 26.58000 11.25000 3851.76 27.08000 11.75000 3852.50 27.58000 12.25000 3853.51 28.08000 12.75000 3854.25 28.58000 13.25000 3855.07 29.08000 13.75000 3855.50 29.83000 14.50000 3856.50 30.58000 15.25000 3857.25 31.33000 16.00000 3857.99 32.08000 16.75000 3858.74 32.83000 17.50000 3859.48 33.58000 18.25000 3859.99 34.33000 19.00000 3860.73 35.08000 19.75000 3860.99 35.83000 20.50000 3861.49 36.58000 21.25000 3862.24 37.58000 22.25000 3862.74 38.58000 23.25000 3863.22 39.58000 24.25000 3863.48 40.58000 25.25000 3863.99 41.58000 26.25000 3864.49 42.58000 27.25000 3864.73 43.83000 28.50000 3865.23 45.33000 30.00000 3865.74

Other pertinent data are as follows:

Producing time before the test = 15.33 hours

Production rate before the test = 174 STB/D

Wellbore radius = 0.29 ft

Formation thickness = 107 ft

Porosity = 25 %

Total compressibility = 4.2x10-6 psi-1

Formation volume factor = 1.06 RB/STB

Oil viscosity = 2.5 cp

Page 243: +Peters Ekwere j. - Petrophysics

3-43

a. Screen the pressure data using appropriate diagnostic plots.

b. Analyze the test data and determine the formation permeability, the

total skin factor and the initial reservoir pressure.

c. Calculate the radius of investigation of the test.

d. Simulate the buildup test assuming a homogeneous reservoir with

wellbore storage and skin. Show the best match graphically by

superimposing the simulated and the test data on the plots of P versus

t, log-log diagnostic plots and the Horner plot. What are the values of k,

S and CD , Pwf(tp) and Pi that gave the best match?

The pertinent data for the diagnostic plots and Horner plot are shown in

Table 3.6.

Table 3.6: Calculated Values of ΔP and ΔP'

t tΔ etΔ Horner ( )wsP tΔ PΔ '

eP wrt tΔ Δ 'P wrt tΔ Δ

(hrs) (hrs) (hrs) Time (psia) (psi) (psi) (psi)

15.33000 0.00000 3086.33

15.33417 0.00417 0.00417 3677.26 3090.57 4.24

15.33833 0.00833 0.00833 1841.34 3093.81 7.48 5.98 5.98

15.34250 0.01250 0.01249 1227.40 3096.55 10.22 9.33 9.32

15.34667 0.01667 0.01665 920.62 3100.03 13.70 13.46 13.45

15.35083 0.02083 0.02080 736.96 3103.27 16.94 16.88 16.85

15.35500 0.02500 0.02496 614.20 3106.77 20.44 20.24 20.20

15.35917 0.02917 0.02911 526.54 3110.01 23.68 22.73 22.69

15.36333 0.03333 0.03326 460.95 3113.25 26.92 25.98 25.93

15.36750 0.03750 0.03741 409.80 3116.49 30.16 23.96 23.91

15.37583 0.04583 0.04569 335.50 3119.48 33.15 27.55 27.46

15.38000 0.05000 0.04984 307.60 3122.48 36.15 37.07 36.95

15.38833 0.05833 0.05811 263.82 3128.96 42.63 47.21 47.03

Page 244: +Peters Ekwere j. - Petrophysics

3-44

15.39667 0.06667 0.06638 230.94 3135.92 49.59 49.03 48.82

15.40500 0.07500 0.07463 205.40 3141.17 54.84 53.03 52.76

15.41333 0.08333 0.08288 184.97 3147.64 61.31 77.43 76.99

15.42583 0.09583 0.09523 160.97 3161.95 75.62 88.83 88.32

15.43833 0.10833 0.10757 142.51 3170.68 84.35 71.73 71.24

15.45083 0.12083 0.11989 127.87 3178.39 92.06 80.09 79.46

15.46333 0.13333 0.13218 115.98 3187.12 100.79 85.25 84.53

15.47583 0.14583 0.14446 106.12 3194.24 107.91 92.29 91.40

15.49250 0.16250 0.16080 95.34 3205.96 119.63 110.52 109.37

15.50917 0.17917 0.17710 86.56 3216.68 130.35 119.27 117.89

15.52583 0.19583 0.19336 79.28 3227.89 141.56 129.06 127.44

15.54250 0.21250 0.20959 73.14 3238.37 152.04 136.87 135.00

15.55917 0.22917 0.22579 67.89 3249.07 162.74 146.06 143.92

15.58000 0.25000 0.24599 62.32 3261.79 175.46 155.10 152.61

15.62167 0.29167 0.28622 53.56 3287.21 200.88 172.44 169.27

15.66333 0.33333 0.32624 46.99 3310.15 223.82 192.65 188.52

15.70500 0.37500 0.36605 41.88 3334.34 248.01 212.55 207.52

15.74667 0.41667 0.40564 37.79 3356.27 269.94 208.64 203.21

15.78833 0.45833 0.44502 34.45 3374.98 288.65 216.24 209.94

15.83000 0.50000 0.48421 31.66 3394.44 308.11 241.12 233.50

15.87167 0.54167 0.52318 29.30 3413.90 327.57 265.11 256.05

15.91333 0.58333 0.56195 27.28 3433.83 347.50 247.95 239.07

15.95500 0.62500 0.60052 25.53 3448.05 361.72 253.29 243.21

15.99667 0.66667 0.63889 23.99 3466.26 379.93 283.06 271.37

16.03833 0.70833 0.67705 22.64 3481.97 395.64 243.76 233.17

16.08000 0.75000 0.71502 21.44 3493.69 407.36 258.62 246.28

16.14250 0.81250 0.77160 19.87 3518.63 432.30 298.44 283.73

16.20500 0.87500 0.82775 18.52 3537.34 451.01 258.25 244.44

16.26750 0.93750 0.88347 17.35 3553.55 467.22 273.98 258.07

16.33000 1.00000 0.93876 16.33 3571.75 485.42 278.26 261.44

16.39250 1.06250 0.99363 15.43 3586.23 499.90 283.74 265.20

16.45500 1.12500 1.04809 14.63 3602.95 516.62 301.05 280.62

16.51750 1.18750 1.10213 13.91 3617.41 531.08 288.60 267.90

16.58000 1.25000 1.15576 13.26 3631.15 544.82 253.30 234.50

16.64250 1.31250 1.20899 12.68 3640.86 554.53 247.56 227.85

Page 245: +Peters Ekwere j. - Petrophysics

3-45

16.70500 1.37500 1.26182 12.15 3652.85 566.52 281.16 258.06

16.76750 1.43750 1.31426 11.66 3664.32 577.99 263.46 241.04

16.83000 1.50000 1.36631 11.22 3673.81 587.48 247.74 225.68

16.95500 1.62500 1.46926 10.43 3692.27 605.94 227.42 206.12

17.08000 1.75000 1.57070 9.76 3705.52 619.19 210.56 188.93

17.20500 1.87500 1.67066 9.18 3719.26 632.93 224.74 200.33

17.33000 2.00000 1.76919 8.67 3732.23 645.90 208.37 184.96

17.58000 2.25000 1.96203 7.81 3749.71 663.38 162.84 142.20

17.70500 2.37500 2.05641 7.45 3757.19 670.86 150.46 130.44

17.83000 2.50000 2.14947 7.13 3763.44 677.11 140.26 120.70

18.08000 2.75000 2.33172 6.57 3774.65 688.32 140.46 119.19

18.33000 3.00000 2.50900 6.11 3785.11 698.78 138.98 116.46

18.58000 3.25000 2.68151 5.72 3794.06 707.73 115.07 95.49

18.83000 3.50000 2.84944 5.38 3799.80 713.47 133.68 108.08

19.08000 3.75000 3.01297 5.09 3809.50 723.17 150.18 121.27

19.33000 4.00000 3.17227 4.83 3815.97 729.64 107.37 85.60

19.58000 4.25000 3.32750 4.61 3820.20 733.87 64.24 50.83

19.83000 4.50000 3.47882 4.41 3821.95 735.62 40.75 31.50

20.08000 4.75000 3.62637 4.23 3823.70 737.37 56.31 42.75

20.33000 5.00000 3.77029 4.07 3826.45 740.12 79.61 59.90

20.58000 5.25000 3.91071 3.92 3829.69 743.36 87.17 65.00

20.83000 5.50000 4.04777 3.79 3832.64 746.31 74.57 55.11

21.08000 5.75000 4.18157 3.67 3834.70 748.37 72.12 52.33

21.33000 6.00000 4.31224 3.56 3837.19 750.86 70.51 50.88

21.58000 6.25000 4.43987 3.45 3838.94 752.61 34.46 25.33

22.08000 6.75000 4.68648 3.27 3838.02 751.69 19.52 12.42

22.58000 7.25000 4.92217 3.11 3840.78 754.45 53.06 36.18

23.08000 7.75000 5.14764 2.98 3843.01 756.68 43.29 28.98

23.58000 8.25000 5.36355 2.86 3844.52 758.19 41.52 26.89

24.08000 8.75000 5.57049 2.75 3846.27 759.94 40.82 26.16

24.58000 9.25000 5.76902 2.66 3847.51 761.18 33.25 20.81

25.08000 9.75000 5.95963 2.57 3848.52 762.19 40.20 24.38

25.58000 10.25000 6.14279 2.50 3850.01 763.68 37.65 22.86

26.08000 10.75000 6.31892 2.43 3850.75 764.42 32.21 18.81

26.58000 11.25000 6.48843 2.36 3851.76 765.43 33.95 19.69

Page 246: +Peters Ekwere j. - Petrophysics

3-46

27.08000 11.75000 6.65168 2.30 3852.50 766.17 36.54 20.56

27.58000 12.25000 6.80901 2.25 3853.51 767.18 38.36 21.44

28.08000 12.75000 6.96074 2.20 3854.25 767.92 36.51 19.89

28.58000 13.25000 7.10716 2.16 3855.07 768.74 30.55 16.56

29.08000 13.75000 7.24854 2.11 3855.50 769.17 27.64 14.43

29.83000 14.50000 7.45173 2.06 3856.50 770.17 32.70 16.92

30.58000 15.25000 7.64495 2.01 3857.25 770.92 30.23 15.15

31.33000 16.00000 7.82892 1.96 3857.99 771.66 32.51 15.89

32.08000 16.75000 8.00429 1.92 3858.74 772.41 34.83 16.64

32.83000 17.50000 8.17164 1.88 3859.48 773.15 30.98 14.58

33.58000 18.25000 8.33152 1.84 3859.99 773.66 33.60 15.21

34.33000 19.00000 8.48442 1.81 3860.73 774.40 27.78 12.67

35.08000 19.75000 8.63077 1.78 3860.99 774.66 23.22 10.01

35.83000 20.50000 8.77100 1.75 3861.49 775.16 40.28 17.08

36.58000 21.25000 8.90548 1.72 3862.24 775.91 39.27 16.70

37.58000 22.25000 9.07644 1.69 3862.74 776.41 26.72 10.90

38.58000 23.25000 9.23853 1.66 3863.22 776.89 21.33 8.60

39.58000 24.25000 9.39243 1.63 3863.48 777.15 24.52 9.34

40.58000 25.25000 9.53875 1.61 3863.99 777.66 33.76 12.75

41.58000 26.25000 9.67803 1.58 3864.49 778.16 25.91 9.71

42.58000 27.25000 9.81077 1.56 3864.73 778.40 23.88 8.48

43.83000 28.50000 9.96817 1.54 3865.23 778.90 30.25 10.62

45.33000 30.00000 10.14560 1.51 3865.74 779.41

The diagnostic plots depicted in Figure 3.19 show the classic response of a

homogeneous reservoir model with wellbore storage and skin. From the

welltest derivative function, it would appear that the effect of wellbore storage

had subsided at about 10 hrs, in which case, the pressure data beyond 10 hrs

might be fitted to the Horner semilog line. Notice the noise in the derivative

function. This is not surprising because numerical differentiation is a noisy

process. The wellbore storage coefficient can be calculated from the unit slope

line as

Page 247: +Peters Ekwere j. - Petrophysics

3-47

( )( )174 1.06 1 0.0096 RB/psi24 800

C ⎛ ⎞= =⎜ ⎟⎝ ⎠

The dimensionless wellbore storage coefficient is given by

( )( )( )( )( )( )

226

5.615 0.00969.09 10

2 0.25 4.20 10 107 0.29DC xxπ −

= =

Figure 3.19. Diagnostic plots for field example.

Figure 3.20 shows the Horner plot. The Horner straight line is given by

( ) 15.333880.60 78.52logwstP t

t+ Δ⎛ ⎞Δ = − ⎜ ⎟Δ⎝ ⎠

Page 248: +Peters Ekwere j. - Petrophysics

3-48

The slope of the Horner line is

78.52 psi/log cyclem = −

from which the permeability is calculated as

( )( )( )( )( )( )

162.6 174 2.5 1.068.92 md

78.52 107k = − =

Figure 3.20. Horner plot for field example.

From the Horner line,

( )1 3785.35wsP hr = psia

The skin factor can be calculated as

Page 249: +Peters Ekwere j. - Petrophysics

3-49

( )( )( )( )26

3086.33 3785.35 8.921.1513 log 3.2378.52 0.25 2.5 4.20 10 0.29

Sx −

⎧ ⎫⎡ ⎤−⎪ ⎪⎛ ⎞ ⎢ ⎥= − +⎨ ⎬⎜ ⎟− ⎢ ⎥⎝ ⎠⎪ ⎪⎣ ⎦⎩ ⎭

( )1.1513 8.90 7.61 3.23 5.21S = − + =

Thus, the well is damaged. The initial reservoir pressure is given by

* 3880.6iP P= = psia.

The radius of investigation for the drawdown portion of the test is given by

( )( )( )( )( )6

8.92 15.330.03248 234 ft

0.25 2.5 4.20 10invrx −

= =

The radius of investigation for the buildup portion of the test is given by

( )( )( )( )( )6

8.92 300.03248 330 ft

0.25 2.5 4.20 10invrx −

= =

The buildup test cannot investigate what the drawdown did not investigate.

Therefore, the radius of investigation of the buildup test test is obtained from

the drawdown portion of the test as 234 ft. The reservoir bulk volume

investigated by the test is given by

( )( )( )2 2 6 3234 107 18.406 10 ftinvV r h xπ π= = =

The estimated permeability of 8.92 md can be assigned to this volume of the

reservoir. This is in contrast to the core permeability, which is measured on a

minute portion of the reservoir.

The last step in a pressure transient analysis is to simulate the test. A

successful simulation of the test, though non-unique, tends to lend credibility

to the results of the analysis. Using the k, S and CD obtained from the semilog

analysis as preliminary estimates, the homogeneous reservoir model with

Page 250: +Peters Ekwere j. - Petrophysics

3-50

wellbore storage and skin can be used to simulate the test by a trial and error

procedure.

Figures 3.21 through 3.23 show graphical comparisons of the simulated

test and the field test. Figure 3.21 compares the pressure buildup responses;

Figure 3.22 compares the log-diagnostic plots; and Figure 3.23 compares the

Horner plots. Clearly, the agreement between the simulated test and the field

test is good. The best match was obtained with the following parameters:

9.4 k = md

6S = 29 10DC x=

2 81.46 10SDC e x=

( ) 3089.46wf pP t = psia

3880.60iP = psia

Thus, the simulation serves to fine tune the parameters obtained by the

conventional semilog analysis of the test.

Figure 3.21. A comparison of the simulated and the measured pressure buildup response for field example.

Page 251: +Peters Ekwere j. - Petrophysics

3-51

Figure 3.22. A comparison of the simulated and the measured

diagnostic plots for field example.

Figure 3.23. A comparison of the simulated and the measured Horner plots for field example.

Page 252: +Peters Ekwere j. - Petrophysics

3-52

3.4.10 Welltest Model for Dry Gas Reservoir

The diffusivity equation for real gas flow is given by

2 1tc M MMk t t

φμα

∂ ∂∇ = =

∂ ∂ (3.97)

The dependent variable in Eq.(3.97), M(P), is defined as

2( )o

P

P

PM P dPZμ

= ∫ (3.98)

where Po is a reference pressure. M(P) is known as the real gas potential or

the real gas pseudo pressure. Eq.(3.97) is similar in structure to Eq.(3.23)

that formed the basis for the welltest model for oil (and water) reservoirs. This

suggests that all the equations used to analyze transient pressure tests in an

oil reservoir can be adapted to analyze transient pressure tests in a dry gas

reservoir provided the pressure data from the gas well test are first

transformed into the real gas potential, M(P).

The dimensionless pressure for real gas flow in oilfield units is defined

as

( ) ( )1422

i wfD

sc

M P M PP

q Tkh

−=

⎛ ⎞⎜ ⎟⎝ ⎠

(3.99)

where

qsc = gas production rate, Mscf/D

T = absolute temperature of the reservoir, ºRankine (ºR = ºF+460)

The semilog drawdown line is given

( ) ( ) 2

1637 log log 3.23 0.87scwf i

t w

q T kM P M P t Skh c rφμ

⎡ ⎤⎛ ⎞= − + − +⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (3.100)

Page 253: +Peters Ekwere j. - Petrophysics

3-53

The gas velocity in the reservoir can be quite high given the low viscosity of

gas. This could lead to non-darcy flow with an additional pressure drop above

that of darcy flow. The effect of non-darcy flow can be expressed as a skin

factor as shown in Section 3.12. As a result, the total skin factor for a gas well

test, S, is normally split into the sum of two skin factors S* and Dqsc, where D

is a non-darcy coefficient and S* is the part of the total skin factor that does

not include the non-darcy effect. With this modification, Eq.(3.100) becomes

( ) ( ) *2

1637 log log 3.23 0.87 0.87scwf i sc

i ti w

q T kM P M P t S Dqkh c rφμ

⎡ ⎤⎛ ⎞= − + − + +⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦(3.101)

The slope of the semilog line is given by

1637 scq Tmkh

= − (3.102)

from which the formation permeability can be determined. The total skin

factor for drawdown test is given by

( ) ( )

121.1513 log 3.23

1637wf ihr

sc i ti w

M P M P kSq T c r

khφμ

⎡ ⎤⎢ ⎥− ⎛ ⎞⎢ ⎥= − +⎜ ⎟⎛ ⎞⎢ ⎥⎝ ⎠−⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.103)

Horner's buildup equation is given by

( ) ( )* 1637 log pscws

t tq TM P M Pkh t

+ Δ⎛ ⎞= − ⎜ ⎟Δ⎝ ⎠

(3.104)

Eq.(3.104) suggests that a graph of ( )wsM P versus log pt tt

+ Δ⎛ ⎞⎜ ⎟Δ⎝ ⎠

will be linear

with a negative slope, m, given by

1637 scq Tmkh

= − (3.96)

from which the formation permeability can be determined. The total skin

Page 254: +Peters Ekwere j. - Petrophysics

3-54

factor for a buildup test is given by

( ) ( )1

* * 21.1513 log 3.231637

wf ws hr

sc t w

M P M P kSq T c r

khφμ

⎡ ⎤⎢ ⎥− ⎛ ⎞⎢ ⎥= − +⎜ ⎟⎛ ⎞⎢ ⎥⎝ ⎠−⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.105)

where M(Pws)1hr is obtained from the semilog straight line or its extrapolation,

and *μ and *tc are obtained at the pressure intercept, P*. Diagnostic plots

should be used to screen the M(P) data for wellbore storage in the manner

described for oil welltests to determine the correct semilog lines if any.

The M(P) function for a dry gas reservoir can be computed in advance

using the gas properties. To do so requires the ability to estimate the gas Z

factor and viscosity as functions of pressure and temperature. Correlations

exist for these functions in the literature. Table 3.7 shows the gas properties

and the M(P) function for a dry gas reservoir with the following properties:

Reservoir temperature = 246 ºF + 459.67 = 705.67 ºR

Pseudocritical pressure = 665.0 psia

Pseudocritical temperature = 374.9 ºR

Apparent molecular weight of gas = 18.870 lb mass/lb mole

Apparent molecular weight of air = 28.9625 lb mass/lb mole

Density of fresh water = 62.368 lb mass/cu ft

Table 3.7: Variation of Gas Properties with Pressure

Pressure P/Z Bg ρg cg μg M(P)

(psia) Z (psia) (ft3/Mscf) (lbm/ft3) (106/psi) (cp) (psia2/cp)

200 0.986 203 98.120 0.505 5066.65 0.01393 2.8933E+06

400 0.974 411 48.428 1.024 2562.91 0.01413 1.1625E+07

600 0.962 624 31.895 1.554 1725.07 0.01437 2.6130E+07

800 0.951 841 23.656 2.095 1303.13 0.01466 4.6298E+07

1000 0.942 1062 18.736 2.646 1047.10 0.01498 7.1964E+07

Page 255: +Peters Ekwere j. - Petrophysics

3-55

1200 0.934 1285 15.477 3.203 873.69 0.01534 1.0292E+08

1400 0.927 1511 13.169 3.764 747.30 0.01573 1.3890E+08

1600 0.921 1737 11.456 4.327 650.20 0.01616 1.7962E+08

1800 0.917 1962 10.140 4.888 572.64 0.01661 2.2476E+08

2000 0.915 2186 9.102 5.446 508.85 0.01710 2.7397E+08

2200 0.914 2406 8.267 5.996 455.22 0.01760 3.2690E+08

2400 0.915 2623 7.583 6.537 409.42 0.01813 3.8321E+08

2600 0.917 2836 7.015 7.066 369.82 0.01868 4.4254E+08

2800 0.920 3043 6.538 7.582 335.30 0.01924 5.0455E+08

3000 0.925 3244 6.133 8.082 305.01 0.01982 5.6894E+08

3200 0.931 3438 5.786 8.567 278.31 0.02040 6.3540E+08

3400 0.938 3626 5.486 9.036 254.68 0.02100 7.0367E+08

3600 0.945 3808 5.225 9.488 233.71 0.02159 7.7350E+08

3800 0.954 3982 4.995 9.923 215.03 0.02220 8.4466E+08

4000 0.964 4150 4.793 10.341 198.36 0.02280 9.1697E+08

4200 0.974 4312 4.614 10.744 183.45 0.02340 9.9024E+08

4400 0.985 4467 4.454 11.130 170.07 0.02400 1.0643E+09

4600 0.997 4616 4.310 11.501 158.03 0.02460 1.1391E+09

4800 1.009 4759 4.181 11.857 147.19 0.02520 1.2144E+09

5000 1.021 4896 4.063 12.199 137.39 0.02579 1.2901E+09

5200 1.034 5028 3.957 12.528 128.52 0.02637 1.3662E+09

5400 1.048 5154 3.859 12.843 120.47 0.02695 1.4426E+09

5600 1.061 5276 3.770 13.147 113.14 0.02753 1.5192E+09

5800 1.075 5393 3.689 13.439 106.46 0.02810 1.5959E+09

6000 1.090 5506 3.613 13.719 100.37 0.02866 1.6727E+09

6200 1.104 5614 3.543 13.990 94.79 0.02921 1.7496E+09

6400 1.119 5719 3.479 14.250 89.67 0.02976 1.8265E+09

6600 1.134 5820 3.418 14.501 84.96 0.03030 1.9033E+09

6800 1.149 5917 3.362 14.743 80.63 0.03084 1.9801E+09

7000 1.165 6011 3.310 14.977 76.64 0.03137 2.0568E+09

Page 256: +Peters Ekwere j. - Petrophysics

3-56

3.5 FACTORS AFFECTING PERMEABILITY

Permeability is affected by compaction, pore size, sorting, cementation,

layering and clay swelling.

3.5.1 Compaction

Just as compaction reduces porosity, it also reduces permeability. As a

result of compaction, the permeability of rocks tends to decrease with depth of

burial.

3.5.2 Pore Size (Grain Size)

In general, for a sand, the permeability is proportional to the square of

the mean pore size. For a well sorted sand, the pore size is proportional to the

grain size. Therefore, for well sorted sands, the permeability is proportional to

the square of the grain size. Thus, a well sorted sand with a larger grain size

will have a higher permeability than a well sorted sand with a smaller grain

size (Table 3.8).

Table 3.8: Measured Porosities and Permeabilities of Artificial Sandpacks (adapted from Beard and Weyl, 1973)

Page 257: +Peters Ekwere j. - Petrophysics

3-57

To see why grain size affects the permeability of a medium, let us

compare the wetted surface area per unit bulk volume (specific surface area)

for flow through a pipe with no grains and through a porous medium with two

grain sizes. Figure 3.24 shows a cube of each medium of volume L3 meter3.

Let the porous medium consist of uniform spherical sand grains of radius r

meter and porosity φ. The wetted surface area of the pipe per unit bulk

volume (S) is

2

2 33

Surface Area 4 4 /Bulk Volume

LS m mL L

⎛ ⎞= = = ⎜ ⎟⎝ ⎠

(3.106)

Figure 3.24. Comparison of flow in a pipe with flow in a porous medium.

For the porous medium, the number of grains, N, contained in the L3

meter3 of bulk volume is given by

( )3

3

1Total Solid Volume4Volume of 1 sand grain3

LN

r

φ

π

−= =

⎛ ⎞⎜ ⎟⎝ ⎠

(3.107)

The wetted surface area is given by

Page 258: +Peters Ekwere j. - Petrophysics

3-58

( ) ( )3 32

3

1 3 1Area of 1sand grain 4

43

L LArea Nx x r

rr

φ φπ

π

− −= = =

⎛ ⎞⎜ ⎟⎝ ⎠

(3.108)

The wetted area per unit bulk volume (specific surface area) is given by

( ) ( )32 3

3

3 1 3 11 /L

S m mr L r

φ φ⎡ ⎤− −⎛ ⎞= =⎢ ⎥ ⎜ ⎟⎝ ⎠⎣ ⎦

(3.109)

Let us examine some numerical values. Let L = 1 meter and let the porous

medium consist of very fine sand with a grain diameter of 6100 10x − meter (r = 650 10x − meter) and a porosity of 28%. Figure 1.1 gives the diameters of the

grains of various clastic rocks. The specific surface area for the pipe flow is

2 34 /S m m=

The specific surface area for the porous medium is

( ) 2 36

3 1 .2843200 /

50 10S m m

x −

−= =

Thus, the specific surface area of the porous medium is 10,800 times that of

the pipe. Since the fluid velocity is zero at the wetted surface during flow (no

slip condition), it is obvious that much more energy will be required to sustain

the same volumetric flow rate through the porous medium than through the

pipe, everything else being equal. Put another way, it is obvious that the

porous medium is much less permeable than the pipe.

Let us examine the effect of grain size on the specific surface area of the

porous medium, and hence the permeability, by changing the grain size to a

fine silt with a grain diameter of 610 10x − meter (r = 65 10x − meter). In this case,

the specific surface area is

( ) 2 36

3 1 .28432000 /

5 10S m m

x −

−= =

Page 259: +Peters Ekwere j. - Petrophysics

3-59

Thus, the specific surface area of the fine silt has increased by a factor of 10

compared to the very fine sand. Using the same argument as before, it is

obvious that the silt will be less permeable than the fine sand. Therefore,

permeability decreases as the grain size decreases.

Example 3.2

An idealized petroleum reservoir consists of an unconsolidated sand of

uniform grain size. The sand grains were deposited in such a way as to give

the closest packing possible with spherical grains. This extremely well-sorted,

homogeneous and isotropic reservoir has the following dimensions:

Length = 3 miles

Width = 2 miles

Thickness = 250 feet

Grain diameter = 1/8 mm (fine sand)

Estimate the following properties for the reservoir:

a. Porosity.

b. Surface area per unit bulk volume (specific surface area).

c. Wetted surface area of the sand grains in square miles.

Notes:

1 mile = 5280 ft

1 ft = 30.48 cm

Solution to Example 3.2

L = 3 miles = (3)(5280)(30.48) = 482803.20 cm

W = 2 miles = (2)(5280)(30.48) = 321868.80 cm

h = 250 ft = (250)(30.48) = 7620.00 cm

Page 260: +Peters Ekwere j. - Petrophysics

3-60

d = 1/8 mm = 0.0125 cm (fine sand)

Rhombohedral packing with a porosity of 25.9% is the closest packing.

The specific surface area based on bulk volume is given by Eq.(3.109) as

( )( )

2 33 1 0.259355.68 /

0.0125 / 2S cm cm

−= =

The wetted surface are is given by

( )( ) 17 2Wetted Surface Area 355.68 482803.20 321868.80 7620.00 4.212 10bSxV x x x cm= = =

( ) ( )17 2

72 2 2

4.212 10Wetted Surface Area 1.626 105280 30.48 /

x cm xx cm mile

= = square miles.

It should be observed that the wetted surface area for the idealized

reservoir is enormous and would be considerably more if the grains were of

clay sized particles. This enormous surface area has implications for the

occurrence of surface phenomena in reservoir rocks.

3.5.3 Sorting

Poor sorting reduces the pore size and consequently reduces the

permeability of a medium (Table 3.8).

3.5.4 Cementation

Cementation reduces the pore size and consequently reduces the

permeability of the rock.

3.5.5 Layering

Permeability is a tensor and can therefore be different in different

directions. Because of layering in sedimentary rocks, horizontal permeabilities

in petroleum reservoirs tend to be higher than vertical permeabilities in the

absence of vertical fractures.

Page 261: +Peters Ekwere j. - Petrophysics

3-61

3.5.6 Clay swelling

Many consolidated sandstones contain clay and silt, e.g., arkoses and

graywackes. Since montmorillonite-type clays absorb fresh water and swell,

the permeabilities of such sandstones will be greatly reduced when measured

with fresh water. The addition of salts, such as sodium chloride or potassium

chloride, will in most cases, eliminate the clay swelling.

3.6 TYPICAL RESERVOIR PERMEABILITY VALUES

Reservoir permeabilities vary widely, from 0.001 md for a tight gas sand

in East Texas to 4000 md for an unconsolidated sand in the Niger Delta.

Reservoir permeabilities may be loosely described as follows:

Very low: k < 1md

Low: 1md < k < 10 md

Fair: 10 md < k < 50 md

Average: 50 md < k < 200 md

Good: 200 md < k < 500 md

Excellent: k > 500 md

Because of low gas viscosity, gas reservoirs with permeabilities less than 1 md

can still produce at economic rates if the reservoir is hydraulically fractured.

Because of the higher oil viscosities compared to gas viscosities, oil reservoirs

with permeabilities less than 1 md are unlikely to produce at economic rates

even with hydraulic fracturing.

3.7 PERMEABILITY-POROSITY CORRELATIONS

Since permeability depends on the continuity of pore space, there is not

in theory, or in fact, a unique relationship between the porosity of a rock and

its permeability. For unconsolidated sands, it is possible to establish

relationships between porosity and either some measure of apparent pore

Page 262: +Peters Ekwere j. - Petrophysics

3-62

diameter or specific surface area and permeability. However, these have very

limited applications.

In the same depositional and diagenetic environment, there is

sometimes a reasonable relationship between porosity and permeability,

although for a given porosity, permeabilities can vary widely. Figure 3.25

shows a typical permeability-porosity correlation obtained from core analysis.

The figure shows a correlation of permeability versus porosity for 500 samples

from a uniform sandstone reservoir in Pennsylvania. The permeabilities were

measured with gas and corrected for Klinkenberg effect. The porosity was

measured by saturating a well machined core with a wetting liquid. The

scatter in the data is not the result of experimental errors. Apart from

showing the general increase of permeability with an increase in porosity, the

correlation shows the wide spread, or lack of a close relationship, between

porosity and permeability. Nelson (1994) has presented a comprehensive

review of permeability - porosity correlations using data from the open

literature.

Attempts have been made to correlate permeability and porosity using

an equation of the form:

ln k a bφ= + (3.110)

where a and b are constants. This relationship is at best applicable at a local

level. Nevertheless, where it exists, it does provide one method of estimating

permeability from well logs and even possibly from drill cuttings. To test the

strength of the linear relationship between lnk and φ, let us examine the data

from sandstone cores shown in Table 3.9. Figure 3.26 shows the regression

line for the permeability - porosity correlation. The regression line is given by

ln 0.327 3.049k φ= −

with the square of the correlation coefficient (R2) equal to 0.3514. The

correlation is weak.

Page 263: +Peters Ekwere j. - Petrophysics

3-63

Figure 3.25. Permeability-porosity correlation for a sandstone reservoir (Ryder, 1948).

Table 3.9: Permeability - Porosity Data for Sandstone Cores (from Kenyon et al., 1986)

Sample φ k

No (%) (md)

1 31.2 2719

2 30.4 1382

3 24.8 3.3

4 28.5 215

5 33 151

Page 264: +Peters Ekwere j. - Petrophysics

3-64

6 31 426

7 30.3 1072

8 9.2 0.0012

9 22 32.1

10 22.8 83.2

11 20.4 4.56

12 19 311

13 18.6 7.47

14 20.9 20.5

15 21.7 52.5

16 14.8 4.17

17 17.6 10.9

18 11.3 4.5

19 11 3

20 10.1 2.8

21 10.4 3.4

22 13.7 2.1

23 12.8 1.3

24 12.2 7.3

25 11.7 4.8

26 17.2 91.9

27 16.7 190.4

28 18.6 424.9

29 18.3 270.3

30 18.3 137.3

31 16.3 28.5

32 15.6 200.5

33 17.2 36.5

34 16.4 243.2

35 18.3 337.3

36 18.1 39.9

37 14.9 64.6

38 19.6 669.5

39 18.3 540.5

40 18.8 890.4

Page 265: +Peters Ekwere j. - Petrophysics

3-65

41 17.8 429.8

42 16.1 0.011

43 18 0.52

44 16.6 0.52

45 16.1 0.087

46 17.2 12.1

47 22.5 5.97

48 19.6 1.72

49 12.7 0.063

50 15.2 0.93

51 19.1 20.3

52 14.8 4.72

53 20.5 45

54 20.5 45

55 23.9 663

56 23.8 591

57 22.9 511

58 21.6 478

59 22.2 131

60 22.3 1305

61 16.8 621

62 6.3 10.4

63 24.6 1425

64 24.3 2590

65 20 0.849

66 21.9 6

67 6.3 0.003

68 19.8 0.499

69 31.3 139

70 25.7 12.6

71 14.7 1.68

Page 266: +Peters Ekwere j. - Petrophysics

3-66

Figure 3.26. Permeability-porosity correlation for sandstone cores (from

Kenyon et al., 1986).

Figure 3.27 shows permeability - porosity correlations in which an effort

was made to classify the grain size of the sandstones. Samples containing

more than 50% of 1-2 mm grain size were classified as very coarse-grained;

0.5-1 mm, coarse-grained; 0.25-0.5 mm, medium-grained; and 0.1-0.25 mm,

fine-grained. Samples containing more than 10% of silt (<0.1 mm) were called

silty, whereas samples with clay (<0.004 mm) content greater than 7% were

classified as clayey. Both the absolute permeabilities and the effective

permeabilities to gas at irreducible water saturations were measured.

Permeabilities also were corrected for Klinkenberg effect. The permeabilities

shown in Figure 3.27 are the effective permeabilities to the nonwetting gas at

irreducible wetting phase saturation. These end point permeabilities are not

functions of the porous medium alone as in the case of the absolute

permeabilities. They are functions of the porous medium and the wetting and

nonwetting fluids. It would appear that accounting for the grain size

Page 267: +Peters Ekwere j. - Petrophysics

3-67

distribution improved the correlation of the end point effective permeabilites

with porosity. The author indicates that the correlations of the absolute

permeabilities with porosity showed the usual scatter regardless of the grain

size classification but the data were not presented.

Figure 3.27. Effect of grain size on the permeability-porosity correlations for various sandstones (from Chillingar, 1964).

The permeability - porosity correlation for unconsolidated sands is even

weaker than for consolidated sandstones. Figure 2.28 shows the correlations

based on the data of Table 3.3 for clean, unconsolidated and well sorted

sandpacks. The data are highly scattered. The R2 for the regression line for

the absolute permeability - porosity correlation is only 0.1742, which is a very

weak correlation. The R2 of 0.2305 for the correlation of the effective

Page 268: +Peters Ekwere j. - Petrophysics

3-68

permeability at irreducible water saturation versus porosity is higher than for

the absolute permeability. Thus, the effective permeability in the presence of

irreducible water saturation correlates better with porosity than the absolute

permeability of the medium. This observation is consistent with the

observation of the author of Figure 3.27.

Figure 3.28. Permeability - porosity correlations for unconsolidated

sandpacks (from Peters, 1979).

Page 269: +Peters Ekwere j. - Petrophysics

3-69

3.8 CAPILLARY TUBE MODELS OF POROUS MEDIA

3.8.1 Carman-Kozeny Equation

Because rock pore space is made up of a network of interconnected,

tortuous, flow conduits, attempts have been made to calculate permeability by

modeling pore-level flow as if through cylindrical capillary tubes. Such

models are sometimes termed “bundle-of-capillary-tubes” models and can

provide some insights into permeability - porosity relationships and the effect

of grain size on permeability. Figure 2.29 shows such a porous medium model

with n cylindrical and tortuous pores (capillary tubes). The Hagen-Poiseuille’s

law for steady flow through a single tortuous, circular capillary tube of radius

r is

( )4 41 2

8 8ie e

p pr r pqL L

π πμ μ

− Δ= = (3.111)

where

q1 = volumetric flow rate for a single capillary tube, cm3/s

r = the radius of the capillary tubes, cm

ΔP = pressure drop across the medium, dynes/cm2

μ = fluid viscosity, poise

Le = tortuous length of the capillary tube, cm

For n capillary tubes,

( )4 41 2

8 8Te e

p pn r n r pqL L

π πμ μ

− Δ= = (3.112)

where qT is the total volumetric flow rate for the porous medium. The porosity

of the medium is given by

Page 270: +Peters Ekwere j. - Petrophysics

3-70

2

p e

b T

V n r LV A L

πφ = = (3.113)

Figure 3.29. Bundle of capillary tubes model of a porous medium with tortuous pores.

where AT is the total cross-sectional area and L is the length of the porous

medium. Solving Eq.(3.113) for AT gives

2

eT

Ln rAL

πφ

⎛ ⎞= ⎜ ⎟⎝ ⎠

(3.114)

The integrated form of Darcy's law for single phase flow through the medium

is given by

TT

kA PqLμ

Δ= (3.115)

A comparison of Eq.(3.115) with Eq.(3.112) shows that the permeability of the

bundle of capillary tube model is given by

Page 271: +Peters Ekwere j. - Petrophysics

3-71

4

8 T e

n r LkA L

π ⎛ ⎞= ⎜ ⎟

⎝ ⎠ (3.116)

Substituting Eq.(3.114) into (3.116) gives the permeability for the model as

( )

2

28 /e

rkL Lφ

= (3.117)

The tortuosity of the porous medium is defined as

2

eLL

τ ⎛ ⎞= ⎜ ⎟⎝ ⎠

(3.118)

Thus, the permeability is given by

2

8rk φτ

= (3.119)

Eq.(3.119) shows once again that permeability is proportional to the square of

a characteristic dimension of the porous medium. In addition, it shows that

the permeability is also proportional to the porosity of the medium. The

tortuosity, τ, is a parameter that is greater than 1. Note that some authors

define tortuosity as (Le/L) instead of (Le/L)2 as defined here.

To enable the bundle of capillary tubes model to describe a porous

medium made of granular material, we introduce specific surface area into the

model. Let Sp be the wetted surface area of the pores per unit pore volume of

the porous medium. For cylindrical pores,

2

2 2ep

e

nrLSnr L rπ

π= = (3.120)

Substituting for r in Eq.(3.119) using Eq.(3.120) gives

22 p

kS

φτ

= (3.121)

Page 272: +Peters Ekwere j. - Petrophysics

3-72

What if the pore is not circular in cross-section? In that case, the 2 in

Eq.(3.121) can be replaced by an empirical numerical factor, ko , to give

2o p

kk S

φτ

= (3.122)

Eq.(3.122) is one version of the Carman-Kozeny equation. The factor, koτ, is

known as the Kozeny constant. Based on laboratory measurements, Carman

suggested that for granular material, the Kozeny constant, koτ, is equal to 5.

Thus, to estimate the permeability of such materials, Eq.(3.122) can be

written as

25 p

kSφ

= (3.123)

The permeability in Eq.(3.123) is in units of length squared. It can be easily

converted to darcy or millidarcy using Eq.(3.13).

Other versions of the Carman-Kozeny equation can be obtained

depending on how the specific surface area is defined. Let S be the wetted

surface area per unit bulk volume of the porous medium. Then

pS Sφ= (3.124)

Substituting Eq.(3.124) into (3.122) gives the following alternative version of

the Carman-Kozeny equation:

3

2o

kk S

φτ

= (3.125)

Eq.(3.125) shows that the permeability is proportional to the porosity cubed.

Let Ss be the wetted surface area per unit grain volume of the porous

medium. S and Ss are related by

( )1 sS Sφ= − (3.126)

Page 273: +Peters Ekwere j. - Petrophysics

3-73

Substituting Eq.(3.126) into (3.122) gives yet another version of the Carman-

Kozeny equation as

( )

3

22 1o s

kk S

φτ φ

=−

(3.127)

The three versions of the Carman-Kozeny equation, Eqs.(3.122), (3.125) and

(3.127) show that permeability is inversely proportional to the square of the

specific surface area of the grains. Since small grains have a higher specific

surface area than large grains, a medium composed of small grains will have

a lower permeability than one composed of large grains, a conclusion we

arrived at previously. All three versions of the Carman-Kozeny equation will

give the same permeability. Which one to use depends on what information is

on hand.

Two more versions of the Carman-Kozeny equation can be obtained by

introducing the concept of hydraulic radius into the model and introducing

the grain diameter in the case of granular material. The hydraulic radius, rH,

is defined as

Volume of Pore 1Wetted Surface Area of PoreH

p

rS

= = (3.128)

For cylindrical pores,

2

2 2e

He

n r L rrnrL

ππ

= = (3.129)

Substituting Eq.(3.129) into (3.119) gives

2

2Hrk φτ

= (3.130)

It is unfortunate that the hydraulic radius of a circular pore is equal to (r/2)

instead of r.

Page 274: +Peters Ekwere j. - Petrophysics

3-74

For granular material with spherical grains of diameter, Dp,

( )( )

2

3

4 / 2 64 / 23

ps

pp

DS

DD

π

π= = (3.131)

Substituting Eq.(3.131) into (3.127) gives

( )

2 3

236 1p

o

Dk

k

φ

τ φ=

− (3.132)

If ko = 2 is substituted into Eq.(3.132), then

( )

2 3

272 1pD

τ φ=

− (3.133)

Eq.(3.133) suggests that permeability is proportional to the square of the

grain size, an assertion we have encountered before.

To apply any of the Carman-Kozeny equations to estimate permeability,

the porosity, specific surface area and the Kozeny constant must be known.

Several techniques have been proposed for the determination of the specific

surface area of porous media. These include (1) a statistical method, (2)

adsorption methods, (3) the heat of wetting method and (4) a method based on

fluid flow developed by Kozeny.

Example 3.3

Estimate the permeability of the idealized petroleum reservoir of Example 3.2.

Solution to Example 3.3

Since the specific surface area with respect to bulk volume (S) is available

from Example 3.2, Eq.(3.125), which contains S is the most convenient

version of the Carman-Kozeny equation to use for the permeability

esstimation. Substituting numerical values into Eq.(3.125) gives

Page 275: +Peters Ekwere j. - Petrophysics

3-75

( )( )( ) ( )

3 8 28 2

2 9 2

0.259 2.7467 102.7467 10 2.78 9.869 10 /5 355.68

x cmk x cm darcysx cm darcy

−−

−= = = =

where the Kozeny constant of 5 has been used.

3.8.2 Tortuosity

Based on the bundle of capillary tube model, we see that tortuosity is a

geometric property of the porous medium that reflects the length of the flow

path at the pore level as the fluid flows around the grain obstacles relative to

the length of the porous medium. Therefore, the lower the porosity, the higher

the tortuosity should be. Winsauer et al. (1952) have measured the tortuosity

of sandstones along with other properties. The tortuosity was measured

electrically based on the analogy between the flow of electrical current and

fluid flow. Their results are shown in Table 3.10. They defined tortuosity as

(Le/L). Their original data have been squared to obtain the more current

definition of tortuosity.

Table 3.10: Tortuosity and Other Data from Winsauer et al., 1952.

Porosity Permeability Resistivity Tortuosity

Core φ k Factor τ

No % md F (Le/L) (Le/L)2

1 17.0 90 23.3 2.30 5.29

2 14.7 7 51.0 3.30 10.89

3 6.7 4 67.0 3.20 10.24

4 17.6 220 16.6 2.00 4.00

5 26.3 1920 8.6 1.60 2.56

6 25.6 4400 9.4 1.60 2.56

7 13.9 145 33.0 2.40 5.76

8 18.6 25 22.9 2.30 5.29

9 18.8 410 18.6 2.00 4.00

Page 276: +Peters Ekwere j. - Petrophysics

3-76

10 16.1 3 42.0 3.20 10.24

11 15.0 9 41.0 2.90 8.41

12 22.1 200 13.1 1.90 3.61

13 20.6 36 16.6 2.10 4.41

14 30.7 70 8.4 1.70 2.89

15 16.4 330 21.1 2.10 4.41

16 18.8 98 19.3 2.20 4.84

17 24.8 1560 10.8 1.80 3.24

18 19.1 36 17.2 2.30 5.29

19 29.8 1180 8.4 1.75 3.06

20 27.1 3200 11.7 2.05 4.20

21 28.2 2100 10.9 2.05 4.20

22 19.4 8 24.0 2.50 6.25

23 19.7 18 20.8 2.35 5.52

24 31.5 2200 6.9 1.50 2.25

25 19.3 19 24.4 2.70 7.29

26 27.3 88 12.4 2.20 4.84

27 25.1 370 11.6 1.97 3.88

28 15.0 115 37.3 2.90 8.41

29 18.4 130 19.0 2.00 4.00

30

31 39.5 4.7 1.37 1.88

Figure 3.30 shows the correlation between tortuosity and porosity. As

expected, there is a negative correlation between the two variables. The lower

the porosity, the longer is the length of the flow path that the fluid particles

must take to flow from one end of the porous medium to the other. The longer

the flow path, the larger the tortuosity of the medium.

Figure 3.31 shows the correlation between tortuosity and the formation

resistivity factor. As expected, there is a strong positive correlation between

the two variables. Resistivity factor represents the obstacle to the flow of

electric current in the medium whereas tortuosity represents the obstacle to

the flow of fluid in the medium. Since there is a close relationship between the

Page 277: +Peters Ekwere j. - Petrophysics

3-77

flow path for electric current and for fluid flow, a property that obstructs one

will also obstruct the other. Hence, tortuosity and formation resistivity factor

should be positively correlated.

Figure 3.30. Tortuosity-porosity correlation (from Winsauer et al., 1952).

Page 278: +Peters Ekwere j. - Petrophysics

3-78

Figure 3.31. Tortuosity-formation resistivity factor correlation (from Winsauer

et al., 1952).

An approximate relationship between tortuosity and formation

resistivity factor can be derived as follows. Consider the flow of electric

current through a core sample of porosity φ, length L and cross-sectional area

A saturated with water of resistivity Rw. The electrical resistance of the core is

given by

o oLr RA

⎛ ⎞= ⎜ ⎟⎝ ⎠

(3.134)

where Ro is the resistivity of the core fully saturated with the water. The

electrical current in the core is conducted by the water alone since the rock

minerals are nonconductors. Let us replace the core with the same volume of

water in a test tube such that the resistance of the water is equal to that of

the saturated core. Since the electrical current in the core follows a tortuous

path in flowing from one end to the other, the length of the equivalent water

Page 279: +Peters Ekwere j. - Petrophysics

3-79

circuit Le will be longer than L. Let Ae be the cross-sectional area of the

equivalent water circuit. Then

e eA L ALφ= (3.135)

The equivalent water circuit must have the same volume as the water in the

core to maintain the same salinity. The resistance of the equivalent water

circuit is given by

ew w

e

Lr RA

⎛ ⎞= ⎜ ⎟

⎝ ⎠ (3.136)

Substituting Eq.(3.135) into (3.136) gives

2e

w wLr RALφ

⎛ ⎞= ⎜ ⎟

⎝ ⎠ (3.137)

Since the equivalent water circuit has the same resistance as the saturated

core, Eqs.(3.134) and (3.137) give

2e

o wLLR R

A ALφ⎛ ⎞⎛ ⎞ = ⎜ ⎟⎜ ⎟

⎝ ⎠ ⎝ ⎠ (3.138)

Eq.(3.138) can be rearranged as

2 1o e

w

R LR L

τφ φ

⎛ ⎞⎛ ⎞= =⎜ ⎟⎜ ⎟⎝ ⎠ ⎝ ⎠

(3.139)

Thus, an approximate relationship between turtuosity and the formation

resistivity factor is

Fτ φ= (3.140)

3.8.3 Calculation of Permeability from Pore Size Distribution

The bundle of capillary tubes model presented in Section 3.7.1 assumed

that the n capillary tubes had the same pore size. Here, we extend the model

to include an arbitrary pore size distribution. Given the probability density

Page 280: +Peters Ekwere j. - Petrophysics

3-80

function (pdf) for the pore size (pore diameter) distribution, we wish to

calculate the permeability of the bundle of capillary tubes model.

Consider the probability density function for pore diameter distribution

as shown in Figure 3.32. The probability density function satisfies

( )0

1f dδ δ∞

=∫ (3.141)

Let nδ be the number of pores with diameter between δ and δ+dδ and n be the

total number of pores. Then

( )n nf dδ δ δ= (3.142)

Figure 3.32. Probability density function for pore size distribution.

The cross-sectional area occupied by pores with diameter between δ and δ+dδ

is given by

Page 281: +Peters Ekwere j. - Petrophysics

3-81

( )2

4cdA nf dδπ δ δ⎛ ⎞

= ⎜ ⎟⎝ ⎠

(3.143)

The cross-sectional area occupied by all the pores is obtained by integrating

Eq.(3.143) as

( )2

2

04 4cn nRA f dπ πδ δ δ

∞= =∫ (3.144)

where 2R is a constant defined as

( )2 2

0R f dδ δ δ

∞= ∫ (3.145)

The cross-sectional area occupied by all the pores is related to the cross-

sectional area of the porous medium by the porosity as

c TA A φ= (3.146)

Substituting Eq.(3.144) into (3.146) and solving for n, one obtains

2

4 TAnR

φπ

= (3.147)

The volumetric flow rate for pores with diameter between δ and δ+dδ is

obtained from Hagen-Poiseuille's law as

( )4

128Te

Pdq nf dL

πδ δ δμ

Δ= ⎡ ⎤⎣ ⎦ (3.148)

The total volumetric flow rate for the porous medium is obtained by

integrating Eq.(3.148) as

( ) 4

0128Te

n Pq f dL

π δ δ δμ

∞Δ= ∫ (3.149)

Substituting Eq.(3.147) into (3.149) gives

Page 282: +Peters Ekwere j. - Petrophysics

3-82

( ) 42 032

TT

e

A Pq f dL R

φ δ δ δμ

∞Δ= ∫ (3.150)

A comparison of Eq.(3.150) with Darcy's law, Eq.(3.115), gives the

permeability of the bundle of capillary tube model as

( ) ( ) 4

2 032 /e

k f dL L R

φ δ δ δ∞

= ∫ (3.151)

Substituting Eq.(3.118) into (3.151) gives

( ) ( ) 4

2 032k f d

Rφ δ δ δτ

∞= ∫ (3.152)

Substituting Eq.(3.145) into (3.152) gives the final result

( )

( )( )

4

0

2

032

f dk

f d

δ δ δφτ δ δ δ

⎡ ⎤⎢ ⎥=⎢ ⎥⎢ ⎥⎣ ⎦

∫∫

(3.153)

Example 3.4

The probability density function for the pore size distribution of a bundle of

capillary tubes model is given by a triangular distribution as shown in Figure

3.33. The porosity of the medium is 15%. If the minimum pore diameter (δ1) =

0 μm, the most likely pore diameter (δ2) = 8 μm and the largest pore diameter

(δ3) = 10 μm, calculate the permeability of the medium. Assume a tortuosity of

1.0.

Page 283: +Peters Ekwere j. - Petrophysics

3-83

Figure 3.33. Probability density function for pore size distribution for Example

Solution to Example 3.4

The required permeability is given by Eq.(3.153). All we need to do is to

evaluate the two integrals on the right of the equation for the triangular

probability density function. Eq.(3.153) then takes the form

( )

( ) ( )

( ) ( )

2 3

1 2

2 3

1 2

4 41 2

2 21 2

32

f d f dk

f d f d

δ δ

δ δδ δ

δ δ

δ δ δ δ δ δφτ δ δ δ δ δ δ

⎡ ⎤+⎢ ⎥= ⎢ ⎥+⎢ ⎥⎣ ⎦

∫ ∫∫ ∫

(3.154)

Eqs.(2.82) and (2.86) give

( ) ( )( )( )

11

3 1 2 1

2f

δ δδ

δ δ δ δ−

=− −

(3.155)

( ) ( )( )( )

32

3 1 3 2

2f

δ δδ

δ δ δ δ−

=− −

(3.156)

Substituting Eqs.(3.155) and (3.156) into Eq.(3.154) and performing the

integrations gives

Page 284: +Peters Ekwere j. - Petrophysics

3-84

3 13 22

1 13 2

0.15 2.241 10 2.583 10 0.2617 261.74.067 10 9.869 10 /32 1.0

x x mk m darcy mdx x m darcy

μ−

⎛ ⎞= = = =⎜ ⎟

⎝ ⎠

3.9 STEADY STATE FLOW THROUGH FRACTURES

To increase the productivities of certain reservoirs, it is often necessary

to hydraulically fracture the reservoirs. Consider the single vertical fracture

shown in Figure 3.34. It can be shown that Hagen-Poiseuille’s law for steady

state flow through the fracture of width w in cgs units is given by

2

12w A Pq

LμΔ

= (3.157)

where A = wh is the cross-sectional area of the fracture. A comparison of

Eq.(3.157) with Darcy’s law, Eq.(3.115), shows that the permeability of the

fracture is given by

2

2

12wk cm= (3.158)

Example 3.5

Calculate the permeability of a fracture of width 2 millimeters.

Solution to Example 3.5

The fracture permeability is given by Eq.(3.158) as

( )2 25

9 2

0.20 1 3.378 1012 9.869 10 /

cmk x x darcys

x cm darcy−= =

The permeability of the fracture is enormous and therein lies the benefit of

fracturing for the improvement of the productivity of a well.

Page 285: +Peters Ekwere j. - Petrophysics

3-85

Figure 3.34. Flow through a fracture (Flow between parallel plates).

3.10 AVERAGING PERMEABILITY DATA

Because of heterogeneity, different portions of the same reservoir may

have different permeabilities. It becomes necessary to be able to estimate the

apparent or average permeability of various permeability combinations.

For linear beds in series (Figure 3.35), the volumetric flow rate is the

same for each bed and the total pressure drop is equal to the sum of the

pressure drop across each bed. For n beds in series, the average permeability

is given by

1

1 1

n

iin n

i i

i ii i

LLk

L Lk k

=

= =

= =∑

∑ ∑ (3.159)

where L Is total length of the porous medium in the flow direction, Li is the

length of bed i and ki is the permeability of bed i.

Page 286: +Peters Ekwere j. - Petrophysics

3-86

Figure 3.35. Linear beds in series For radial beds in series (Figure 3.36), the volumetric flow rate is the

same for each bed and the total pressure drop is equal to the sum of the

pressure drop across each bed. For n beds in series, the average permeability

is given by

( )( )1

1

ln /ln /

e wn

i i

i i

r rk

r rk

=

=

∑ (3.160)

where re is the well's drainage radius and r0 is equal to the wellbore radius, rw

Figure 3.36. Radial beds in series

Page 287: +Peters Ekwere j. - Petrophysics

3-87

For linear beds in parallel (Figure 3.37), the total volumetric flow rate is

the sum of the flow rate of each bed and the pressure drop is the same for all

the beds. For n beds in parallel, the average permeability is given by

1

1

n

i ii

n

ii

k Ak

A

=

=

=∑

∑ (3.161)

where Ai is the area of bed i and ki is the permeability of bed i. If the beds

have the same width, then

1

1

n

i ii

n

ii

k hk

h

=

=

=∑

∑ (3.162)

Figure 3.37. Linear beds in parallel.

For radial beds in parallel (Figure 3.38), the total volumetric flow rate is

the sum of the flow rate of each bed and the pressure drop is the same for all

the beds. For n beds in parallel, the average permeability is given by

Eq.(3.162).

Page 288: +Peters Ekwere j. - Petrophysics

3-88

Figure 3.38. Radial beds in parallel.

3.11 DARCY'S LAW FOR INCLINED FLOW

For inclined flow, Darcy's Law in differential form can be written as

61.0133 10kA dP g dzq

ds x dsρ

μ⎛ ⎞= − ±⎜ ⎟⎝ ⎠

(3.163)

where

q = volumetric flow rate, cm3/s

k = absolute permeability of the porous medium, darcy

A = total area of the medium normal to the flow direction, cm2

μ = fluid viscosity, centipoise

P = absolute pressure, atm

s = distance in the direction of flow and is always positive, cm

dPds

= pressure gradient along s, atm/cm

ρ = fluid density, gm/cm3

g = acceleration of gravity, 981 cm/s2

1 atm = 1.0133x106 dynes/cm2

The sign convention for applying Darcy's law for inclined flow is as follows. If

the z direction is positive upwards as shown in Figure 3.39, then the positive

Page 289: +Peters Ekwere j. - Petrophysics

3-89

sign is used in the term in parenthesis in Eq.(3.163) and the equation

becomes

Figure 3.39. Coordinate system for applying Darcy's law for z positive upwards.

61.0133 10kA dP g dzq

ds x dsρ

μ⎛ ⎞= − +⎜ ⎟⎝ ⎠

(3.164)

On the other hand, if the z coordinate is positive downwards as shown in

Figure 3.40, then the negative sign is used in the term in parenthesis in

Eq.(3.163) to obtain

61.0133 10kA dP g dzq

ds x dsρ

μ⎛ ⎞= − −⎜ ⎟⎝ ⎠

(3.165)

Page 290: +Peters Ekwere j. - Petrophysics

3-90

Figure 3.40. Coordinate system for applying Darcy's law for z positive downwards.

dzds

in Eqs.(3.163) to (3.165) could be written as sinθ, where θ is the angle the

porous medium along s makes with the horizontal plane. Alternatively, it

could be written as cosα, where α is the angle that the porous medium makes

with the upward vertical z axis and is equal to 2π θ⎛ ⎞−⎜ ⎟

⎝ ⎠.

Eq.(3.163) can be written in oilfield units as

0.001127 0.433kA dP dzqds ds

γμ

⎛ ⎞= − ±⎜ ⎟⎝ ⎠

(3.166)

where γ is the specific gravity of the fluid and 0.433 is the pressure gradient of

fresh water in psi/ft.

Eq.(3.163) also can be written in terms of a velocity potential as

kA dqdsμΦ

= − (3.167)

Page 291: +Peters Ekwere j. - Petrophysics

3-91

where the velocity potential is defined by

61.0133 10gzPx

ρΦ = ± (3.168)

in Darcy units or

0.433P zγΦ = ± (3.169)

in oilfield units. Eq.(3.167) shows that flow will occur in the direction of

decreasing velocity potential. Note that for inclined systems, flow does not

necessarily occur in the direction of decreasing pressure. For example, for a

static system in which the porous medium is oriented vertically, the pressure

at the top of the medium is less than at the bottom and yet there is no flow.

Eq.(3.156) may also be written as

6

6

1.0133 101.0133 10s

q k g d x Pv zA x ds g

ρμ ρ

⎛ ⎞= = − ±⎜ ⎟

⎝ ⎠ (3.170)

where vs is known as the Darcy velocity or superficial velocity. An interstitial

velocity can be defined as

si

vvφ

= (3.171)

Eq.(3.170) can be further written in terms of the hydraulic conductivity, K,

and hydraulic or piezometric head, h, as

sq dhv KA ds

= = − (3.172)

where

61.0133 10k gK

μ= (3.173)

and

Page 292: +Peters Ekwere j. - Petrophysics

3-92

Ph zgρ

= ± (3.174)

Eq.(3.172) is the form of Darcy's law normally used in the groundwater

community. Note that hydraulic head has dimension of length and the

hydraulic conductivity has dimension of length/time. Possible units of

hydraulic conductivity are cm/s, m/s, m/d, US gal/day ft2 and US gal/min

ft2. Useful conversion factors are as follows:

2 5 21 gal/day ft 4.75 10 cm/s 4.08 10 /US x x m d− −= = (3.175)

41 9.613 10 cm/s (for water at 20 )darcy x C−= (3.176)

2 21 1.4156 10 gal/min ft (for water at 20 )darcy x US C−= (3.177)

Table 3.11 shows typical hydraulic conductivity values for various rock types

and soils. Figure 3.41 shows the range of hydraulic conductivity and

permeability for various rock types and soils. Clearly, rocks and soils have a

wide range of permeability and hydraulic conductivity values.

Sometimes, it is easier to solve single phase flow problems in terms of

hydraulic head (piezometric head) rather than in terms of pressure. This is

especially true for inclined or vertical flow in which gravity plays a role.

Figure 3.42 shows the definition of the hydraulic head (piezometric head) for a

point O in an inclined porous medium located above a datum level at which z

is zero. The hydraulic head is given by

h zψ= + (3.178)

Page 293: +Peters Ekwere j. - Petrophysics

3-93

Table 3.11: Typical Hydraulic Conductivity for Various Rock Types

Page 294: +Peters Ekwere j. - Petrophysics

3-94

Figure 3.41. Range of permeability and hydraulic conductivity of various rock types and soils.

where h is the hydraulic head, z is the height of O above (or below) an

arbitrary datum and ψ is the pressure head. It should be emphasized that z in

Eq.(3.178) is not a cartesian coordinate in the flow direction but rather the

elevation of the point P above or below a reference datum. In a sense, z is a

gravity head that reflects the potential energy of the fluid at O in the earth’s

gravitational field. The datum at which z = 0 is arbitrary and can be selected

anywhere in the system. If the point O is below the datum, then z is negative.

The pressure head is given by

Pg

ψρ

= (3.179)

Page 295: +Peters Ekwere j. - Petrophysics

3-95

It should be observed in Figure 3.42 that the hydraulic head at point O is

given by the elevation of a liquid manometer at point O above (or below) the

arbitrary datum.

Figure 3.42. Definition of hydraulic head h, pressure head Ψ and elevation head z for a laboratory manometer.

Consider the flow in an inclined porous medium shown in Figure 3.43.

Darcy’s law for 1D flow is given by

dhq KA h KAds

= − ∇ = − (3.180)

Examination of the manometer levels (hydraulic heads) immediately shows

the direction of flow. Flow always occurs from a high hydraulic head to a low

hydraulic head. Eq.(3.180) may be integrated to obtain

2 1

2 1

h hq KAs s

⎛ ⎞−= − ⎜ ⎟−⎝ ⎠

(3.181)

Eq.(3.174) can be rewritten as

1 2

2 1

h h hq KA KAs s L

⎛ ⎞− Δ⎛ ⎞= =⎜ ⎟ ⎜ ⎟− Δ⎝ ⎠⎝ ⎠ (3.182)

Page 296: +Peters Ekwere j. - Petrophysics

3-96

Figure 3.43. Darcy’s experiment for inclined flow.

where Δh is the positive difference in the hydraulic heads at points 1 and 2

and ΔL is the positive length of the porous medium between points 1 and 2 in

the flow direction. It should be observed that the elevations of points 1 and 2

(z1 and z2) do not appear explicitly in Eq.(3.182) although they do appear

implicitly in the hydraulic heads. However, Eq.(3.182) can be rewritten such

that z1 and z2 appear explicitly as follows. From Figure 3.43,

1 2cosL z zαΔ = − (3.183)

where α is the angle the porous medium makes with the vertical direction.

Substituting Eq.(3.183) into Eq.(3.182) gives

1 2

1 2

cosh hq KAz z

α⎛ ⎞−

= ⎜ ⎟−⎝ ⎠ (3.184)

Page 297: +Peters Ekwere j. - Petrophysics

3-97

In the special case in which the porous medium is vertical, α is zero, cosα is

1.0 and z1 – z2 = ΔL. Substituting these values into Eq.(3.184) gives

hq KAL

Δ⎛ ⎞= ⎜ ⎟Δ⎝ ⎠ (3.185)

which is identical to Eq.(3.182). Thus, Eq.(3.182) applies for all angles of

inclination. The hydraulic conductivity is related to the permeability of the

porous medium in consistent units by

k gK ρμ

= (3.186)

Figure 3.44 shows schematic diagrams of two types of liquid

permeameters normally used to measure permeability in groundwater

hydrology. Figure 3.44 (a) is a constant head permeameter whereas Figure

3.44 (b) is a falling head permeameter. Using the principles of inclined flow

outline above, the working equations for the permeameters can easily be

derived as follows. For the constant head permeameter, Eq.(3.182) gives

h k g hq KA AL L

ρμ

⎛ ⎞= = ⎜ ⎟

⎝ ⎠ (3.187)

in consistent units, where A is the cross-sectional area of the porous medium.

For the falling head permeameter, Darcy's law at time t is

hq KAL

= (3.188)

The volumetric balance for flow in the manometer gives

dhq adt

= − (3.189)

where a is the cross-sectional area of the manometer. Substituting Eq.(3.187)

into (3.186) gives the differential equation for the falling head, h, as

Page 298: +Peters Ekwere j. - Petrophysics

3-98

Figure 3.44. Schematic diagram of liquid permeameters; (a) constant head

permeameter, (b) falling head permeameter.

dh KA hdt aL

⎛ ⎞= −⎜ ⎟⎝ ⎠

(3.190)

with the initial condition

( )0 oh h= (3.191)

Eq.(3.190) along with (3.191) can easily be solved to obtain the working

equation for the falling head permeameter as

Page 299: +Peters Ekwere j. - Petrophysics

3-99

lno

h KA k gAt th aL aL

ρμ

⎛ ⎞ ⎛ ⎞⎛ ⎞= − = −⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠⎝ ⎠

(3.192)

in consistent units.

3.12 VALIDITY OF DARCY'S LAW

Darcy's law is valid for slow, laminar flow in porous media without

chemical reaction. Figure 3.45 shows the Fanning friction factor versus

Reynolds number for single phase flow in a porous medium. For flow through

porous media, Reynolds number is defined as

Re pv Dρμ

= (3.193)

where

v = Darcy velocity, cm/s

Dp = mean grain diameter of the granular porous medium, cm

ρ = fluid density, gm/cm3

μ = fluid viscosity, poise

Figure 3.45. Fanning friction factor factor for flow in porous media (Rose,

1945)

Page 300: +Peters Ekwere j. - Petrophysics

3-100

For laminar flow, in which Darcy's law is valid, the friction factor versus

Reynolds number is given by the line shown in Figure 3.45 having the

equation

1000Re

f = (3.194)

It can seen in the figure that the experimental data begin to deviate from the

line at a Reynolds number of about 1.0. Thus, Darcy's law is valid for flow in

porous media for Reynolds number up to about 1. At Reynolds numbers

greater than 1, Darcy's law is no longer valid. Therefore, flow at Reynolds

number greater than 1 can be characterized as non-darcy flow.

Based on analogy with flow in pipes, the flow regimes for non-darcy flow

can be classified as shown in Figure 3.46 as transition flow for Reynolds

number in the range 1 to 100, and turbulent flow for Reynolds number

greater than 100.

Figure 3.46. Classification of flow in porous media.

Page 301: +Peters Ekwere j. - Petrophysics

3-101

3.13 NON-DARCY FLOW

Darcy’s law is inadequate for modeling high velocity flow in porous

media such as high velocity gas flow near the wellbore in a natural gas

reservoir in which the Reynolds number is greater than 1. For this non-darcy

flow, Forchheimer has proposed a flow equation of the form

2dP v vdx k

μ βρ− = + (3.195)

where the variables are in Darcy units. The first term on the right-side of

Eq.(3.195) gives the viscous component of the pressure gradient (Darcy flow)

whereas the second term gives the inertial-turbulent component of the

pressure gradient, which constitutes the non-darcy flow effect. Because the

non-darcy flow effect is most severe near the wellbore, this additional effect is

usually incorporated into the gas flow equation as a rate-dependent skin

factor. The velocity coefficient (β) is a function of the permeability and porosity

of the porous medium and is independent of rate. Figure 3.47 shows a

correlation of the velocity coefficient with permeability from core analysis.

Consider radial steady state flow of a real gas from the reservoir into the

well. For this case, Eq.(3.195) can be written as

2dP v vdr k

μ βρ= + (3.196)

Let us multiply Eq.(3.196) by (2p/μZ) and integrate to get

22 2 2e e e

wf w w

P r r

P r r

P P v PdP dr v drZ Z k Z

μ βρμ μ μ

= +∫ ∫ ∫ (3.197)

The gas density in Eq.(3.197) is at reservoir conditions. We would like to

evaluate the density at standard conditions. The density at reservoir

conditions and at standard conditions are related by the gas formation

volume factor as

Page 302: +Peters Ekwere j. - Petrophysics

3-102

Figure 3.47. Velocity coefficient in Forchheimer equation (Firoozabadi and Katz, 1979.

sc

gBρρ = (3.198)

where Bg is the gas formation volume factor and ρsc is the gas density at

standard conditions. For real gas flow, the gas formation volume factor is

given by

scg

sc

P ZTBT P

= (3.199)

The Darcy velocity at radius r is given by

Page 303: +Peters Ekwere j. - Petrophysics

3-103

2

qvrhπ

= (3.200)

The equation of state for real gas flow is given by

sc sc

sc

P qPqZT Z T

= (3.201)

Substituting Eqs.(3.198) to (3.201) into (3.197) gives

2

2 22

e wf e e

o o w w

P P r rsc sc sc sc sc scP P r r

sc sc

q P T q P T kqP P dr drdP dPZ Z khT r khT h r

βρμ μ π π πμ

⎛ ⎞⎛ ⎞− = + ⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠∫ ∫ ∫ ∫ (3.202)

where Po is an arbitrary base pressure. In Eq.(3.202), the left–hand side of

Eq.(3.197) has been split into two integrals for convenience. Integration of

Eq.(3.202) gives

( ) ( ) 1 1ln2

sc sc e sc sc sc sce w

sc w sc w e

q P T r q P T kqM P M PkhT r khT h r r

βρπ π πμ

⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎛ ⎞− = + −⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠ ⎝ ⎠ ⎝ ⎠ (3.203)

where M(P) is the real gas potential or real gas pseudo pressure. At field scale,

(1/re) is normally negligibly small compared to (1/rw). With this

approximation, Eq.(3.203) becomes

( ) ( ) 1ln2

sc sc e sc sc sc sce wf

sc w sc w

q P T r q P T kqM P M PkhT r khT h r

βρπ π πμ

⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎛ ⎞− = +⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠ ⎝ ⎠ ⎝ ⎠ (3.204)

Eq.(3.204) can be written in dimensionless form as

ln eD sc

w

rP Dqr

⎛ ⎞= +⎜ ⎟

⎝ ⎠ (3.205)

where the dimensionless pressure in terms of M(P) is defined as

( ) ( )e wf

Dsc sc

sc

M P M PP q P T

khTπ

−= (3.206)

Page 304: +Peters Ekwere j. - Petrophysics

3-104

and a non-darcy flow coefficient is defined as

2

sc

w

kDhr

βρπμ

= (3.207)

Eq.(3.205) can be modified to include the mechanical skin factor S* to give

*ln eD sc

w

rP S Dqr

⎛ ⎞= + +⎜ ⎟

⎝ ⎠ (3.208)

Eq.(3.208) can be rewritten in the form of an inflow performance relationship

in Darcy units as

( ) ( )

*ln

e wfscsc

sc esc

w

M P M PkhTqP T r S Dq

r

π⎡ ⎤⎢ ⎥−⎢ ⎥=⎢ ⎥⎛ ⎞

+ +⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.209)

In oilfield units, Eq.(3.209) becomes

( ) ( )

*1422ln

e wfsc

esc

w

M P M PkhqT r S Dq

r

⎡ ⎤⎢ ⎥−⎢ ⎥=⎢ ⎥⎛ ⎞

+ +⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(3.210)

It should be noted that the skin factor normally determined from well

test analysis is the total skin factor given by

*scS S Dq= + (3.211)

Therefore, the total skin factor determined from a gas well test should be

investigated further to determine the contribution of the non-darcy flow

component. The non-darcy flow component of the total skin factor cannot be

eliminated by well stimulation. It can be estimated by conducting drawdown

tests at two different flow rates.

Page 305: +Peters Ekwere j. - Petrophysics

3-105

It can be shown that the additional potential drop due to non-darcy flow

occurs mostly near the wellbore. This additional potential drop at any radius r

can be obtained from Eq.(3.203) as

( ) 1 12

sc sc sc sc

sc w

q P T kqM rkhT h r r

βρπ πμ

⎛ ⎞ ⎛ ⎞⎛ ⎞Δ = −⎜ ⎟ ⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠ ⎝ ⎠ (3.212)

At r =re,

( ) 12

sc sc sc sce

sc w

q P T kqM rkhT h r

βρπ πμ

⎛ ⎞ ⎛ ⎞⎛ ⎞Δ = ⎜ ⎟ ⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠ ⎝ ⎠ (3.213)

where (1/re) is negligibly small compared to (1/rw). Thus,

( )( ) 1 w

e

rM rM r r

Δ= −

Δ (3.214)

Table 3.12 shows the variation of Δm(r)/Δm(re) with r as calculated from

Eq.(3.214). It can be observed that 90% of the total non-darcy flow potential

drop occurs within r = 10rw. Thus, if rw is 0.3 ft, then 90% of the total non-

darcy flow potential drop occurs within 3 ft of the well. This is the justification

for treating the non-darcy flow effect as a flow-dependent skin factor.

Table 3.12: Non-Darcy Potential Drop

w

rr

( )( )

e

M rM r

ΔΔ

1 0 2 0.50 5 0.80 10 0.90

Page 306: +Peters Ekwere j. - Petrophysics

3-106

3.14 DARCY’S LAW FOR ANISOTROPIC POROUS MEDIA

3.14.1 Definition of Homogeneity and Anisotropy

A medium is homogeneous with respect to a certain property if that

property is independent of position within the medium. Otherwise, the

medium is heterogeneous. A medium is isotropic with respect to a certain

property if that property is independent of direction within the medium. If at a

point in the medium, a property of the medium varies with direction, the

medium is said to be anisotropic with respect to that property. Figure 3.48

shows the principal values of permeability anisotropy at two locations (x1, z1)

and (x2, z2) in a vertical cross-section of a porous medium. Although arrows

Figure 3.48. Diagrams showing heterogeneity and anisotropy (Freeze and Cherry, 1979).

Page 307: +Peters Ekwere j. - Petrophysics

3-107

and arrow lengths have been used to indicate the directions and magnitudes

of the principal values of the permeability anisotropy, they should not be

misinterpreted as vectors. Permeability is a tensor, not a vector. The special

form of the permeability tensor shown in the figure is

( )0

,0x

z

kk x z

k⎡ ⎤

= ⎢ ⎥⎣ ⎦

(3.215)

and not

( ), x

z

kk x z

k⎡ ⎤

= ⎢ ⎥⎣ ⎦

(3.216)

3.14.2 Darcy's Law for Homogeneous and Anisotropic Medium

For a homogeneous and anisotropic porous medium, Darcy’s Law is

given by

.kvμ

= − ∇Φ (3.217)

Eq.(3.217) can be expanded in 3D Cartesian coordinates as

1x xx xy xz

y yx yy yz

z zx zy zz

xv k k kv k k k

yv k k k

z

μ

⎡ ⎤∂Φ⎢ ⎥∂⎡ ⎤ ⎢ ⎥⎡ ⎤∂Φ⎢ ⎥ ⎢ ⎥⎢ ⎥ = − ⎢ ⎥ ⎢ ⎥⎢ ⎥ ∂⎢ ⎥ ⎢ ⎥⎢ ⎥⎣ ⎦ ⎣ ⎦ ∂Φ⎢ ⎥

⎢ ⎥∂⎣ ⎦

(3.218)

The permeability tensor given by

( ), ,xx xy xz

yx yy yz

zx zy zz

k k kk x y z k k k

k k k

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(3.219)

Page 308: +Peters Ekwere j. - Petrophysics

3-108

is a second rank tensor with nine elements. The tensor is symmetric and as

such kxy = kyx, kxx = kzx and kyz = kzy. The corresponding equations for 2D flow

are

1x xx xy

y yx yy

v k k xv k k

∂Φ⎡ ⎤⎢ ⎥⎡ ⎤ ⎡ ⎤ ∂⎢ ⎥= −⎢ ⎥ ⎢ ⎥ ∂Φ⎢ ⎥⎣ ⎦ ⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.220)

and

( ), xx xy

yx yy

k kk x y

k k⎡ ⎤

= ⎢ ⎥⎣ ⎦

(3.221)

Darcy's law can also be written in terms of hydraulic conductivity and

hydraulic head as

.v K h= − ∇ (3.222)

where K is the hydraulic conductivity tensor. All the properties of the

permeability tensor are applicable to the hydraulic conductivity tensor.

An important property of a symmetric tensor is that by a suitable

change of axes, it is possible to transform the full tensor into another tensor

consisting of only diagonal terms with the off-diagonal terms being zero. This

is a considerable simplification. The axes that permit this transformation of

the tensor to diagonal form are called the principal axes of the permeability

anisotropy of the porous medium. These principal axes contain the maximum

and minimum values of the directional permeabilities which are the diagonal

terms of the transformed tensor. When viewed in the coordinates of the

principal axes of anisotropy, the permeability (hydraulic conductivity) tensor

in 3D becomes

Page 309: +Peters Ekwere j. - Petrophysics

3-109

( )0 0

, 0 00 0

u

v

z

kk u v k

k

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(3.223)

where u, v, w are the coordinates in the directions of the principal axes of the

anisotropy. When viewed in the coordinates of the principal axes of the

anisotropy, Darcy's law for homogeneous and anisotropic media in 3D

simplifies to

1

u

u

v v

w

w

kuv

v kv

vk

w

μ

∂Φ⎡ ⎤⎢ ⎥∂⎡ ⎤ ⎢ ⎥

∂Φ⎢ ⎥ ⎢ ⎥= −⎢ ⎥ ⎢ ⎥∂⎢ ⎥ ⎢ ⎥⎣ ⎦ ∂Φ⎢ ⎥

⎢ ⎥∂⎣ ⎦

(3.224)

In this special coordinate system, the Darcy velocity vector and the velocity

potential gradient vector are collinear and the equation can be written as

1u v w u v wv i v j v w k i k j k w

u v uμ∂Φ ∂Φ ∂Φ⎛ ⎞+ + = − + +⎜ ⎟∂ ∂ ∂⎝ ⎠

(3.225)

where i, j and w are unit vectors. The corresponding equations for 2D flow are

( )0

,0u

v

kk u v

k⎡ ⎤

= ⎢ ⎥⎣ ⎦

(3.226)

1 ux

yv

kv uv

kv

μ

∂Φ⎡ ⎤⎢ ⎥⎡ ⎤ ∂= − ⎢ ⎥⎢ ⎥ ∂Φ⎢ ⎥⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.227)

and

1u v u vv i v j k i k j

u vμ∂Φ ∂Φ⎛ ⎞+ = − +⎜ ⎟∂ ∂⎝ ⎠

(3.228)

Page 310: +Peters Ekwere j. - Petrophysics

3-110

Example 3.5

Steady state flow of a single phase liquid occurs in a horizontal, 2D,

homogeneous and anisotropic reservoir, which has a permeability tensor

given by

k(x,y)= 200 –50–50 750

md

The pressure in the reservoir has been measured at three observations wells

whose coordinates are given in Table 3.13. The liquid viscosity is 1 cp. Do the

negative entries in the permeability tensor bother you?

Table 3.13: Data from Observation Wells

Observation

Well

x – Coordinate

(ft)

y – Coordinate

(ft)

Pressure

(psia)

1 0 0 2000

2 0 800 1800

3 800 0 1500

1. Calculate the magnitude (in ft/day) and the direction of the Darcy

velocity for the flow. Give the direction in terms of the angle the Darcy

velocity vector makes with the positive x- axis.

2. Sketch two vectors with one pointing in the flow direction and the other

pointing in the direction of the pressure gradient. What is the angle

between the two vectors in degrees?

3. Sketch the flow field and the pressure gradient field.

Solution to Example 3.5

Figure 3.49 shows the positions of the wells and their bottomhole pressures.

Page 311: +Peters Ekwere j. - Petrophysics

3-111

Figure 3.49. Positions of the observation wells and their bottomhole pressures.

We apply Eq.(3.217) in oilfield units to obtain

( )( )0.001127 5.615 .kv Pμ

= − ∇ (3.229)

For the 2D flow, Eq.(3.229) can be expanded to give

( )( )0.001127 5.615x xx xy

y yx yy

Pv k k x

Pv k ky

μ

∂⎡ ⎤⎢ ⎥⎡ ⎤ ⎡ ⎤ ∂⎢ ⎥= −⎢ ⎥ ⎢ ⎥ ∂⎢ ⎥⎣ ⎦ ⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.230)

From the given data,

1500 2000 5800 8

Px

∂ − ⎛ ⎞= = −⎜ ⎟∂ ⎝ ⎠ psi/ft

1800 2000 1800 4

Py

∂ − ⎛ ⎞= = −⎜ ⎟∂ ⎝ ⎠ psi/ft

Page 312: +Peters Ekwere j. - Petrophysics

3-112

5 1 0.625 0.2508 4

P i j i j∇ = − − = − −

( ) ( )2 20.625 0.250 0.6731P∇ = − + − = psi/ft

Let the pressure gradient make an angle θ with the negative x-axis. Then

( ) ( ) ( ) ( )( )( )2 20.625 0.250 . 0.625 0.250 1 cosi j i θ− − − = − + −

2 2

0.625cos 0.92850.625 0.250

θ = =+

21.80θ =

The components of the Darcy velocity are given by

( )( )5

200 500.001127 5.615 850 750 11

4

x

y

vv

⎡ ⎤−⎢ ⎥−⎡ ⎤ ⎡ ⎤= − ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎢ ⎥⎣ ⎦ −⎢ ⎥⎣ ⎦

( )( ) ( ) ( )5 10.001127 5.615 200 50 0.71208 4xv ⎡ ⎤⎛ ⎞ ⎛ ⎞= − =⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦

ft/D

( )( ) ( ) ( )5 10.001127 5.615 50 750 0.98888 4yv ⎡ ⎤⎛ ⎞ ⎛ ⎞= − + =⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦

ft/D

The Darcy velocity vector is

0.7120 0.9888v i j= +

( ) ( )2 20.7120 0.9888 1.2185v = + = ft/D

Let the Darcy velocity make an angle α with the positive x-axis.

0.9888tan 1.38880.7120

α = =

Page 313: +Peters Ekwere j. - Petrophysics

3-113

54.24α =

Figure 3.50 shows the magnitudes and the directions of the darcy velocity and

the pressure gradient vectors. The figure shows that the Darcy velocity and

pressure gradient vectors are not collinear for flow in the anisotropic porous

medium.

Figure 3.50. Directions and magnitudes of Darcy velocity and pressure gradient vectors for flow in an anisotropic reservoir.

The angle between the Darcy velocity vector and the pressure gradient vector

is 147.56º. This angle can also be calculated directly from vector calculus as

. . cosv P v P β∇ = ∇

Page 314: +Peters Ekwere j. - Petrophysics

3-114

where β is the angle between the two vectors. Thus,

( ) ( )( )( )

0.7120 0.9888 . 0.625 0.25 0.6922cos 0.84401.2185 0.6731 0.8202

i j i jβ

+ − −= = − = −

147.56β =

Figure 3.51 shows sketches of the flow field and the pressure gradient field.

Figure 3.51. Flow and pressure gradient fields for Example 3.5.

3.14.3 Transformation of Permeability Tensor from one Coordinate System to Another

The objective is to derive the equations for transforming the

permeability tensor (hydraulic conductivity tensor) for an anisotropic reservoir

from one coordinate system to another coordinate system that makes and

angle θ with the first coordinate system, where θ is considered positive for

Page 315: +Peters Ekwere j. - Petrophysics

3-115

anticlockwise rotation. This can be achieved by first establishing the

geometric relationship between the two coordinate systems and then

transforming Darcy’s Law from the first coordinate system to the second

coordinate system. Such a transformation can be used to (1) determine the

principal values of the permeability anisotropy, (2) determine the principal

axes of the permeability anisotropy and (3) examine the flow field in any

coordinate system.

Figure 3.52 shows the relationship between the two coordinate systems.

It should be emphasized that the (u,v) coordinates shown in Figure 3.52 are

not the principal axes of the permeability anisotropy. The two coordinate

systems are related by the following equations:

cos sinx u vθ θ= − (3.231)

sin cosy u vθ θ= + (3.232)

Eqs.(3.231) and (3.232) can be written in matrix form as

cos sinsin cos

x uy v

θ θθ θ

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥

⎣ ⎦ ⎣ ⎦ ⎣ ⎦ (3.233)

Using the inverse of the rotation matrix of Eq.(3.233), the relationship

between the two coordinate systems also can be written as

cos sinsin cos

u xv y

θ θθ θ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(3.234)

Darcy’s law in the xy coordinate system is given by

Page 316: +Peters Ekwere j. - Petrophysics

3-116

Figure 3.52. Relationship between the two coordinate systems.

1x xx xy

y yx yy

v k k xv k k

∂Φ⎡ ⎤⎢ ⎥⎡ ⎤ ⎡ ⎤ ∂⎢ ⎥= −⎢ ⎥ ⎢ ⎥ ∂Φ⎢ ⎥⎣ ⎦ ⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.220)

The Darcy velocity in the uv coordinate system is related to the Darcy velocity

in the xy coordinate system by Eq.(3.234) as

Page 317: +Peters Ekwere j. - Petrophysics

3-117

cos sinsin cos

xu

yv

vvvv

θ θθ θ

⎡ ⎤⎡ ⎤ ⎡ ⎤= ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎣ ⎦⎣ ⎦ ⎣ ⎦

(3.235)

Substituting Eq.(3.220) into (3.235) gives

cos sin1sin cos

xx xyu

yx yyv

k kv xk kv

y

θ θθ θμ

∂Φ⎡ ⎤⎢ ⎥⎡ ⎤⎡ ⎤ ⎡ ⎤ ∂⎢ ⎥= − ⎢ ⎥⎢ ⎥ ⎢ ⎥ ∂Φ− ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.236)

The flow potential in the xy coordinate system is related to the flow potential

in the uv coordinate system by Eq.(2.233) as

cos sinsin cos

x u

y v

θ θθ θ

∂Φ⎡ ⎤ ∂Φ⎡ ⎤⎢ ⎥ ⎢ ⎥−⎡ ⎤∂ ∂⎢ ⎥ = ⎢ ⎥⎢ ⎥∂Φ ∂Φ⎢ ⎥ ⎣ ⎦ ⎢ ⎥⎢ ⎥ ⎢ ⎥∂ ∂⎣ ⎦⎣ ⎦

(3.237)

Substituting Eq.(3.237) into (3.236) gives

cos sin cos sin1sin cos sin cos

xx xyu

yx yyv

k kv uk kv

v

θ θ θ θθ θ θ θμ

∂Φ⎡ ⎤⎢ ⎥−⎡ ⎤⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ∂= − ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥− ∂Φ⎣ ⎦ ⎣ ⎦⎣ ⎦ ⎢ ⎥⎣ ⎦⎢ ⎥∂⎣ ⎦

(3.238)

Eq.(3.238) is Darcy’s law in the uv coordinate system and is of the form

1u uu uv

v vu vv

v k k uv k k

∂Φ⎡ ⎤⎢ ⎥⎡ ⎤ ⎡ ⎤ ∂= − ⎢ ⎥⎢ ⎥ ⎢ ⎥ ∂Φ⎣ ⎦ ⎣ ⎦ ⎢ ⎥⎢ ⎥∂⎣ ⎦

(3.239)

A comparison of Eqs.(3.238) and (3.239) shows that the permeability tensor in

the uv coordinate system is related to the tensor in the xy coordinate system

as follows:

cos sin cos sinsin cos sin cos

xx xyuu uv

yx yyvu vv

k kk kk kk k

θ θ θ θθ θ θ θ

−⎡ ⎤⎡ ⎤ ⎡ ⎤ ⎡ ⎤= ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦⎣ ⎦ ⎣ ⎦

(3.240)

Page 318: +Peters Ekwere j. - Petrophysics

3-118

Carrying out the matrix multiplications on the right side of Eq.(3.240) and

equating the result to the left side matrix gives

2 2cos sin cos sin cos sinuu xx yx xy yyk k k k kθ θ θ θ θ θ= + + + (3.241)

2 2sin cos sin cos sin cosuv xx yx xy yyk k k k kθ θ θ θ θ θ= − − + + (3.242)

2 2sin cos cos sin sin cosvu xx yx xy yyk k k k kθ θ θ θ θ θ= − + − + (3.243)

2 2sin sin cos sin cos cosvv xx yx xy yyk k k k kθ θ θ θ θ θ= − − + (3.244)

From trigonometry,

2 12cos (1 cos 2 )θ θ= + (3.245)

2 12sin (1 cos 2 )θ θ= − (3.246)

12sin cos sin 2θ θ θ= (3.247)

Substituting Eqs.(3.245) to (3.247) into Eq.(3.241) gives

1 1 1 12 2 2 2(1 cos 2 ) sin 2 sin 2 (1 cos 2 )uu xx yx xy yyk k k k kθ θ θ θ= + + + + − (3.248)

Since kxy = kyx, Eq.(3.248) can be simplified to

cos 2 sin 22 2

xx yy xx yyuu xy

k k k kk kθ θ

+ −= + + (3.249)

A comparison of Eqs.(3.242) and (3.243) shows that kuv = kvu. Substituting

Eqs. Eqs.(3.245) to (3.247) into Eq.(3.242) gives

1 1 1 12 2 2 2sin 2 (1 cos 2 ) (1 cos 2 ) sin 2uv xx yx xy yyk k k k kθ θ θ θ= − − − + + + (3.250)

….Since kxy = kyx and kuv = kvu, Eq.(3.250) can be simplified to

Page 319: +Peters Ekwere j. - Petrophysics

3-119

sin 2 cos 22

xx yyuv vu xy

k kk k kθ θ

−⎛ ⎞= = − +⎜ ⎟

⎝ ⎠ (3.251)

Substituting Eqs.(3.245) to (3.247) into Eq.(3.244) gives

1 1 1 12 2 2 2(1 sin 2 ) sin 2 sin 2 (1 cos 2 )vv xx yx xy yyk k k k kθ θ θ θ= − − − + + (3.252)

which can be simplified to

cos 2 sin 22 2

xx yy xx yyvv xy

k k k kk kθ θ

+ −⎛ ⎞= − −⎜ ⎟

⎝ ⎠ (3.253)

One of the principal axes of the permeability anisotropy is obtained from

0uv vuk k= = (3.254)

Substituting Eq.(3.254) into Eq.(3.251) gives

sin 2 cos 2 02

xx yyxy

k kkθ θ

−⎛ ⎞− + =⎜ ⎟

⎝ ⎠ (3.255)

Eq.(3.255) yields

2sin 2tan 2

cos 2xy

xx yy

kk k

θθθ

= =−

(3.256)

or

112

2tan xy

xx yy

kk k

θ −⎛ ⎞

= ⎜ ⎟⎜ ⎟−⎝ ⎠ (3.257)

The other principal axis of the permeability anisotropy is given by (θ +90)

degrees.

Example 3.6

For the flow field of Example 3.5,

1. Calculate the principal axes of the permeability anisotropy.

Page 320: +Peters Ekwere j. - Petrophysics

3-120

2. Calculate the principal values of the permeability anisotropy. Show the

transformed permeability tensor for the principal coordinate system.

3. Calculate the components of the Darcy velocity vector in the principal

directions of the permeability anisotropy.

Solution to Example 3.6

One of the principal axes of the permeability anisotropy is given by Eq.(3.257)

as

( ) ( ) ( )1 1 12 50tan 2 tan tan 0.1818 10.3048

200 750θ− − −−⎛ ⎞

= = =⎜ ⎟−⎝ ⎠

10.3048 5.152

θ = =

The principal values of the permeability anisotropy are given by Eqs.(3.249)

and (3.253) as

( ) ( )200 750 200 750 cos 10.3048 50sin 10.30482 2uk + −

= + −

475 270.5642 8.9443 195.49uk md= − − =

( ) ( )200 750 200 750 cos 10.3048 50sin 10.30482 2vk + −⎛ ⎞= − +⎜ ⎟

⎝ ⎠

475 270.5642 8.9443 754.51vk md= + + =

The transformed permeability tensor is given by

( )0 195.5 0

,0 0 754.5u

v

kk u v md

k⎡ ⎤ ⎡ ⎤

= =⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

, where ku and kv are the principal values of

the permeability anisotropy. Figure 3.53 shows that the coordinate system

along the principal axes of the permeability anisotropy (u,v) makes an angle of

+5.15º with the x,y coordinate system.

Page 321: +Peters Ekwere j. - Petrophysics

3-121

Figure 3.53. Relationship between the principal axes of the permeability anisotropy and the (x,y) coordinates.

Also shown in Figure 3.53 is the Darcy velocity vector in relation to the x,y

and the u,v coordinate systems, where the u,v coordinates are the principal

axes of the permeability anisotropy.

The components of the Darcy velocity in the u,v coordinates can be

calculated as follows.

1.2185cos(54.24 5.15 ) 1.2185cos 49.09 0.7980uv = − = = ft/D

( )1.2185sin 54.24 5.15 1.2185sin 49.09 0.9209vv = − = = ft/D

The components of the Darcy velocity in the u,v coordinates also can be

calculated from Eq.(3.234) as

Page 322: +Peters Ekwere j. - Petrophysics

3-122

0.7120 0.7980cos5.15 sin 5.15 cos5.15 sin 5.150.7120 0.9888

0.9888 0.9209sin 5.15 cos5.15 sin 5.15 cos5.15u

v

vv

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤ ⎡ ⎤= = + =⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎣ ⎦ ⎣ ⎦⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦

0.7980uv = ft/D

0.9209vv = ft/D

These are the results obtained previously. It should be noted that because the

flow is invariant with a rotation of the axes, the magnitude of the Darcy

velocity (1.2185 ft/D) remains the same in the new coordinate system. The

direction of the Darcy velocity also remains the same. Only the components of

the Darcy velocity change in the new coordinate system.

3.14.4 Alternative Calculation of the Principal Values and the Principal Axes of the Permeability Anisotropy

From linear algebra, the principal values of the permeability tensor are

given by the eigenvalues of the permeability tensor whereas the principal axes

of the permeability anisotropy are given by the eigenvectors of the

permeability tensor. This can be demonstrated by reworking Example 3.6 as

an eigenvalue problem.

Example 3.7

Rework Example 3.6 by calculating the eigenvalues and eigenvectors of the

permeability tensor.

Solution to Example 3.7

Let the eigenvalues of the permeability tensor be λ. The characteristic

equation is given by

200 50det 0

50 750λ

λ− −⎡ ⎤

=⎢ ⎥− −⎣ ⎦

( )( ) ( )( )200 750 50 50 0λ λ− − − − − =

Page 323: +Peters Ekwere j. - Petrophysics

3-123

2 950 147500 0λ λ− + =

( )( )( )2950 950 4 1 147500 950 559.01702 2

λ± − ±

= =

1 195.5mdλ =

2 754.5mdλ =

These principal values are the same as obtained in Example 3.6. For

1 195.5λ = , the eigenvector is obtained by solving the following simultaneous

equations:

200 195.5 50 050 750 195.5 0

xy

− −⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎣ ⎦ ⎣ ⎦ ⎣ ⎦

4.5085 50 050 554.5085 0

xy

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦ ⎣ ⎦

4.5085 50 0x y− =

50 554.5085 0x y− + =

11.09021

xy

y⎡ ⎤ ⎡ ⎤

=⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

The eigenvector is

11.09021

u ⎡ ⎤= ⎢ ⎥

⎣ ⎦

This eigenvector makes an angle θ with the positive x-axis given by

1tan 0.090211.0902

θ = =

5.15θ =

Page 324: +Peters Ekwere j. - Petrophysics

3-124

For 2 754.5λ = , the eigenvector is obtained by solving the following

simultaneous equations:

200 754.5 50 050 750 754.5 0

xy

− −⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎣ ⎦ ⎣ ⎦ ⎣ ⎦

554.5 50 050 4.5 0

xy

− −⎡ ⎤ ⎡ ⎤ ⎡ ⎤=⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦ ⎣ ⎦

554.5 50 0x y− − =

50 4.5 0x y− − =

0.09021

xy

y−⎡ ⎤ ⎡ ⎤

=⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

The eigenvector is

0.09021

v−⎡ ⎤

= ⎢ ⎥⎣ ⎦

We can easily show that the two eigenvectors are orthogonal.

( )( ). 11.0902 0.0902 1 0u v = − − =

The eigenvectors point in the directions of the principal axes of the

permeability anisotropy. These are the same results obtained in Example 3.6.

3.14.5 Directional Permeability

For an isotropic medium, the Darcy velocity vector and the velocity

potential gradient vector are collinear as shown in Figure 3.54. Because the

two vectors lie along the same line with an angle of 180º between them, we

ordinarily do not specify whether the permeability is in the direction of flow or

in the direction of the velocity potential gradient. For this case, the

"directional" permeability in the direction of flow (kdf) is equal to the

Page 325: +Peters Ekwere j. - Petrophysics

3-125

"directional" permeability in the direction of the velocity potential gradient

(kdp) and is calculated in Darcy units as

df dp

vk k

μ= =

∇Φ (3.258)

Figure 3.54. Collinear relationship between the direction of flow and the direction of the velocity potential gradient for an isotropic medium.

For an anisotropic medium, in general, the Darcy velocity vector and

the velocity potential gradient vector are not collinear as shown in Figure

3.55. The concept of directional permeability seeks to extend the definition of

permeability of an isotropic medium expressed in Eq.(3.258) to an anisotropic

medium. Accordingly, the directional permeability in the direction of flow is

defined in Darcy units by the equation

cosdf

vk

μα

=∇Φ

(3.259)

where cosα∇Φ is the component of the velocity potential gradient vector that

is collinear with the Darcy velocity. Similarly, the directional permeability in

the direction of the velocity potential gradient is defined in Darcy units by the

equation

Page 326: +Peters Ekwere j. - Petrophysics

3-126

cos

dp

vk

μ α=

∇Φ (3.260)

where cosv α is the component of the Darcy velocity vector that is collinear

with the velocity potential gradient vector.

Figure 3.55. Relationship between the direction of flow and the direction of the velocity potential gradient in an anisotropic porous medium.

Example 3.8

Calculate the directional permeabilities for the flow field of Example 3.5.

Solution to Example 3.8

The directional permeability in the direction of flow is calculated with

Eq.(3.259) in oilfield units as

( )( )( ) ( )

(0.001127(5.615) cos

1 1.2185 338.96 md

(0.001127(5.615) 0.6731 cos 180 147.56

df

vk

μα

=∇Φ

= =−

Page 327: +Peters Ekwere j. - Petrophysics

3-127

The directional permeability in the direction of the velocity potential gradient

is calculated with Eq.(3.260) in oilfield units as

( )( ) ( )( )

cos(0.001127(5.615)

1 1.2185 cos 180 147.56 241.46 md

(0.001127(5.615) 0.6731

dp

vk

μ α=

∇Φ

−= =

We now derive the relationship between the directional permeabilities

and the principal values of the permeability anisotropy. Let kx and ky be the

principal values of the permeability anisotropy where the x,y coordinates are

the principal axes of the anisotropy in a 2D reservoir. Let kdf be the directional

permeability in the direction of flow (in the direction of the Darcy velocity

vector) that makes an angle θ with the positive x-axis as shown in Figure

3.56. The permeability tensor for this medium in the x,y coordinate system is

( )0

,0x

y

kk x y

k⎡ ⎤

= ⎢ ⎥⎣ ⎦

(3.261)

Figure 3.56. Darcy velocity vector in the flow direction that makes an angle θ with the positive x axis.

Page 328: +Peters Ekwere j. - Petrophysics

3-128

Darcy’s law can be written in the x, y and s directions as

xx

kvxμ

∂Φ= −

∂ (3.262)

yy

kv

yμ∂Φ

= −∂

(3.263)

dfs

kv

sμ∂Φ

= −∂

(3.264)

where s is the flow direction. Figure 3.56 shows the relationships between the

Darcy velocity in the flow direction vs and its x and y components vx and vy.

From Figure 3.56, we find

cosx sv v θ= (3.265)

siny sv v θ= (3.266)

Eqs. (3.362) and (3.263) give the potential gradients in the x and y directions

as

cossx

vx k

μ θ∂Φ= −

∂ (3.267)

sinsy

vy k

μ θ∂Φ= −

∂ (3.268)

Because Φ = Φ(x,y), from Calculus,

dx dys x ds x ds

∂Φ ∂Φ ∂Φ= +

∂ ∂ ∂ (3.269)

From Figure 3.56,

cosdxds

θ= (3.270)

Page 329: +Peters Ekwere j. - Petrophysics

3-129

sindyds

θ= (3.271)

Substituting Eqs.(3.267) to (3.271) into Eq.(3.264) and rearranging gives the

directional permeability in the direction of flow as

2 21 cos sin

df x yk k kθ θ

= + (3.272)

Given the principal values of the permeability anisotropy (kx, ky) and the

principal coordinate system (x,y), Eq.(3.272) can be used to calculate the

directional permeability in the flow direction that makes an angle θ with the

positive x-axis.

Eq.(3.272) can be put in rectangular coordinates by choosing a vector r

in the s direction (flow direction) and noting that

cosx r θ= (3.273)

siny r θ= (3.274)

2

22cos x

rθ = (3.275)

2

22sin y

rθ = (3.276)

Substituting Eqs.(3.275) and (3.276) into (3.272) gives

2 2 2

df x y

r x yk k k

= + (3.277)

If we make the magnitude of the vector dfr k= , then Eq.(3.277) becomes

2 2

1x y

x yk k

= + (3.278)

Page 330: +Peters Ekwere j. - Petrophysics

3-130

or

( ) ( )

2 2

2 21x y

x y

k k= + (3.279)

Eq.(3.279) is the cannonical equation of an ellipse with major semiaxes

xk and yk . Thus, by constructing the permeability ellipse as shown in

Figure 3.57, the directional permeability can easily be determined graphically

for any given flow direction.

Figure 3.57. Permeability ellipse for calculating the directional permeability in the direction of flow.

For flow 3D, the directional permeability in the direction of flow will be given

by the permeability ellipsoid

( ) ( ) ( )

2 2 2

2 2 21x y z

x y z

k k k= + + (3.280)

with semiaxes xk , yk and zk .

Page 331: +Peters Ekwere j. - Petrophysics

3-131

Let the velocity potential gradient vector make an angle α with the

positive x-axis. It can be shown that the directional permeability in the

direction of the gradient vector is given by

2 21 cos sin

11 1xdp ykk k

α α= +

⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠⎝ ⎠ ⎝ ⎠

(3.281)

The permeability ellipse in 2D is

2 2

11 1x y

x y

k k

= +⎛ ⎞ ⎛ ⎞⎜ ⎟ ⎜ ⎟⎜ ⎟⎝ ⎠ ⎝ ⎠

(3.282)

or

2 2

2 211 1

x y

x y

k k

= +⎛ ⎞ ⎛ ⎞⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠

(3.283)

with semiaxes ( )1/ xk and ( )1/ yk . The permeability ellipsoid in 3D is

2 2 2

2 2 2111 1

zx y

x y y

kk k

= + +⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎜ ⎟⎜ ⎟ ⎜ ⎟ ⎝ ⎠⎝ ⎠ ⎝ ⎠

(3.284)

with semiaxes ( )1/ xk , ( )1/ xk and ( )1/ zk .

In an anisotropic medium, the directional permeability in the direction

of the velocity potential gradient is distinct from the directional permeability

in the direction of flow. Thus, there are two distinct definitions of directional

permeability for an anisotropic medium. It should be noted that principal

values of the permeability anisotropy (kx, ky and kz) are the true permeabilities

in the principal directions (x, y and z). They are also the directional

Page 332: +Peters Ekwere j. - Petrophysics

3-132

permeabilities in the direction of flow and the direction of the potential

gradient because in the principal coordinate system, the Darcy velocity and

the velocity potential gradient are collinear.

Figures 3.58 shows a polar plot of permeability data from Scheidegger

(1954) for permeabilities measured around a full-sized core. The measured

permeability data appear to fit a complicated curve. However, when the same

data are plotted as k and 1k

as shown in Figures 3.59 and 3.60, they fit on

ellipses. This observation verifies the tensorial nature of permeability. These

data can be used to identify the principal axes and the principal values of the

permeability anisotropy.

Figure 3.58. Permeability data measured around a cylindrical core (Data from Scheidegger, 1954).

Page 333: +Peters Ekwere j. - Petrophysics

3-133

Figure 3.59. Permeability ellipse for permeability in the direction of flow for the data of Figure 3.58.

Page 334: +Peters Ekwere j. - Petrophysics

3-134

Figure 3.60. Permeability ellipse for permeability in the direction of the

velocity potential gradient for the data of Figure 3.58.

Page 335: +Peters Ekwere j. - Petrophysics

3-135

Example 3.9

Calculate and plot the permeability ellipses for the flow field of Example 3.5.

Solution to Example 3.9

The permeability ellipses are given by Eqs.(3.279) for the flow direction and

(3.283) for the potential gradient direction, where x and y are the principal

axes of the permeability anisotropy. These equations can be adapted for our

purpose as follows:

( ) ( ) ( ) ( )2 2 2 2

2 2 2 21195.5 754.5u v

u v u v

k k= + = +

and

2 2 2 2

2 2 2 211 11 1

195.5 754.5u v

u v u v

k k

= + = +⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞

⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟ ⎝ ⎠ ⎝ ⎠⎝ ⎠ ⎝ ⎠

These ellipses are plotted in Figures 3.61 and 3.62.

Page 336: +Peters Ekwere j. - Petrophysics

3-136

Figure 3.61. Permeability ellipse for permeability in the direction of flow for Example 3.9.

Page 337: +Peters Ekwere j. - Petrophysics

3-137

Figure 3.62. Permeability ellipse for permeability in the direction of the

velocity potential gradient for Example 3.9.

3.14.6 Measurement of Transverse Permeability of a Cylindrical Core

If the transverse permeabilities around a full-sized core could be

measured, then the permeability anisotropy of the core could be determined.

Figure 3.63 shows a schematic diagram of an apparatus that can be used to

make the measurements. After a permeability measurement, the sleeve is

rotated to a new angle and the measurement is repeated. Using conformal

mapping, Collins (1961) has derived the following equation for calculating the

transverse permeability from a gas flow experiment (in Darcy units):

( )1 1

2 1

Pqk GbPL PP

μ α=⎛ ⎞+ Δ⎜ ⎟⎝ ⎠

(3.285)

Page 338: +Peters Ekwere j. - Petrophysics

3-138

Figure 3.63. Apparatus for measuring transverse permeability of a full-sized cylindrical core.

where P1 is the inlet pressure, q1 is the gas volumetric flow rate measured at

P1, P is the mean pressure 1 2

2P P+⎛ ⎞

⎜ ⎟⎝ ⎠

, PΔ is the pressure drop between the

inlet and the outlet ( )1 2P P− , b is the Klinkenberg parameter, L is the height of

the core, G(α) is a geometric correction factor to account for the complex flow

Page 339: +Peters Ekwere j. - Petrophysics

3-139

geometry of the gas from inlet to outlet and α is equal to 2θ⎛ ⎞

⎜ ⎟⎝ ⎠

. Figure 3.64

shows the plot of G(α) versus α.

Figure 3.64. Geometric factor for transverse permeability calculation (from Collins, 1961).

By making such measurements around the core, a polar plot of the

permeability can be made as shown in Figure 3.65. Such a plot can be used

to determine the principal values and the principal directions of the

permeability anisotropy. The permeability tensor of the medium can then be

constructed from this information for any coordinate system.

Page 340: +Peters Ekwere j. - Petrophysics

3-140

Figure 3.65. Polar plot of transverse permeabilities around a core (from Scheidegger, 1954).

3.15 EXAMPLE APPLICATIONS OF PERMEABILITY

Permeability is used in numerical reservoir simulations to predict

reservoir performance. Permeability (absolute and relative permeability) must

be assigned to every one of the approximately one million grid blocks in a full-

field simulation, often with a full tensor representation. The purpose of this

Page 341: +Peters Ekwere j. - Petrophysics

3-141

section is to present simple examples of how permeability could be used to

predict reservoir performance. We examine the productivity indices of a

horizontal well and a hydraulically fractured well with infinite fracture

conductivity. The analysis shows that a horizontal well and a fractured well

are similar in the manner in which they affect reservoir performance.

3.15.1 Productivity of a Horizontal Well Introduction

Horizontal wells are increasingly being drilled for petroleum recovery.

One of the major advantages of a horizontal well over a vertical well is the

larger contact area between a horizontal well and the reservoir compared to a

vertical well in the same reservoir. This larger areal contact can significantly

enhance the productivity of a horizontal well compared to a vertical well

draining the same reservoir volume. However, the productivity of a horizontal

well can be affected significantly by the permeability anisotropy of the

reservoir.

The objectives of this section are: (1) to provide a simple formula for

estimating the productivity index of a horizontal well for single phase flow in

terms of the reservoir permeability and other relevant factors, (2) to compare

the productivities of a horizontal well and a vertical well in the same reservoir

to determine if a horizontal well is superior to a vertical well under the

particular circumstance, and (3) to examine how the permeability anisotropy

of the reservoir affects the productivity of a horizontal well.

Homogeneous and Isotropic Reservoirs

The steady state productivity index (PI) for a horizontal well in a

homogeneous and isotropic reservoir is given by Giger et al. (1984) as

Page 342: +Peters Ekwere j. - Petrophysics

3-142

2

2

1 12

ln ln2

2

H D

e

w

e

kLBPI F

LrL h

h rLr

πμ

π

⎛ ⎞⎜ ⎟⎝ ⎠=

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟⎢ ⎥ ⎛ ⎞⎝ ⎠ +⎢ ⎥ ⎜ ⎟⎛ ⎞ ⎝ ⎠⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥⎣ ⎦

(3.286)

where

PIH = productivity index of a horizontal well, surface rate/unit pressure drop

FD = units conversion constant

k = effective permeability to oil

L = total length of the horizontal well

μ = oil viscosity

B = oil formation volume factor, reservoir volume/surface volume

h = formation thickness

rw = wellbore radius

re = well drainage radius

Figure 3.66 shows the flow geometry. The horizontal well of length L is located

half-way between the top and the bottom of the reservoir of thickness h. L is

assumed to be small relative to 2re. The units conversion constant, FD, is

given in Table 3.14.

Page 343: +Peters Ekwere j. - Petrophysics

3-143

Figure 3.66: Flow geometry for steady state productivity index equation.

Table 3.14: Units and Dimensions

Dimension

Darcy Units

SI Units

Oilfield Units

Field Metric Units

Distance L cm m ft m Area L2 cm2 m2 ft2 m2

Pressure ML-1T-2 atm Pa psi MPa Permeability L2 darcy m2 millidarcy

(md) μm2

Fluid Viscosity

ML-1T-1 cp Pa.s cp mPa.s

Liquid Flow Rate

L3T-1 cm3/s m3/s B/D m3/D

FD 1 1 0.001127 86.4

The steady state productivity index of a vertical well can be obtained by

integration of Darcy’s Law as

Page 344: +Peters Ekwere j. - Petrophysics

3-144

2

lnV D

e

w

khBPI F

rr

πμ

⎛ ⎞⎜ ⎟⎝ ⎠=⎛ ⎞⎜ ⎟⎝ ⎠

(3.287)

Therefore, the ratio of the productivity index of a horizontal well to a vertical

well draining the same reservoir volume is obtained from Eqs.(3.286) and

(3.287) as

2

ln

1 12

ln ln2

2

e

wHD

V

e

w

e

rrPI F

PI LrL h

h rLr

π

⎛ ⎞⎜ ⎟⎝ ⎠=

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟⎢ ⎥ ⎛ ⎞⎝ ⎠ +⎢ ⎥ ⎜ ⎟⎛ ⎞ ⎝ ⎠⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥⎣ ⎦

(3.288)

Note that because the right side of Eq.(3.288) involves ratios of lengths,

Eq.(3.288) is valid in any system of units.

Example 3.10

Given the following data for an isotropic reservoir, calculate (a) the

productivity index for a horizontal well, (b) the productivity index for a vertical

well, and (c) the ratio of the productivity indices of the horizontal and the

vertical wells.

Length of horizontal well (L) = 500 m

Formation thickness (h) = 40 m

Drainage radius (re) = 1000 m

Wellbore radius (rw) = 0.12 m

Horizontal permeability (kH) = 400 md

Vertical permeability (kV) = 400 md

Page 345: +Peters Ekwere j. - Petrophysics

3-145

Oil viscosity (μo) = 5 cp

Oil formation volume factor (B)= 1.2 RB/STB

Solution to Example 3.10

In oilfield units, FD = 0.001127. Substituting the numerical values into

Eq.(3.286) gives the productivity index for the horizontal well as

2

400 500 3.280825 1.20.001127 26.02 STB/D/psi

5001 12 1000500 40ln ln

50040 2 0.122 1000

H

x xxPI

xx

x

π

π

⎛ ⎞⎜ ⎟⎝ ⎠= =

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟ ⎛ ⎞⎝ ⎠⎢ ⎥ + ⎜ ⎟⎢ ⎥⎛ ⎞ ⎝ ⎠⎢ ⎥⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

Substituting the numerical values into Eq.(3.287) gives the productivity index

for the vertical well as

400 40 3.280825 1.20.001127 6.86 STB/D/psi1000ln0.12

V

x xxPI

π ⎛ ⎞⎜ ⎟⎝ ⎠= =

⎛ ⎞⎜ ⎟⎝ ⎠

The ratio of the horizontal to vertical well productivity indices is

26.02 3.86.86

H

V

PIPI

= =

Clearly, in this isotropic reservoir, the productivity of the horizontal well is

much higher than that of the vertical well.

Homogeneous and Anisotropic Reservoirs

Many petroleum reservoirs are anisotropic and have different

permeabilities in the different directions. For example, in a layered reservoir,

the vertical permeability is usually much less than the horizontal

permeability. Also, a reservoir that is intersected by a large number of vertical

Page 346: +Peters Ekwere j. - Petrophysics

3-146

fractures will have a higher vertical permeability than the horizontal

permeability. Permeability anisotropy can have a significant effect on the

productivity of a horizontal well. This effect is examined in this section.

Consider an anisotropic reservoir with horizontal and vertical

permeabilities kH and kV, respectively. These are the principal values of the

permeability anisotropy and x and z are the principal axes of the anisotropy.

The diffusivity equation for steady state single phase flow for the anisotropic

medium is given by

2 2 0H VP Pk kx z

∂ ∂+ =

∂ ∂ (3.289)

By a suitable change of coordinates, Eq.(3.289) can be transformed into a

form that describes the flow in an equivalent isotropic medium having a

permeability *k . The required transformation equations are as follows:

*H Vk k k= (3.290)

*

H

kX xk

= (3.291)

and

*

V

kZ zk

= (3.292)

where X and Z are the transformed spatial coordinates. With these

transformations, Eq.(3.289) becomes

*2 2 0P Pk

X Z∂ ∂⎛ ⎞+ =⎜ ⎟∂ ∂⎝ ⎠

(3.293)

Since *k is nonzero, Eq.(3.293) yields

Page 347: +Peters Ekwere j. - Petrophysics

3-147

2 2 0P PX Z∂ ∂⎛ ⎞+ =⎜ ⎟∂ ∂⎝ ⎠

(3.294)

which is the diffusivity equation for an equivalent isotropic medium with a

permeability of *k . Eqs.(3.290) to (3.292) can be used in conjunction with

Eq.(3.286) to calculate the productivity index of a horizontal well in an

anisotropic reservoir as shown in the following example.

Example 3.11

Given the following data for an anisotropic reservoir, calculate (a) the

productivity index for a horizontal well, (b) the productivity index for a vertical

well, and (c) the ratio of the productivity indices of the horizontal and the

vertical wells.

Length of horizontal well (L) = 500 m

Formation thickness (h) = 40 m

Drainage radius (re) = 1000 m

Wellbore radius (rw) = 0.12 m

Horizontal permeability (kH) = 400 md

Vertical permeability (kV) = 25 md

Oil viscosity (μo) = 5 cp

Oil formation volume factor (B) = 1.2 RB/STB

Solution to Example 3.11

From Eq.(3.290),

* 400 25 100 mdk x= =

From Eq.(3.291),

* 100500 250 m400

L = =

Page 348: +Peters Ekwere j. - Petrophysics

3-148

* 1001000 500 m400er = =

From Eq.(3.292),

* 10040 80 m25

h = =

Substituting these numerical values into Eq.(3.286) gives the productivity

index for the horizontal well in the anisotropic reservoir as

2

100 250 3.280825 1.20.001127 8.71 STB/D/psi

2501 12 500250 80ln ln

25080 2 0.122 500

H

x xxPI

xx

x

π

π

⎛ ⎞⎜ ⎟⎝ ⎠= =

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟ ⎛ ⎞⎝ ⎠⎢ ⎥ + ⎜ ⎟⎢ ⎥⎛ ⎞ ⎝ ⎠⎢ ⎥⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

The productivity index for the vertical well in the anisotropic reservoir is given

by Eq.(3.287) with k = kH. Thus, for the vertical well,

400 40 3.280825 1.20.001127 6.86 STB/D/psi1000ln0.12

V

x xxPI

π ⎛ ⎞⎜ ⎟⎝ ⎠= =

⎛ ⎞⎜ ⎟⎝ ⎠

The ratio of the horizontal to the vertical well productivity indices in the

anisotropic medium is

8.71 1.276.86

H

V

PIPI

= =

Thus, the productivity advantage of a horizontal well is reduced considerably

in an anisotropic reservoir in which the vertical permeability is significantly

lower than the horizontal permeability.

Page 349: +Peters Ekwere j. - Petrophysics

3-149

In a naturally fractured reservoir, if the azimuth of the horizontal is

such that it intersects the vertical fractures, then the productivity of the well

will enhanced as shown in the following example.

Example 3.12

A horizontal well has intersected a large number of vertical fractures as

shown in Figure 3.67 such that the vertical permeability is larger than the

horizontal permeability. Given the following data for this anisotropic reservoir,

calculate (a) the productivity index for a horizontal well, (b) the productivity

index for a vertical well, and (c) the ratio of the productivity indices of the

horizontal and the vertical wells.

Length of horizontal well (L) = 500 m

Formation thickness (h) = 40 m

Drainage radius (re) = 1000 m

Wellbore radius (rw) = 0.12 m

Horizontal permeability (kH) = 400 md

Vertical permeability (kV) = 800 md (Due to the vertical fractures)

Oil viscosity (μo) = 5 cp

Oil formation volume factor (B) = 1.2 RB/STB

Page 350: +Peters Ekwere j. - Petrophysics

3-150

Figure 3.67. A horizontal well that intersects vertical fractures.

Solution to Example 3.12

From Eq.(3.290),

* 400 800 565.69 mdk x= =

From Eq.(3.291),

* 565.69500 594.60 m400

L = =

* 565.691000 1189.21 m400er = =

From Eq.(3.292),

* 565.6940 33.64 m800

h = =

Substituting these numerical values into Eq.(3.286) gives the productivity

index for the horizontal well in the fractured anisotropic reservoir as

Page 351: +Peters Ekwere j. - Petrophysics

3-151

2

565.69 594.60 3.280825 1.20.001127 32.34 STB/D/psi

594.601 12 1189.21594.60 33.64ln ln25033.64 2 0.12

2 1189.21

H

x xxPI

xx

x

π

π

⎛ ⎞⎜ ⎟⎝ ⎠= =

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟ ⎛ ⎞⎝ ⎠⎢ ⎥ + ⎜ ⎟⎢ ⎥⎛ ⎞ ⎝ ⎠⎢ ⎥⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

The productivity index for the vertical well in the anisotropic reservoir is given

by Eq.(3.286) with k = kH. Thus, for the vertical well,

400 40 3.280825 1.20.001127 6.86 STB/D/psi1000ln0.12

V

x xxPI

π ⎛ ⎞⎜ ⎟⎝ ⎠= =

⎛ ⎞⎜ ⎟⎝ ⎠

The ratio of the horizontal to the vertical well productivity indices in the

anisotropic medium with vertical fractures is

32.34 4.716.86

H

V

PIPI

= =

Clearly, the advantage of a horizontal well that intersects vertical fractures in

an anisotropic reservoir compared to a vertical well is apparent from this

example. Note, however, that if the horizontal well was wrongly drilled parallel

to the fractures instead of normal to the fractures as shown in Figure 3.67,

the productivity of the horizontal well will not be enhanced by the fractures.

The horizontal well placement relative to the fracture azimuth is therefore

important to take advantage of the fractures.

The change in coordinates given by Eqs.(3.290) to (3.292) also

transforms the circular wellbore into an elliptical wellbore with the same cross

sectional area as the circular wellbore. This change in the shape of the

wellbore has been neglected in the calculations in Examples 3.11 and 3.12

because its effect is negligibly small.

Page 352: +Peters Ekwere j. - Petrophysics

3-152

There are more elaborate equations for calculating the productivity

index of a horizontal well in the literature. These more complicated equations

and their associated assumptions may be found in the references.

3.14.2 Productivity of a Vertically Fractured Well

The productivity index of a vertically fractured well with an infinite

fracture conductivity in a homogeneous and isotropic reservoir as shown in

Figure 3.68 is given by Karcher et al. (1986) as

2

2

1 12

ln

2

H

F D

e

e

k LBPI F

LrL

h Lr

πμ

⎛ ⎞⎜ ⎟⎝ ⎠=

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥

⎛ ⎞⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥⎣ ⎦

(3.295)

where L is the total length of the fracture. The assumptions underlying the

above equation are:

• Fracture height is equal to the reservoir pay thickness

• The fracture is plane and is centered on the well

• The well drainage radius is large compared to the fracture length

• The fracture has infinite conductivity

The fracture conductivity is defined as

C fF k w= (3.296)

where kf is the fracture permeability and w is the fracture width. A

dimensioless fracture conductivity is defined as

Page 353: +Peters Ekwere j. - Petrophysics

3-153

fCD

H f

k wF

k L= (3.297)

where Lf is the fracture half-length 2L⎛ ⎞

⎜ ⎟⎝ ⎠

. A hydraulic fracture behaves as if it

has an infinite conductivity for

300CDF ≥ (3.298)

Figure 3.68. Vertically fractured well.

Page 354: +Peters Ekwere j. - Petrophysics

3-154

If 12 e

Lr

⎛ ⎞⎜ ⎟⎝ ⎠

, then Eq.(3.295) simplifies to

2

4ln

H

V De

k hBPI FrL

πμ

⎛ ⎞⎜ ⎟⎝ ⎠=⎛ ⎞⎜ ⎟⎝ ⎠

(3.299)

which is the productivity index of a vertical well with an equivalent wellbore

radius given by

4wLr = (3.300)

The ratio of the productivity index of the fracture well to the unfractured well

is given by

2

ln

1 12

ln

2

e

wFD

V

e

e

rrPI F

PI LrL

h Lr

⎛ ⎞⎜ ⎟⎝ ⎠=

⎡ ⎤⎛ ⎞⎢ ⎥+ − ⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥

⎛ ⎞⎢ ⎥⎜ ⎟⎢ ⎥⎝ ⎠⎢ ⎥⎣ ⎦

(3.301)

Eq.(3.301) can be used to demonstrate the increased productivity of a

fractured well compared to the unfractured well.

A comparison of Eqs.(3.295) and (3.286) shows that the productivity

index of a vertical fracture with an infinite conductivity is similar to but

somewhat higher than the productivity index of a horizontal well. The

horizontal well has a positive pseudo skin represented by the term ln2 w

hrπ

⎛ ⎞⎜ ⎟⎝ ⎠

in

Eq.(3.286) compared to the fractured well. The pseudoskin is caused by the

convergence of flow into the horizontal wellbore whereas the flow into the

fracture occurs over the entire fracture face with no convergence of flow. The

Page 355: +Peters Ekwere j. - Petrophysics

3-155

similarity in the productivity index equations for a fracture and a horizontal

well indicates that fracturing a vertical well is an alternative to drilling a

horizontal well.

NOMENCLATURE

A = cross sectional area in the flow direction

Ai = Area of bed i

AT = total area

b = Klinkenberg empirical constant dependent on rock and gas

d = core diameter

B = oil formation volume factor

Bg = gas formation volume factor

cf = compressibility of porous medium

ct = total compressibility

C = wellbore storage coefficient

CD = dimensionless wellbore storage coefficient

D = non-Darcy coefficient

Dp = grain diameter, mean grain diameter

Ei = exponential integral function

f(δ) = probability density function for pore diameter for a bundle of capillary tubes model

F = formation resisitivity factor

FC = fracture conductivity

FCD= dimensionless fracture conductivity

g = gravitational acceleration

G(α) = geometric factor for calculating transverse permeability of a cylindrical porous medium

h = pay thickness, hydraulic head

hi = thickness of bed i

K = hydraulic conductivity

k = absolute permeability of the medium

kdf = directional permeability in the direction of flow

Page 356: +Peters Ekwere j. - Petrophysics

3-156

kdp = directional permeability in the direction of the velocity potential gradient

ki = permeability of bed i

k = permeability tensor, average permeability

kij = individual entries in the permeability tensor

kg = measured gas permeability

kL = permeability to liquid (absolute permeability of medium)

kor = permeability to nonwetting phase at irreducible wetting phase saturation

koτ = Kozeny constant

kx, ky, kz = principal values of permeability anisotropy where xyz is the principal coordinate system

ku, kv, kw = principal values of permeability anisotropy where uvw is the principal coordinate system

L = length

Li = length of bed i

Le = tortuous length of flow path

ln = natural logarithm (log to base e)

log = log to base 10

M = real gas potential or real gas pseudo pressure

P = pressure

PD = dimensionless pressure

'DP = dimensionless welltest derivative function

Pi = initial reservoir pressure

P* = extrapolated pressure from a Horner plot

Psc = reference pressure

Pe = pressure at external radius

Pw = pressure at the wellbore

PwD= dimensionless wellbore pressure

wDP = dimensionless wellbore pressure in Laplace space

'wDP = Calculus derivative of dimensionless wellbore pressure in Laplace

space

Page 357: +Peters Ekwere j. - Petrophysics

3-157

Pwf = flowing bottomhole pressure

Pws = shut-in bottomhole pressure

P1 = inlet pressure

P2 = outlet pressure

P = mean pressure 1 2

2P P+⎛ ⎞

⎜ ⎟⎝ ⎠

dPds

= pressure gradient

PI = productivity index

PIH = productivity index of a horizontal well

PIV = productivity index of a vertical well

q = volumetric flowrate

qsf = sandface rate

qsc = volumetric rate at a reference pressure, Psc

qT = total volumetric flow rate

r = radial distance from the wellbore

rD = dimensionless radius

re = external drainage radius

rH = hydraulic radius

ri = radius of bed i

rinv = radius of investigation of a welltest

ro = resistance of a porous medium fully saturated with water

rw = resistance of water

rw = wellbore radius

Re = Reynolds Number

Ro = resistivity of a porous medium fully saturated with water

Rw = resistivity of water

s = cartesian coordinate in the flow direction.

S = skin factor

S = storativity

S = surface area per unit bulk volume (specific surface area)

Page 358: +Peters Ekwere j. - Petrophysics

3-158

Sp = surface area per unit pore volume (specific surface area)

Ss = surface area per unit grain volume (specific surface area)

Swirr = irreducible wetting phase saturation

S* = mechanical skin facor

t = time

tD = dimensionless time

tp = producing time before shut-in

T = absolute temperature

T = transmissibility

Tsc = reference temperature

v = flux vector, Darcy velocity vector

vx, vy, vz = components of Darcy velocity in the xyz Cartesian coordinate system.

vu, vv, vw = components of Darcy velocity in the uvw Cartesian coordinate system.

w = fracture width

z = height above (or below) an arbitrary datum

z = Cartesian coordinate in the vertical direction

Z = gas Z-factor

β = non-darcy velocity coefficient

ρ = fluid density.

μ = viscosity

φ = porosity, fraction

τ = tortuosity

γ = liquid specific gravity

Φ = velocity flow potential function

Ψ = pressure head

∇ = gradient operator

ΔP = pressure change

ΔP' = welltest derivative function

Δt = shut-in time

Δte = effective shut-in time

Page 359: +Peters Ekwere j. - Petrophysics

3-159

Δh = positive difference in the hydraulic heads at points 1 and 2.

ΔL = positive length of the porous medium between points 1 and 2 in the flow direction.

REFEREENCES

Amyx, J.W., Bass, D.M., Jr. and Whiting, R.L. : Petroleum Reservoir Engineering, McGraw-Hill Book Company, New York, 1960.

Anderson, G. : Coring and Core Analysis Handbook, Petroleum Publishing Company, Tulsa, Oklahoma, 1975.

Archer, J.S. and Wall, C.G. : Petroleum Engineering, Graham & Trotman, London, England, 1986.

Archie, G.E. :“The Electrical Resistivity Log as an Aid in Determining Some Reservoir Characteristics,” Trans. AIME (1942) 146, 54-62.

Archie, G.E. :“Introduction to Petrophysics of Reservoir Rocks,” AAPG Bull., Vol. 34, No. 5 (May 1950) 943-961.

Babu, D.K. and Odeh, A.S. : "Productivity of a Horizontal Well," SPE Reservoir Engineering (November, 1989) 417-421.

Bear, J. : Dynamics of Fluids in Porous Media, Elsevier, New York, 1972.

Berg, R.R. : “Method for Determining Permeability from Reservoir Rocks,” Trans., Gulf Coast Assoc. of Geol. Soc., Vol. XX (1970) 303-335.

Bradley, H.B. (Editor-in-Chief) : Petroleum Engineering Handbook, SPE, Richardson, Texas, 1987.

Butler, R.M. : Horizontal Wells for the Recovery of Oil, Gas and Bitumen, Petroleum Society, Canadian Inst. Min., Met. and Pet., Calgary, 1994.

Carslaw, H.S. and Jaeger, J.C. : Conduction of Heat in Solids, 2nd edition, Clarendon Press, Oxford (1959).

Carman, P.C. : "Fluid Flow Through A Granular Bed,” Trans. Inst. Chem. Eng. London (1937) 15, 150-156.

Carman, P.C. : "Determination of the Specific Surface of Powders,” J. Soc. Chem. Indus (1938) 57, 225-234.

Carpenter, C.B. and Spencer, G.B. : “Measurements of the Compressibility of Consolidated Oil-Bearing Sandstones,” RI 3540, USBM (Oct. 1940).

Chilingar, G.V.: “Relationship Between Porosity, Permeability and Grain Size Distribution of Sands and Sandstones,” Proceedings of the Sixth International Sedimentological Congress, The Netherlands and Belgium,

Page 360: +Peters Ekwere j. - Petrophysics

3-160

1963, Deltaic  and  Shallow  Marine  Deposits,  L.M.J.U  van  Straatan,  Ed., Elsevier, New York, 72‐75. 

Clarke, R.H. :"Reservoir Properties of Conglomerates and Conglomeratic Sandstones," AAPG Bulletin, Vol. 63 (1979) 799-809.

Collins, R.E. : Flow of Fluids Through Porous Materials, Van Nostrand Reinhold Company, 1961. Reprinted by the Petroleum Publishing Company, 1976. Reprinted by Research & Engineering Consultants Inc., 1990.

Core Laboratories Inc. : Special Core Analysis, 1976.

Corey, A.T. : Mechanics of Heterogeneous Fluids in Porous Media, Water Resources Publications, Fort Collins, Colorado, 1977.

Cosse, R. : Basics of Reservoir Engineering, Editions Technip, Paris, 1993.

De Boodt, M.F. and Kirkham, D. : "Anisotropy and Measurement of Air Permeability of Soil Clods," Soil Sci., Vol. 76 (1953) 127-313.

Dobrynin, V.M. : "The Effect of Overburden Pressure on Some Properties of Sandstones,” Soc. Pet. Eng. J. (Dec. 1962) 360-366.

Dullien, F.A.L. : Porous Media - Fluid Transport and Pore Structure, Academic Press, New York, 1979.

Firoozabadi, A. and Katz, D.L. : "An Aanlysis of High Velocity Gas Flow Through Porous Media," JPT (Feb. 1979) 221.

Freeze, R.A. and Cherry, J.A. : Goundwater, Prentice-Hall, Englewood Cliffs, 1979.

Giger, F.M., Reiss, L.H. and Jourdan, A.P. : "The Reservoir Engineering Aspects of Horizontal Wells," paper SPE 13024, presented at the 59th SPE Annual Technical Meeting, Dallas, Texas, September 16-19, 1984.

Gilman, J.R. and Jargon, J.R. : "Evaluating Horizontal Versus Vertical Well Performance," World Oil (April, 1992) 67-72.

Gilman, J.R. and Jargon, J.R. : "Evaluating Horizontal Versus Vertical Well Performance: Part 2," World Oil (June, 1992) 55-60.

Goode, P.A. and Kuchuk, F.J. : "Inflow Performance for Horizontal Wells," SPE Reservoir Engineering (August, 1991) 319-323.

Hutta, J.J. and Griffiths, : "Directional Permeability of Sandstones: A Test Technique," Producers Monthly, Nov. 26, Vol.34 and Oct.24, Vol. 31, 1955.

Jennings, H.J. : “How to Handle and Process Soft and Unconsolidated Cores,” World Oil (June 1965) 116-119.

Page 361: +Peters Ekwere j. - Petrophysics

3-161

Johnson, W.E. and Hughes, R.V.: "Directional Permeability Measurements and their Significance," Producers Monthly, No. 1, Vol. 18 (November 1948) 17-25.

Joshi, S. D. : "A Review of Horizontal Well and Drainhole Technology," paper SPE 16868, presented at the 62nd SPE Annual Technical Meeting, Dallas, Texas, September 27-30, 1987.

Joshi, S. D. : "Production Forecasting Methods for Horizontal Wells," paper SPE 17580, presented at the SPE International Meeting on Petroleum Engineering, Tianjin, China, November 1-4, 1988.

Joshi, S. D. : Horizontal Well Technology, PennWell Publishing Co., Tulsa, 1991.

Karcher, B.J., Giger, F.M. and Combe, J. : "Some Practical Formulas to Predict Horizontal Well Behavior," paper SPE 15430, presented at the 61st SPE Annual Technical Meeting,New Orleans, LA, October 5-8, 1986.

Keelan, D.K. : "A Critical Review of Core Analysis Techniques” The Jour. Can. Pet. Tech. (April-June 1972) 42-55.

Klinkenberg, L.J. : “The Permeability of Porous Media to Liquids and Gases,” Drilling and Production Practices, American Petroleum Institute (1941) 200.

Krumbein, W.C. and Monk, G.D. : “Permeability as a Function of the Size Parameters of Unconsolidated Sand,” Amer. Int. Mining and Met. Tech. Pub. 1492, 1942.

Kuchuk, F.J., Goode, P.A., Brice, B.W., Sherrard, D.W. and Thambynayagam, R.K.M.: "Pressure Transient Analysis and Inflow Performance for Horizontal Wells," paper SPE 18300, presented at the 63rd SPE Annual Technical Meeting, Houston, Texas, October 2-5, 1988.

Leva, M., Weintraub, M., Grummer, M. Pollchick, M. and Storch, H.H. : "Fluid Flow Through Packed and Packed and Fluidized Systems," US Bureau of Mines Bull. No. 504, 1951.

Maasland, M. and Kirkham, D. : "Theory and Measurement of Anisotropic Air Permeability in Soil Clods," Proc. Soil Sci. Soc. Amer., No. 4, Vol. 19 (1955) 395-400.

Mayer-Gurr, A. : Petroleum Engineering, John Wiley & Sons, New York, 1976.

Monicard, R.P. : Properties of Reservoir Rocks, Gulf Publishing Company, Houston, TX, 1980.

Nelson, P.H. :”Permeability-Porosity Relationships in Sedimentary Rocks,” The Log Analyst (May-June 1994) 38-62.

Page 362: +Peters Ekwere j. - Petrophysics

3-162

Ozkan, E. : "Performance of Horizontal Wells," PhD Dissertation, University of

Tulsa, Tulsa, Oklahoma, 1988.

Peters, E.J. : Stability Theory and Viscous Fingering in Porous Media, PhD Dissertation, University of Alberta, Edmonton, Alberta, Canada, January 1979.

Peters, E.J. and Afzal, N. : “Characterization of Heterogeneities in Permeable Media with Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 7, No. 3/4, (May 1992) 283-296.

Pirson, S.J. : Oil Reservoir Engineering, McGraw-Hill Book Company, Inc., New York, 1958.

Poston, S.W., Ysreal, S.C., Hossain, A.K.M.S., Montgomery, E.F., III, and Ramey, H.J., Jr. : “The Effect of Temperature on Irreducible Water Saturation and Relative Permeability of Unconsolidated Sands,” Soc. Pet. Eng. J. (June 1970) 171-180; Trans., AIME (1970) 249.

Richardson, J.G. : “Flow Through Porous Media,” Section 16, Handbook of Fluid Dynamics, Edited by V.I. Streeter, McGraw-Hill Book Company, Inc., New York, 1961.

Rose, H.E. : “ An Investigation into the the Laws of Flow of Fluids Through Granular Material,” Proc. Inst. Mech. Eng. (1945) 153, 141-148.

Ryder, H.M. : "Permeability, Absolute, Effective, Measured” World Oil (May 1948) 173-177.

Scheidegger, A.E. : The Physics of Flow Through Porous Media, University of Toronto Press, Toronto, 1960.

Stehfest, H. : “Algorithm 368, Numerical Inverse of Laplace Transforms,” D-5, Communications of the ACM, Vol. 13, No.1 (1970) 47- 48.

Vose, D. : Quantitative Risk Analysis: A Guide to Monte Carlo Simulation Modeling, John Wiley and Sons, New York, 1998.

Winsauer, W.O., Shearin, H.M., Jr., Masson, P.H. and Williams, M. : "Resistivity of Brine-Saturated Sands in Relation to Pore Geometry," AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 253-277.

Wyllie, M.R.J. and Spangler, M.B. : "Application of Electrical Resistivity Measurements to Problems of Fluid Flow in Porous Media,” AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 359-403.

Wyllie, M.R.J. and Gregory, A.R.: "Fluid Flow Through Unconsolidated Porous Aggregates” Industrial and Engineering Chem. Vol. 47 (1955) 1379-1388.

Page 363: +Peters Ekwere j. - Petrophysics

4-1

CHAPTER 4

HETEROGENEITY

4.1 INTRODUCTION

Petroleum reservoir rocks are, in general, heterogeneous and very

often anisotropic in nature. This means that the petrophysical properties

of interest such as permeability, porosity, fluid saturation, lithology and

others vary in space both vertically and laterally. Evidence of

heterogeneity was seen in the results of the core analysis presented in

Table 2.2. It was found that the permeability, porosity and fluid

saturations varied from top to bottom within the same reservoir.

Variability of rock and fluid properties is a reality that must be dealt with

in petroleum resource assessment and reservoir performance prediction.

Suppose it has been decided to model reservoir performance

through numerical simulation. In the construction of the simulation

model, every one of the nearly one million grid blocks requires the

assignment of porosity, absolute permeability, relative permeability,

capillary pressure and various fluid properties. How can one generate the

values of the rock and fluid properties to be used in the simulation

model? One may assume that the reservoir is homogeneous and use the

same value of each property in every grid block. Such a simulation model

Page 364: +Peters Ekwere j. - Petrophysics

4–2

ignores the geology of the reservoir and will consequently yield

misleading and optimistic results. One may resort to populating the grid

blocks with property values from a random number generator. This

model also ignores the geology of the reservoir and the general

observation that data from nearby locations tend to be similar whereas

data from locations that are far apart tend to be dissimilar. In other

words, petrophysical data tend to be correlated. Of course, some of these

properties can be estimated from core analysis and well logs. However,

because wells sample only a very minute volume of the reservoir, the

properties in the bulk of the reservoir remain unknown and must be

estimated by other means. There is therefore a need to devise a rational

way to estimate the petrophysical properties for the reservoir simulation

model.

The estimation of reservoir properties at locations for which no

measurements have been made can be accomplished with a new type of

statistics known as geostatistics (compare with geochemistry and

geophysics). Geostatistics is the practical applications of the Theory of

Regionalized Variables developed by Georges Matheron in Fontainebleau.

The main difference between ordinary statistics and geostatistics is that

ordinary statistics is typically based on random, independent and

uncorrelated data whereas geostatistics is based on random and spatially

correlated data.

The basic concepts of geostatistics are presented in this chapter.

Before estimating the values of the properties of the heterogeneous

reservoir, one must first characterize the heterogeneity from the samples

of the property that have been collected from core analysis, well logs and

other sources. After characterizing the nature of the heterogeneity, an

Page 365: +Peters Ekwere j. - Petrophysics

4-3

estimation is undertaken that honors both the measured data (samples)

and the structure of the heterogeneity observed from the samples.

4.2 MEASURES OF CENTRAL TENDENCY AND VARIABILITY (HETEROGENEITY)

4.2.1 Measures of central tendency Mean

The mean (arithmetic mean) is the best known measure of central

tendency. Given N sample data 1 2 3, , ,..., NΦ Φ Φ Φ , the mean (arithmetic

mean) of the sample data is defined as

1

1 i N

iiN

=

=

Φ = Φ∑ (4.1)

Geometric Mean

The geometric mean is a better measure of central tendency for

data from a log normal distribution than the arithmetic mean. The

geometric mean is defined as

1

1ln lni N

geom iiN

=

=

Φ = Φ∑ (4.2)

Taking antilog in Eq.(4.2) gives the geometric mean as

1 2 3...Ngeom NΦ = Φ Φ Φ Φ (4.3)

Median

The median of a sample data set is found by determining that

value of iΦ which has equal number of values above it and below it. To

determine the median, the data are arranged in ascending (or

Page 366: +Peters Ekwere j. - Petrophysics

4–4

descending) order. If the number of data N is odd, then the median is

determined from the sorted data as

( 1) / 2 for N oddmedian N +Φ = Φ (4.4)

If N is even, then the mean of the two most central values is used to

calculate the median as

( )( )/ 2 / 2 11 for N even2median N N +Φ = Φ + Φ (4.5)

Mode

The mode of a probability distribution ( )p Φ is the value of Φ at

which the probability distribution is a maximum. The mode is useful

primarily when there is a single sharp maximum, in which case the mode

estimates the central value. Sometimes, a distribution is bimodal, with

two relative maxima. In such a case, one may wish to know the two

modes individually. For a bimodal distribution, the mean and median are

not very useful as measures of central tendency.

4.2.2 Measures of Variability (Heterogeneity or Spread) Variance

The most useful measure of variability around the central value is

the variance defined for sample data as

( ) ( )2

21

1

1...1

i N

N ii

Var sN

=

=

Φ Φ = = Φ − Φ− ∑ (4.6)

The standard deviation is the positive square root of the variance and is

given by

Page 367: +Peters Ekwere j. - Petrophysics

4-5

( ) ( )1 1... ...N Ns VarΦ Φ = Φ Φ (4.7)

Dykstra-Parsons Coefficient of Variation

A measure of permeability variability that is widely used in the

petroleum industry is the Dykstra-Parsons (1950) coefficient of variation,

V. This coefficient of variation is determined based on the assumption

that permeability data are drawn from a log normal distribution. The

calculation of the Dykstra-Parsons coefficient of permeability variation

involves plotting the frequency distribution of the permeability data on a

log-normal probability graph paper. This is done by arranging the

permeability values in descending order and then calculating for each

permeability, the percent of the samples with permeabilities greater than

or equal to that value. Table 4.1 shows an example calculation. Note

that to avoid values of 0 or 100%, which are not present on the

probability scale, the percent greater than or equal to value is normalized

by N+1, where N is the number of samples.

The data are plotted on a log-normal probability graph paper as

shown in Figure 4.1. Normally, such a plot gives a straight line, at least

when the central portion is used. The midpoint of the permeability

distribution ( )% 50≥ is the median permeability. The Dykstra-Parsons

coefficient of permeability variation, V, is defined as

50 84.1

50

k kVk−

= (4.8)

where k50 is the permeability value at ( )% 50≥ , which is the log mean

permeability and k84.1 is the permeability value at ( )% 84.1≥ , which is one

Page 368: +Peters Ekwere j. - Petrophysics

4–6

standard deviation from the mean. For the data of Table 4.1 and Figure

4.1,

N+1 = 20

k50 = 17.8 md

k84.1 = 6.1 md

50 84.1

50

17.8 6.1 0.6617.8

k kVk− −

= = =

Table 4.1. Calculation of Dykstra-Parsons Coefficient of Variation (V)

k (md)

No of Samples >= k % >= k

100 1 5 65 2 10 50 3 15 32 4 20 29 5 25 27 27 7 35 23 23 9 45 18 10 50 17 17 12 60 16 13 65 8 14 70 7 7 16 80 6 17 85 4 18 90 1 19 95

Page 369: +Peters Ekwere j. - Petrophysics

4-7

Figure 4.1. Log normal permeability distribution.

Dykstra-Parsons coefficient of variation is a dimensionless number

that ranges from 0 to 1. A homogeneous reservoir has a coefficient of

permeability variation that approaches 0 whereas an extremely

heterogeneous reservoir has a coefficient of permeability variation that

approaches 1. Petroleum reservoirs typically have Dykstra-Parsons

coefficients of permeability variation between 0.5 and 0.9. Figure 4.2

shows theoretical log normal permeability distributions and their

corresponding Dykstra-Parsons coefficients.

Page 370: +Peters Ekwere j. - Petrophysics

4–8

Figure 4.2. Theoretical log normal permeability distributions and their corresponding Dykstra-Parsons coefficients (Carlson, 2003).

Lorenz Coefficient

Another measure of heterogeneity used in the petroleum industry

is the Lorenz coefficient . The Lorenz coefficient of variation is obtained

by plotting a graph of cumulative kh versus cumulative φh, sometimes

called a flow capacity plot. Table 4.2 shows an example calculation

whereas Figure 4.3 shows the plot of cumulative kh versus cumulative

φh for determining the Lorenz coefficient. The Lorenz coefficient is

defined from Figure 4.2 as

area ABCALorenz Coefficientarea ADCA

= (4.9)

Page 371: +Peters Ekwere j. - Petrophysics

4-9

where ABCA is the cross-hatched area in the figure and ADCA is the

triangle below the cross-hatched area. From Figure 4.2, the Lorenz

coefficient of the data of Table 4.2 is about 0.65.

The Lorenz coefficient of variation also varies from 0 to 1.

Unfortunately, the Lorenz coefficient is not a unique measure of reservoir

heterogeneity. Several different permeability distributions can give the

same value of Lorenz coefficient. For log-normal permeability

distribtutions, the Lorenz coefficient is very similar to the Dykstra-

Parsons coefficient of permeability variation.

Table 4.2 . Calculation of Lorenz Coefficient of Variation

h k φ kh Σkh Σkh/Sumk

h

φh Σφh Σφh/Sumφh

(ft) (md) (fraction) (md-ft) (md-ft) (ft) (ft)

0.000 0.000

8.1 4388 0.22 35542.

8

35542.

8

0.399 1.782 1.782 0.062

2.0 2640 0.22 5280.0 40822.

8

0.459 0.440 2.222 0.078

3.7 2543 0.21 9409.1 50231.

9

0.564 0.777 2.999 0.105

5.0 1579 0.20 7895.0 58126.

9

0.653 1.000 3.999 0.140

4.0 930 0.20 3720.0 61846.

9

0.695 0.800 4.799 0.168

9.6 662 0.19 6355.2 68202.

1

0.766 1.824 6.623 0.232

6.4 441 0.20 2822.4 71024.

5

0.798 1.280 7.903 0.277

2.7 402 0.16 1085.4 72109.

9

0.810 0.432 8.335 0.292

Page 372: +Peters Ekwere j. - Petrophysics

4–10

5.6 401 0.20 2245.6 74355.

5

0.836 1.120 9.455 0.331

8.0 378 0.18 3024.0 77379.

5

0.870 1.440 10.895 0.381

4.0 267 0.21 1068.0 78447.

5

0.882 0.840 11.735 0.411

15.1 250 0.21 3775.0 82222.

5

0.924 3.171 14.906 0.522

5.9 249 0.17 1469.1 83691.

6

0.940 1.003 15.909 0.557

2.8 232 0.22 649.6 84341.

2

0.948 0.616 16.525 0.578

7.4 200 0.17 1480.0 85821.

2

0.964 1.258 17.783 0.622

9.2 136 0.20 1251.2 87072.

4

0.978 1.840 19.623 0.687

7.6 98 0.19 744.8 87817.

2

0.987 1.444 21.067 0.737

10.1 47 0.19 474.7 88291.

9

0.992 1.919 22.986 0.804

9.6 30 0.18 288.0 88579.

9

0.995 1.728 24.714 0.865

3.5 28 0.17 98.0 88677.

9

0.997 0.595 25.309 0.886

6.6 16 0.16 105.6 88783.

5

0.998 1.056 26.365 0.923

15.8 13 0.14 205.4 88988.

9

1.000 2.212 28.577 1.000

Sumkh = 88988.

9

Sumφh = 28.577

Page 373: +Peters Ekwere j. - Petrophysics

4-11

Figure 4.3. Flow capacity distribution.

While the Dykstra-Parsons and Lorenz coefficients give

quantitative measures of the permeability variation, they provide no

information on the spatial relationship between the permeability values.

It is well known that permeability and other petrophysical properties of

reservoir rocks are not randomly distributed but are spatially correlated.

There is a need for other measures of heterogeneity that take into

account the spatial correlation of the data.

4.3 MEASURES OF SPATIAL CONTINUITY

Figure 4.4 shows the spatial distributions of a petrophysical

property, Φ , measured at equally spaced coordinates in two linear

reservoirs A and B. Which of the two reservoirs is more heterogeneous

with respect to the property Φ ? Most professionals will say that reservoir

Page 374: +Peters Ekwere j. - Petrophysics

4–12

A is more heterogeneous than reservoir B. However, a careful

examination of the numerical values of the property shows them to be

the same in both reservoirs. For both reservoirs, the mean of Φ is 5.0,

the variance is 6.67 and the standard deviation is 2.58. Thus, by the

usual measure of ordinary statistics, the degree of heterogeneity of the

two reservoirs is the same. Yet, there is something about the two

reservoirs that leads one to conclude that A is more heterogeneous than

B. It is the spatial arrangement of the values of the property relative to

each that leads one to conclude that A is more heterogeneous than B. In

reservoir A, the property appears to be randomly distributed in space

whereas in reservoir B, it is distributed in an orderly and continuous

fashion. Thus, to fully characterize heterogeneity, the spatial correlation

structure of the property must be taken into account.

Three related functions are normally used to characterize the

spatial continuity of the data from a heterogeneous reservoir. These are

(1) the variogram (semi-variogram), (2) the covariance function and (3)

the correlation coefficient function.

Page 375: +Peters Ekwere j. - Petrophysics

4-13

Figure 4.4. Spatial distribution of a petrophysical property in two linear reservoirs A and B.

Page 376: +Peters Ekwere j. - Petrophysics

4–14

4.3.1. Variogram Definition

The variogram is a function obtained by plotting the semivariance

of the differences between the properties at two locations separated by a

distance h versus h. The variogram is defined as

( ) ( ) ( ) ( ) ( ) 2

2 2 h

x x hVar x x hh

Φ − Φ +⎡ ⎤⎣ ⎦Φ − Φ +⎡ ⎤⎣ ⎦= =∑

(4.10)

where

γ = semivariance

h = lag distance

( )xΦ = value of property at location x

( )x hΦ + = value of property at location x+h

Nh = number of data pairs separated by the distance h.

Examination of Eq.(4.10) shows that each numerical value of γ is the

variance of ( ) ( )x x hΦ − Φ +⎡ ⎤⎣ ⎦ divided by 2, where the mean of

( ) ( )x x hΦ − Φ +⎡ ⎤⎣ ⎦ is normally assumed to be zero. Thus, each numerical

value of γ is the semivariance of ( ) ( )x x hΦ − Φ +⎡ ⎤⎣ ⎦ for a lag distance h and

the function γ (h) is the variogram or semivariogram.

Figure 4.5 shows an ideal variogram. It starts at zero and increases

with increasing lag distance until a certain distance is reached at which

it levels off and becomes constant. The lag distance at which the

variogram levels off (a in the figure) is defined as the correlation length,

Page 377: +Peters Ekwere j. - Petrophysics

4-15

or the range of influence, and the value of the variogram at this point is

called the sill. The sill is the semivariance of the entire data set. Thus,

Figure 4.5. Typical variogram

hidden in the variogram are the variance and standard deviation of the

data set, the usual measures of heterogeneity of ordinary statistics. If the

correlation length is zero, the spatial distribution of the property is fully

random. With increasing correlation length, the range of influence of one

value on its neighbors increases up to the correlation length. At lag

distances beyond the correlation length, the data are no longer

correlated.

The variogram and correlation length are directional quantities

(anisotropy again), and in general, will be different in different directions.

Page 378: +Peters Ekwere j. - Petrophysics

4–16

Sometimes, the variogram has a discontinuity at the origin as shown in

Figure 4.5. The value of the variogram at the discontinuity (Co in the

figure) is known as the nugget effect, a term that originates from the

mining industry which was the first industry to widely apply geostatistics

to estimate ore grade.

It should be emphasized that not all variograms have a nugget

effect, a sill or even a correlation length. Variograms come in different

shapes depending on the underlying geological structure of the

heterogeneity. Figure 4.5 was presented as one possible variogram shape

to introduce the general features of the function and the associated

nomenclature.

In order to demonstrate that the variogram does capture the

spatial continuity of the data from a heterogeneous reservoir, let us

compare the variograms for reservoirs A and B as shown in Figure 4.6.

The variogram for reservoir A is cyclical with a constant average value of

about 5.5 for all lag distances. The constant average value can be viewed

as a pure nugget effect, which indicates that the values of the property

are randomly distributed in space as is apparent from Figure 4.4. By

contrast, the variogram for reservoir B is continuous for all lag distances,

an indication of the orderly and continuous spatial distribution of the

property as evident in Figure 4.4. Therefore, the variogram (and later the

covariance and correlation functions) does capture the spatial continuity

of the data.

Page 379: +Peters Ekwere j. - Petrophysics

4-17

Figure 4.6. Comparison of the variograms for reservoirs A and B of Figure 4.4.

How to Calculate the Variogram

Let us consider the simple case of a linear reservoir in which the

property Φ is measured at 10 equally spaced locations with a distance of

Δx between each datum. Let the values of the property at locations 1

through 10 be 1 2 3 10, , ,...,Φ Φ Φ Φ . To compute the semivariance at each lag

distance, one can generate a table of ( )xΦ and ( )x hΦ + for each lag

distance as shown in Table 4.3. Using the entries in the table, the

semivariances can easily be calculated with Eq.(4.10). In fact, the

semivariance for h = 0 can be obtained by inspection as γ(0) = 0. The

Page 380: +Peters Ekwere j. - Petrophysics

4–18

semivariance at h = 0 is always equal to 0 even if the variogram has a

nugget effect.

Table 4.3. Table of Φ(x) and Φ(x+h) for Computing Semivariances of

Sample Data

h = 0 h = Δx h = 2Δx h = 3Δx h = 4Δx h = 5Δx

Nh = 10 Nh = 9 Nh = 8 Nh = 7 Nh = 6 Nh = 5

Φ(x

) Φ (x+h)

Φ(x

) Φ (x+h)

Φ(x

) Φ (x+h)

Φ(x

) Φ (x+h) Φ(x) Φ (x+h) Φ(x) Φ (x+h)

Φ1 Φ1 Φ1 Φ2 Φ1 Φ3 Φ1 Φ4 Φ1 Φ5 Φ1 Φ6

Φ2 Φ2 Φ2 Φ3 Φ2 Φ4 Φ2 Φ5 Φ2 Φ6 Φ2 Φ7

Φ3 Φ3 Φ3 Φ4 Φ3 Φ5 Φ3 Φ6 Φ3 Φ7 Φ3 Φ8

Φ4 Φ4 Φ4 Φ5 Φ4 Φ6 Φ4 Φ7 Φ4 Φ8 Φ4 Φ9

Φ5 Φ5 Φ5 Φ6 Φ5 Φ7 Φ5 Φ8 Φ5 Φ9 Φ5 Φ10

Φ6 Φ6 Φ6 Φ7 Φ6 Φ8 Φ6 Φ9 Φ6 Φ10

Φ7 Φ7 Φ7 Φ8 Φ7 Φ9 Φ7 Φ10

Φ8 Φ8 Φ8 Φ9 Φ8 Φ10

Φ9 Φ9 Φ9 Φ10

Φ10 Φ10

The following observations can be made about the calculation of

semivariances and the variogram.

1. The number of data pairs, Nh, decreases as h increases. Beyond h

= NΔx/2, the reduction in Nh causes the variogram to fluctuate

excessively and unmeaningfully. Therefore, the variogram is

typically truncated beyond h = NΔx /2.

2. The variogram is a non-negative function.

Page 381: +Peters Ekwere j. - Petrophysics

4-19

3. The variogram can be computed even for irregularly spaced data.

In this case, a different computational strategy must be used. Here

is one possible algorithm. Consider the non-uniformly distributed

data at ten locations in a linear reservoir as shown in Figure 4.7. If

there are N data points, there will be NC2 data pairs in the set

where NC2 is the combination of N things taken 2 at a time and is

given by

( )

( )2

1!2 !2! 2N

N NNCN

−= =

− (4.11)

For 10 data points, there will be 45 data pairs. The first task is to

compute and store the lag distances, hij, and the corresponding

values of ( )2

2i jΦ − Φ

for all the data pairs. The algorithm is as

follows:

1. Create two one-dimensional arrays of size equal to the

number of data pairs. Let these be H(M) and B(M) where M is

the number of data pairs as determined from Eq.(4.11).

2. Visit location 1, compute and store sequentially, the

following data pairs:

( )

( )

( )

22 1

2 1

23 1

3 1

210 1

10 1

,2

,2

...

,2

x x

x x

x x

Φ − Φ−

Φ − Φ−

Φ − Φ−

Page 382: +Peters Ekwere j. - Petrophysics

4–20

Figure 4.7. Irregularly spaced data.

where the first number is stored in array H and the second

number is stored in array B.

3. Eliminate 1Φ from the data set to prevent duplication.

4. Move to location 2 and continue to compute and store the

following data in the arrays

Page 383: +Peters Ekwere j. - Petrophysics

4-21

( )

( )

( )

23 2

3 2

24 2

4 2

210 2

10 2

,2

,2

...

,2

x x

x x

x x

Φ − Φ−

Φ − Φ−

Φ − Φ−

5. Eliminate 2Φ from the data set to avoid duplication.

6. Continue to compute and fill the arrays by visiting each

location in the manner described above until the last data

pair is added to the array. The last data pair for our example

will be

( )210 9

10 9 ,2

x xΦ − Φ

7. Perform a scatter plot of B(i) versus H(i), where array H now

contains the lag distances, hij, that were computed as

i jx x− . Such a scatter plot is shown in Figure 4.8.

8. Divide the data in the scatter diagram into bins as shown in

Figure 4.8. For each bin, compute and plot

( )21 1 versus

2i j

ijn n

hn n

Φ − Φ∑ ∑

where n is the number of data points in the bin, which can

be different for each bin. In constructing the bins, there

should be enough data points in each bin to prevent the

Page 384: +Peters Ekwere j. - Petrophysics

4–22

variogram from fluctuating excessively. Several bin sizes

may be tried to determine the optimum bin sizes.

9. The plot in step 8 is the experimental variogram as shown in

Figure 4.8.

Figure 4.8. Scatter plot for computing experimental variogram.

Page 385: +Peters Ekwere j. - Petrophysics

4-23

4. If there is an underlying trend in the data, the trend should be

subtracted from the data before computing the variogram. For

example, the following data set from a linear reservoir C has an

underlying trend or drift of the form 1trend xΦ = + :

3 6 9 12 15 15 14 13 12

This data set is presented in Figure 4.9. This trend or drift should

be subtracted from the data before the variogram is computed.

Figure 4.10 compares the variograms for reservoir C with and

without the underlying trend or drift. The variogram with the drift

is an ever increasing function of h because of the underlying trend

and will never reach the sill, whereas after removing the drift, the

resulting variogram is lower and reaches the sill at h = 4 km. This

is the true experimental variogram for reservoir C. In this example,

after subtracting the drift, the resulting distribution is the same as

in reservoir B. Thus, the true variogram for reservoir C is the same

as for reservoir B.

5. If data are missing from some locations, those locations should be

skipped over. Resist the temptation to interpolate and fill in the

missing data. Table 4.4 shows a data set in which data are

missing from location 7. Calculate the semivariance of the

permeability (NOT the natural log of permeability) at lag distances

of 1 foot and 2 feet. Start your calculations from the top of the

reservoir and work your way down.

Page 386: +Peters Ekwere j. - Petrophysics

4–24

Figure 4.9. Spatial distribution of a petrophysical property in a linear reservoir C with an underlying trend or drift.

Figure 4.10. Variograms for reservoir C with and without the underlying trend or drift.

Page 387: +Peters Ekwere j. - Petrophysics

4-25

Table 4.4. Porosity and Permeability Distributions with Missing Data.

Relativ

e

Depth

Porosit

y

Permeabilit

y

(ft) (%) (md)

1 20 93

2 16.3 18

3 9.7 8.4

4 16.2 21

5 14.9 10

6 12.7 1.7

7

8 5.5 25

9 5.8 17

10 6.5 4.8

11 4.7 22

12 7.3 5.6

6. If the data are distributed in 2D or 3D, the variogram can still be

computed using the basic method outlined for 1D data. In this

case, the variogram should be computed in several directions to

reveal any anisotropy that may be present. Figure 4.11 shows the

distribution of a petrophysical property in a 2D reservoir. In this

case, variograms should be computed in the following directions:

N-S, E-W, NE-SW and NW-SE.

Page 388: +Peters Ekwere j. - Petrophysics

4–26

Figure 4.11. Distribution of a petrophysical property in a two-dimensional reservoir.

7. Unless you are already familiar with the variogram, you may not

see what it has to do with spatial correlation of the data. We can

see that the entries in Table 4.3 have a lot to do with spatial

correlation by plotting the scatter diagrams of Φ(x+h) versus Φ(x)

for each value of h. For h = 0, Φ(x+h) and Φ(x) are perfectly

correlated and the data will follow the 45º line on the scatter plot

as shown in Figure 4.12. As h increases, the cloud of data points

scattered about the 45º line increases, indicating less and less

correlation between Φ(x+h) and Φ(x) as shown in Figures 4.13 and

4.14.

Page 389: +Peters Ekwere j. - Petrophysics

4-27

Figure 4.12. Scatter plot of Φ(x+h) versus Φ(x) for h = 0 km for reservoir B.

Figure 4.13. Scatter plot of Φ(x+h) versus Φ(x) for h = 1 km for reservoir B.

Page 390: +Peters Ekwere j. - Petrophysics

4–28

Figure 4.14. Scatter plot of Φ(x+h) versus Φ(x) for h = 3 km for reservoir B.

Physical Meaning of the Variogram

The scatter plot of Φ(x+h) versus Φ(x) at a fixed h can be used to

derive the physical meaning of the variogram. Consider the scatter plot

shown in Figure 4.15. The distance d from a datum point to the 45º line

is given by

( ) ( ) ( ) ( )cos 45

2

x h xd x h x

Φ + − Φ⎡ ⎤⎣ ⎦= Φ + − Φ =⎡ ⎤⎣ ⎦ (4.12)

( ) ( ) ( ) ( ) 222 2cos 45

2x h x

d x h xΦ + − Φ⎡ ⎤⎣ ⎦= Φ + − Φ =⎡ ⎤⎣ ⎦ (4.13)

From statistics, the expectation of d2 is given by

Page 391: +Peters Ekwere j. - Petrophysics

4-29

( )( ) ( ){ } ( ) ( )

( )

22

2

2 2 h

x h xE x h xE d h

Φ + − Φ⎡ ⎤Φ + − Φ⎡ ⎤ ⎣ ⎦⎣ ⎦= = =

∑ (4.14)

Thus, the variogram is the mean of d2 about the 45º line at each lag

distance as a function of the lag distance h.

Figure 4.15. The h-scatter plot.

Variogram Models

The variogram is a means to an end not an end in itself. The

variogram is used to quantify the correlation structure of the variable of

interest for the purpose of estimation and conditional simulation. The

variogram of the sample data is known as the experimental variogram.

After computing the experimental variogram, a smooth theoretical

Page 392: +Peters Ekwere j. - Petrophysics

4–30

variogram model is usually fitted to the experimental variogram and the

model is then used for estimation. As will be shown later, the estimation

process involves the solution of a set of linear simultaneous algebraic

equations, whose coefficients are derived from the variogram. Unless the

variogram is well behaved, the simultaneous equations may not have a

solution. Hence, the need to fit a smooth and well behaved theoretical

model to the rough experimental variogram for the purpose of estimation.

Popular variogram models include (1) the spherical model, (2) the

exponential model, (3) the guassian model, (4) the linear model, (5) the

generalized linear model, (6) the nugget effect model, and (7) the cardinal

sine model (also known as the hole effect model).

1. The Spherical Model

The spherical model is given by

( )

( )

3

3

3 1 for 2 2

for

o

o

h hh C C h a

a a

h C C h a

γ

γ

⎛ ⎞= + ⎜ − ⎟ <

⎜ ⎟⎝ ⎠

= + ≥

(4.15)

where a is the correlation length or range, Co is the nugget effect if

present and C is the sill minus Co.

2. The Exponential Model

The exponential model is given by

( ) 1 ha

oh C C eγ⎛ ⎞

−⎜ ⎟⎜ ⎟⎝ ⎠

⎛ ⎞⎜ ⎟= + −⎜ ⎟⎝ ⎠

(4.16)

Page 393: +Peters Ekwere j. - Petrophysics

4-31

Figure 4.16 compares the spherical and the exponential models

with the same range and sill whereas Figure 4.17 compares the

two models with the same initial slope and sill. Note that there is

no nugget effect in these figures.

Figure 4.16. A comparison of the spherical and exponential models with the same range and sill.

Page 394: +Peters Ekwere j. - Petrophysics

4–32

Figure 4.17. A comparison of the spherical and exponential models with the same initial slope and sill.

3. The Gaussian Model

The Gaussian model is given by

( )2

2

1 ha

oh C C eγ

⎛ ⎞⎜ ⎟−⎜ ⎟⎝ ⎠

⎛ ⎞⎜ ⎟= + −⎜ ⎟⎜ ⎟⎝ ⎠

(4.17)

Figure 4.18 shows an example of a Gaussian model with Co = 0, C

= 10 and a = 4 units.

Page 395: +Peters Ekwere j. - Petrophysics

4-33

Figure 4.18. Example Gaussian model.

4. The Linear Model

The linear model is given by

( ) oh C m hγ = + (4.18)

where m is the slope. This model does not have a sill.

5. The Generalized Linear Model

The generalized linear model is given by

( ) 0 2oh C m h forαγ α= + < ≤ (4.19)

Page 396: +Peters Ekwere j. - Petrophysics

4–34

where m is a constant and α is an exponent between 0 and 2.

When α = 1, the model degenerates to the linear model. This model

does not have a sill. Figure 4.19 compares the linear model to the

generalized linear model.

Figure 4.19. The linear model and the generalized linear model.

5. The Nugget Effect Model

The nugget effect model is given by

( )( )0 0

for 0oh C h

γ

γ

=

= > (4.20)

This model gives the variogram of a property that has a random

spatial distribution. It is basically a spherical model with a very

Page 397: +Peters Ekwere j. - Petrophysics

4-35

small range of influence. Figure 4.20 shows an example nugget

effect model.

Figure 4.20. The nugget effect model.

6. The Cardinal Sine Model (Hole Effect Model)

The cardinal sine model or hole effect model is given by

( ) ( )sin /1

/o

h ah C C

h aγ

⎡ ⎤= + −⎢ ⎥

⎣ ⎦ (4.21)

Figure 4.21 shows an example cardinal sine model with Co = 0, C =

10 and a = 1 unit. The sinusoidal nature of the variogram is an

indication of the periodic nature of the underlying heterogeneity.

Page 398: +Peters Ekwere j. - Petrophysics

4–36

Figure 4.21. The cardinal sine model (hole effect model).

Fitting a Theoretical Variogram Model to an Experimental Variogram

Every experimental variogram can be fitted with a theoretical

variogram model. If one model does not fit the experimental variogram,

several models can be combined or nested to fit the experimental

variogram. The only restriction is that each of the combined models must

be applied to all the lag distances. It is not permissible to apply a model

up to a certain lag distance and then switch to a different model for the

remaining lag distances. The procedure for fitting a theoretical model to

an experimental variogram is a trial and error procedure.

To demonstrate the procedure for fitting a theoretical model to an

experimental variogram, let us compute the experimental variogram for

the permeability data from the results of the core analysis of Table 2.2

and then fit a theoretical model to it. To keep the magnitude of the

Page 399: +Peters Ekwere j. - Petrophysics

4-37

variogram manageable, we compute the variogram for the natural log of

the permeability rather than the permeability itself. Such a

transformation is frequently done in geostatistics to make the data more

manageable. The transformation affects the magnitude of the variogram

but not its shape. It is not unusual to transform the data, compute the

variogram, fit a theoretical model to the experimental variogram, use the

model to perform estimation and then transform the estimated data back

to its original units. It should be noted that the transformation in this

example does not imply that the permeability data is log normally

distributed. The transformation has been done purely for convenience.

The results of the calculations are summarized in Table 4.5.

Table 4.5. Variograms for Permeability Data of Table 2.2.

1 2 3 4 5 6 7 8

Uncorrecte

d

Correcte

d

γ(h) γ(h)

Depth X-

Coord

k lnk lnk h (ft) With Drift Without Drift

4807.

5

0 2.5 0.916 -3.812 0 0.000 0.000

4808.

5

1 59 4.078 -0.688 1 0.038 0.362

4809.

5

2 221 5.398 0.596 2 0.049 0.488

4810.

5

3 211 5.352 0.512 3 0.051 0.468

4811.

5

4 275 5.617 0.740 4 0.057 0.618

4812.

5

5 384 5.951 1.037 5 0.057 0.591

4813. 6 108 4.682 -0.269 6 0.050 0.545

Page 400: +Peters Ekwere j. - Petrophysics

4–38

5

4814.

5

7 147 4.990 0.002 7 0.054 0.574

4815.

5

8 290 5.670 0.644 8 0.061 0.630

4816.

5

9 170 5.136 0.073 9 0.060 0.623

4817.

5

10 278 5.628 0.528 10 0.065 0.632

4818.

5

11 238 5.472 0.335 11 0.065 0.641

4819.

5

12 167 5.118 -0.057 12 0.068 0.728

4820.

5

13 304 5.717 0.505 13 0.075 0.780

4821.

5

14 98 4.585 -0.664 14 0.064 0.605

4822.

5

15 191 5.252 -0.034 15 0.074 0.684

4823.

5

16 266 5.583 0.260 16 0.077 0.628

4824.

5

17 40 3.689 -1.672 17 0.057 0.485

4825.

5

18 260 5.561 0.163 18 0.084 0.632

4826.

5

19 179 5.187 -0.248 19 0.086 0.642

4827.

5

20 312 5.743 0.271 20 0.097 0.720

4828.

5

21 272 5.606 0.097 21 0.098 0.655

4829.

5

22 395 5.979 0.432 22 0.105 0.769

4830. 23 405 6.004 0.420 23 0.115 0.787

Page 401: +Peters Ekwere j. - Petrophysics

4-39

5

4831.

5

24 275 5.617 -0.004 24 0.109 0.701

4832.

5

25 852 6.748 1.089 25 0.135 0.946

4833.

5

26 610 6.413 0.718 26 0.136 0.830

4834.

5

27 406 6.006 0.274 27 0.132 0.880

4835.

5

28 535 6.282 0.513 28 0.149 0.965

4836.

5

29 663 6.497 0.690 29 0.164 1.170

4837.

5

30 597 6.392 0.548

4838.

5

31 434 6.073 0.192

4839.

5

32 339 5.826 -0.092

4840.

5

33 216 5.375 -0.580

4841.

5

34 332 5.805 -0.188

4842.

5

35 295 5.687 -0.343

4843.

5

36 882 6.782 0.715

4844.

5

37 600 6.397 0.292

4845.

5

38 407 6.009 -0.133

4847.

5

40 479 6.172 -0.044

Page 402: +Peters Ekwere j. - Petrophysics

4–40

4848.

5

41 0

4849.

5

42 139 4.934 -1.356

4850.

5

43 135 4.905 -1.422

4851.

5

44 0

Figure 4.22 shows the experimental variogram for the natural log of

permeability, lnk. The graph is the plot of the data in columns 6 and 7 of

Table 4.5. The ever increasing nature of the variogram indicates an

underlying trend or drift in the permeability data. Before addressing the

problem of the underlying trend, let us fit a theoretical model to the

experimental variogram to demonstrate the procedure.

Figure 4.22. Experimental variogram for natural log of permeability for core analysis data.

Page 403: +Peters Ekwere j. - Petrophysics

4-41

Figure 4.23 shows the fitted theoretical model from trial and error.

The nested model is

( )3

23

3 10.027 0.025 0.0001012 3 2 3

h hh hγ

⎛ ⎞= + ⎜ − ⎟ +

⎜ ⎟⎝ ⎠

The nested model is of the form

( ) h Spherical Model Generalized Linear Modelγ = +

Figure 4.23. Theoretical variogram model fit to the experimental

variogram of core analysis data.

Page 404: +Peters Ekwere j. - Petrophysics

4–42

Let us now deal with the underlying trend in the permeability data.

Figure 4.24 shows a scatter plot of the log permeability data versus

relative depth together with the regression line. Clearly, there is an

underlying trend or drift in the permeability data. The trend line is

ln 0.0372 4.7281k x= +

where x is the relative depth. This trend was subtracted from the lnk

data to obtain the corrected lnk data shown in column 5 of Table 4.5.

Figure 4.24. Permeability data showing underlying trend or drift.

Page 405: +Peters Ekwere j. - Petrophysics

4-43

Figure 4.25 shows the scatter plot for the corrected permeability data

together with the regression line. Clearly, the underlying trend has been

removed.

Figure 4.26 shows the experimental variogram and the theoretical

model fit after the drift has been removed. The experimental variogram is

the graph of the data in columns 6 and 8 of Table 4.5. We see that after

removing the underlying trend, the experimental variogram can be fitted

with a spherical model of the form

( )3

3

3 10.31 0.352 9 2 9

h hhγ

⎛ ⎞= + ⎜ − ⎟

⎜ ⎟⎝ ⎠

.

Figure 4.25. Permeability data after the drift has been removed.

Page 406: +Peters Ekwere j. - Petrophysics

4–44

The true variogram of the natural log of permeability has a nugget effect

of 0.31, a sill of 0.66 and a correlation length of 9 ft.

Variogram Anisotropy

Variograms computed in different directions can show anisotropy.

Figure 4.27 shows two such anisotropies. In Figure 4.27a, the sills of the

variograms in the two directions are the same but the correlations

lengths are different. In Figure 4.27b, the slopes of the variograms in the

two directions are different. We can compute the correlation lengths or

the slopes of the variograms in different directions and plot them as

Figure 4.26. Variogram of natural log of permeability after removing the

underlying trend.

Page 407: +Peters Ekwere j. - Petrophysics

4-45

shown in Figure 4.28. If the plot turns out to be an ellipse as shown in

the figure, then a simple coordinate transformation can be used to

compute an equivalent isotropic variogram. If γ1(h) is the variogram in

direction 1, then the equivalent isotropic variogram is given by

( )2

2 211 1 2

2

ah h ha

γ γ⎛ ⎞⎛ ⎞⎜ ⎟= + ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(4.22)

for the case of the correlation length anisotropy and

( )2

2 211 1 2

2

slopeh h hslope

γ γ⎛ ⎞⎛ ⎞⎜ ⎟= + ⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

(4.23)

for the case of the slope anisotropy.

Figure 4.27. Geometric anisotropy.

Page 408: +Peters Ekwere j. - Petrophysics

4–46

Figure 4.28. Ellipses showing variogram anisotropy.

Figure 4.28b shows the importance of computing the variograms in at

least four directions in a 2D data set to reveal possible anisotropy. If the

variograms were computed only in the horizontal and vertical directions,

the anisotropy would be missed because, in the case shown in the figure,

the principal axes of the anisotropy make an angle of 45º with the

vertical and horizontal axes. As a result, the correlation lengths in the

vertical and horizontal directions are the same and would not reveal the

presence of the anisotropy.

In sedimentary rocks, because of layering, the sill of the variogram

in the vertical direction is usually different from that in the horizontal

direction because the degree of heterogeneity normal to the layers is

higher than along the layers. In this case, the variogram can be split into

two components, an isotropic component given by ( )2 2 21 2 3o h h hγ + + and a

vertical component given by ( )3v hγ . The overall variogram is then given

by

( ) ( ) ( )2 2 21 2 3 3o vh h h h hγ γ γ= + + + (4.24)

Page 409: +Peters Ekwere j. - Petrophysics

4-47

Example Experimental Variograms

Presented in this section are the experimental variograms for the

sandpack and Berea sandstone cores whose porosity distributions were

presented in Figures 2.7 to 2.10. Experimental variograms can be

computed for any property that varies in space. Here, we present the

variograms for the CT numbers of the dry scans of the sandpack and the

Berea sandstone. The CT number is proportional to the bulk density of

the sample and is similar to the measurements from a density log. Each

3D image data set consisted of 128x128x50 voxels (volume elements) for

a total of 812,200 data points. For each sample, variograms were

computed in three orthogonal directions (Z, Y and X) as shown in Figure

4.29.

Figure 4.29. Orthogonal directions in which variograms were computed for CT data.

Page 410: +Peters Ekwere j. - Petrophysics

4–48

Figure 4.30 shows CT images for a vertical slice and a typical

cross-sectional slice of the sandpack. The packing method introduced

radial and longitudinal heterogeneities into the sandpack. Figures 4.31

and 4.32 show the experimental variograms for the sandpack in the

transverse and longitudinal directions. Figure 4.31 shows that the

vertical and horizontal variograms are essentially the same, confirming

the radial symmetry observed in the CT images of the sandpack. The

variogram in the longitudinal direction (Figure 4.32) has a wavy or

sinusoidal characteristic. This is caused by the longitudinal

heterogeneity introduced into the sandpack by the packing method. This

characteristic is known as a trend surface.

Page 411: +Peters Ekwere j. - Petrophysics

4-49

Figure 4.30. Vertical and cross-sectional slices of CT image of dry

sandpack (Peters and Afzal, 1992).

Figure 4.31. Experimental variograms for sandpack in the transeverse

directions (Y, Z) (Peters and Afzal, 1992).

Page 412: +Peters Ekwere j. - Petrophysics

4–50

Figure 4.32. Experimental variogram for sandpack in the longitudinal

direction (X) (Peters and Afzal, 1992).

Figure 4.33 shows CT images for a vertical slice and a typical

cross-sectional slice of the Berea sandstone. Figures 4.34 and 4.35 show

the experimental variograms for the Berea sandstone in the transverse

and longitudinal directions. Figure 4.34 shows a wavy characteristic in

the vertical direction (Y-direction). This is caused by the layering in that

direction. As expected, the magnitudes of the variograms show that the

medium is more heterogeneous in the Y-direction (across the layers) than

in the Z-direction (along the layers). Figure 4.35 shows that the

correlation length in the longitudinal direction (X-direction) is about 40

cm.

Clearly, the variograms are able to characterize the spatial

structure of the heterogeneity in the property of interest, in this case, the

x-ray absorption coefficient of a sandpack and a Berea sandstone.

Page 413: +Peters Ekwere j. - Petrophysics

4-51

Page 414: +Peters Ekwere j. - Petrophysics

4–52

Figure 4.33. Vertical and cross-sectional slices of CT image of dry Berea sandstone (Peters and Afzal, 1992).

Figure 4.34. Experimental variograms for Berea sandstone in the transeverse directions (Y, Z) (Peters and Afzal, 1992).

Page 415: +Peters Ekwere j. - Petrophysics

4-53

Figure 4.35. Experimental variogram for Berea sandstone in the longitudinal direction (X) (Peters and Afzal, 1992).

4.3.2. Covariance (Autocovariance) Function Definition

For sample data, the covariance function at a lag distance h is

defined as

( )( ) ( ) ( ) ( )

1h

x x x h x hC h

N

⎡Φ − Φ ⎤ ⎡Φ + − Φ + ⎤⎣ ⎦ ⎣ ⎦=

∑ (4.25)

The following observations can be made about the covariance function:

1. The covariance function gives the strength of the linear

relationship between Φ(x) and Φ (x+h).

2. The covariance function can be positive or negative. Recall that the

variogram function was non-negative.

Page 416: +Peters Ekwere j. - Petrophysics

4–54

3. Because the covariance function depends on the units of

measurement of Φ (x), it is not always possible to tell from the

magnitude of the covariance if the linear correlation is strong or

weak. A better measure of the strength of the linear correlation is

a dimensionless covariance function known as correlation

coefficient to be discussed next.

4. The table of data used to calculate the variogram can also be used

to calculate the covariance function. The algorithm for computing

the covariance function for unequally spaced data is similar to that

for computing the variogram. In this case, three one-dimensional

arrays should be used to store hij, iΦ and jΦ as the locations are

visited.

5. It will be shown that for a stationary random field, the variogram

and the covariance function are related by the equation

( ) ( ) ( )0h C C hγ = − (4.26)

where C(0) is the covariance at a zero lag distance and is equal the

variance of the data, s2. The fact that C(0) is the variance of the

data set is apparent from the entries in Table 4.3 for h = 0. The

variance of the data is given by

( )( ) ( ) ( ) ( ) ( ) ( ) 2

201 1

x x x x x xC s

N N

⎡Φ − Φ ⎤ ⎡Φ − Φ ⎤ ⎡Φ − Φ ⎤⎣ ⎦ ⎣ ⎦ ⎣ ⎦= = =

− −

∑ ∑ (4.27)

Thus, like the variogram, the covariance function also has hidden

in it the usual measure of variability by ordinary statistics.

Page 417: +Peters Ekwere j. - Petrophysics

4-55

Figure 4.36 shows a comparison of the covariance functions for

reservoirs A and B. The covariance function for reservoir A is periodic

indicating the random spatial distribution of the property in reservoir A

whereas the covariance function for reservoir B is smooth and orderly,

indicating the continuous nature of the spatial distribution of the

property in reservoir B. These are the same features observed in the

variograms. In fact, a comparison of Figures 4.36 and 4.6 shows that the

covariance function is roughly the variogram turned up-side-down about

the h-axis in the spirit of Eq.(4.26). This observation is more apparent in

Figures 4.37 and 4.38, which compare the variogram and the covariance

function for each reservoir. Thus, like the variogram, the covariance

function captures the correlation structure of the spatial distribution of

the property of interest and can be used for the purpose of estimation

instead of the variogram.

Figure 4.36. A comparison of the covariance functions for reservoirs A and B.

Page 418: +Peters Ekwere j. - Petrophysics

4–56

Figure 4.37. A comparison of the variogram and the covariance function for reservoir A.

Page 419: +Peters Ekwere j. - Petrophysics

4-57

Figure 4.38. A comparison of the variogram and the covariance function for reservoir B.

Physical Meaning of Covariance Function

The covariance function measures the covariation of Φ(x) and

Φ(x+h) about their respective means. Figure 4.39 shows the scatter plot

for the covariation of ( ) ( )x h x h⎡Φ + − Φ + ⎤⎣ ⎦ versus ( ) ( )x x⎡Φ − Φ ⎤⎣ ⎦ at a fixed

lag distance h. The sum of the shaded areas divided by (Nh-1) is the

covariance function at that lag distance. Because the areas can be

positive or negative depending on which quadrant the scatter plot data

fall in, the covariance can be positive or negative. If the scatter plots are

concentrated along the axis AB in Figure 4.40, then all the areas will be

positive and the covariance will show a strong positive correlation. If the

scatter plots are concentrated along the axis CD, then all the areas are

negative and the covariance will show a strong negative correlation.

Page 420: +Peters Ekwere j. - Petrophysics

4–58

Figure 4.39. Scatter plot of covariation at a fixed lag distance.

Page 421: +Peters Ekwere j. - Petrophysics

4-59

Figure 4.40. Patterns of covariation at a fixed lag distance.

Figure 4.41 shows the scatter plot of the covariation for reservoir A at a

lag distance of 1 unit. The data are concentrated in the quadrants in

which all the areas are negative. Therefore, the covariance function will

be negative at this lag distance as can be confirmed in Figure 4.36 or

4.37.

Page 422: +Peters Ekwere j. - Petrophysics

4–60

Figure 4.41. Covariation for reservoir A at a lag distance of 1 unit.

A cross covariance function can be computed between two different

variables such as permeability and porosity. In that case, the cross

covariance is defined as

( )( )( )

,1

i ik kC k

N

φ φφ

− −=

∑ (4.28)

4.3.3. Correlation Coefficient Function (Autocorrelation Function) Definition

The correlation coefficient function is the dimensionless version of

the covariance function and is defined as

Page 423: +Peters Ekwere j. - Petrophysics

4-61

( ) ( )( ) ( )x x h

C hh

s sρ

Φ Φ +

= (4.29)

where ( )xsΦ is the standard deviation of ( )xΦ and ( )x hsΦ + is the standard

deviation of ( )x hΦ + .

The following observations can be made about the correlation

coefficient function.

1. Like the covariance function, the correlation coefficient function

gives the strength of the linear relationship between Φ(x) and

Φ(x+h).

2. The correlation coefficient function is dimensionless and can vary

in value from –1 to +1 as may be seen in Figure 4.42, which shows

the correlation coefficient function for reservoir A.

Figure 4.42. Correlation coefficient function for reservoir A.

Page 424: +Peters Ekwere j. - Petrophysics

4–62

3. The magnitude of the correlation coefficient function is a measure

of the strength of the linear relationship between Φ(x) and Φ (x+h).

4. The sign of the correlation coefficient indicates the nature of the

correlation. If low Φ (x) are paired with low Φ (x+h) or high Φ (x) are

paired with high Φ (x+h), the correlation coefficient will be positive.

If low Φ (x) are paired with high Φ (x+h) or high Φ (x) are paired

with low Φ (x+h), the correlation coefficient will be negative.

5. If Φ (x) and Φ (x+h) are independent, then the correlation

coefficient (and the covariance) will be zero.

6. A zero correlation coefficient (or covariance) does not necessarily

mean that Φ (x) and Φ (x+h) are independent. They could be related

as a quadratic or other nonlinear function. Remember that the

correlation coefficient is a test for a linear relationship not a

quadratic relationship.

A cross correlation function between two variables such as

permeability and porosity can be computed as

( ) ( ),,

k

C kk

s sφ

φρ φ = (4.30)

4.4 PROBABILITY DISTRIBUTIONS

There are several probability distributions in statistics, but only

the normal and the log normal distributions are presented in this section

because they are the two distributions most commonly observed in

petrophysical properties. Based on core analyses and field

Page 425: +Peters Ekwere j. - Petrophysics

4-63

measurements, it has been observed that porosity tends to follow a

normal distribution (see Figures 2.9 and 2.10) whereas permeability

tends to follow a log normal distribution.

4.4.1 Normal (Gaussian) Distribution

The probability density function for a normal distribution, also

commonly known as the Gaussian, is given by

( ) ( )2

2

1 exp for 22

xP x x

μσσ π

⎡ ⎤−= − − ∞ ≤ ≤ ∞⎢ ⎥

⎢ ⎥⎣ ⎦ (4.31)

where x is a random variable and μ is the mean, σ2 is the variance and σ

is the standard deviation of the distribution. The distribution is

characterized by two parameters, the mean (μ) and the standard

deviation (σ). Figure 4.43 shows a normal distribution with μ = 5 and

standard σ = 4. The distribution is centered on the mean and is

symmetric about the mean. For this distribution, all the measures of the

central tendency of the distribution (mean, median and mode) have the

same numerical value. Note that the distribution extends from −∞ to +∞ .

Using the transformation

xz μσ−

= (4.32)

all normal distributions can be can standardized as

( )21 exp for

22zP z z

π⎛ ⎞

= − − ∞ ≤ ≤ +∞⎜ ⎟⎝ ⎠

(4.33)

Page 426: +Peters Ekwere j. - Petrophysics

4–64

where z is a random variable with μ = 0 and σ = 1. Figure 4.44 shows a

standard normal distribution with the horizontal axis calibrated in units

Figure 4.43. Normal (Gaussian) distribution with μ = 5 and σ = 4.

of standard deviation. Also shown are the areas under the curve for

certain ranges of the horizontal axis. For example, about 68% of the data

are contained within one standard deviation below and above the mean.

Thus, for a normal distribution, the following are true

( ) 0.6826P z σ≤ = (4.34)

( )1.96 0.9500P z σ≤ = (4.35)

( )2 0.9544P z σ≤ = (4.36)

Page 427: +Peters Ekwere j. - Petrophysics

4-65

( )3 0.9973P z σ≤ = (4.37)

( ) 0.1588P z σ−∞ ≤ ≤ − = (4.38)

( ) 0.8413P z σ−∞ ≤ ≤ + = (4.39)

Figure 4.44. Standard normal distribution with baseline in standard deviation units.

The cumulative distribution function for a normal distribution is

defined as

( ) ( )2

2

1 exp22

x xF x dx

μσσ π −∞

⎡ ⎤−= −⎢ ⎥

⎢ ⎥⎣ ⎦∫ (4.40)

Page 428: +Peters Ekwere j. - Petrophysics

4–66

Figure 4.45 shows the cumulative distribution plot on linear graph

paper. When potted on a special probability graph paper, the cumulative

distribution function plots as a straight line as shown in Figure 4.46.

Figure 4.45. Cumulative normal distribution, mean μ and standard deviation σ.

The cumulative distribution function is difficult to compute because the

integration called for in Eq.(4.40) cannot be performed analytically. Only

numerical integration is possible. It is not particularly convenient to

integrate any function numerically from to +−∞ ∞ . To overcome this

difficulty, we undertake a transformation that allows the integration of

Eq.(4.40) using the complementary error function, which is tabulated in

mathematical handbooks.

Page 429: +Peters Ekwere j. - Petrophysics

4-67

Figure 4.46. Cumulative normal distribution plotted on probability paper.

Page 430: +Peters Ekwere j. - Petrophysics

4–68

Let

2

xu μσ

−= (4.41)

2

dxduσ

= (4.42)

Substituting Eqs.(4.41) and (4.42) into (4.40) gives the cumulative

distribution function as

( ) 21 u uF u e duπ

−∞= ∫ (4.43)

The function 2ue

π

is symmetric about u = 0. Therefore,

2 21 11

u u u

ue du e du

π π∞− −

−∞= −∫ ∫ (4.44)

Substituting Eq.(4.44) into (4.43) and rearranging gives

( ) 21 222

u

uF u e du

π∞ −⎛ ⎞= −⎜ ⎟

⎝ ⎠∫ (4.45)

Eq.(4.45) can be written as

( ) ( )1 22

F u erfc u= −⎡ ⎤⎣ ⎦ (4.46)

where

Page 431: +Peters Ekwere j. - Petrophysics

4-69

( ) 22 u

uerfc u e du

π∞ −= ∫ (4.47)

The function, erfc(u), is the complementary error function, which arises

frequently in the solutions of certain partial differential equations of

mathematical physics such as the diffusivity equation, heat conduction

equation, diffusion equation and convection-dispersion equation. The

error function, erf(u), is defined as

( ) 2

0

2 u uerf u e duπ

−= ∫ (4.48)

Thus,

( ) ( )1erf u erfc u= − (4.49)

Table 4.6 gives the error and the complementary error functions whereas

Figure 4.48 shows graphs of the two functions. Using the information in

Table 4.6, the cumulative distribution function F(x) can be computed

with Eq.(4.46) instead of Eq.(4.40) and plotted against x. Figure 4.48

shows the F(x) obtained from such a computation for the normal

distribution of Figure 4.43.

Table 4.6. Error and Complementary Error Functions

x erf(x) erfc(x)

0.00 0 1

0.05 0.05637

2

0.94362

8

0.10 0.11246

3

0.88753

7

Useful

relationships

0.15 0.16799 0.83200

Page 432: +Peters Ekwere j. - Petrophysics

4–70

6 4

0.20 0.22270

3

0.77729

7

erfc(-x) = 1 + erf(x)

0.25 0.27632

6

0.72367

4

0.30 0.32862

7

0.67137

3

erf(-x) = -erf(x)

0.35 0.37938

2

0.62061

8

0.40 0.42839

2

0.57160

8

erfc(x) = 1 - erf(x)

0.45 0.47548

2

0.52451

8

0.50 0.52050

0

0.47950

0

0.55 0.56332

3

0.43667

7

0.60 0.60385

6

0.39614

4

0.65 0.64202

9

0.35797

1

0.70 0.67780

1

0.32219

9

0.75 0.71115

6

0.28884

4

0.80 0.74210

1

0.25789

9

0.85 0.77066

8

0.22933

2

0.90 0.79690

8

0.20309

2

0.95 0.82089

1

0.17910

9

1.00 0.84270 0.15729

Page 433: +Peters Ekwere j. - Petrophysics

4-71

1 9

1.10 0.88020

5

0.11979

5

1.20 0.91031

4

0.08968

6

1.30 0.93400

8

0.06599

2

1.40 0.95228

5

0.04771

5

1.50 0.96610

5

0.03389

5

1.60 0.97634

8

0.02365

2

1.70 0.98379

0

0.01621

0

1.80 0.98909

1

0.01090

9

1.90 0.99279

0

0.00721

0

2.00 0.99532

2

0.00467

8

2.10 0.99702

1

0.00297

9

2.20 0.99813

7

0.00186

3

2.30 0.99885

7

0.00114

3

2.40 0.99931

1

0.00068

9

2.50 0.99959

3

0.00040

7

2.60 0.99976

4

0.00023

6

2.70 0.99986 0.00013

Page 434: +Peters Ekwere j. - Petrophysics

4–72

6 4

2.80 0.99992

5

0.00007

5

2.90 0.99995

9

0.00004

1

3.00 0.99997

8

0.00002

2

Figure 4.47. Graphs of error and complementary error functions.

Page 435: +Peters Ekwere j. - Petrophysics

4-73

Figure 4.48. Cumulative distribution function calculated with the complementary error function for the normal distribution of Figure 4.43.

Generation of Random Deviates from a Normal (Gaussian) Distribution

Sometimes, we wish to generate sample data from a normal

distribution. Two methods are presented for generating such data. The

first method is based on the Central Limit Theorem (CLT) and the second

method is the Box Muller method.

The Central Limit Theorem shows that the mean of a group of

numbers drawn randomly from any distribution tends to a normal

(Gaussian) distribution as the number of means increases. Thus, if we

calculate many times the sums of N variates drawn from a uniform

distribution, we should expect the sums to fall into a truncated normal

(Gaussian) distribution bounded by 0 and N, with a mean value of N/2. If

we generate N values of a random number ri from a uniform distribution

( )0 1ir≤ ≤ and calculate

Page 436: +Peters Ekwere j. - Petrophysics

4–74

1 2

N

ii

Nz r=

= −∑ (4.50)

the random variable z will be drawn from an approximately Gaussian

distribution with a mean μ and a standard deviation σ given by

0μ = (4.51)

12Nσ = (4.52)

The choice of N = 12 is particularly convenient because z will be a normal

(Gaussian) distribution with a mean of 0 and a standard deviation of 1.

Such a random variate can be used to draw a random sample, x, from a

normal (Gaussian) distribution of mean μ and standard deviation σ as

x zμ σ= + (4.53)

Eqs.(4.51) and (4.52) can be proved as follows. For a uniform

distribution of r between a and b, the mean or expectation of the

distribution is given by

[ ]2

b

a

r a bE r drb a

μ += = =

−∫ (4.54)

The variance or second moment is given by

( ) ( )( ) [ ]( )222 2Var r E r E r E r E rσ ⎡ ⎤ ⎡ ⎤= = − = −⎣ ⎦⎣ ⎦ (4.55)

But

Page 437: +Peters Ekwere j. - Petrophysics

4-75

2 2 2

2

3b

a

r a ab bE r drb a

+ +⎡ ⎤ = =⎣ ⎦ −∫ (4.56)

Substituting Eqs.(4.54) and (4.56) into (4.55) gives

( ) ( ) ( )2 22 22

3 4 12a b b aa ab bVar rσ

+ −+ += = − = (4.57)

The mean of the variable z is given by

[ ] [ ] [ ] [ ]1 21

...2 2

i N

i Ni

N NE z E r E r E r E r Eμ=

=

⎡ ⎤ ⎡ ⎤= = − = + + + −⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦∑ (4.58)

For the case of a = 0 and b = 1, Eq.(4.54) gives E[r]=1/2. Thus,

substituting this into Eq.(5.58) gives

[ ] 1 1 1... 02 2 2 2 2 2

N N NE zμ = = + + + − = − = (4.59)

The variance of z is given by

( )21 2 ...

2NNVar z Var r r rσ ⎡ ⎤= = + + + −⎢ ⎥⎣ ⎦

(4.60)

Since r1, r2, …, rN are independent, then

( ) ( ) ( ) ( )21 2 ...

2NNVar z Var r Var r Var r Varσ ⎛ ⎞= = + + + − ⎜ ⎟

⎝ ⎠ (4.61)

Substituting Eq.(4.57) with a = 0 and b = 1 into Eq.(4.61) gives

( )2 1 1 1... 012 12 12 12

NVar zσ = = + + + − = (4.62)

Page 438: +Peters Ekwere j. - Petrophysics

4–76

The standard deviation for z is

12Nσ = (4.63)

Eqs.(4.59) and (4.63) are the same as Eqs.(4.51) and (4.52).

The method of Box and Muller for generating normal deviates from

a uniform random number generator is as follows. Select two random

numbers r1 and r2 from a uniform distribution ( )0 1ir≤ ≤ . Calculate

( )1 1 22 ln cos 2z r rπ= − (4.64)

( )2 1 22 ln sin 2z r rπ= − (4.65)

It can be shown that z1 and z2 are random deviates from a normal

(Gaussian) distribution with a mean of 0 and a standard deviation of 1.

The deviates can be used to draw two random variates (x1 and x2) with

mean μ and standard deviation σ as

1 1x zμ σ= + (4.66)

2 2x zμ σ= + (4.67)

4.4.2 Log Normal Distribution

The probability density function for a log normal distribution is

given by

Page 439: +Peters Ekwere j. - Petrophysics

4-77

( ) ( )2

2

ln1 1 exp 22x

P xx

μσσ π

⎡ ⎤−= −⎢ ⎥

⎢ ⎥⎣ ⎦ (4.68)

where x is a random variable and μ is the mean, σ2 is the variance and σ

is the standard deviation of lnx. The mean, τ, and the variance, ω2, of the

log normal distribution are related to the parameters of lnx as

2

exp2

στ μ⎛ ⎞

= +⎜ ⎟⎝ ⎠

(4.69)

( ) ( )22 21 exp 2eσω μ σ= − + (4.70)

It should be noted that x has a log normal distribution whereas lnx has a

normal distribution.

Figure 4.49 shows a log normal distribution. The distribution is

positively skewed with a long tail towards the high values of x. Note that

x is always positive. Because of the high values in the tail, the mean is

larger than the median, which in turn is larger than the mode or

geometric mean. Note that in this case, the geometric mean is a better

measure of central tendency than the mean or the median.

Page 440: +Peters Ekwere j. - Petrophysics

4–78

Figure 4.49. Log normal distribution.

Figure 4.50 shows the cumulative distribution function for a log

normal distribution plotted on a log normal probability graph paper. On

this scale, the log normal distribution plots as a straight line. Such a plot

was used in Figure 4.1 to determine the Dykstra-Parsons coefficient of

permeability variation on the assumption the data were drawn from a log

normal distribution.

Page 441: +Peters Ekwere j. - Petrophysics

4-79

Figure 4.50. Cumulative log normal distribution plotted on log normal probability paper.

4.5 ESTIMATION

4.5.1 Introduction

Estimation is one of the major applications of geostatistics, the

other being conditional simulation. The objective is to estimate the

Page 442: +Peters Ekwere j. - Petrophysics

4–80

variable of interest at a location xo for which no datum has been

measured, using the data measured at other locations xi distributed in

space. Suppose you are given some petrophysical data from a linear

reservoir as shown in Figure 4.51 and requested to estimate the values at

locations 3, 5 and 8 at which no measurements were made. What would

you do? Most professionals would construct the diagram shown

Figure 4.51. Sample data from a linear reservoir.

in Figure 4.52 to estimate * *3 5and Φ Φ but would be at a loss about how to

estimate *8 Φ , where the asterisk is used in the symbols to distinguish

the estimated from the measured values. In fact, some would claim that

there was not enough information to estimate *8Φ because no data were

measured beyond location 8.

Page 443: +Peters Ekwere j. - Petrophysics

4-81

Figure 4.52. Estimation of unmeasured values by linear interpolation.

Let us write the equation for the estimator *3Φ . From linear

interpolation

( )* 3 23 2 4 2

4 2

x xx x

⎛ ⎞−Φ = Φ + Φ − Φ⎜ ⎟−⎝ ⎠

(4.71)

Let

3 2 234

4 2 42

x x hx x h

λ⎛ ⎞−

= =⎜ ⎟−⎝ ⎠ (4.72)

where hij are lag distances. Substituting Eq.(4.72) into (4.71) and

rearranging gives the estimator as

Page 444: +Peters Ekwere j. - Petrophysics

4–82

*3 2 2 4 4λ λΦ = Φ + Φ (4.73)

where

2 41λ λ= − (4.74)

or

2 4 1λ λ+ = (4.75)

We can easily generalize the estimator of Eq.(4.73) and the constraint of

Eq.(4.75) as follows:

*

1

i N

o i ii

λ=

=

Φ = Φ∑ (4.76)

subject to the constraint

1

1N

ii

λ=

=∑ (4.77)

In Eq.(4.76), the subscript o has been used to indicate the location at

which an estimate is to be made and N could be the entire available data

set or a subset of the available data depending on the structure of the

heterogeneity. Let us now use Eqs.(4.76) and (4.77) to estimate * * *3 5 8, and Φ Φ Φ . For *

3Φ , λ2 = λ4 = 0.5 and

( )( ) ( )( )*3 0.50 30 0.50 50 40Φ = + =

Similarly

Page 445: +Peters Ekwere j. - Petrophysics

4-83

( )( ) ( )( )*5 0.50 50 0.50 20 35Φ = + =

To estimate *8Φ , we proceed as follows. The weights, λi, in Eq.(4.76)

should be assigned to each measured value used in the estimation

depending on how close the measured value is to the point xo at which

the estimate is to be made. Unfortunately, there is no mathematical

expression for closeness. We can measure how far away each sample is

from the point of estimation but we cannot measure how close it is to the

point of estimation. Let us measure the lag distances of the measured

data from the point of estimation as

2

4

6

642

= 12

o

o

o

x x xx x xx x xTotal x

− = Δ− = Δ

− = ΔΔ

Based on the above lag distances, it is clear that 6Φ should have the most

weight, followed by 4Φ and 2Φ . We can adopt several different strategies

to accomplish this. Here is one strategy. Let us make the weights

proportional to the inverse lag distances as follows:

2

4

6

0.16676

0.25004

0.50002 = 0.9167Total

αλ α

αλ α

αλ α

α

= =

= =

= =

where α is to be chosen to satisfy the constraint of Eq.(4.77). In this case,

α = 1.0909. Thus,

Page 446: +Peters Ekwere j. - Petrophysics

4–84

( )( )

( )( )

( )( )

2

4

6

0.1667 0.1667 1.0909 0.18196

0.2500 0.2500 1.0909 0.27274

0.5000 0.5000 1.0909 0.54552 = 1.0000Total

αλ α

αλ α

αλ α

= = = =

= = = =

= = = =

Therefore,

( )( ) ( )( ) ( )( )*8 0.1819 30 0.2727 50 0.5455 20 30Φ = + + =

It should be noted that in estimating * *3 5and Φ Φ , only a subset of the

measured data was used whereas in estimating *8Φ , all the measured

data were used. However, there was no good reason to use the subset of

the measured data to estimate * *3 5and Φ Φ . All the measured data could

have been used to estimate * *3 5and Φ Φ in the manner similar to the

estimation of *8Φ .

Having completed the estimation, the next question is what is the

reliability of the estimates? What are the error bounds on the estimates?

Most professionals would give up at this point. This is where geostatistics

can help. Before presenting the geostatistical estimation technique

known as ordinary kriging, it is worthwhile to mention the major

limitation of the estimation undertaken above. It is that the structure of

the heterogeneity as represented by either the variogram, the covariance

function or the correlation coefficient function has not be taken into

account in the estimation.

Geostatistics uses a probabilistic framework to make estimations

and in so doing provides estimates that honor the correlation structure of

Page 447: +Peters Ekwere j. - Petrophysics

4-85

the heterogeneity. It also provides quantitative assessment of the

reliability of the estimates. Let us demonstrate a simple geostatistical

estimation using a probabilistic framework. Suppose we wish to estimate

the missing datum in Figure 4.53. Proceeding in the spirit of Eqs.(4.76)

and (4.47), let us use the first "ring" of data around the missing datum

(or rather first "square" of data) to obtain an estimate as

*1 2 3 4 5 6 7 819 18 17 15 18 15 15 18o λ λ λ λ λ λ λ λΦ = + + + + + + + (4.78)

The error of the estimation, which is unknown, is given by

8

*1 2 3 4 5 6 7 8

119 18 17 15 18 15 15 18

i

o o o o ii

e λ λ λ λ λ λ λ λ λ=

=

= Φ − Φ = + + + + + + + − Φ ∑ (4.79)

where oΦ is the unknown missing datum. Notice that the constraint of

Eq.(4.47) has been used in Eq.(4.79). The error can be expanded as

( ) ( ) ( ) ( )( ) ( ) ( ) ( )

*1 2 3 4

5 6 7 8

19 18 17 15

18 15 15 18

o o o o o o o

o o o o

e λ λ λ λ

λ λ λ λ

= Φ − Φ = − Φ + − Φ + − Φ + − Φ

+ − Φ + − Φ + − Φ + − Φ (4.80)

Page 448: +Peters Ekwere j. - Petrophysics

4–86

Figure 4.53. Missing data from a 2D reservoir.

To simplify the calculations, let us choose λ2 = λ3 = λ4 = λ5 = λ6 = λ7 = λ8 =

0 and λ1 = 1. This is not the optimum choice of the weights and will not

lead to the best estimate. However, because the weights sum to 1, we

have not violated the estimation model given by Eqs.(4.46) and (4.47) and

as such the estimation is legitimate. With this simplification, our

estimate is

* 19oΦ = (4.81)

and the error is now given by

Page 449: +Peters Ekwere j. - Petrophysics

4-87

( )* 19o o o oe = Φ − Φ = − Φ (4.82)

Although the error given by Eq.(4.82) is unknown, there is enough

information in the available data for us to determine its statistics. If we

accept the premise that the error is a function of the lag distance, 2xΔ ,

then we can look at the field of data in the SW-NE direction to find

similar errors and from them determine the statistics of our error. There

are in fact 47 such errors in the data set. They are tabulated in column 4

of Table 4.7.

Table 4.7. Errors in the SW-NE Direction of the Data Set at a Lag

Distance of 2xΔ .

1 2 3 4 5

Error

#

Φ(x) Φ(x+h) Φ(x+h - Φ(x) [Φ(x+h)-Φ(x)]2

1 23 22 -1 1

2 22 20 -2 4

3 20 19 -1 1

4 21 17 -4 16

5 17 17 0 0

6 17 14 -3 9

7 19 15 -4 16

8 15 18 3 9

9 18 20 2 4

10 20 16 -4 16

11 18 18 0 0

12 19 14 -5 25

13 14 19 5 25

14 17 16 -1 1

15 16 15 -1 1

16 15 18 3 9

Page 450: +Peters Ekwere j. - Petrophysics

4–88

17 18 18 0 0

18 18 23 5 25

19 23 16 -7 49

20 15 14 -1 1

21 14 10 -4 16

22 10 15 5 25

23 15 20 5 25

24 20 25 5 25

25 25 21 -4 16

26 21 14 -7 49

27 13 10 -3 9

28 10 16 6 36

29 16 18 2 4

30 18 20 2 4

31 20 20 0 0

32 20 17 -3 9

33 11 13 2 4

34 13 14 1 1

35 14 23 9 81

36 23 18 -5 25

37 18 19 1 1

38 10 13 3 9

39 13 18 5 25

40 18 22 4 16

41 22 13 -9 81

42 17 15 -2 4

43 15 20 5 25

44 20 20 0 0

45 16 14 -2 4

46 14 18 4 16

47 15 17 2 4

Mean 0.128

Variance 15.766 15.447

Page 451: +Peters Ekwere j. - Petrophysics

4-89

Std Dev 3.971 3.930

Semivarianc

e

7.723

As shown at the bottom of Table 4.7, the mean, variance and standard

deviation of the error are

2

0.12815.766

3.791

μσσ

=

==

(4.83)

If we assume that the error is normally distributed, then we are 95%

confident that the error of our estimation is given by

1.96 1.96oeμ σ μ σ− ≤ ≤ + (4.84)

Thus,

( ) ( ) ( )0.128 1.96 3.791 19 0.128 1.96 3.791o− ≤ − Φ ≤ + (4.85)

( )7.272 19 7.558o− ≤ − Φ ≤ (4.86)

or

11.442 26.272o≤ Φ ≤ (4.87)

Given our knowledge of the missing datum, is the statement of Eq.(4.87)

true? You bet. Of course, it would have been nice if the 95% confidence

interval was narrower than that given by Eq.(4.87). It would have been

narrower if we had chosen the weights to minimize the variance of the

Page 452: +Peters Ekwere j. - Petrophysics

4–90

error of Eq.(4.80) instead of choosing them arbitrarily as we did to

simplify the calculations. Figure 4.54 shows the histogram of the errors

of Table 4.7, which indicates that the assumption of a normal

distribution of the errors in the calculation of the 95% confidence

interval is reasonable.

Figure 4.54. Histogram of errors in the SW-NE direction at h = 2xΔ .

It should be noted that the data in columns 2 to 5 of Table 4.7 are

the same data needed to calculate the semivariance at a lag distance of

( )2h x= Δ in the SW-NE direction. From the calculations, we have

( )2

2 7.7232

xσ γ= Δ = (4.88)

Page 453: +Peters Ekwere j. - Petrophysics

4-91

Thus, geostatistics gives the mean, variance and standard deviation of

the error as

2

0.00015.447

3.930

μσσ

=

==

(4.89)

These statistics of the error are very similar to those given in Eq.(4.83).

Applying Eq.(4.89) gives the following 95% confidence interval:

( ) ( ) ( )0.000 1.96 3.930 19 0.000 1.96 3.930o− ≤ − Φ ≤ + (4.90)

( )7.703 19 7.703o− ≤ − Φ ≤ (4.91)

or

11.297 26.703o≤ Φ ≤ (4.92)

The result given in Eq.(4.92) is essentially the same as that given by

Eq.(4.87). Therefore, in geostatistics, the variogram (or the covariance

function) is used to calculate the error variance. Having now

demonstrated a simple geostatistical estimation that gave both the

estimate and the confidence limits for the estimate, we are ready to

formally derive ordinary kriging equations for geostatistical estimation.

4.5.2 Ordinary Kriging Equations

The objective is to estimate an unknown variable at the location x0 at

which no measurement has been made using the measured data at

locations xi as shown in Figure 4.55. In geostatistical estimation, the

variable of interest is treated as a stationary random function with a

Page 454: +Peters Ekwere j. - Petrophysics

4–92

normal probability distribution. If the measured data do not exhibit a

normal distribution, they must first be transformed into a normal

distribution before proceeding further. Such a transformation can always

be done by using the cumulative distribution function of the data and

the cumulative distribution function of a standard normal distribution

with a mean of zero and a standard deviation of 1.0 as shown

schematically in Figure 4.56. After the estimation, the estimated value is

transformed back to its original distribution by using the same two

cumulative distribution functions in reverse order.

Figure 4.55. Schematic showing the locations sample data and the location at which an estimation is to be made.

Figure 4.57 shows two random functions ( )AZ x and ( )BZ x together

with their probability density functions at four locations, x1, …, x4. The

mean of random function ( )AZ x is constant throughout the field. As a

result, ( )AZ x is described as a stationary random function. By contrast,

Page 455: +Peters Ekwere j. - Petrophysics

4-93

the mean of random function ( )BZ x is not constant throughout the field.

In fact, it increases as x increases. As a result, random function ( )BZ x is

described as a non-stationary random function.

Figure 4.56. Transformation of sample data to a standard normal distribution.

The kriging estimate is calculated as the weighted average of the

measured data as in Eq.(4.76). The challenge is to determine the weights

λi in such a way as to obtain the best estimate in some sense. The

criteria used in ordinary kriging to obtain the best estimate are (1) the

estimate should be unbiased and (2) the estimation error should have

Page 456: +Peters Ekwere j. - Petrophysics

4–94

minimum variance. Thus, ordinary kriging is a Best Linear Unbiased

Estimator (BLUE).

Figure 4.57. Stationary ( ( )AZ x ) and non-stationary ( ( )BZ x ) random functions.

Derivation in Terms of the Covariance Function

The estimated value at the unsampled location xo is given by

Page 457: +Peters Ekwere j. - Petrophysics

4-95

*

1( ) ( )

N

o i ii

Z x Z xλ=

= ∑ (4.93)

where Z*(xo) is the estimated value at location xo, Z(xi) are the measured

data, Z(x) is the assumed random function model, λi are the weights to

be determined and xi are the locations where the variable of interest has

been measured (i.e., the data). Note that in this model, the Z(xi) are just

samples drawn from the random function Z(x). Moreover, Z(x) is a

stationary random function with a constant mean, m, which is

independent of the locations xo and xi. Also, Eq.(4.93) is valid in 1D, 2D

or 3D.

Let the true but unknown value at xo be Z(xo). The estimation error

is given by

*( ) ( ) ( )o o oe x Z x Z x= − (4.94)

This error is a random variable with a probability distribution. In order to

obtain an unbiased estimate, on average, the mean estimation error

must be zero. Thus, for an unbiased estimate, the expectation of the

error must be zero:

[ ] *( ) ( ) ( ) 0o o oE e x E Z x Z x⎡ ⎤= − =⎣ ⎦ (4.95)

What is an unbiased or biased estimator? Figure 4.58 shows three

estimators for a petrophysical property whose value is know to μ.

Estimator 1 is unbiased but not very precise. Estimator 2 is unbiased

Page 458: +Peters Ekwere j. - Petrophysics

4–96

and more precise than estimator 1. Estimator 3 is biased and more

precise than estimator 1.

Figure 4.58. Biased and unbiased estimators.

Substituting Eq.(4.93) into Eq.(4.94) gives the condition for an

unbiased estimate as

1

( ) ( ) 0N

i i oi

E Z x Z xλ=

⎡ ⎤− =⎢ ⎥

⎣ ⎦∑ (4.96)

Changing the order of the summation and the expectation gives

Page 459: +Peters Ekwere j. - Petrophysics

4-97

[ ] [ ]1

( ) ( ) 0N

i i oi

E Z x E Z xλ=

− =∑ (4.97)

But the expectations of the random variables, Z(xi) and Z(xo) are equal

and is given by the mean of Z(x), m, which is a constant for the

stationary random function model used in the estimation. Eq.(4.97) can

now be written as

1 1

1 0N N

i ii i

m m mλ λ= =

⎛ ⎞− = − =⎜ ⎟

⎝ ⎠∑ ∑ (4.98)

The condition for an unbiased estimate is therefore given by

1

1N

ii

λ=

=∑ (4.99)

which is the same as Eq.(4.77).

The error variance is given by

( )22 *( ) ( )e o oE Z x Z xσ ⎡ ⎤= −⎢ ⎥⎣ ⎦ (4.100)

Substituting Eq.(4.93) into Eq.(4.100) gives the error variance as

2

2

1

( ) ( )N

e i i oi

E Z x Z xσ λ=

⎡ ⎤⎛ ⎞= −⎢ ⎥⎜ ⎟

⎝ ⎠⎢ ⎥⎣ ⎦∑ (4.101)

Let us add and subtract the mean, m, from the inner bracket in

Eq.(4.101) and make use of Eq.(4.99) (∑λi = 1) to obtain

Page 460: +Peters Ekwere j. - Petrophysics

4–98

( )2

2

1

( ) ( ( ) )N

e i i oi

E Z x m Z x mσ λ=

⎡ ⎤⎛ ⎞= − − −⎢ ⎥⎜ ⎟

⎝ ⎠⎢ ⎥⎣ ⎦∑ (4.102)

Expanding the right side of Eq.(4.102) gives the estimation variance as

( ) ( )2

22

1 1

( ) ( ) 2 ( ( ) )( ( ) )N N

e i i o i i oi i

E Z x m Z x m Z x m Z x mσ λ λ= =

⎡ ⎤⎛ ⎞= − + − − − −⎢ ⎥⎜ ⎟

⎝ ⎠⎢ ⎥⎣ ⎦∑ ∑ (4.103)

Algebraically, the square of the simple summation in Eq.(4.103) can be

rewritten in terms of a double summation to obtain

( ) ( )( )22

1 1 1

( ( ) )( ( ) ) ( ) 2 ( ) ( )N N N

e i j i j o i i oi j i

E Z x m Z x m Z x m Z x m Z x mσ λ λ λ= = =

⎡ ⎤= − − + − − − −⎢ ⎥

⎣ ⎦∑ ∑ ∑

…………….(4.104)

Changing the order of the summation and the expectation in Eq.(4.104)

gives

( )

( )( )

22

1 1

1

( ( ) )( ( ) ) ( )

2 ( ) ( )

N N

e i j i j oi j

N

i i oi

E Z x m Z x m E Z x m

E Z x m Z x m

σ λ λ

λ

= =

=

⎡ ⎤⎡ ⎤= − − + −⎣ ⎦ ⎣ ⎦

⎡ ⎤− − −⎣ ⎦

∑ ∑

∑ (4.105)

The expectations in Eq.(4.105) can be expressed in terms of the

covariance function as follows:

2

1 1 1

( ) (0) 2 ( )N N N

e i j ij i ioi j i

C h C C hσ λ λ λ= = =

= + −∑ ∑ ∑ (4.106)

Page 461: +Peters Ekwere j. - Petrophysics

4-99

where C(hij) is the covariance function, hij is the lag distance between

locations i and j, and hio is the lag distance between locations o and i.

Recall that the location of the point at which an estimate is to be made is

o and the locations of the data to be used in the estimation are at points

i. Eq.(4.106) can be rearranged as

2

1 1 1

(0) ( ) 2 ( )N N N

e i j ij i ioi j i

C C h C hσ λ λ λ= = =

= + −∑ ∑ ∑ (4.107)

But C(0), the covariance function at a lag distance of 0, is the same as

the variance of the data, σ2. Therefore, Eq.(4.107) can be written in the

following final form:

2 2

1 1 1( ) 2 ( )

N N N

e i j ij i ioi j i

C h C hσ σ λ λ λ= = =

= + −∑ ∑ ∑ (4.108)

The estimation problem now boils down to determining the values

of λi, which will minimize the estimation variance given by Eq.(4.108)

subject to the unbiasedness constraint given by Eq.(4.99). This is a

classical problem of optimization, which can be solved by the method of

Lagrange multipliers.

Derivation in Terms of the Variogram

Eq.(4.108) can be derived in terms of the variogram instead of the

covariance function. To do so, we first derive the relationship between

the variogram and the covariance function for a stationary random

function, Z(x). By definition, the variogram is given by

( )212( ) ( ) ( )h E Z x Z x hγ ⎡ ⎤= − +⎣ ⎦ (4.109)

Page 462: +Peters Ekwere j. - Petrophysics

4–100

Adding and subtracting the mean of the random function from the inner

bracket of Eq.(4.109) gives

( ) ( )( )212( ) ( ) ( )h E Z x m Z x h mγ ⎡ ⎤= − − + −⎢ ⎥⎣ ⎦

(4.110)

Expanding the right side of Eq.(4.110) gives

( ) ( ) ( )( )2 212( ) ( ) ( ) 2 ( ) ( )h E Z x m Z x h m Z x m Z x h mγ ⎡ ⎤= − + + − − − + −⎣ ⎦ (4.111)

Eq.(4.111) can be rewritten as

( )( ) ( )( )( )( )

1 12 2

22

( ) ( ) ( ) ( ) ( )

( ) ( )

h E Z x m Z x m E Z x h m Z x h m

E Z x h m Z x m

γ ⎡ ⎤ ⎡ ⎤= − − + + − + −⎣ ⎦ ⎣ ⎦⎡ ⎤− + − −⎣ ⎦

(4.112)

The expectations in Eq.(4.112) can be expressed in terms of the

covariance function as follows:

1 12 2( ) (0) (0) ( )h C C C hγ = + − (4.113)

Finally, Eq.(4.113) yields the required relationship as

( ) (0) ( )h C C hγ = − (4.114)

which is the same as Eq.(4.26). For large lag distances, Eq.(4.114)

becomes

( ) (0) ( )C Cγ ∞ = − ∞ (4.115)

For a stationary random function, γ(∞) is the sill of the variogram, C(∞) is

0 because there is no more correlation beyond the correlation length and

Page 463: +Peters Ekwere j. - Petrophysics

4-101

C(0) is the variance of the data (σ2) as previously stated. Substituting

these facts into Eq.(4.115) gives

2( ) (0) sill of the variogramCγ σ∞ = = = (4.116)

Figure 4.59 compares the variogram and the covariance function for a

stationary random function.

Figure 4.59. A comparison of the variogram and the covariance function for a stationary random function.

We are now ready to rewrite the error variance equation, Eq.(4.108), in

terms of the variogram. From Eq.(4.114), we have

( ) (0) ( )ij ijC h C hγ= − (4.117)

( ) (0) ( )io ioC h C hγ= − (4.118)

Page 464: +Peters Ekwere j. - Petrophysics

4–102

Substituting Eqs.(4.117) and (4.118) into Eq.(4.108) gives the error

variance as

[ ]2

1 1 1(0) (0) ( ) 2 (0) ( )

N N N

e i j ij i ioi j i

C C h C hσ λ λ γ λ γ= = =

⎡ ⎤= + − − −⎣ ⎦∑ ∑ ∑ (4.119)

Eq.(4.119) can be rearranged as

2

1 1 1 1 1 1(0) (0) ( ) 2 (0) 2 ( )

N N N N N N

e i j i j ij i i ioi j i j i i

C C h C hσ λ λ λ λ γ λ λ γ= = = = = =

= + − − +∑ ∑ ∑ ∑ ∑ ∑ (4.120)

Eq.(4.120) can be simplified to

2

1 1 1

(0) (0) ( ) 2 (0) 2 ( )N N N

e i j ij i ioi j i

C C h C hσ λ λ γ λ γ= = =

= + − − +∑ ∑ ∑ (4.121)

Further simplification of Eq.(4.121) gives the error variance in terms of

the variogram in final form as

2

1 1 1

( ) 2 ( )N N N

e i j ij i ioi j i

h hσ λ λ γ λ γ= = =

= − +∑ ∑ ∑ (4.122)

Solution of the Kriging Equations in terms of the Covariance Function

The problem to be solved is to choose the weights (λi) to minimize

the error variance given by Eq.(4.108) subject to the constraint given by

Eq.(4.99). The problem statement is as follows:

2 2

1 1 1

( ) ( ) 2 ( )N N N

e i i j ij i ioi j i

Minimize C h C hσ λ σ λ λ λ= = =

= + −∑ ∑ ∑ (4.123)

Page 465: +Peters Ekwere j. - Petrophysics

4-103

1

to 1 0N

ii

Subject λ=

⎛ ⎞− =⎜ ⎟

⎝ ⎠∑ (4.124)

This problem can be solved by the method of Lagrange multipliers. The

solution steps are as follows:

1. Convert all constraints into equality constraints.

2. Multiply each equality constraint by a new variable μi, where μi is

the Lagrange multiplier for the ith constraint.

3. Add (or subtract) the resulting constraint equations to the original

objective function to obtain the Lagrangian function, L. This step

relaxes the constraint and converts the constrained optimization to

an unconstrained optimization.

4. Differentiate the Lagrangian function and equate to zero to

determine the stationary points, which constitute the required

solution.

Applications of steps 1 to 3 to the problem at hand will result in the

following Lagrangian function:

2

1 1 1 1

( , ) ( ) 2 ( ) 2 1N N N N

i i j ij i io ii j i i

L C h C hλ μ σ λ λ λ μ λ= = = =

⎛ ⎞= + − + −⎜ ⎟

⎝ ⎠∑ ∑ ∑ ∑ (4.125)

Consider the case for N = 2. The Lagrangian can be expanded to obtain

2 21 2 1 11 1 2 12 2 1 21

22 22 1 10 2 20

1 2

( , , ) ( ) ( ) ( )

( ) 2 ( ) 2 ( ) 2 2 2

L C h C h C h

C h C h C h

λ λ μ σ λ λ λ λ λ

λ λ λλ μ λ μ μ

= + + +

+ − −+ + −

(4.126)

Differentiating Eq.(4.126) and equating to zero will give the following

linear simultaneous equations:

Page 466: +Peters Ekwere j. - Petrophysics

4–104

1 11 2 12 2 21 101

2 ( ) ( ) ( ) 2 ( ) 2 0L C h C h C h C hλ λ λ μλ

∂= + + − + =

∂ (4.127)

1 12 1 21 2 22 202

( ) ( ) 2 ( ) 2 ( ) 2 0L C h C h C h C hλ λ λ μλ

∂= + + − + =

∂ (4.128)

1 22 2 2 0L λ λμ

∂= + − =

∂ (4.129)

Eqs.(4.127) to (4.129) can be rewritten in matrix form as

11 12 21 1 10

12 21 22 2 20

2 ( ) ( ) ( ) 2 2 ( )( ) ( ) 2 ( ) 2 2 ( )

2 2 0 2

C h C h C h C hC h C h C h C h

λλμ

+⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥+ =⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.130)

Because C(h12) = C(h21), Eq.(4.130) can be rewritten as

11 12 1 10

21 22 2 20

2 ( ) 2 ( ) 2 2 ( )2 ( ) 2 ( ) 2 2 ( )

2 2 0 2

C h C h C hC h C h C h

λλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.131)

Upon dividing by 2, Eq.(4.131) simplifies to

11 12 1 10

21 22 2 20

( ) ( ) 1 ( )( ) ( ) 1 ( )1 1 0 1

C h C h C hC h C h C h

λλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.132)

Eq.(4.132) can be solved by standard techniques such as matrix

inversion or Gaussian elimination.

Page 467: +Peters Ekwere j. - Petrophysics

4-105

After solving Eq.(4.132), λ1, λ2 and μ can be substituted into

Eq.(4.123) to calculate the minimum estimation error variance. Let us

derive an algebraic expression for the minimum error variance.

Expanding Eq.(4.123) for N = 2 gives the error variance as

2 2 21 2 1 11 1 2 12 2 1 21

22 22 1 10 2 20

( , , ) ( ) ( ) ( )

( ) 2 ( ) 2 ( )

e C h C h C h

C h C h C h

σ λ λ μ σ λ λ λ λ λ

λ λ λ

= + + +

+ − − (4.133)

Eq.(4.133) can be rearranged as

2 21 2 1 1 11 2 12 2 1 21 2 22

1 10 2 20

( , , ) ( ) ( ) ( ) ( )

2 ( ) 2 ( )

e C h C h C h C h

C h C h

σ λ λ μ σ λ λ λ λ λ λ

λ λ

⎡ ⎤ ⎡ ⎤= + + + +⎣ ⎦ ⎣ ⎦− − (4.134)

From Eq.(4.132), it can be seen that

1 11 2 12 10( ) ( ) ( )

C h C h C hλ λ μ+ = − (4.135)

and

2 21 2 22 20( ) ( ) ( )

C h C h C hλ λ μ+ = − (4.136)

Substituting Eqs.(4.135) and (4.136) into Eq.(4.134) gives the minimum

estimation error variance as

( ) ( )2 2

min 1 10 2 20

1 10 2 20

( ) ( ) 2 ( ) 2 ( )

e C h C hC h C h

σ σ λ μ λ μ

λ λ

= + − + −

− − (4.137)

Page 468: +Peters Ekwere j. - Petrophysics

4–106

which simplifies to

22 2min

1

( )

e i ioi

C hσ σ μ λ=

= − − ∑ (4.138)

Eq.(4.138) can be used to calculate the minimum error variance directly

instead of substituting λ1, λ2 and μ into Eq.(4.123) to calculate the

minimum estimation error variance. Eq.(4.132) and (4.138) can easily be

generalized to any value of N. For example, for N = 4, Eqs.(4.132) and

(4.138) will become

11 12 13 14 1 10

21 22 23 24 2 20

31 32 33 34 3 30

41 42 43 44 4 40

( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )1 1 1 1 0 1

C h C h C h C h C hC h C h C h C h C hC h C h C h C h C hC h C h C h C h C h

λλλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.139)

and

42 2min

1

( )

e i ioi

C hσ σ μ λ=

= − − ∑ (4.140)

For any value of N, Eq.(4.132) can be generalized to the following N+1

linear simultaneous equations:

1

1

( ) ( ) for i = 1, 2, ..., N

1

N

j ij ioj

N

jj

C h C hλ μ

λ

=

=

+ =

=

∑ (4.141)

Page 469: +Peters Ekwere j. - Petrophysics

4-107

The minimum estimation error variance can be generalized to

2 2min

1

( )

N

e i ioi

C hσ σ μ λ=

= − − ∑ (4.142)

The following observations can be made about the ordinary kriging

model in terms of the covariance function.

1. In order to set up the simultaneous equations to be solved, one

must first compute the lag distance matrix, hij. For N = 4, the

lag distance matrix will look like this:

11 12 13 14 12 13 14

21 22 23 24 21 23 24

31 32 33 34 31 32 34

41 42 43 44 41 42 43

00

00

ij

h h h h h h hh h h h h h h

hh h h h h h hh h h h h h h

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎡ ⎤ = =⎣ ⎦ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

(4.143)

2. Except for the one zero at the bottom corner, the diagonal

entries of the matrix equation to be solved are given by

211 22( ) ( ) ... ( ) (0)NNC h C h C h C σ= = = = (4.144)

3. All the off diagonal entries of the matrix equation are less than

the diagonal entries. Thus, the matrix is diagonally dominant.

This is a desirable structure for solving the system of linear

simultaneous equations.

4. The matrix to be inverted is full. Therefore, a lot of calculations

could be involved, depending on the value of N.

5. The solution λ1, λ2, …, λN and μ depends only on the spatial

coordinates of the data and not on the values of the data.

Page 470: +Peters Ekwere j. - Petrophysics

4–108

6. The estimation error variance depends only on the spatial

coordinates of the data and not on the values of the data.

7. Kriging is an exact estimator and will return the measured

value if it is applied to the location for which datum was

measured.

Solution of the Kriging Equations in terms of the Variogram

In this case, the problem statement is

2

1 1 1 ( ) ( ) 2 ( )

N N N

e i i j ij i ioi j i

Minimize h hσ λ λ λ γ λ γ= = =

= − +∑ ∑ ∑ (4.145)

1

to 1 0N

ii

Subject λ=

⎛ ⎞− =⎜ ⎟

⎝ ⎠∑ (4.146)

The Lagrangian is given by

1 1 1 1

( , ) ( ) 2 ( ) 2 1N N N n

i i j ij i io ii j i i

L h hλ μ λ λ γ λ γ μ λ= = = =

⎛ ⎞= − + + −⎜ ⎟

⎝ ⎠∑ ∑ ∑ ∑ (4.147)

For the case of N = 2, the Lagrangian can be expanded to obtain

21 2 1 11 1 2 12 2 1 21

22 22 1 10 2 20

1 2

( , , ) ( ) ( ) ( )

( ) 2 ( ) 2 ( ) 2 2 2

L h h h

h h h

λ λ μ λ γ λ λ γ λ λ γ

λ γ λ γ λ γλ μ λ μ μ

= − − −

− + ++ + −

(4.148)

Differentiating the Lagrangian and equating to zero will give the following

linear simultaneous equations:

Page 471: +Peters Ekwere j. - Petrophysics

4-109

1 11 2 12 2 21 101

2 ( ) ( ) ( ) 2 ( ) 2 0L h h h hλ γ λ γ λ γ γ μλ

∂= − − − + + =

∂ (4.149)

1 12 1 21 2 22 202

( ) ( ) 2 ( ) 2 ( ) 2 0L h h h hλ γ λ γ λ γ γ μλ

∂= − − − + + =

∂ (4.150)

1 22 2 2 0L λ λμ

∂= + − =

∂ (4.151)

Eqs.(4.149) to (4.151) can be rewritten in matrix form as

[ ]

[ ]11 12 21 1 10

12 21 22 2 20

2 ( ) ( ) ( ) 2 2 ( )( ) ( ) 2 ( ) 2 2 ( )

2 2 0 2

h h h hh h h h

γ γ γ λ γγ γ γ λ γ

μ

⎡ ⎤− − + −⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥− + − = −⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦⎣ ⎦

(4.152)

Upon dividing by 2 and noting that γ(h12) = γ(h21), Eq.(4.152) can be

rewritten as

11 12 1 10

21 22 2 20

( ) ( ) 1 ( )( ) ( ) 1 ( )1 1 0 1

h h hh h h

γ γ λ γγ γ λ γ

μ

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥− =⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.153)

After solving Eq.(4.153), λ1, λ2 and μ can be substituted into

Eq.(4.145) to calculate the minimum estimation error variance. Let us

derive an algebraic expression for the minimum error variance.

Expanding Eq.(4.145) for N = 2 gives the error variance as

2 21 2 1 11 1 2 12 2 1 21

22 22 1 10 2 20

( , , ) ( ) ( ) ( )

( ) 2 ( ) 2 ( )

e h h h

h h h

σ λ λ μ λ γ λ λ γ λ λ γ

λ γ λ γ λ γ

= − − −

− + + (4.154)

Page 472: +Peters Ekwere j. - Petrophysics

4–110

Eq.(4.154) can be rearranged as

21 2 1 1 11 2 12 2 1 21 2 22

1 10 2 20

( , , ) ( ) ( ) ( ) ( )

2 ( ) 2 ( )

e h h h h

h h

σ λ λ μ λ λ γ λ γ λ λ γ λ γ

λ γ λ γ

⎡ ⎤ ⎡ ⎤= − + − +⎣ ⎦ ⎣ ⎦+ + (4.155)

From Eq.(4.153), it can be seen that

1 11 2 12 10( ) ( ) ( )

h h hλ γ λ γ γ μ+ = + (4.156)

and

2 21 2 22 20( ) ( ) ( )

h h hλ γ λ γ γ μ+ = + (4.157)

Substituting Eqs.(4.156) and (4.157) into Eq.(4.155) gives the minimum

estimation error variance as

( ) ( )2

min 1 10 2 20

1 10 2 20

( ) ( ) 2 ( ) 2 ( )

e h hh h

σ λ γ μ λ γ μλ γ λ γ

= − + − +

+ + (4.158)

which simplifies to

22min

1

( )

e i ioi

hσ μ λ γ=

= − + ∑ (4.159)

Eqs.(4.153) and (4.159) can easily be generalized to any value of N. For

example, for N = 4, Eqs.(4.153) and (4.159) will become

Page 473: +Peters Ekwere j. - Petrophysics

4-111

11 12 13 14 1 10

21 22 23 24 2 20

31 32 33 34 3 30

41 42 43 44 4 40

( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )( ) ( ) ( ) ( ) 1 ( )1 1 1 1 0 1

h h h h hh h h h hh h h h hh h h h h

γ γ γ γ λ γγ γ γ γ λ γγ γ γ γ λ γγ γ γ γ λ γ

μ

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=−⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.160)

and

42min

1

( )

e i ioi

hσ μ λ γ=

= − + ∑ (4.161)

For any value of N, Eq.(4.153) can be generalized to the following N+1

linear simultaneous equations:

1

1

( ) ( ) for i = 1, 2, ..., N

1

N

j ij ioj

N

jj

h hλ γ μ γ

λ

=

=

− =

=

∑ (4.162)

The minimum estimation error variance can be generalized to

2min

1

( )

N

e i ioi

hσ μ λ γ=

= − + ∑ (4.163)

The following observations can be made about the kriging model in

terms of the variogam.

1. The diagonal entries of the matrix equation to be solved are given

by

11 22( ) ( ) ... ( ) 0NNh h hγ γ γ= = = (4.164)

Page 474: +Peters Ekwere j. - Petrophysics

4–112

All the diagonal entries of the matrix equation are zero. This is an

undesirable structure for the solution of the system of linear

simultaneous equations. This is why the formulation and solution

of the estimation problem in terms of the covariance function is

usually preferred over the formulation in terms of the variogram.

2. Most of the entries in the matrix equation are numbers computed

from the variogram. It is essential that these numbers be

consistent and well behaved for the system of equations to have a

solution. Inconsistent numbers from an experimental variogram

could lead to a system of equations without a solution. This is why

a well behaved theoretical variogram model is usually fitted to the

experimental variogram and the theoretical variogram model is

then used instead of the experimental variogram for the estimation

calculations.

Example 4.1

The porosities at locations 1 and 4 in a linear reservoir have been

measured as shown Figure 4.60. The locations are evenly spaced 10

meters apart. The variogram for the porosity distribution in this reservoir

is shown in Figure 4.61. We are required to estimate the porosity at

location 3 at which no measurement was made.

Page 475: +Peters Ekwere j. - Petrophysics

4-113

Figure 4.60. Measured porosity values and their locations.

1. Calculate the best estimate of the porosity at location 3 based on

the available information and determine the 95% confidence

interval assuming a normal distribution.

2. Assuming a stationary random function model, carefully sketch on

Figure 4.60 the covariance function for the porosity distribution for

this reservoir showing important features of your sketch.

Page 476: +Peters Ekwere j. - Petrophysics

4–114

Figure 4.61. Variogram for Example 4.1.

3. Show that kriging is an exact interpolator. An exact interpolator

will return the measured value if it is applied at a location

containing a measured value. In other words, if kriging is applied

to estimate the porosity at location 1, it should return a value of

10% with a minimum error variance of 0.

Solution to Example 4.1

The best estimate of the porosity at location 3 can be obtained using

ordinary kriging. The kriging equations to be solved can be written by

inpection in terms of the covariance function as

11 12 1 10

21 22 2 20

11

1 1 0 1

C C CC C C

λλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.165)

Page 477: +Peters Ekwere j. - Petrophysics

4-115

Let us adopt the following subscripts: Location 1 = 1, location 3 = 0 and

location 4 = 3. The matrix of lag distances for location 3 is (subscript 0) is

11 12 10

21 22 20

0 30 2030 0 10

h h hmeters

h h h⎡ ⎤ ⎡ ⎤

=⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

(4.166)

For a stationary random function, the variogram and the covariance

function are related by Eq.(4.114). Therefore, the covariances needed in

Eq.(4.165) can easily be computed from the variogram. Thus, the matrix

equation to be solved is

1

2

5 2 1 32 5 1 41 1 0 1

λλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.167)

Eq.(4.167) can easily be solved by elimination to obtain

1

2

1323

0

λ

λ

μ

=

=

=

The estimated porosity is given by

( ) ( )* *0 3 1 4

1 2 1 210 25 20%3 3 3 3

φ φ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞= = Φ + Φ = + =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠

The minimum error variance is computed from Eq.(4.140) as

22 2min

1

1 2 4( ) 5 0 3 43 3 3

e i ioi

C h x xσ σ μ λ=

⎛ ⎞= − − = − − + =⎜ ⎟⎝ ⎠

Page 478: +Peters Ekwere j. - Petrophysics

4–116

The minimum standard deviation is

min4 1.15473

eσ = =

The 95% confidence interval is given by

( )( ) ( ) ( )( )30 1.96 1.1547 20 0 1.96 1.1547φ+ ≤ − ≤ +

( )32.26 20 2.26φ≤ − ≤

or

317.74 22.26 or 20% 2.26%φ≤ ≤ ±

The problem could also be solved in terms of the variogram. In this case,

the corresponding equations are

11 12 1 10

21 22 2 20

11

1 1 0 1

γ γ λ γγ γ λ γ

μ

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥− =⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.168)

1

2

0 3 1 23 0 1 11 1 0 1

λλμ

−⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥− =⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.169)

The solution of Eq.(4.169) gives

Page 479: +Peters Ekwere j. - Petrophysics

4-117

1

2

1323

0

λ

λ

μ

=

=

=

The estimated porosity is given by

( ) ( )* *0 3 1 4

1 2 1 210 25 20%3 3 3 3

φ φ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞= = Φ + Φ = + =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠

The minimum error variance is obtained from Eq.(4.159) as

22min

1

1 2 4( ) 0 2 13 3 3

e i ioi

h x xσ μ λ γ=

⎛ ⎞= − + = − + + =⎜ ⎟⎝ ⎠

The minimum standard deviation is

min4 1.15473

eσ = =

The 95% confidence interval is given by

( )( ) ( ) ( )( )30 1.96 1.1547 20 0 1.96 1.1547φ+ ≤ − ≤ +

( )32.26 20 2.26φ≤ − ≤

or

317.74 22.26 or 20% 2.26%φ≤ ≤ ±

Page 480: +Peters Ekwere j. - Petrophysics

4–118

Thus, as expected, the kriging equations based on the covariance

function and the variogram give the same results. The sketch of the

covariance function on Figure 4.61 is left as an exercise for the reader.

To demonstrate that kriging is an exact interpolator, we solve the

kriging equation at location 1, which contains a sample datum. We

should recover the sample value of 10% with an error variance of 0. The

matrix of lag distances is

11 12 10

21 22 20

0 30 030 0 30

h h hmeters

h h h⎡ ⎤ ⎡ ⎤

=⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

(4.170)

Substituting these numbers into Eq.(4.165) gives the matrix equation to

be solved as

1

2

5 2 1 52 5 1 21 1 0 1

λλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.171)

Eq.(4.171) can be solved to obtain

1

2

100

λλμ

===

( ) ( ) ( )( ) ( )( )* *0 1 1 41 0 1 10 0 25 10%φ φ= = Φ + Φ = + =

The minimum error variance is computed from Eq.(4.140) as

( )2

2 2min

1

( ) 5 0 1 5 0 2 0

e i ioi

C h x xσ σ μ λ=

= − − = − − + =∑

Page 481: +Peters Ekwere j. - Petrophysics

4-119

Therefore, kriging is an exact interpolator.

Example 4.2

Estimate the petrophysical property shown in Figure 4.51 at locations 1,

3, 5, 7 and 8 at which no measurements were made. The correlation

structure of the heterogeneous property is given by

( ) 0.3100 hC h e−= (4.172)

Solution to Example 4.2

To begin the calculations, we must determine the order in which the

estimates will be made. This order is determined by a random drawing.

Using a random number generator for integer values from 1 to 8, it has

been determined that the order for the estimations is 8, 1, 5, 7 and 3.

After estimating *8Φ , this value is treated as a known sample and is used

along with the measured data for the subsequent estimations. Thus, the

number of equations to be solved increases as the estimation progresses.

We begin the calculations by visiting location (node) 8. In order to

generate a compact matrix equation, we renumber the known values as

shown in Figure 4.62. Next, we construct the lag distance matrix for

location 8 as shown in Table 4.8. The entries in Table 4.8 should be read

as follows. The indices 1, 2, 3 represent the locations of the renumbered

known values. The index 0 represents the location at which an estimate

is to be made. The other entries are lag distances. For example, h10 is the

lag distance (6Δx) from the point of estimation to the renumbered sample

1, h20 is the lag distance (4Δx) from the point of estimation to the

renumbered sample 2 and h30 is the lag distance (2Δx) from the point of

estimation to the renumbered sample 3. These are the lag distances

needed to construct the right hand side vector of the system of equations

Page 482: +Peters Ekwere j. - Petrophysics

4–120

to be solved. Therefore, they have been transferred into the last column

of the table under the heading hi0. The other entries in the table should

be self explanatory. For example, h11 is the lag distance between the

renumbered sample 1 and itself, which is zero.

Figure 4.62. Renumbered sample values for the estimation at location 8.

Table 4.8. Lag Distance Matrix hij for Location 8

1 2 3

0 6Δx 4Δx 2Δx hi0

1 0Δx 2Δx 4Δx 6Δx

2 2Δx 0Δx 2Δx 4Δx

3 4Δx 2Δx 0Δx 2Δx

The matrix equation to be solved is

Page 483: +Peters Ekwere j. - Petrophysics

4-121

11 12 13 1 10

21 22 23 2 20

31 32 33 3 30

111

1 1 1 0 1

C C C CC C C CC C C C

λλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.173)

In this problem, Δx = 1 unit. Using the matrix of lag distances in Table

4.8 and Eq.(4.172), the entries of the covariance matrix of Eq.(4.173) can

easily be computed. For example, C12 is given by

( ) ( ) 0.3 212 12 2 100 54.8812C h C e−= = =

The results of the other calculations are shown in Eq.(4.174).

1

2

3

100.0000 54.8812 30.1194 1 16.529954.8812 100.0000 54.8812 1 30.119430.1194 54.8812 100.0000 1 54.8812

1 1 1 0 1

λλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.174)

Eq.(4.174) can be solved by any standard method of solving a system of

linear simultaneous equations, for example, by matrix inversion. The

inverse matrix for Eq.(4.174) is

1

0.0117 0.0090 0.0026 0.40800.0090 0.0181 0.0090 0.18410.0026 0.0090 0.0117 0.4080

0.4080 0.1841 0.4080 63.1862

ijC−

− −⎡ ⎤⎢ ⎥− −⎢ ⎥=⎢ ⎥− −⎢ ⎥−⎣ ⎦

(4.175)

Multiplying the right hand side of Eq.(4.174) into the inverse matrix of

Eq.(4.175) gives the solution vector as

Page 484: +Peters Ekwere j. - Petrophysics

4–122

1

2

3

0.18410.08300.732928.5088

λλλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦

(4.176)

The kriged value is calculated as

( )( ) ( )( ) ( )( )* *0 8 0.1841 30 0.0830 50 0.7329 20 24.3321Φ = Φ = + + =

The minimum error variance (estimation variance) is computed from

Eq.(4.140) as

( )

( )

32 2min

1

( ) 100 28.5089

0.1841 16.5299 0.0830 30.1194 0.7329 54.8812 82.7434

e i ioi

C h

x x x

σ σ μ λ=

= − − = − −

− + + =

The minimum standard deviation (estimation standard deviation) is

min 82.7435 9.0963

eσ = =

The 95% confidence interval is given by

( )( )*8 24.3321 1.96 9.0963 24.33 17.83Φ = ± = ±

The estimated value at location 8 is then added to the sample data set in

preparation for the estimation at location 1 as shown in Figure 4.63.

Page 485: +Peters Ekwere j. - Petrophysics

4-123

Figure 4.63. Sample data set for the estimation at location 1.

Next, we visit location 1 and construct the lag distance matrix shown in

Table 4.9.

Table 4.9. Lag Distance Matrix hij for Location 1

1 2 3 4

0 1Δx 3Δx 5Δx 7Δx hi0

1 0Δx 2Δx 4Δx 6Δx 1Δx

2 2Δx 0Δx 2Δx 4Δx 3Δx

3 4Δx 2Δx 0Δx 2Δx 5Δx

4 6Δx 4Δx 2Δx 0Δx 7Δx

Page 486: +Peters Ekwere j. - Petrophysics

4–124

The matrix equation to be solved is

11 12 13 14 1 10

21 22 23 24 2 20

31 32 33 34 3 30

41 42 43 44 4 40

1111

1 1 1 1 0 1

C C C C CC C C C CC C C C CC C C C C

λλλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.177)

Using the matrix of lag distances in Table 4.9 and Eq.(4.170), Eq.(4.177)

becomes

1

2

3

4

100.0000 54.8812 30.1194 16.5299 1 74.081854.8812 100.0000 54.8812 30.1194 1 40.657030.1194 54.8812 100.0000 54.8812 1 22.313016.5299 30.1194 54.8812 100.0000 1 12.2456

1 1 1 1 0 1

λλλλμ

⎡ ⎤ ⎡ ⎤ ⎡⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ =⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣

⎤⎥⎥

⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎦

(4.178)

The inverse matrix for Eq.(4.178) is

1

0.0121 0.0089 0.0010 0.0022 0.34450.0089 0.0182 0.0083 0.0010 0.15550.0010 0.0083 0.0182 0.0089 0.15550.0022 0.0010 0.0089 0.0121 0.3445

0.3445 0.1555 0.1555 0.3445 53.3636

ijC−

− − −⎡ ⎤⎢ ⎥− − −⎢ ⎥⎢ ⎥= − − −⎢ ⎥− − −⎢ ⎥⎢ ⎥−⎣ ⎦

(4.179)

Multiplying the right hand side of Eq.(4.178) into the inverse matrix of

Eq.(4.179) gives the solution vector as

1

2

3

4

0.83010.04030.04030.089313.8309

λλλλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦

(4.180)

Page 487: +Peters Ekwere j. - Petrophysics

4-125

The kriged value is calculated as

( )( ) ( )( ) ( )( ) ( )( )* *0 1 0.8301 30 0.0403 50 0.0403 20 0.0893 24.3321 29.8968Φ = Φ = + + + =

The estimation variance is computed from Eq.(4.140) as

( )

( )

32 2min

1

( ) 100 13.8309

0.8301 74.0819 0.0403 40.6570 0.0403 22.3130 0.0893 12.2456 48.7035

e i ioi

C h

x x x x

σ σ μ λ=

= − − = − −

− + + + =

The estimation standard deviation is

min 48.7035 6.9788

eσ = =

The 95% confidence interval is given by

( )( )*8 29.8968 1.96 6.9788 29.99 13.68Φ = ± = ±

The estimated value at location 1 is then added to the sample data set in

preparation for the estimation at location 5 as shown in Figure 4.64.

Next, we visit location 5 and construct the lag distance matrix shown in

Table 4.10.

Page 488: +Peters Ekwere j. - Petrophysics

4–126

Figure 4.64. Sample data set for the estimation at location 5.

Table 4.10. Lag Distance Matrix hij for Location 5

1 2 3 4 5

0 3Δx 1Δx 1Δx 3Δx 4Δx hi0

1 0Δx 2Δx 4Δx 6Δx 1Δx 3Δx

2 2Δx 0Δx 2Δx 4Δx 3Δx 1Δx

3 4Δx 2Δx 0Δx 2Δx 5Δx 1Δx

4 6Δx 4Δx 2Δx 0Δx 7Δx 3Δx

5 1Δx 3Δx 5Δx 7Δx 0Δx 4Δx

Page 489: +Peters Ekwere j. - Petrophysics

4-127

The matrix equation to be solved is

11 12 13 14 15 1 10

21 22 23 24 25 2 20

31 32 33 34 35 3 30

41 42 43 44 45 4 40

51 52 53 54 55 5 50

11111

1 1 1 1 1 0 1

C C C C C CC C C C C CC C C C C CC C C C C CC C C C C C

λλλλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥

=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦

(4.181)

Using the matrix of lag distances in Table 4.10 and Eq.(4.170), Eq.(4.181)

becomes

100.0000 54.8812 30.1194 16.5299 74.0818 154.8812 100.0000 54.8812 30.1194 40.6570 130.1194 54.8812 100.0000 54.8812 22.3130 116.5299 30.1194 54.8812 100.0000 12.2456 174.0818 40.6570 22.3130 12.2456 100.0000 1

1 1 1 1 1 0

⎡⎢⎢⎢⎢⎢⎢⎢⎣

1

2

3

4

5

40.657074.081874.081840.657030.1194

1

λλλλλμ

⎤ ⎡ ⎤ ⎡ ⎤⎥ ⎢ ⎥ ⎢ ⎥⎥ ⎢ ⎥ ⎢ ⎥⎥ ⎢ ⎥ ⎢ ⎥

=⎥ ⎢ ⎥ ⎢ ⎥⎥ ⎢ ⎥ ⎢ ⎥⎥ ⎢ ⎥ ⎢ ⎥⎥ ⎢ ⎥ ⎢ ⎥⎦ ⎣ ⎦ ⎣ ⎦

(4.182)

The inverse matrix for Eq.(4.182) is

1

0.0262 0.0082 0.0003 0.0007 0.0170 0.10880.0082 0.0182 0.0083 0.0009 0.0008 0.14400.0003 0.0083 0.0182 0.0088 0.0008 0.14400.0007 0.0009 0.0088 0.0122 0.0018 0.31920.0170 0.0008 0.0008 0.0018 0.0205

ijC−

− − − −− − − −− − − −

=− − − −− − − − 0.28400.1088 0.1440 0.1440 0.3192 0.2840 49.4359

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥

−⎣ ⎦

(4.183)

Multiplying the right hand side of Eq.(4.182) into the inverse matrix of

Eq.(4.183) gives the solution vector as

Page 490: +Peters Ekwere j. - Petrophysics

4–128

1

2

3

4

5

0.00470.48460.48460.01380.01232.1441

λλλλλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥

=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥

−⎣ ⎦ ⎣ ⎦

(4.184)

The kriged value is calculated as

( )( ) ( )( ) ( )( ) ( )( )( )( )

* *0 5 0.0047 30 0.4846 50 0.4846 20 0.0138 24.3321

0.0123 29.8968 34.7659

Φ = Φ = + + +

+ =

The estimator variance is computed from Eq.(4.140) as

( )

32 2min

1( ) 100

0.0047 40.6570 0.4846 74.0818 0.4846 74.0818 2.1441 29.2243

0.0138 40.6570 0.0123 30.1194

e i ioi

C h

x x xx x

σ σ μ λ=

= − − =

+ +⎛ ⎞− − − =⎜ ⎟+ +⎝ ⎠

The estimator standard deviation is

min 29.2243 5.4059

eσ = =

The 95% confidence interval is given by

( )( )*8 34.7659 1.96 5.4059 34.77 10.60Φ = ± = ±

The estimated value at location 5 is then added to the sample data set in

preparation for the estimation at location 7 as shown in Figure 4.65.

Page 491: +Peters Ekwere j. - Petrophysics

4-129

Figure 4.65. Sample data set for the estimation at location 7.

Next, we visit location 7 and construct the lag distance matrix shown in

Table 4.11.

Table 4.11. Lag Distance Matrix hij for Location 7

1 2 3 4 5 6

0 5Δx 3Δx 1Δx 1Δx 6Δx 2Δx hi0

1 0Δx 2Δx 4Δx 6Δx 1Δx 3Δx 5Δx

2 2Δx 0Δx 2Δx 4Δx 3Δx 1Δx 3Δx

3 4Δx 2Δx 0Δx 2Δx 5Δx 1Δx 1Δx

4 6Δx 4Δx 2Δx 0Δx 7Δx 3Δx 1Δx

5 1Δx 3Δx 5Δx 7Δx 0Δx 4Δx 6Δx

6 2Δx 1Δx 1Δx 3Δx 4Δx 0Δx 2Δx

Page 492: +Peters Ekwere j. - Petrophysics

4–130

The matrix equation to be solved is

11 12 13 14 15 16 1 10

21 22 23 24 25 26 2 20

31 32 33 34 35 36 3 30

41 42 43 44 45 46 4 40

51 52 53 54 55 56 5 50

61 62 63 64 65 66 6 60

111111

1 1 1 1 1 1 0 1

C C C C C C CC C C C C C CC C C C C C CC C C C C C CC C C C C C CC C C C C C C

λλλλλλμ

⎡ ⎤ ⎡ ⎤ ⎡⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ =⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣

⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎦

(4.185)

Using the matrix of lag distances in Table 4.11 and Eq.(4.170), Eq.(4.185)

becomes

100.0000 54.8812 30.1194 16.5299 74.0818 40.6570 154.8812 100.0000 54.8812 30.1194 40.6570 74.0818 130.1194 54.8812 100.0000 54.8812 22.3130 74.0818 116.5299 30.1194 54.8812 100.0000 12.2456 40.6570 174.0818 40.6570 22.3130 12.

1

2

3

4

5

6

22.313040.657074.081874.0818

2456 100.0000 30.1194 1 16.529954.8812 74.0818 74.0818 40.6570 30.1194 100.0000 1 54.8812

1 1 1 1 1 1 0 1

λλλλλλμ

⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ……... (4.186)

The inverse matrix for Eq.(4.186) is

1

0.0322 0.0023 0.0073 0.0006 0.0207 0.0159 0.09930.0023 0.0239 0.0008 0.0009 0.0044 0.0155 0.1348

0.0073 0.0008 0.0279 0.0087 0.0055 0.0202 0.13200.0006 0.0009 0.0087 0.0122 0.0019 0.0002 0.31910.0207

ijC−

− − − −− − − − −

− − − −= − − − − −

− −0.0044 0.0055 0.0019 0.0228 0.0097 0.28980.0159 0.0155 0.0202 0.0002 0.0097 0.0420 0.0251

0.0993 0.1348 0.1320 0.3191 0.2898 0.0251 49.4209

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥− −⎢ ⎥− − − −⎢ ⎥

⎢ ⎥−⎣ ⎦

(4.187)

Page 493: +Peters Ekwere j. - Petrophysics

4-131

Multiplying the right hand side of Eq.(4.186) into the inverse matrix of

Eq.(4.187) gives the solution vector as

1

2

3

4

5

6

0.00470.00580.48410.49210.01230.00092.1422

λλλλλλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎣ ⎦ ⎣ ⎦

(4.188)

The kriged value is calculated as

( )( ) ( )( ) ( )( ) ( )( )( )( ) ( )( )

* *0 7 0.0047 30 0.0058 50 0.4841 20 0.4921 24.3321

0.0123 29.8968 0.0009 34.7659 22.4886

Φ = Φ = + + +

+ + =

The estimation variance is computed from Eq.(4.140) as

( )

32 2min

1( ) 100

0.0047 22.3130 0.0058 40.6570 0.4841 74.0818 2.1422 29.2242

0.4921 74.0818 0.0123 16.5299 0.0009 54.8812

e i ioi

C h

x x xx x x

σ σ μ λ=

= − − =

+ +⎛ ⎞− − − =⎜ ⎟+ + +⎝ ⎠

The estimation standard deviation is

min 29.2242 5.4059

eσ = =

The 95% confidence interval is given by

( )( )*7 22.4886 1.96 5.4059 22.4886 10.60Φ = ± = ±

Page 494: +Peters Ekwere j. - Petrophysics

4–132

The estimated value at location 7 is then added to the sample data set in

preparation for the estimation at location 3 as shown in Figure 4.66.

Figure 4.66. Sample data set for the estimation at location 3.

Next, we visit location 3 and construct the lag distance matrix shown in

Table 4.12.

Table 4.12. Lag Distance Matrix hij for Location 3

1 2 3 4 5 6 7

0 1Δx 1Δx 3Δx 5Δx 2Δx 2Δx 4Δx hi0

1 0Δx 2Δx 4Δx 6Δx 1Δx 3Δx 3Δx 1Δx

2 2Δx 0Δx 2Δx 4Δx 3Δx 1Δx 1Δx 1Δx

3 4Δx 2Δx 0Δx 2Δx 5Δx 1Δx 1Δx 3Δx

4 6Δx 4Δx 2Δx 0Δx 7Δx 3Δx 3Δx 5Δx

5 1Δx 3Δx 5Δx 7Δx 0Δx 4Δx 4Δx 2Δx

6 2Δx 1Δx 1Δx 3Δx 4Δx 0Δx 0Δx 2Δx

7 2Δx 1Δx 1Δx 3Δx 4Δx 0Δx 0Δx 4Δx

Page 495: +Peters Ekwere j. - Petrophysics

4-133

The matrix equation to be solved is

11 12 13 14 15 16 17 1

21 22 23 24 25 26 27 2

31 32 33 34 35 36 37 3

41 42 43 44 45 46 47 4

51 52 53 54 55 56 57 5

61 62 63 64 65 66 67 6

71 72 73 74 75 76 77

1111111

1 1 1 1 1 1 1 0

C C C C C C CC C C C C C CC C C C C C CC C C C C C CC C C C C C CC C C C C C CC C C C C C C

λλλλλλ

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

10

20

30

40

50

60

7 70

1

CCCCCCCλ

μ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

(4.189)

Using the matrix of lag distances in Table 4.12 and Eq.(4.170), Eq.(4.189)

becomes

100.0000 54.8812 30.1194 16.5299 74.0818 40.6570 22.3130 154.8812 100.0000 54.8812 30.1194 40.6570 74.0818 40.6570 130.1194 54.8812 100.0000 54.8812 22.3130 74.0718 74.0818 116.5299 30.1194 54.8812 100.0000 12.2456 40.6570 74.0

1

2

3

4818 174.0818 40.6570 22.3130 12.2456 100.0000 30.1194 16.5299 154.8812 74.0818 74.0818 40.6570 30.1194 100.0000 54.8812 122.3130 40.6570 74.0818 74.0818 16.5299 54.8812 100.0000 1

1 1 1 1 1 1 1 0

λλλλλ

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

5

6

7

74.081874.081840.657022.313054.881254.881230.1194

1

λλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

…… (4.190)

The inverse matrix for Eq.(4.190) is

Page 496: +Peters Ekwere j. - Petrophysics

4–134

1

0.0322 0.0023 0.0074 0.0006 0.0207 0.0159 0.0001 0.09900.0023 0.0239 0.0007 0.0008 0.0044 0.0155 0.0002 0.1343

0.0074 0.0007 0.0359 0.0006 0.0053 0.0202 0.0166 0.09650.0006 0.0008 0.0006 0.0205 0.0017

ijC−

− − − − −− − − − − −

− − − − −− − − −

=0.0001 0.0168 0.2830

0.0207 0.0044 0.0053 0.0017 0.0228 0.0097 0.0004 0.28880.0159 0.0155 0.0202 0.0001 0.0097 0.0420 0.00004 0.02500.0001 0.0002 0.0166 0.0168 0.0004 0.00004 0.0342 0.0733

0.0990 0.1343 0.0965

− −− − − − −− − − − −− − − − − −

0.2830 0.2888 0.2306 0.0733 49.2637

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥

− −⎢ ⎥⎣ ⎦

(4.191)

Multiplying the right hand side of Eq.(4.190) into the inverse matrix of

Eq.(4.191) gives the solution vector as

1

2

3

4

5

6

7

0.48410.59720.11700.01510.01510.2325

0.00392.6342

λλλλλλλμ

⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥=⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥

−⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥

−⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦

(4.192)

The kriged value is calculated as

( )( ) ( )( ) ( )( ) ( )( )( )( ) ( )( ) ( )( )

* *0 3 0.4841 30 0.5972 50 0.1170 20 0.0151 24.3321

0.0151 29.8968 0.2325 34.7659 0.0039 22.4886 39.5490

Φ = Φ = + + +

+ − + =

The estimation variance is computed from Eq.(4.140) as

( )3

2 2min

1( ) 100 2.6342

0.4841 74.0818 0.5972 74.0818 0.1170 40.657029.2455

0.0.0151 22.3130 0.0151 54.8812 0.2325 54.8812 0.0039 30.1194

e i ioi

C h

x x xx x x x

σ σ μ λ=

= − − = − −

+ +⎛ ⎞− =⎜ ⎟+ + − +⎝ ⎠

The estimation standard deviation is

Page 497: +Peters Ekwere j. - Petrophysics

4-135

min 29.2455 5.4079

eσ = =

The 95% confidence interval is given by

( )( )*7 39.5490 1.96 5.4079 39.55 10.60Φ = ± = ±

Table 4.13 presents a summary of the results of the estimation. Figure

4.67 shows the estimated values along with the measured values. It

should be noted that kriging gives the estimated values of *3Φ and *

5Φ that

are essentially the same as those obtained from the linear interpolation

of Figure 4.52.

Table 4.13. Results of Estimations for Example 4.2.

Location iΦ *iΦ 2

mineσ mineσ 1.96 mineσ *min1.96i eσΦ − *

min1.96i eσΦ +

1 29.90 48.72 6.98 13.68 16.22 43.58

2 30

3 39.55 29.27 5.41 10.60 28.95 50.15

4 50

5 34.77 29.27 5.41 10.60 24.17 45.37

6 20

7 22.49 29.27 5.41 10.60 11.89 33.09

8 24.33 82.81 9.10 17.84 6.49 42.17

Page 498: +Peters Ekwere j. - Petrophysics

4–136

Figure 4.67. Graph of the measured and the kriged values for Example 4.2.

Using kriging, we are now able to estimate the petrophysical

properties for all the grid blocks of our reservoir simulation model that

honor the measured data and the correlation structure of the

heterogeneous properties. Further, we are able to estimate the

uncertainty associated with the estimates at each grid block. We have

made significant progress in dealing with the heterogeneities of our

reservoir rock.

We have studied ordinary kriging. There are other types of kriging

such as simple kriging, universal kriging, block kriging and ordinary

Page 499: +Peters Ekwere j. - Petrophysics

4-137

cokriging. These other types of kriging are usually covered in a

geostastistics course.

4.6 CONDITIONAL SIMULATION

4.6.1 Introduction

Kriging gives a smooth estimate because the estimate is a weighted

average of the sample data. Such a weighted average can never be larger

than the largest sample value nor can it be smaller than the smallest

sample value. Thus, kriging eliminates local variability. If such local

variability is important, then it can be incorporated into the estimated

values using conditional simulation. Conditional simulation also is

referred to as stochastic simulation or Monte Carlo simulation is the

other major application of geostatistics. The simulation is conditioned on

the measured data.

The idea behind conditional simulation is as follows. Each of the

estimates obtained from kriging was associated with an uncertainty in

the estimated value measured by the estimation variance or the

estimation standard deviation. Thus, the kriged estimate is a random

variable with a known variance or standard deviation. If the kriged value

comes from a normal distribution, then it is possible to draw a simulated

value from this distribution that is centered on the kriged value and has

a variance that is equal to the estimation variance and a standard

deviation that is equal to the estimation standard deviation.

4.6.2 Sequential Gaussian Simulation

The objective is to perform a stochastic simulation to estimate the

values of a petrophysical property in a heterogeneous reservoir at

Page 500: +Peters Ekwere j. - Petrophysics

4–138

locations for which no sample data have been measured. The simulated

values should honor the measured sample data and the correlation

structure of the heterogeneity as revealed by the analysis of the sample

data. Further, it should retain the local variability. Here is an algorithm

for performing such a stochastic simulation known as Sequential

Gaussian Simulation (SGS).

1. Select at random a location or node not yet simulated in the grid.

2. Use ordinary kriging to compute the local estimate at the node

along with the estimation variance and the estimation standard

deviation.

3. Draw a random value from a normal distribution with a mean

equal to the kriged estimate and a standard deviation equal to the

estimation standard deviation at that node. This is the simulated

value at that node. This step can be accomplished by using either

Eq.(4.50, (4.64) or (4.65) to draw a sample from a standard normal

distribution and then applying either Eq.(4.53), (4.66) or (4.67) to

compute the simulated value as

*si i ei izσΦ = Φ + (4.193)

where siΦ is the simulated value at location i, *iΦ is the kriged value

at location i, eiσ is the standard deviation or the square root of the

estimation variance at location i and iz is a standard normal

variate with μ = 0 and σ = 1 drawn for location i. Note that upon

moving to a new location, a new iz must be drawn for that

location.

Page 501: +Peters Ekwere j. - Petrophysics

4-139

4. Include the newly simulated value in the set of conditioning data

and use the expanded data set to simulate the value at the next

location.

5. Proceed to the next location, which of course was selected

randomly at the start, and repeat the calculations to simulate the

value at this location.

6. Repeat the calculations until all grid nodes have a simulated value.

7. This completes one realization of the simulation. Other realizations

can be obtained by repeating the calculations from step 1. To do

so, the order of the simulation is first determined using a random

number generator. This order should be different from all previous

orders. The calculations will then yield a new realization, whose

simulated values will be different from the previous realizations. If

there are N nodes to be simulated, there will be N! (N factorial)

possible realizations.

Example 4.3

Simulate values at the locations 1, 3, 5, 7 and 8 of Figure 4.51 at which

no samples were taken using Sequential Gaussian Simulation.

Solution to Example 4.3.

Let us simulate one realization. The first part of the calculations was

done in Example 4.2 in which kriged estimates were computed in the

following random order: 8, 1, 5, 7 and 3. The estimation variance and

standard deviations also were computed and presented in Table 4.13. To

simulate values at the five nodes, we draw five variates from a standard

normal distribution using Eq.(4.50, (4.64) or (4.65). For example, using

Page 502: +Peters Ekwere j. - Petrophysics

4–140

Eq.(4.64) the following five variates were drawn from a standard normal

distribution:

0.8821

-1.1679

0.6419

2.8546

0.91009

Using these numbers, we can simulate the values at nodes 8, 1, 5, 7 and

3 as follows:

( )( )* *8 8 8 8 24.33 9.10 0.8821 32.36s e zσΦ = Φ + = + =

( )( )* *1 1 1 1 29.90 6.98 1.1679 21.75s e zσΦ = Φ + = + − =

( )( )* *5 5 5 5 34.77 5.41 0.6419 38.24s e zσΦ = Φ + = + =

( )( )* *7 7 7 7 22.49 5.41 2.8546 37.92s e zσΦ = Φ + = + =

( )( )* *3 3 3 3 39.55 5.41 0.9009 44.42s e zσΦ = Φ + = + =

Figure 4.68 shows the measured values, kriged values and the simulated

values.

Page 503: +Peters Ekwere j. - Petrophysics

4-141

Figure 4.68. A comparison of the measured, kriged and simulated values for Example 4.3.

4.6.3 A Practical Application of Sequential Gaussian Simulation

A laboratory waterflood experiment was conducted to determine

the oil recovery curve for a viscous oil reservoir (μo = 100 cp). The

coreflood experiment was performed in an unconsolidated sandpack.

The task at hand is to use the laboratory test in a sandpack to forecast

the oil recovery curve for a heterogeneous reservoir. Figure 4.69 shows

the CT images of the laboratory waterflood at three times just before

water breakthrough. It is shows a fairly uniform displacement of oil by

Page 504: +Peters Ekwere j. - Petrophysics

4–142

the water. Figure 4.70 shows the oil recovery curve from the experiment.

It should observed that even for a fairly uniform sandpack, the oil

recovery after 3 pore volumes of water injection is less than 60% of the

initial oil in place. What will be the oil recovery in a reservoir with

significant permeability heterogeneity?

Figure 4.69. CT images of a laboratory waterflood in a sandpack: (a) 0.05 pore volume injected, (b) 0.10 pore volume injected, ( c) 0.25 pore volume

injected (Gharbi and Peters, 1993)

.

Page 505: +Peters Ekwere j. - Petrophysics

4-143

Figure 4.70. Oil recovery curve for a laboratory waterflood in a sandpack (Gharbi and Peters, 1993)

.

The problem of forecasting the oil recovery curve in the field based

on a laboratory waterflood boils down to investigating the effect of

permeability heterogeneity on the waterflood performance. To address

this question, we generated twelve synthetic reservoirs with varying

degrees of permeability heterogeneity and correlation structures and then

scaled the laboratory waterflood to the synthetic reservoirs through

numerical simulation. Figure 4.71 shows the twelve heterogeneous, 2D

Page 506: +Peters Ekwere j. - Petrophysics

4–144

permeability fields generated to cover a wide range of Dykstra-Parsons

coefficient and correlation lengths using Sequential Gaussian

Simulation. Permeability fields were generated at Dykstra-Parsons

coefficients (V) of 0.01, 0.55 and 0.87 and dimensionless correlation

lengths of 0, 0.2, 0.7 and 2.0. The correlation length was made

dimensionless by dividing the correlation length by the length of the

reservoir in that direction. A Dykstra-Parsons coefficient of 0.01

represents a nearly homogeneous medium whereas a Dyktra-Parsons

coefficient of 0.89 represents an extremely heterogeneous medium.

Petroleum reservoirs typically have Dykstra-Parsons coefficients that

range from 0.5 to 0.9. A dimensionless correlation length of zero

represents an uncorrelated or random permeability distribution; a

correlation length of 0.2 represents mild correlation; a correlation length

of 2.0 represents extremely strong correlation. Depending on the

depositional environment, petroleum reservoirs can have widely different

correlation lengths. The value of dimensionless correlation length in the

y direction (Ly) was constant at 0.2 for the permeability fields shown in

Figure 4.71.

Two observations can be made from the permeability distributions

of Figure 4.71. First, as the correlation length (Lx) increases, the

permeability distributions become more and more stratified. The

number of layers is inversely proportional to Ly. In fact, as Lx

approaches infinity, for Ly = 0.2, the permeability distribution will

consist of exactly five (1/0.2) distinct homogeneous layers. Second, with

increasing Dykstra-Parsons coefficient at a given Lx, the contrast in the

permeability values increases while their spatial arrangements remain

similar. Figure 4.72 shows the permeability histograms, which indicate

that the permeability fields are log-normally distributed in accordance

Page 507: +Peters Ekwere j. - Petrophysics

4-145

with observations in sedimentary rocks. Figure 4.73 shows the

variograms for the twelve permeability fields, which give a visual

impression of the degree of correlation in each permeability field.

Figure 4.71. Simulated permeability distributions (Gharbi and Peters, 1993)

Figure 4.74 shows a comparison of the recovery curve for the waterflood

experiment and the numerical simulation of the experiment. The

agreement between the two recovery curves is good. The simulation was

used to determine the relative permeability curves to be used to scale the

laboratory waterflood to the synthetic heterogeneous reservoirs.

Page 508: +Peters Ekwere j. - Petrophysics

4–146

Figure 4.72. Simulated permeability histograms (Gharbi and Peters,

1993).

Page 509: +Peters Ekwere j. - Petrophysics

4-147

Figure 4.73. Simulated permeability variograms in the x-direction

(Gharbi and Peters, 1993).

Page 510: +Peters Ekwere j. - Petrophysics

4–148

Figure 4.74. A comparison of the oil recovery curves of the experiment and the numerical simulation of the experiment (Gharbi and Peters,

1993).

Figure 4.75 compares the simulated oil recovery curves for each of

the twelve heterogeneous synthetic reservoirs with that of the laboratory

waterflood experiment. The following observations can be made from

these results. If the heterogeneous reservoir is characterized by low

variability in the permeability distribution (low Dykstra-Parsons

coefficient), the waterflood response will be essentially the same as in the

laboratory sandpack regardless of the correlation structure of the

Page 511: +Peters Ekwere j. - Petrophysics

4-149

heterogeneity. This is indicated by the agreement between the simulated

and the experimental recovery curves in the first column of Fig. 4.75 (V =

0.01). If the heterogeneous reservoir is characterized by a low correlation

length in the permeability distribution (low Lx), the waterflood response

will be essentially the same as in the laboratory sandpack regardless of

the variability in the permeability distribution. This is indicated by the

agreement between the simulated and the experimental recovery curves

in the first row of Figure 4.75 (Lx = 0). If the heterogeneous reservoir is

characterized by high variability and high correlation length in the

permeability distribution, the waterflood response could be significantly

different from that of the laboratory sandpack. This is most clearly

shown by the response in the last permeability field in Figure 4.75 (Lx =

2 and V = 0.87). In this case, the waterflood effeciency is significantly

less in the heterogeneous reservoir than in the laboratory sandpack.

To investigate the reason for the significant disparity in

performance between the laboratory waterflood in a relatively

homogeneous sandpack and in certain kinds of heterogeneous reservoirs,

we examine the simulated water saturation maps. Figures 4.76 and 4.77

show the simulated water saturation maps for each of the twelve

heterogeneous reservoirs at 0.10 and 0.25 pore volume injected. We see

that the displacements in the heterogeneous media with high Dykstra-

Parsons coefficient and high correlation length are dominated by

channeling of the injected water due to the permeability stratification.

This results in significant bypassing of the oil in some layers, resulting in

a low oil recovery. These channels provide easy pathways for the water to

flow from the injection well to the producing well, essentially leaving

much of the reservoir unswept. By contrast, the displacements in the

reservoirs with low Dykstra-Parsons coefficients are characterized by

Page 512: +Peters Ekwere j. - Petrophysics

4–150

excellent sweep comparable with that observed in the CT images of the

laboratory waterflood experiment (Figure 4.69). This results in a

displacement performance that is comparable to the laboratory

waterflood experiment in the sandpack.

We conclude from this study that the performance of an enhanced

oil recovery (EOR) displacement in a heterogeneous reservoir could be

significantly lower than in a laboratory experiment depending on the

degree and structure of the heterogeneity of the reservoir. This

conclusion underscores the need for proper scaling when using the

results of laboratory coreflood experiments in relatively homogeneous

porous media to forecast the expected performance of an EOR process in

heterogeneous reservoirs. The methodology developed and presented in

this study can be used to accomplish this scaling and prevent erroneous

performance forecasts.

Page 513: +Peters Ekwere j. - Petrophysics

4-151

Figure 4.75. A comparison of the experimental and simulated oil recovery

curves (Gharbi and Peters, 1993).

Page 514: +Peters Ekwere j. - Petrophysics

4–152

Figure 4.76. Simulated water saturation maps at 0.10 pore volume

injected (Gharbi and Peters, 1993).

Figure 4.77. Simulated water saturation maps at 0.25 pore volume

injected (Gharbi and Peters, 1993).

NOMENCLATURE

a = correlation length (range of influence)

C = covariance function

Page 515: +Peters Ekwere j. - Petrophysics

4-153

C0 = nugget effect

erf = error function

erfc = complementary error function

F = cumulative probability distribution

h = pay thickness

h = lag distance

k = absolute permeability of the medium

ln = natural logarithm (log to base e)

Lx = dimensionless correlation length in the x-direction

Ly = dimensionless correlation length in the y-direction

P = probability density function

s = standard deviation of sample data

s2 = variance of sample data

V = Dykstra-Parsons coefficient of variation

x = linear coordinate

x = random variable

z = variate from a standard normal distribution

ρ = correlation coefficient function

μ = population mean

μ = Lagrange parameter

σ = population standard deviation

σ2 = population variance

φ = porosity, fraction

τ = mean of log normal distribution

γ = variogram

λ = kriging weights

Φ = sample data

Φ = sample mean

ω = standard deviation of a log normal distribution

ω2 = variance of a log normal distribution

Page 516: +Peters Ekwere j. - Petrophysics

4–154

REFERENCES AND SUGGESTED READINGS

Armstrong, M. : Basic Linear Geostatistics, Springer-Verlag, New York, 1998.

Caers, J. : Petroleum Geostatistics, Society of Petroleum Engineers, Richardson, 2005.

Carlson, M.R. : Practical Reservoir Simulation, PennWell, Tulsa, 2003.

Chambers, R.L., Yarus, J.M. and Hird, K.B. : “Petroleum Geostatistics for Nongeostatisticians – Part 1,” The Leading Edge (May 2000) 474-479.

Chambers, R.L., Yarus, J.M. and Hird, K.B. : “Petroleum Geostatistics for Nongeostatisticians – Part 2,” The Leading Edge (June 2000) 592-599.

Clark, I. : Practical Geostatistics, Applied Science Publishers, London, 1979.

Clark, I. and Harper, W.V. : Practical Geostatistics 2000, Ecosse North America Llc, Columbus, 2000.

Clark, I. : "Does Geostatistics Work? ", Proc. 16th APCOM, Thomas J O'Neil, Ed., Society of Mining Engineers of AIME Inc, New York, 1979, 213-225.

Deutsch, C.V.: "What in the Reservoir is Geostatistics Good For ?", Jour. Cand. Pet. Tech. (April 2006) 14-20.

Deutsch, C.V.: Geostatistical Reservoir Modeling, Oxford University Press, New York, 2002.

Deutsch, C.V. and Journel, A.G. : GSLIB Geostatistical Software Library and User’s Guide, Oxford University Press, New York, 1992.

Dykstra, H. and Parsons, R.L. : “The Prediction of Oil Recovery by Waterflood,” Secondary Recovery of Oil in the United States, American Petroleum Institute (1950) 160-175.

Gharbi, R. and Peters, E.J. : “Scaling Coreflood Experiments to Heterogeneous Reservoirs,” Journal of Petroleum Science and Engineering, 10, (1993) 83-95.

Gharbi, R.: Numerical Modeling of Fluid Displacements in Porous Media Assisted by Computed Tomography Imaging, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1993.

Page 517: +Peters Ekwere j. - Petrophysics

4-155

Gotway, C.A. and Hergert, G.W. : "Incorporating Spatial Trends and Anisotropy in Geostatistical Mapping of Soil Properties," Soil Science of America Journal, Vol. 61 (1977) 298-309.

Hohn, M.E. : Geostatistics and Petroleum Geology, Van Nostrand Reinhold, New York, 1988.

Hirsche, K., Porter-Hische, J., Mewhort, L. and Davis, R. : "The Use and Abuse of Geostatistics," The Leading Edge (March 1997) 253-260.

Holbrook, P. : Pore Pressure Through Earth Mechanical Systems, Force Balanced Petrophysics, 2001.

Isaaks, E.H. and Srivastava, R.M. : An Introduction to Applied Geostatistics, Oxford University Press, New York, 1989.

Jensen, J.L., Lake, L.W., Corbett, P.W.M. and Goggin, D.J. : Statistics for Petroleum Engineers and Geoscientists, 2nd Edition, Elsevier, New York, 2000.

Kerbs, L. : “GEO-Statistics: The Variogram,” Computer Oriented Geological Society Computer Contributions (August 1986) 2, No. 2, 54-59.

Metheron, G. : "Principles of Geostatistics," Economic Geology, Vol. 58 (1963) 1446-1266.

Peters, E.J., Afzal, N. and Gharbi, R. : “On Scaling Immiscible Displacements in Permeable Media,” Journal of Petroleum Science and Engineering, 9, (1993) 183-205.

Peters, E.J. and Gharbi, R. : “Numerical Modeling of Laboratory Corefloods,” Journal of Petroleum Science and Engineering, 9, (1993) 207-221.

Peters, E.J. and Afzal, N. : “Characterization of Heterogeneities in Permeable Media with Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 7, No. 3/4, (May 1992) 283-296.

Peters, E.J. and Hardham, W.D. : “Visualization of Fluid Displacements in Porous Media Using Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 4, No. 2, (May 1990) 155-168.

Zirschy, J.H. and Harris, D.J. : "Geostatistical Analysis of Hazardous Waste Site Data," J. of Environmental Engineering, Vol. 112 (1986) 770784.

Page 518: +Peters Ekwere j. - Petrophysics

4–156

Page 519: +Peters Ekwere j. - Petrophysics

CHAPTER 5

DISPERSION IN POROUS MEDIA

5.1 INTRODUCTION

When a miscible fluid displaces another in a porous medium, the

displacing fluid tends to mix with the displaced fluid. The result is that a

mixing or transition zone develops at the front in which the

concentration of the injected fluid decreases from one to zero.

Experiment shows that the mixing zone grows as the displacement

progresses. This mixing and spreading of the injected fluid is known as

dispersion.

Bear (1972) describes dispersion as the "macroscopic outcome of

actual movement of individual tracer particles through pores...".

Essentially, dispersion is the mixing caused by single-phase fluid

movement through a porous medium. What is "mixed" is usually called

a tracer, but can be thought of as a concentration of any chemical

component within a given phase that is transported through the system.

5-1

Page 520: +Peters Ekwere j. - Petrophysics

Dispersion has practical consequences in contaminant transport in

aquifers and in improved oil recovery from petroleum reservoirs. If a

miscible contaminant is accidentally introduced into an aquifer at a site,

dispersion will cause the contaminant to spread to a larger area as it is

being transported by groundwater flow. Even though the concentration

of the contaminant is reduced by dispersion, a much larger area of the

aquifer will become contaminated as a result of dispersion than the

original spill area. Thus, a much larger area than the original spill will

need to be cleaned up by any contaminant remediation measure.

Miscible displacement is the most efficient improved oil recovery

method. Because there is no capillary force to trap the displaced oil, it is

theoretically possible to recover 100% of the oil by miscible displacement.

However, because the injected solvent is usually more expensive than the

oil that is to be displaced, it is usually injected in small quantities as

slugs and chased by a less expensive fluid such as water or gas.

Dispersion will dilute and reduce the effectiveness of the miscible slug as

it is propagated through the reservoir. In this case, dispersion is

detrimental to the recovery process. On the other hand, dispersion

causes a solvent to mix, spread and contact the displaced fluid even after

it had been originally bypassed by the injected solvent. In this case,

dispersion improves the displacement efficiency.

Other industrial processes that involve dispersion include (1) use

of tracers such as dyes, electrolytes and radioactive isotopes to

characterize reservoir and aquifer properties, (2) development of a

transition zone between salt water and fresh water in coastal aquifers, (3)

radioactive and reclaimed sewage waste disposals into aquifers, (4) use of

reactors packed with granular material in the chemical industry, and (5)

5–2

Page 521: +Peters Ekwere j. - Petrophysics

movement of fertilizers in the soil and the leaching of salts from the soil

in agriculture.

5.2 LABORATORY FIRST-CONTACT MISCIBLE DISPLACEMENTS

Before embarking on the mathematics of dispersion in porous

media, it is instructive and illuminating to view images of miscible

displacements that show dispersion at work. The images presented in

this section are for first-contact miscible displacements. This means that

the two fluids used in the experiment are fully miscible upon first

contact. This is in contrast to developed miscibility encountered in

certain enhanced oil recovery process in which the injected fluid is not

initially miscible with the displaced fluid. However, after a certain time

has elapsed, mass transfer between the injected and the displaced fluids

causes the two fluids to become miscible. First-contact miscible

displacements are the most efficient type of displacements.

The efficiency of a miscible displacement or an immiscible

displacement for that matter, is controlled by, among other factors, the

mobility ratio, the effect of gravity, dispersion and the heterogeneity of

the porous medium. Mobility ratio is a dimensionless number that is

characteristic of a displacement. The mobility of a fluid phase in a porous

medium is defined as

phasephase

phase

μ= (5.1)

where is the effective permeability to that phase and phasek phaseμ is the

viscosity of that phase. Mobility ratio is defined as

5-3

Page 522: +Peters Ekwere j. - Petrophysics

displacing fluid

displaced fluid

Mλλ

= (5.2)

For a first-contact miscible displacement, we have single phase flow and

as a result, the effective permeability to each fluid is equal to the

absolute permeability of the porous medium. Therefore, for first-contact

miscible displacements, the mobility ratio simplifies to

displacing fluid displaced fluid

displacing fluid

displaced fluid

k

Mk

μ μμ

μ

⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠= =⎛ ⎞⎜ ⎟⎜ ⎟⎝ ⎠

(5.3)

Thus, for a first-contact miscible displacement, the mobility ratio is the

viscosity ratio given by the viscosity of the displaced fluid divided by the

viscosity of the displacing fluid. A mobility ratio of 1 or less is favorable

to the displacement efficiency whereas a mobility ratio greater than 1 is

unfavorable to the displacement efficiency. The lower the mobility ratio,

the higher is the displacement efficiency. The higher the mobility ratio,

the lower is the displacement efficiency.

Gravity can be as a double-edged sword in displacements. Gravity

override and gravity underride or tonguing can significantly reduce the

displacement efficiency. However, by careful engineering, gravity override

can be used to enhance the displacement efficiency if the less dense

solvent is injected up dip while the denser displaced fluid is withdrawn

down dip.

Dispersion can also be as a double-edged sword in displacements.

If a miscible slug is used in a displacement process to reduce cost,

5–4

Page 523: +Peters Ekwere j. - Petrophysics

dispersion will dilute and degrade the effectiveness of the slug as the

process progresses. On the other hand, mixing caused by dispersion can

allow the injected solvent to spread and contact more of the displaced

fluid than if dispersion was absent. In this case, dispersion is favorable

to the displacement efficiency.

Permeability heterogeneity of the porous medium tends to reduce

the displacement efficiency for the most part. Because the heterogeneity

has a correlation structure, high permeabilities tend to occur next to

high permeabilities and low permeabilities tend to occur next to low

permeabilities as the porous medium is traversed from one point to

another. These permeability arrangements cause the injected fluid to

channel through the high permeability layers, thereby leaving the low

permeability layers unswept.

Figure 5.1 shows CT images of a first-contact miscible

displacement in an unconsolidated sandpack at a favorable mobility ratio

of 0.82. In the experiment, a brine containing sodium chloride was used

to displace another brine containing barium chloride, which is an x-ray

doppant. The images show the concentrations of the injected solvent in a

vertical slice of the 3D sandpack at dimensionless times of 0.13, 0.50

and 1.0 pore volume injected. Mixing caused by dispersion is clearly

visible at the displacement front. The mixing zone length, defined as the

distance between the solvent concentrations of 0.90 and 0.10, is slightly

tilted in Figure 5.1B because of a small density difference between the

injected brine and the displaced brine. The injected brine was slightly

less dense that the displaced brine. Thus, the tilt is a small gravity

override. Because of the favorable mobility ratio, the displacement is very

efficient with almost 100% displacement efficiency at 1 pore volume

injected. The displacement efficiency is slightly less than 100% because

5-5

Page 524: +Peters Ekwere j. - Petrophysics

of the small gravity override problem. This experiment shows that a

miscible displacement can be quite susceptible to gravity instability in

the form of gravity override if the injected fluid is less dense than the

displaced fluid and gravity tonguing if the injected fluid is denser than

the displaced fluid.

Figure 5.1. CT images of solvent concentration for a first-contact miscible displacement in a sandpack. A: tD = 0.13, B: tD = 0.50, C: tD = 1.0 pore

volume injected. M = 0.82, φ = 35.26%, k = 15.76 darcies, v = 0.154x10-2 cm/s, BT recovery = 96.53% (Peters and Hardham, 1990).

5–6

Page 525: +Peters Ekwere j. - Petrophysics

Figure 5.2 shows the 1D profiles of the 3D solvent concentration

distributions at the three dimensionless times. The profiles have the

classical shapes predicted by the convection-dispersion model of

dispersion to be presented later in this chapter. The profiles can be used

to characterize the dispersion phenomenon in the direction of flow.

Figure 5.2. Average solvent concentration profiles for the miscible

displacement experiment of Figure 5.1 (Peters and Hardham, 1990).

Figure 5.3 shows the images of a first-contact miscible

displacement at an unfavorable mobility ratio of 74. Here, brine with an

x-ray doppant was used to displace a mixture of glycerine and brine. The

injected brine was less dense than the displaced fluid. It can be seen that

the density difference results in a significant gravity override. Such an

override is detrimental to the displacement efficiency. Mixing caused by

dispersion is also clearly evident. The transverse or lateral dispersion is

quite orderly as one moves from the pure solvent to the pure displaced

5-7

Page 526: +Peters Ekwere j. - Petrophysics

fluid. There is also viscous instability or viscous fingering caused by the

adverse mobility ratio. At a mobility ratio greater than 1, the displacing

fluid is more mobile than the displaced fluid. As a result, at the interface

between the two fluids, the displacing fluid tends to penetrate the

displaced fluid due to the inevitable imperfections in the porous medium.

These perturbations may grow to form viscous fingers as shown in Figure

5.3B. Figure 5.4 shows the corresponding solvent concentration profiles

for this displacement. The shapes of the profiles are significantly different

from those of Figure 5.2 and cannot be predicted by the idealized

convection-dispersion model. Clearly, this displacement is less efficient

than that of Figure 5.1.

5–8

Page 527: +Peters Ekwere j. - Petrophysics

Figure 5.3. CT images of solvent concentration for a first-contact miscible

displacement in a sandpack. A: tD = 0.13, B: tD = 0.50, C: tD = 1.0 pore volume injected. M = 74, φ = 35.68%, k = 16.69 darcies, v = 0.014x10-2

cm/s, BT recovery = 31.23% (Peters and Hardham, 1990).

Figure 5.4. Average solvent concentration profiles for the miscible

displacement experiment of Figure 5.3 (Peters and Hardham, 1990).

5-9

Page 528: +Peters Ekwere j. - Petrophysics

Figure 5.5 shows the images of a first-contact miscible

displacement at an unfavorable mobility ratio of 54 with gravity override.

The images show that portions of the porous medium that were initially

bypassed by the solvent due to gravity override, were subsequently

displaced with continued solvent injection. This is a testament to the

efficiency of miscible displacements. By contrast, in an immiscible

displacement, portions of the porous medium initially bypassed for

whatever reason typically remain unswept with continued injection.

Figure 5.6 shows the corresponding solvent concentration profiles. Notice

how the disturbances in the profiles propagate in time and space. This is

characteristic of miscible displacements. Such disturbances typically

remain stationary in immiscible displacements.

Figure 5.7 shows a first-contact miscible displacement in a quarter

five-spot pattern at a mobility ratio of 1. Here, dyed water was used to

displace clear water in a thin porous medium consisting of uniform glass

beads. The images were obtained with a home-built imaging system

described by Peters and Reid (1990). As expected, the displacement is

efficient. The mixing zone length caused by dispersion is also visible. The

areal sweep efficiency can easily be measured. The areal sweep efficiency

is the ratio of the area contacted by the injected fluid and the total area

of the pattern. Note that because of the geometry of the displacement

pattern, there are dead spots between the injector and the producing

wells. In a field flood, such dead spots would be potential candidates for

infill drilling. In this experiment, the thin porous medium was oriented

horizontally. As a result, gravity effect was negligibly small.

5–10

Page 529: +Peters Ekwere j. - Petrophysics

Figure 5.5. CT images of solvent concentration for a first-contact miscible

displacement in a sandpack. M = 54, φ = 31.70%, k = 9.5 darcies, v = 0.850x10-2 cm/s, BT recovery = 28.80% (Peters and Afzal, 1992).

5-11

Page 530: +Peters Ekwere j. - Petrophysics

Figure 5.6. Average solvent concentration profiles for the miscible displacement experiment of Figure 5.5 (Peters and Afzal, 1992).

Figure 5.7. Video images of solvent concentration in a miscible

displacement in a glass bead pack. M = 1(Peters and Reid, 1990).

5–12

Page 531: +Peters Ekwere j. - Petrophysics

Figure 5.8 shows the images of another first-contact miscible

displacement in the same porous medium as in Figure 5.7 at a mobility

ratio of 100. The porous medium was oriented horizontally. There is

significant mixing due to dispersion and viscous fingering due to the

adverse mobility ratio. Clearly, this displacement is less efficient than

that of Figure 5.7.

Figure 5.8. Video images of solvent concentration in a miscible displacement in a glass bead pack. M = 100 (Peters and Reid, 1990).

5-13

Page 532: +Peters Ekwere j. - Petrophysics

Figure 5.9 shows images of a first-contact miscible displacement in

the same porous medium and at the same mobility ratio as in Figure 5.8

but with the medium oriented vertically to take advantage of the

beneficial effect of gravity override. This experiment was designed to

simulate a horizontal injection well at the top left corner and a horizontal

producing well at the bottom right corner of the medium. With this

arrangement, gravity override is beneficial to the displacement efficiency

in contrast to the displacement of Figure 5.3 in which gravity segregation

was detrimental to the displacement efficiency. It can be seen that here,

Figure 5.9. Video images of solvent concentration in a gravity-assisted miscible displacement experiment in a glass bead pack (Peters and Reid,

1990).

5–14

Page 533: +Peters Ekwere j. - Petrophysics

gravity override delays the solvent breakthrough thereby significantly

enhancing the displacement efficiency compared to the displacement of

Figure 5.8. There is viscous fingering due to the adverse mobility ratio as

well as significant mixing due to dispersion. Clearly, this displacement is

more efficient than that of Figure 5.8.

Figure 5.10 shows NMR images of a first-contact miscible

displacement in the layered sandstone core of Figure 2.57 at a favorable

mobility ratio of 0.84. One can see mixing due to dispersion. However,

the dominate effect is channeling due to the permeability heterogeneity of

the layered medium. It can be concluded from these images that

permeability stratification in a porous medium can have a significant

adverse effect on any displacement, miscible or immiscible. Figure 5.11

shows the corresponding solvent concentration profiles.

Figure 5.12 shows a miscible displacement is the same layered

sandstone core as in Figure 5.10 but at an unfavorable mobility ratio of

95. Here, the effect of viscous fingering due to the adverse mobility ratio

is superimposed on the channeling due to the heterogeneity of the core.

Notice that the channels created by the injected solvent are thinner than

in Figure 5.10 indicating a less efficient displacement than in Figure

5.10. Figure 5.13 shows the corresponding solvent concentration

profiles.

5-15

Page 534: +Peters Ekwere j. - Petrophysics

Figure 5.10. NMR images of a first-contact miscible displacement in a layered sandstone core at a mobility ratio of 0.84 (Peters and Li, 1996).

5–16

Page 535: +Peters Ekwere j. - Petrophysics

Figure 5.11. Average solvent concentration profiles for the miscible displacement of Figure 5.9 (Peters and Li, 1996).

5-17

Page 536: +Peters Ekwere j. - Petrophysics

Figure 5.12. NMR images of a first-contact miscible displacement in a layered sandstone core at a mobility ratio of 95 (Peters and Li, 1996).

5–18

Page 537: +Peters Ekwere j. - Petrophysics

Figure 5.13.Average solvent concentration profiles of a first-contact miscible displacement in a layered sandstone core at a mobility ratio of

95 (Peters and Li, 1996).

Finally, Figure 5.14 shows the effect of mobility ratio on the

efficiency of miscible displacements. The figure shows the recovery

curves for miscible displacements in the same layered sandstone core as

in Figure 5.10 for mobility ratios ranging from 0.84 to 95. Clearly,

mobility ratio has a significant impact on the efficiency and the timing of

a miscible displacement. Displacements at adverse mobility ratios suffer

from early solvent breakthroughs and require more pore volumes of

5-19

Page 538: +Peters Ekwere j. - Petrophysics

injection to attain the same displacement efficiency as favorable mobility

ratio displacements.

Figure 5.14. Recovery curves for a first-contact miscible displacement in a layered sandstone core at various mobility ratios (Peters and Li, 1996).

5.3 ORIGINS OF DISPERSION IN POROUS MEDIA

Dispersion is the net result of (a) molecular diffusion, (b) local

velocity gradients within given pores, (c) locally heterogeneous streamline

lengths and velocities, and (d) mechanical mixing in pore bodies.

5–20

Page 539: +Peters Ekwere j. - Petrophysics

Dispersion can be viewed as consisting of two components: molecular

diffusion and mechanical dispsersion.

5.3.1 Molecular Diffusion.

Molecular diffusion is a physiochemical dispersion caused by

chemical potential gradient, which is correlated to the chemical

concentration of the solute being transported. It is mixing caused by

random motions of the fluid particles due to the thermal kinetic energy of

the solute. This motion is known as Brownian motion. Molecular

diffusion is isotropic and occurs equally in all directions. Molecular

diffusion can easily be demonstrated in the laboratory. If a drop of blue

ink is carefully added to a beaker of water and allowed to sit, after

sometime, the water in the beaker will turn blue as a result of molecular

diffusion. Thus, molecular diffusion occurs whether there is flow or not.

Molecular diffusion will contribute to both longitudinal and transverse

dispersions. Molecular diffusion in a porous medium is less than it

would be in the absence of the porous medium. The solid grains hinder

diffusion just as they hindered the flow of electrical current and fluid

flow. Because of the larger molecular spacing, molecular diffusion in a

gas is much larger than in a liquid.

5.3.2 Mechanical Dispersion.

The second component of dispersion may be described as

mechanical dispersion. The origins of mechanical dispersion can be seen

in Figure 5.15. The figure shows marked fluid particles at time t and at

time t+Δt. When a fluid flows in a porous medium, its velocity

distribution within a pore is not uniform, due to boundary effects acting

in three different ways as shown in Figure 5.15. In Figure 5.15a, the no

slip condition at the pore wall creates a velocity gradient in the fluid. This

5-21

Page 540: +Peters Ekwere j. - Petrophysics

velocity gradient causes the marked fluid particles to be spread out in

the flow direction. This is longitudinal dispersion in the flow direction. In

Figure 5.15b, the variation in pore dimensions (recall k ∝ r2) causes flow

to occur faster (further) in some pores than in others. This causes the

marked fluid particles to be spread out in the direction of flow. This is

longitudinal dispersion in the direction of flow. In Figure 5.15c, the

streamlines fluctuate with respect to the mean flow direction as the fluid

particles navigate around the solid grains. This effect of the tortuosity of

the porous medium causes the fluid particles to be spread out in the

transverse direction with respect to the mean flow direction. This is

transverse dispersion, perpendicular to the direction of flow. Finally,

there is local mixing of the fluids within the pores as shown in Figure

5.16. This contributes to the mechanical dispersion.

Figure 5.15.Origins of mechanical dispersion.

5–22

Page 541: +Peters Ekwere j. - Petrophysics

Figure 5.16.Local mechanical mixing of fluid particles.

5.4 CONVECTION-DISPERSION EQUATION

5.4.1 Generalized Equation in Vector Notation

Chemical species are transported in a flowing system by two

transport mechanisms: (a) advection or convection and (b) dispersion.

Advection is mass transport due to the bulk motion of the carrying fluid

and is given by

aJ uCφ= (5.4)

where is the mass flux vector (mass/area/time) of species i due to

advection, φ is the porosity of the porous medium, u

aJ

is the interstitial

velocity vector (Darcy velocity vector/porosity) and C is the concentration

(mass/unit volume of solution) of species i in the solution. The mass flux

due to dispersion is given by

dJ Dφ C= − ∇ (5.5)

5-23

Page 542: +Peters Ekwere j. - Petrophysics

where is the mass flux vector (mass/area/time) of species i due to

dispersion and

dJ

D is the dispersion coefficient tensor. The dispersion

coefficient tensor characterizes molecular diffusion and mechanical

dispersion and is given by

dD D D= + m (5.6)

where is the molecular diffusion coefficient and dD mD is the mechanical

dispersion tensor. The continuity equation for mass transport is given by

( ) .C

Jtφ∂

0+∇ =∂

(5.7)

where t is time and is the total mass flux vector of species i due to

advection and dispersion and is given by the vector sum

J

aJ J Jd= + (5.8)

Substituting Eqs.(5.4) and (5.5) into (5.7) gives the mass transport

equation for a constant porosity medium as

( ) ( ). .C uC D Ct

∂ 0+∇ −∇ ∇ =∂

(5.9)

Eq.(5.9) is a well known equation of mathematical physics, which

is known by a variety of names such the advection-dispersion equation,

convection-dispersion equation, first-contact miscible displacement

equation, solute transport equation and mass transport equation. It is a

linear, second order, parabolic partial differential equation. It can be

5–24

Page 543: +Peters Ekwere j. - Petrophysics

used to solve a variety of transport problems. If the transport is by pure

convection with no dispersion, then Eq.(5.9) simplifies to

( ).C uCt

0∂+∇ =

∂ (5.10)

If there is no convection (static fluid), then there will be no mechanical

dispersion. In this case, transport is by molecular diffusion only and

Eq.(5.9) becomes

2 0dC D Ct

∂− ∇ =

∂ (5.11)

which is the diffusion or diffusivity equation.

5.4.2 One-Dimensional Convection-Dispersion Equation

For 1D transport in the x direction, Eq.(5.9) becomes

2

2 0x LC C Cu Dt x x

∂ ∂ ∂+ −

∂ ∂ ∂= (5.12)

where vx is the interstitial velocity in the x direction and DL is the

principal value of the dispersion coefficient in the x direction known as

the longitudinal dispersion coefficient. In this case, the flow direction is a

principal axis of the dispersion coefficient anisotropy. Eq.(5.12) can be

written in terms of Darcy velocity instead the interstitial velocity as

2

2 0xL

vC C CDt x xφ

∂ ∂ ∂+ −

∂ ∂ ∂= (5.13)

5-25

Page 544: +Peters Ekwere j. - Petrophysics

where ux is the Darcy velocity also known as the superficial velocity. It is

possible to extend the mass transport equation to include retardation of

the solute due to adsorption, chemical reaction, biological

transformations or radioactive decay. In the case of retardation due to

adsorption of the chemical species on the surface of the porous medium,

Eq.(5.12) becomes

2

2 0x L

f f

u DC C Ct R x R x

∂ ∂ ∂+ −

∂ ∂ ∂= (5.14)

where Rf is a retardation factor that accounts for adsorption. If Rf is

equal to 1, there is no adsorption whereas if Rf is greater than 1, there is

adsorption and the transport of the chemical species will be retarded.

This means that in the presence of adsorption, the concentration profiles

of the chemical species will travel at a speed that is lower than if there

was no adsorption. This fact is obvious from Eq.(5.14) in which the speed

of convection has been reduced from vx to vx/Rf. Also, retardation

reduces the effective dispersion coefficient as shown in Eq.(5.14).

5.4.3 Solution of the One-Dimensional Convection-Dispersion Equation

The initial-boundary value problem for 1D transport consists of

Eq.(5.12) together with appropriate initial and boundary conditions. For

transport in a semi infinite medium, the initial-boundary value problem

consists of the following equations:

2

2 0x LC C Cu Dt x x

∂ ∂ ∂+ −

∂ ∂ ∂= (5.12)

( ),0 iC x C= (5.15)

5–26

Page 545: +Peters Ekwere j. - Petrophysics

( )0, jC t C= (5.16)

( ), iC t C∞ = (5.17)

Eq.(5.15) is the initial condition, which specifies the initial concentration

of the injected species in the solution domain at time zero. A convenient

initial concentration in laboratory experiments is Ci equal to zero, which

means there is no injected chemical species in the initial saturating fluid.

Eq.(5.16) is the inlet boundary condition, which specifies the

concentration of the injected species at all times. A convenient inlet

boundary condition is Cj is equal to 1 at all times, although this is

difficult to arrange in an actual laboratory experiment. Eq.(5.17) is the

external boundary condition, which specifies that far away from the inlet,

the concentration of the injected species is equal to its initial value at

time zero. Such a boundary condition was used in the welltest model of

Chapter 3 for an infinite acting reservoir.

The initial-value problem can be put in dimensionless form as

follows:

2

2

1 0D D D

D D Pe D

C C Ct x N x

∂ ∂ ∂+ −

∂ ∂ ∂= (5.18)

( ),0 0DC x = (5.19)

( )0, 1D DC t = (5.20)

( ), DC t 0∞ = (5.21)

5-27

Page 546: +Peters Ekwere j. - Petrophysics

where

( ), iD

j i

C x t CC

C C−

=−

(5.22)

xD

u t qttL A Lφ

= = (5.23)

DxxL

= (5.24)

xPe

L L

u L qLND A Dφ

= = (5.25)

Eq.(5.23) gives the dimensionless time in pore volumes injected. Eq.(5.25)

defines a relevant dimensionless number for the transport known as the

Peclet Number. It is the ratio of the two transport mechanisms involved,

advection and dispersion. In the definition of Peclet Number in Eq.(5.25),

the length of the porous medium, L, has been used as a characteristic

dimension of the system. Of course, in a semi infinite medium, the length

of the porous medium would not be a convenient characteristic

dimension of the system. In general, Peclet Number is defined as

x pPe

L

u DN

D= (5.26)

where Dp is a characteristic dimension of the porous medium. A

convenient characteristic dimension could be the grain diameter in the

case in which the porous medium is composed of granular material.

5–28

Page 547: +Peters Ekwere j. - Petrophysics

Another possible characteristic dimension of a porous medium could be

in the spirit of Eq.(3.119).

Eqs.(5.12), (5.15), (5.16) and (5.17) or their dimensionless versions,

Eqs.(5.18) through (5.21), can be solved by Laplace transformation.

Ogata and Banks (1961) give the solution of Eqs.(5.12), (5.15), (5.16) and

(5.17) as

( ),2 2 2

x

L

v xDi x

L L

C x u t x u tC x t erfc e erfcD t D t

x⎡ ⎤⎛ ⎞ ⎛ ⎞−

= ++

⎢ ⎥⎜ ⎟ ⎜⎜ ⎟ ⎜ ⎟⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦ (5.27)

where erfc is the complementary error function defined by Eq.(4.47) and

reproduced here for convenience as

( ) 22 u

uerfc u e du

π∞ −= ∫ (5.28)

The error function and the complementary error function are related by

( ) ( ) 1erf u erfc u+ = (5.29)

( ) ( )1erfc u erf u− = + (5.30)

Also,

( ) ( )erf u erf u− = − (5.31)

The solution given by Eq.(5.27) can be written in dimensionless form as

5-29

Page 548: +Peters Ekwere j. - Petrophysics

( ) 1, e2

2 2

D Pex ND D D DD D D

D D

Pe Pe

x t x tC x t erfc erfct t

N N

⎡ ⎤⎛ ⎞ ⎛ ⎞⎢ ⎥⎜ ⎟ ⎜ ⎟

− +⎢ ⎥⎜ ⎟ ⎜= + ⎟⎢ ⎥⎜ ⎟ ⎜ ⎟⎢ ⎥⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦

(5.32)

Ogata and Banks (1961) demonstrated that the second term in the

square bracket of the right hand sides of Eqs.(5.27) and Eq.(5.32) can be

neglected in comparison to the first term. The solution given by Eq.(5.27)

then becomes

( ),2 2

i

L

C x u tC x t erfcD t

x⎡ ⎤⎛ ⎞−

= ⎢ ⎥⎜⎜ ⎟⎟⎢ ⎥⎝ ⎠⎣ ⎦ (5.33)

A careful examination of the solution given by Eq.(5.33) shows that it is

related to the cumulative normal probability distribution with mean xv t ,

variance 2 and standard deviation LD t 2 LD t . Eq.(5.33) in dimensionless

form is

( ) 1,2

2

D DD D D

D

Pe

x tC x t erfct

N

⎡ ⎤⎛ ⎞⎢ ⎥⎜ ⎟

−⎢ ⎥⎜= ⎟⎢ ⎥⎜ ⎟⎢ ⎥⎜ ⎟⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(5.34)

or

( ) 1,2 2

D DD D D Pe

D

x tC x t erfc Nt

⎡ ⎤⎛ ⎞−= ⎢ ⎥⎜⎜ ⎟⎟⎢ ⎥⎝ ⎠⎣ ⎦

(5.35)

5–30

Page 549: +Peters Ekwere j. - Petrophysics

Eq.(5.34) or (5.35) also is related to the cumulative normal probability

distribution with mean Dt , variance 2 D

Pe

tN

and standard deviation 2 D

Pe

tN

.

Figure 5.17 shows the evolution of the solution predicted by

Eq.(5.34) or (5.35) for a Peclet Number of 100. It should be observed that

the solution at each dimensionless time is given by

(5.36) ( ), 1 Pr D D DC x t Cumulative Normal obability Distribution= −

with mean Dt and standard deviation 2 D

Pe

tN

. The shape of the solution is

that of an inverted S. Note also that because of dispersion, the

breakthrough time is less than 1 pore volume injected. The concentration

CD = 0.5 travels at the advection speed all times. Although it has not be

stated explicitly, the solution given by Eq.(5.34) or (5.35) is valid for an

idealized system under very restrictive conditions. The conditions are (1)

homogeneous and isotropic porous medium, (2) the injected fluid has the

same density as the displaced fluid, and (3) the injected fluid has the

same viscosity as the displaced fluid. These conditions preclude

channeling due to gross permeability heterogeneity, gravity override or

gravity tonguing and viscous fingering, all of which phenomena are not

included in the convection-dispersion equation. If these restrictions are

met, then Eq.(5.34) or (5.35) gives a good representation of the solvent

concentration profiles observed in laboratory experiments. For example,

although the experiment shown in Figures 5.1 and 5.2 does not quite

meet all the restrictions because of the small gravity override,

nevertheless Eq.(5.34) or (5.35) can be used to approximately reproduce

5-31

Page 550: +Peters Ekwere j. - Petrophysics

the laboratory measured solvent concentration profiles of Figure 5.1.

Figure 5.18 shows the result of a first attempt to reproduce the solvent

concentration profiles of Figure 5.1 using Eq.(5.34). The agreement

between laboratory measured profiles and those predicted by Eq.(5.34) is

good. Of course, the experimental profiles are distorted by the small

gravity override observed in the experiment. As a result, the experimental

profiles and the calculated profiles do not match exactly at the

displacement front. There is also evidence of retardation in the

experiment as the experimental profiles seem to lag behind the

calculated profiles.

Let us focus on the solution at a dimensionless time of 0.50 pore

volume injected shown in Figure 5.19. Marked on the figure are the

Figure 5.17. Solutions of the convection-dispersion equation for a Peclet Number of 100.

5–32

Page 551: +Peters Ekwere j. - Petrophysics

Figure 5.18. A comparison of the simulated and the measured solvent concentration profiles for the miscible displacement of Figure 5.1 and

5.2.

advection front, the dimensionless mixing zone length between CD = 0.90

and 0.10 and the dimensionless mixing zone length equal to twice the

standard deviation between CD = 0.8413 and 0.1588. It should be

observed that at tD = 0.5 pore volume injected, the advection front has

traveled exactly half the distance of the porous medium. This is as it

should be because the advection front travels at the advection speed or

at the interstitial velocity. Both the mixing zone length and the length

that represents twice the standard deviation grow in proportion to Dt ,

although this is not obvious from looking at the solution at one time. In

5-33

Page 552: +Peters Ekwere j. - Petrophysics

fact, it can be shown that the dimensionless mixing length between CD =

0.90 and CD = 0.10 is given by

3.625 DD

Pe

txN

Δ = (5.37)

or in dimensional form by

3.625 Lx D tΔ = (5.38)

Eq.(5.38) shows that if the mixing zone length can be measured as a

function of time during an experiment, then a graph of Δx versus

t could be used to determine the longitudinal dispersion coefficient.

Figure 5.19. Solution of the convection-dispersion equation for a Peclet Number of 100 at tD = 0.50 pore volume injected.

5–34

Page 553: +Peters Ekwere j. - Petrophysics

Let us replot the solutions of Figure 5.19 as CD versus z, where z is

defined as

D D

D

x tzt−

= (5.39)

A remarkable thing happens to the solutions at the various

dimensionless times. They collapse into one dimensionless curve that is

characteristic of the displacement as shown in Figure 5.20. The slope of

this curve reflects the Peclet Number or the longitudinal dispersion

coefficient. The steeper the curve, the higher is the Peclet Number or the

lower is the longitudinal dispersion coefficient. This transformation can

be used to test if the concentration profiles measured in an experiment

obey the complementary error function solution of the convection-

dispersion equation. The transformation was derived from Eq.(5.35),

which shows that CD is a function of the parameter NPe and the

independent variable D D

D

x tt− .

Figure 5.21 shows the effect of Peclet Number or longitudinal

dispersion coefficient on the solutions of the convection-dispersion

equation. At low Peclet Numbers or high longitudinal dispersion

coefficient, the mixing zone length is large whereas at high Peclet

Numbers or low longitudinal dispersion coefficient, the mixing zone

length is small.

5-35

Page 554: +Peters Ekwere j. - Petrophysics

Figure 5.20. Transformation of the solutions at a fixed Peclet Number.

5–36

Page 555: +Peters Ekwere j. - Petrophysics

Figure 5.21. The effect of Peclet Number or longitudinal dispersion coefficient on the solution of the convection-dispersion equation.

The breakthrough curve obtained by observing the concentration

at the outlet end of the porous medium (where 1Dx = ) for the case of a

finite length medium is given by

( ) 111,2 2

DD D Pe

D

tC t erfc Nt

⎡ ⎤⎛ ⎞−= ⎢ ⎥⎜⎜ ⎟⎟⎢ ⎥⎝ ⎠⎣ ⎦

(5.40)

Eq.(5.40) is the cumulative normal distribution with mean 1, variance

2

PeN and standard deviation 2

PeN. Figure 5.22 shows the breakthrough

curve given by Eq.(5.40) for a Peclet Number of 100. It is S-shaped and a

mirror image of the solvent concentration profiles. Thus, Eq.(5.40) can be

used in conjunction with the breakthrough curve measured in an

experiment to determine the Peclet Number ( )PeN and hence, the

longitudinal dispersion coefficient ( )LD .

5-37

Page 556: +Peters Ekwere j. - Petrophysics

Figure 5.22. Breakthrough curve.

Figure 5.23 shows the transformed breakthrough curve, where the

transformation equation is

1 D

D

tzt−

= (5.41)

It is similar to the solvent concentration profiles. This curve, if plotted on

a linear normal probability graph paper, will be a straight line as shown

in Figure 5.24. Such a plot can be used to fit the breakthrough curve to

the normal distribution for the purpose of determining the Peclet

Number, and hence, the longitudinal dispersion coefficient.

5–38

Page 557: +Peters Ekwere j. - Petrophysics

Figure 5.23. Transformed breakthrough curve.

5-39

Page 558: +Peters Ekwere j. - Petrophysics

Figure 5.24. Transformed breakthrough curve plotted on normal

probability scale.

5–40

Page 559: +Peters Ekwere j. - Petrophysics

In the presence of retardation due to adsorption, the solution given

by Eq.(5.34) can be modified to include the retardation factor as

( ) 1,2

2

DD

fD D D

D

f Pe

txR

C x t erfct

R N

⎡ ⎤⎛ ⎞−⎢ ⎥⎜ ⎟

⎢ ⎥⎜= ⎟⎢ ⎥⎜ ⎟⎢ ⎥⎜ ⎟⎜ ⎟⎢ ⎥⎝ ⎠⎣ ⎦

(5.42)

5.5 DISPERSION COEFFICIENT AND DISPERSIVITY

The dispersion coefficient is a second rank tensor. The dispersion

coefficient tensor in the xyz coordinate system is given by

( ), ,xx xy xz

yx yy yz

zx zy zz

D D DD x y z D D D

D D D

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(5.43)

in 3D, and by

( ), xx xy

yx yy

D DD x y

D D⎡ ⎤

= ⎢ ⎥⎣ ⎦

(5.44)

in 2D. The dispersion coefficient tensor like the permeability or hydraulic

conductivity tensor is symmetric. Using the same transformation

equations presented in Chapter 3 for permeability and hydraulic

conductivity, the principal values of the dispersion coefficient tensor and

the principal axes of the dispersion coefficient anisotropy can be

determined. Along the principal coordinates of the anisotropy uvw, the

dispersion coefficient tensor becomes

5-41

Page 560: +Peters Ekwere j. - Petrophysics

( )0 0

, , 0 00 0

L

T

W

DD u v w D

D

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(5.45)

in 3D, and

( )0

,0

L

T

DD u v

D⎡ ⎤

= ⎢ ⎥⎣ ⎦

(5.46)

in 2D, where DL is the longitudinal dispersion coefficient in the direction

of mean flow and DT is the transverse dispersion coefficient in the

direction perpendicular to the mean flow.

The relative magnitudes of DL and DT and the tensorial nature of

the dispersion can be demonstrated qualitatively by the following

experiment. Suppose a tracer is injected at time zero as a point source

into a homogeneous and isotropic reservoir in which there is steady flow

as shown in Figure 5.25. At times t1 and t2, the concentration of the

tracer would be as shown in Figure 5.25. The tracer has spread and the

concentration distribution has become elliptical. There is more spreading

or dispersion in the direction of mean flow than in the direction

perpendicular to it. Thus, the direction of mean flow and the direction

perpendicular to the mean flow are principal axes of the dispersion

coefficient anisotropy. The principal value of the dispersion coefficient

tensor in the direction of mean flow is DL and the one perpendicular to

the direction of mean flow is DT. In general, . L TD D≥

5–42

Page 561: +Peters Ekwere j. - Petrophysics

Figure 5.25. Variation in tracer concentration in 2D for a constant velocity flow system.

Using Eq.(5.6), the longitudinal and transverse dispersion

coefficients can be written as

L Ld LD D D m= + (5.47)

T Td TD D D m= + (5.48)

where the two terms on the right sides of Eqs.(5.46) and (5.47) represent

the contributions of molecular diffusion and mechanical dispersion to

the dispersion coefficients. Based on experimental observations,

Eqs.(5.47) and (5.48) may be written in dimensionless form as

1 2pL

o o

uDD C CD D

β⎛ ⎞

= + ⎜ ⎟⎝ ⎠

(5.49)

1 3pT

o o

uDD C CD D

β⎛ ⎞

= + ⎜ ⎟⎝ ⎠

(5.50)

where C1, C2, C3 and β are properties of the porous medium and the flow

regime, Do is the effective binary molecular diffusion coefficient between

the miscible displacing fluid and the displaced fluid, and Dp is the mean

5-43

Page 562: +Peters Ekwere j. - Petrophysics

particle diameter of the porous medium. It has been found also from

experiments that C1 is given by

11 1CFφ τ

= = (5.51)

where F is the formation resistivity factor and τ is the tortuosity of the

porous medium. Therefore, Eqs.(5.49) and (5.50) can be written as

21 pL

o o

uDD CD D

β

τ⎛ ⎞

= + ⎜ ⎟⎝ ⎠

(5.52)

31 pT

o o

uDD CD D

β

τ⎛ ⎞

= + ⎜ ⎟⎝ ⎠

(5.53)

Perkins and Johnson (1963) have shown that β is of the order of 1 to

1.25.

Figure 5.26 shows the correlations for L

o

DD

and T

o

DD

with the Peclet

Number, p

o

vDD

, for unconsolidated porous media obtained by Perkins and

Johnson (1963). Figure 5.26a shows that at Peclet Numbers less than

0.02, molecular diffusion dominates the longitudinal dispersion

coefficient and the dispersion coefficient is equal to the diffusion

coefficient in the porous medium. In this regime, the mechanical

dispersion term on the right side of Eq.(5.52) is negligible compared to

the molecular diffusion term. In this regime, the dimensionless molecular

term is about 0.67. This means that the diffusion coefficient for a tracer

in a porous medium is less than the diffusion coefficient in the same

5–44

Page 563: +Peters Ekwere j. - Petrophysics

liquid in the absence of the porous medium. This difference is caused by

the tortuosity of the porous medium. In fact, 1/0.67 or 1.5 is an estimate

of the tortuosity of unconsolidated porous media. At the transition Peclet

Numbers between 0.02 and 6, both molecular diffusion and mechanical

dispsersion contribute to the dispersion coefficient. At Peclet Numbers

above 6, the dispersion coefficient is dominated by mechanical dispersion

and the effect of molecular diffusion can be neglected.

The shape of the correlation for the dimensionless transverse

dispersion coefficient is similar to that of the dimensionless longitudinal

dispersion coefficient. However, it should be noted that the scale of the

Peclet Number in Figure 5.26b starts at 0.1 compared to 0.001 in Figure

5.26a. This means that the regime dominated by molecular diffusion

occurs over a larger range of Peclet Numbers for transverse dispersion

than for longitudinal dispersion. Thus, molecular diffusion is much more

important to transverse dispersion than to longitudinal dispersion. At

high Peclet Numbers, the longitudinal dispersion coefficient is greater

than the transverse dispersion coefficient. This is clearly evident by

plotting the two correlations for dimensionless dispersion coefficients on

the same scale as shown in Figure 5.27. Other authors have obtained

correlations for longitudinal dispersion coefficient similar to that of

Perkins and Johnston as shown in Figure 5.28.

At normal reservoir velocities, the Peclet Number is normally

greater than 6. Also, as pointed out earlier, β is approximately 1. Under

these conditions, molecular diffusion can be neglected and Eqs.(5.52)

and (5.53) can be approximated as

2p

Lo

uDD C D

D

β

o Luα⎛ ⎞

= ⎜ ⎟⎝ ⎠

≅ (5.54)

5-45

Page 564: +Peters Ekwere j. - Petrophysics

3p

To

uDD C D

D

β

o Tuα⎛ ⎞

= ⎜ ⎟⎝ ⎠

≅ (5.55)

where Lα is known as the longitudinal dispersivity and Tα is the

transverse dispersivity of the porous medium. Eqs.(5.54) and (5.55) can

be used to estimate the dispersivities if the dispersion coefficients have

been measured independently. In terms of the longitudinal dispersivity,

the Peclet Number defined in Eq.(5.25) becomes

Figure 5.26. Correlations for dimensionless longitudinal and transverse

5–46

Page 565: +Peters Ekwere j. - Petrophysics

dispersion coefficients. (a) dimensionless longitudinal dispersion coefficient, (b) transverse dispersion coefficient (Perkins and Johnston,

1963).

Figure 5.27. Correlations for dimensionless longitudinal and transverse dispersion coefficients plotted on the same scale (Perkins and Johnston,

1963).

5-47

Page 566: +Peters Ekwere j. - Petrophysics

Figure 5.28. Correlation for dimensionless longitudinal dispersion coefficient by various authors (Pfannkuch, 1963; Saffman, 1960).

PeL

LNα

= (5.56)

Dispersivity is, in general, not a function of fluid velocity, making it

a property of the porous medium. However, dispersivities are highly scale

dependent. Figures 5.29 and 5.30 show values of measured

dispersitivities as functions of the measurement scale. Note the

logarithmic scales. In general, dispersitivities measured at laboratory

scale are much smaller than those measured at field scale. Reservoir

heterogeneity is the cause of this scale effect. In field measurements, the

effect of heterogeneity is to stretch out the solvent concentration profile,

5–48

Page 567: +Peters Ekwere j. - Petrophysics

which is interpreted as a large dispersion coefficient or a large

dispersivity.

5-49

Page 568: +Peters Ekwere j. - Petrophysics

Figure 5.29. Longitudinal dispersivity versus measurement scale (Arya, 1988).

Figure 5.30. Field measured longitudinal dispersivity versus measurement scale (Gelhar, 1986).

5–50

Page 569: +Peters Ekwere j. - Petrophysics

5.6 MEASUREMENT OF DISPERSION COEFFICIENT AND DISPERSIVITY

5.6.1 Traditional Laboratory Method with Breakthrough Curve

A laboratory coreflood experiment can be used to measure the

longitudinal dispersion coefficient. Figure 5.31 shows a schematic

diagram for such a test. A tracer or solvent is used to displace a miscible

fluid from the porous medium and the solvent concentration of the

effluent at the outlet end of the core is measured as a function of time.

The figure at the upper left of Figure 5.31 shows the injected tracer or

solvent concentration as a function of time whereas the figure at the

upper right shows the concentration of the solvent in the effluent. The

injected solvent concentration is a step function in which pure solvent is

injected throughout the experiment. At the outflow end, initially, the

concentration of the injected solvent will be zero because it has not yet

broken through. After the injected solvent has arrived at the outlet end

ahead of the advection front because of dispersion, its concentration will

increase from zero to 1 over a finite period of time as shown in the

sketch. The curve of concentration versus time at the outlet end of the

core is known as the breakthrough curve and can be fitted to Eq.(5.40) to

estimate the Peclet Number and the longitudinal dispersion coefficient.

5-51

Page 570: +Peters Ekwere j. - Petrophysics

Figure 5.31. Laboratory coreflood experiment for measuring longitudinal dispersion coefficient.

In order for the solvent concentration distribution in the coreflood

to satisfy the mathematical model of Eq.(5.34) or (5.35), the restrictions

outlined previously must be implemented in the experiment. The core

should be fairly uniform, the density of the injected solvent and the

displaced fluid must be matched and the viscosity of the injected solvent

and the displaced fluid also must be matched. When these restrictions

are implemented, the breakthrough curve will approximate the

cumulative normal probability distribution. A graph of CD versus 1 D

D

tt−

on a linear probability graph paper will be a straight line as shown in

Figure 5.24. The mixing zone length between CD = 0.9 and CD = 0.1 read

from the graph is related to the Peclet Number as

0.9 0.1

1 13.625

D D

D D

Pe D DC C

t tN t t

= =

⎛ ⎞ ⎛ ⎞− −= −⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠

(5.57)

5–52

Page 571: +Peters Ekwere j. - Petrophysics

Eq.(5.57) can be rearranged to calculate the longitudinal dispersion

coefficient as

2

0.9 0.1

1 1

3.625D

D D

D DC CL

t tt t

D uL = D =

⎡ ⎤⎛ ⎞ ⎛ ⎞− −−⎢ ⎥⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠

= ⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

(5.58)

Alternatively, the mixing zone length corresponding to CD = 0.84 and CD

= 0.16 will be equal to twice the standard deviation of the normal

distribution. Thus,

0.84 0.16

1 122D D

D D

Pe D DC C

t tN t t

= =

⎛ ⎞ ⎛ ⎞− −= −⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠

(5.59)

Eq.(5.59) can be rearranged to calculate the longitudinal dispersion

coefficient as

2

0.84 0.16

1 18

D D

D DL

D DC C

t tuLDt t

= =

⎡ ⎤⎛ ⎞ ⎛ ⎞− −⎢ ⎥= −⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦ (5.60)

It is much more difficult to measure transverse dispersion coefficient

experimentally. As a result, there are very few transverse dispersion

coefficient data reported in the literature beside those of Perkins and

Johnson (1963).

5.6.2 Laboratory Method of Peters et al. (1996)

Peters et al. (1996) have presented a method for measuring

longitudinal dispersion coefficient and dispersivity by imaging the

5-53

Page 572: +Peters Ekwere j. - Petrophysics

laboratory tracer test. CT or NMR imaging of the coreflood gives the

solvent concentration profiles in time and space similar to those of Figure

5.17. These profiles can be fitted to the solution of the convection-

dispersion equation, Eq.(5.34) or (5.35), by trial and error to estimate the

Peclet Number and the longitudinal dispersion coefficient. The

longitudinal dispersion coefficient so determined will be the same as that

determined from the breakthrough curve. The dispersivity is calculated

from the dispersion coefficient using Eq.(5.54). Furthermore, the growth

of the mixing zone length with time can easily be calculated from the 3D

image data such that the graph of Δx versus t can be used to calculate

the longitudinal dispersion coefficient. The advantage of this approach is

that the contribution of heterogeneity to the dispersion coefficient also

can be measured.

To demonstrate the methods, we measured the longitudinal

dispersion coefficient and the longitudinal dispsersivity of the Berea

sandstone core of Figure 4.33 and the unconsolidated sandpack of

Figure 5.30 using these methods.

A summary of the experimental conditions for the tracer test in the

Berea sandstone core is presented in Table 5.1. Figure 5.32 shows the

CT images of the tracer test in the Berea sandstone. The growth of the

mixing zone is clearly visible. The mixing zone is distorted by

permeability heterogeneity of the core. The lower half of the core is more

permeable than the upper half. This is evident in the permeability image

of the core shown in Figure 5.33. The permeability distribution of the

core was determined by a technique described by Peters and Afzal (1991).

Figure 5.34 shows a preliminary history match of the solvent

concentration profiles using Eq.(5.34). The slopes of the profiles from

5–54

Page 573: +Peters Ekwere j. - Petrophysics

Eq.(5.34) and the experiment are in good agreement indicating that the

Peclet Number is correct. However, at late dimensionless times, the

profiles obtained from Eq.(5.34) have traveled further than those from

the experiment. This is evidence of retardation caused by adsorption.

Figure 5.35 shows the history match using Eq.(5.42) with a retardation

factor, Rf, of 1.11. The agreement between the two sets of solvent

concentration profiles is good. The parameters that resulted in this

history match are , and 159PeN = 5 2600 10 /LD x cm−= s 0.379L cmα = .These are

the longitudinal dispersion coefficient and longitudinal dispersivity that

would have been obtained from a breakthrough curve. It should be noted

that the effect of the distortion in the mixing zone caused by permeability

heterogeneity is interpreted as increased dispersion coefficient and

increased dispersivity of the porous medium.

Table 5.1. Experimental Conditions for Tracer Test in Berea Sandstone

Type of Displacement First-Contact Miscible

Porous Medium Berea Sandstone

Length (cm) 60.2

Diameter (cm) 5.1

Absolute Permeability (md) 160.4

Average Porosity from CT (%) 17.3

Fluids

Displacing Fluid Distilled Water + 10% NaI

Density of Displacing Fluid (g/cm3) 1.078

Viscosity of Displacing Fluid (cp) 1.029

Displaced Fluid Distilled Water+1.4% NaCl + 10% KCl

Density of Displaced Fluid (g/cm3) 1.078

Viscosity of Displaced Fluid (cp) 1.028

Mobility Ratio (Viscosity Ratio) 1.0

Darcy Velocity (cm/s) 2.742x10-3

5-55

Page 574: +Peters Ekwere j. - Petrophysics

Interstitial Velocity (cm/s) 1.714x10-2

Figure 5.32. CT images of a tracer test in a Berea sandstone core. A: tD = 0.20; B: tD = 0.50; C: tD = 0.80 (Peters et al., 1996)

5–56

Page 575: +Peters Ekwere j. - Petrophysics

Figure 5.33. Permeability distribution for Berea sandstone core (Peters and Afzal, 1991).

Figure 5.36 shows the growth of the mixing zone length with time

for the Berea sandstone core. The mixing zone length was obtained for

the 3D CT data by calculating the distances between CD = 0.90 and CD =

0.10 for each dimensionless time. This resulted in several thousand

values of mixing zone length at each dimensionless time. The several

thousand values were then averaged to obtain one mixing zone length for

each time. Each point in Figure 5.36 is the average mixing zone length

plotted against the corresponding time. It can be seen that the mixing

zone length grows as the square root of time as predicted by Eq.(5.38).

The longitudinal dispersion coefficient and longitudinal dispersivity are

calculated from the slope of the straight line of Figure 5.36 as

5-57

Page 576: +Peters Ekwere j. - Petrophysics

5 2431 10 /LD x cm−= s and 0.272 L cmα = . It should be noted that with this

method, the dispersion coefficient and the dispersivity are less than

calculated previously because the effect of the permeability heterogeneity

was excluded from the calculations.

Figure 5.34. A comparison of the simulated and experimental solvent concentration profiles for Berea sandstone core for Rf = 0 (Peters et al.,

1996).

5–58

Page 577: +Peters Ekwere j. - Petrophysics

Figure 5.35. A comparison of the simulated and experimental solvent concentration profiles for Berea sandstone core for Rf = 1.11 (Peters et

al., 1996).

5-59

Page 578: +Peters Ekwere j. - Petrophysics

Figure 5.36. Growth of mixing zone length with time for the Berea sandstone core (Peters et al., 1996).

A summary of the experiment for the unconsolidated sandpack is

presented in Table 5.2. Figure 5.37 shows the CT images of the tracer

test in the sandpack. The growth of the mixing zone also is clearly

visible. The mixing zone is distorted by the heterogeneity of the

sandpack. It should be recalled that this sandpack contained radial

heterogeneities based on the packing method that was used in preparing

it. Figure 5.38 shows the permeability distribution for the sandpack.

Figure 5.39 shows a preliminary history match of the solvent

concentration profiles using Eq.(5.34). The slopes of the profiles from

Eq.(5.34) and the experiment are in good agreement indicating that the

Peclet Number is correct. However, at late dimensionless times, the

profiles obtained from Eq.(5.34) have traveled further than those from

the experiment. There is retardation caused by adsorption in the

sandpack. Figure 5.40 shows the history match using Eq.(5.42) with a

retardation factor of 1.04. The agreement between the two sets of solvent

concentration profiles is good. The parameters that resulted in this

history match are , and 554PeN = 5 2100 10 /LD x cm−= s 0.098 L cmα = .

Figure 5.41 shows the growth of the mixing zone length with time

for the sandpack. The longitudinal dispersion coefficient and longitudinal

dispersivity are calculated from the mixing zone length as

and 5 282 10 /LD x cm−= s 0.080 L cmα = .

Figure 5.42 shows the similarity transformations of the solvent

concentration profiles from the two tracer tests. It can be seen that there

is more dispersion in the Berea sandstone core than in the

5–60

Page 579: +Peters Ekwere j. - Petrophysics

unconsolidated sandpack. The results of the tracer tests for both porous

media are summarized in Table 5.3.

Table 5.2. Experimental Conditions for Tracer Test in Unconsolidated

Sandpack

Type of Displacement First-Contact Miscible

Porous Medium Unconsolidated Sandpack

Length (cm) 54.2

Diameter (cm) 4.8

Absolute Permeability (md) 6400

Average Porosity from CT (%) 29.7

Fluids

Displacing Fluid Distilled Water + 13% NaCl

Density of Displacing Fluid

(g/cm3) 1.089

Viscosity of Displacing Fluid (cp) 1.262

Displaced Fluid Distilled Water + 10% BaCl2

Density of Displaced Fluid (g/cm3) 1.089

Viscosity of Displaced Fluid (cp) 1.127

Mobility Ratio (Viscosity Ratio) 0.9

Darcy Velocity (cm/s) 3.037x10-3

Interstitial Velocity (cm/s) 1.023x10-2

Breakthrough Recovery (%) 95

5-61

Page 580: +Peters Ekwere j. - Petrophysics

Figure 5.37. CT images of a tracer test in unconsolidated sandpack. A: tD = 0.20; B: tD = 0.50; C: tD = 0.80 (Peters et al., 1996)

5–62

Page 581: +Peters Ekwere j. - Petrophysics

Figure 5.38. Permeability distribution for unconsolidated sandpack (Peters and Afzal, 1991).

5-63

Page 582: +Peters Ekwere j. - Petrophysics

Figure 5.39. A comparison of the simulated and experimental solvent concentration profiles for unconsolidated sandpack for Rf = 0 (Peters et

al., 1996).

5–64

Page 583: +Peters Ekwere j. - Petrophysics

Figure 5.40. A comparison of the simulated and experimental solvent concentration profiles for unconsolidated sandpack for Rf = 1.04 (Peters

et al., 1996).

5-65

Page 584: +Peters Ekwere j. - Petrophysics

Figure 5.41. Growth of mixing zone length with time for unconsolidated sandpack (Peters et al., 1996).

5–66

Page 585: +Peters Ekwere j. - Petrophysics

Figure 5.42. Similarity transformation of solvent concentration profiles (Peters et al., 1996).

Table 5.3. Summary of Results

5-67

Page 586: +Peters Ekwere j. - Petrophysics

Porous Medium Berea

Sandstone

Unconsolidated

Sandpack

Longitudinal Dispersion Coefficient

with

Heterogeneity (cm2/s) 600x10-5 100x10-5

Longitudinal Dispersivity with

Heterogeneity (cm) 0.379 0.098

Longitudinal Dispersion Coefficient

without Heterogeneity (cm2/s) 431x10-5 82x10-5

Longitudinal Dispersivity without

Heterogeneity (cm) 0.272 0.080

Retardation Factor 1.11 1.04

Peclet Number 159 554

5.6.3 Field Measurement of Dispersion Coefficient and Dispersivity

Longitudinal dispersion coefficient and dispersivity are measured

in the field by tracer tests either in a single well or between two or more

wells. Eq.(5.9) in radial coordinates can be used to calculate the

dispersion coefficient and the dispersitivity from a single-well tracer test.

Eq.(5.9) can be written in radial coordinates as

2

2 0r L rC C Cu ut r r

α∂ ∂ ∂+ −

∂ ∂ ∂= (5.61)

Gelhar and Collins (1971) give the solution of Eq.(5.61) with appropriate

initial and boundary conditions for a single-well tracer test as

5–68

Page 587: +Peters Ekwere j. - Petrophysics

( )1

1 22

163

112

2 1 1

D DiD

L D D

Di Di

t tC erfc

t tR t tα

⎡ ⎤⎢ ⎥⎢ ⎥

− −⎢ ⎥= ⎢ ⎥⎧ ⎫⎡ ⎤⎢ ⎥⎛ ⎞ ⎛ ⎞⎪ ⎪⎛ ⎞ − − −⎨ ⎬⎢ ⎥⎢ ⎥⎜ ⎟ ⎜ ⎟⎜ ⎟⎝ ⎠ ⎝ ⎠ ⎝ ⎠⎣ ⎦⎪ ⎪⎢ ⎥⎩ ⎭⎣ ⎦

(5.62)

where

qtRhπ φ

= (5.63)

In Eq.(5.62), Dt is the pore volume of fluid produced at various times, Dit is

the total pore volume of tracer fluid injected at the beginning of the test.

The longitudinal dispersitivity can be calculated by fitting Eq.(5.62) to the

concentration profile obtained from the tracer test. Figure 5.43 shows the

fit of Eq.(5.62) to two tracer tests in the same well in a water-bearing

aquifer. The aquifer was 8.2 meters thick, with an average hydraulic

conductivity of 1.4x10-2 cm/s and a porosity of 38%. In Test SW1, the

volume of the tracer fluid injected was such that the tracer fluid

extended to a radius of 3.13 meters from the well. In Test SW2, the

volume of tracer fluid was such that the tracer fluid extended to a radius

of 4.99 meters. Thus, the two tests had different measurement scales.

Based on the history match shown in Figure 5.42, the dispersitivity of

the aquifer from the first test was 3.0 cm and that from the second test

was 9.0 cm. The results show the dependence of the dispersivity on the

scale of the measurement. In Test SW1, the scale of measurement was

3.13 meters whereas in Test SW2, it was 4.99 meters. The two scales of

measurement resulted in two different estimates of dispersivity with the

larger measurement scale resulting in a larger dispersitivity than the

smaller measurement scale.

5-69

Page 588: +Peters Ekwere j. - Petrophysics

Figure 5.43. Comparison of tracer concentration and Eq.(5.62)for single-well injection-withdrawal test (Pickens and Grisak, 1981).

A multi-well tracer test also can be used to estimate dispersivity in

the field. However, in this case, a numerical simulator will be needed to

5–70

Page 589: +Peters Ekwere j. - Petrophysics

interpret the test. Figure 5.44 shows attempts to interpret a two-well

tracer test using a finite element numerical simulator. In the test, a slug

of tracer was injected into one well and chased by water while the tracer

concentration was measured as a function of time at the second well. A

2D homogeneous aquifer model gave a longitudinal dispersitivity of 4.0 m

whereas a 3D heterogeneous aquifer model gave a much smaller

dispersivity of 0.15 m. Of course, the history match of the breakthrough

curve for a heterogeneous aquifer does not give a unique solution to the

problem. Different configurations of the aquifer heterogeneity can result

in good history matches with widely different values of dispersivity.

Figure 5.44. A comparison of three model fits to the breakthrough data from a two-well tracer test (Huyakorn et al., 1986).

5-71

Page 590: +Peters Ekwere j. - Petrophysics

5.7 FACTORS THAT COULD AFFECT DISPERSION COEFFICIENT AND DISPERSIVITY

A poll was taken in a previous class in which the students present

were requested to name the factors they believed could possibly affect the

dispersion coefficient and the dispersivity of a porous medium. There

were no restrictions. All factors that came to mind were listed. Here are

the factors that were listed, organized into rock characteristics, fluid

properties and process characteristics.

1. Rock Characteristics

Porosity (φ)

Permeability (k)

Pore size (δ)

Pore size distribution (frequency)

Heterogeneity

Dykstra–Parsons coefficient (V)

Variogram (γ)

Correlation length (range of influence, a)

Chemical reaction

Adsorption (retardation factor, Rf)

Biological transformation

Radioactive decay

Morphology of the porous medium (Γ)

5–72

Page 591: +Peters Ekwere j. - Petrophysics

Pore structure

Cementation

Dead end pores

Tortuosity

Mean grain diameter (Dp)

Specific surface area (S)

2. Fluid Properties

Viscosity of the displaced fluid (μo)

Viscosity of the displacing fluid (μs)

Density of the displaced fluid (ρo)

Density of the displacing fluid (ρs)

Binary diffusion coefficient between the displaced and displacing fluids (Do)

3. Process (Displacement) Characteristics

Interstitial velocity field (u )

Gravitational acceleration field ( g )

After extensive discussion, it was decided that the longitudinal

dispersion coefficient is a function of some of these variables and that the

functional relationship is of the form

( )1 , , , , ,L s p o oD f D u D gμ μ ρ= Δ (5.64)

The objective is to design an experimental program to determine the

nature of the function f1. To plan the experimental program, dimensional

5-73

Page 592: +Peters Ekwere j. - Petrophysics

analysis can be used to derive the set of complete and independent

dimensionless groups that can be used in the experimental program to

determine the nature of f1. Upon performing the dimensional analysis as

presented in Appendix A, Eq.(5.64) can be written in dimensionless form

as

3

2 , ,p poL

o o s o

uD D gD fD D D s

ρμμ μ

⎛ ⎞Δ= ⎜⎜

⎝ ⎠⎟⎟ (5.65)

where f2 is a new function to be determined from experiments. In the

laboratory experiments for determining the longitudinal dispersion

coefficient, the viscosity of the injected fluid should be matched with that

of the displaced fluid. If this is done, the mobility ratio, o

s

μμ

⎛ ⎞⎜⎝ ⎠

⎟ , will be a

constant equal to 1 and its effect will be eliminated from the experiment.

Also, the density of the injected fluid should be matched with that of the

displaced fluid. In this case, the dimensionless group, 3p

o s

D gD

ρμ

⎛ ⎞Δ⎜⎜⎝ ⎠

⎟⎟ , will be a

constant equal to zero and it will not be a factor in the experiment. If

these restrictions are implemented in the experiments, then Eq.(5.65)

becomes

3pL

o o

uDD fD D

⎛ ⎞= ⎜

⎝ ⎠⎟ (5.66)

A similar analysis gives the functional relationship for the transverse

dispersion coefficient as

4pT

o o

uDD fD D

⎛ ⎞= ⎜

⎝ ⎠⎟ (5.67)

5–74

Page 593: +Peters Ekwere j. - Petrophysics

The experimental program will then consist of measuring L

o

DD

and T

o

DD

as

functions of the Peclet Number, p

o

uDD

, to determine the nature of f3 and f4.

The results of such an experimental program are shown in Figures 5.27

and 5.28.

5.8 NUMERICAL MODELING OF FIRST-CONTACT MISCIBLE DISPLACEMENT

5.8.1 Introduction

First-contact miscible displacement is sufficiently well understood

to be modeled accurately with a finite difference numerical simulator.

Such a numerical simulator can be used to interpret tracer tests or to

model field processes that can be adequately described as first-contact

miscible processes.

5.8.2 Mathematical Model of First-Contact Miscible Displacement

The mathematical model for first-contact miscible displacement in

a heterogeneous and anisotropic reservoir consists of the following mass

transport, continuity, flow and mixing equations:

( ) .( ) .( ) 0C vC D Ctφ φ∂

+∇ −∇ ∇ =∂

(5.68)

( ) .( ) 0vtφρ ρ∂

+∇ =∂

(5.69)

(kv P ρμ

= − ∇ ± )gz (5.70)

5-75

Page 594: +Peters Ekwere j. - Petrophysics

(1 )mixture s s o sC Cρ ρ ρ= + − (5.71)

4

1/ 4 1/ 4

1s smixture

s o

C Cμμ μ

−⎛ ⎞−

= +⎜⎝ ⎠

⎟ (5.72)

where

xx xy xz

yx yy yz

zx zy zz

D D DD D D D

D D D

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(5.73)

and

xx xy xz

yx yy yz

zx zy zz

k k kk k k k

k k k

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(5.74)

Eq.(5.68) is the convection-dispersion equation for a heterogeneous

reservoir. It is a modified version of Eq.(5.9), which was for a

homogeneous reservoir with a constant porosity. Note that the advection

term in Eq.(5.68) is in terms of Darcy velocity instead of interstitial

velocity. Eq.(5.69) is the continuity equation for flow. Eq.(5.70) is Darcy's

law for a heterogeneous and anisotropic reservoir. Eq.(5.71) is the mixing

rule for density. It specifies how to calculate the fluid density as a

function of the solvent concentration. It is a linear mixing rule, which

has been verified experimentally. Figure 5.45 compares the linear mixing

rule with experimental density measurements for mixtures of glycerol

and brine. The agreement is good. Eq.(5.72) is the mixing rule for

viscosity. It specifies how the viscosity is to be calculated as a function of

the solvent concentration. It is known as the quarter-power viscosity

5–76

Page 595: +Peters Ekwere j. - Petrophysics

mixing rule. This mixing rule also has been verified in laboratory

measurements as shown in Figure 5.46 for glycerol and brine.

Figure 5.45. Verification of linear density mixing rule for two first-contact miscible fluids: Fluid 1 is brine; Fluid 2 is glycerol and water.

5-77

Page 596: +Peters Ekwere j. - Petrophysics

Figure 5.46. Verification of quarter-power viscosity mixing rule for two

first-contact miscible fluids: Fluid 1 is brine; Fluid 2 is glycerol and water.

The mathematical model for first-contact miscible displacement,

Eqs.(5.68) through (5.74), can be solved by finite difference using

appropriate initial and boundary conditions. A possible sequence for the

calculations at one time-step is as follows. Substitute Eq.(5.70) into

(5.69) to derive a generalized diffusivity equation for the pressure field.

Solve the pressure equation by finite difference. Use the pressure field

and Eq.(5.70) to calculate the Darcy velocity field. Substitute the Darcy

velocity field into Eq.(5.68) and solve Eq.(5.68) by finite difference to

calculate the solvent concentration distribution. Of course, the

properties of the porous medium and the fluids must be specified in

advance.

5–78

Page 597: +Peters Ekwere j. - Petrophysics

5.8.3 Numerical Modeling of Laboratory Experiments

In this section, we present the results of numerical simulations of

first-contact miscible displacements experiments based on the numerical

solution of Eqs.(5.68) through (5.74). The porous media used in the

experiments and the solvent concentration distributions in time and

space were imaged by NMR. Each simulation entailed the

characterization of the porous medium by constructing its permeability

distribution using Sequential Gaussian Simulation. The porosity

distribution was obtained directly from by NMR imaging. The pertinent

dimensionless groups for the displacements are mobility ratio (M), the

Peclet Number (NPe) as defined in Eq.(5.25), the gravity or buoyancy

number (Ng), and the density number (Nρ).

The gravity number is the ratio of the gravity force to the viscous

force and is given by

( )s og

o

k gN

uρ ρμ−

= (5.75)

The higher the gravity number, the higher the potential for gravity

segregation in the experiment. If the gravity number is negative, the

gravity segregation will be in the form of gravity override, whereas if the it

is positive, the gravity segregation will be in the form of gravity tonguing.

The density number is defined as

s o

o

Nρρ ρρ−

= (5.76)

5-79

Page 598: +Peters Ekwere j. - Petrophysics

The higher the density number, the higher the potential for gravity

segregation. A negative density number correlates with gravity override

whereas a positive density number correlates with gravity tonguing.

All the simulations were performed with UTCHEM, a finite

difference chemical flooding simulator developed at The University of

Texas at Austin. A 3D Cartesian coordinate system was used to simulate

the displacements in a cylindrical core. Figure 5.47 shows how the

rectangular grids were adapted to simulate a cylindrical system. All the

simulations were performed with 40x40x20 grid blocks.

Figure 5.47. Three-dimensional simulation grids.

Experiment 1

This experiment was conducted in homogeneous Berea sandstone

core at a favorable mobility ratio of 0.84 and a Darcy velocity of 0.00014

5–80

Page 599: +Peters Ekwere j. - Petrophysics

cm/s (0.40 ft/day). Figure 5.48 shows an artist impression of the core

along with the nomenclature used in the core characterization. The core

properties were: L = 10 cm, d = 5 cm, k = 622 md and φ = 24%. Figure

5.49 shows the porosity distribution of the core obtained by NMR

imaging. Figure 5.50 shows the NMR images of the solvent concentration

distributions in time and space. The mixing zone is tilted because of

gravity tonguing. Figure 5.51 shows the solvent concentration profiles for

the experiment.

Figure 5.48. Artist impression of homogeneous Berea sandstone core.

5-81

Page 600: +Peters Ekwere j. - Petrophysics

Figure 5.49. Porosity distribution for homogeneous Berea sandstone core (Majors et al., 1997).

5–82

Page 601: +Peters Ekwere j. - Petrophysics

Figure 5.50. Solvent concentration images for Experiment 1. M = 0.84, Ng = 0.3656, Nρ = 0.0899, NPe = 45.5 (Li, 1997)

5-83

Page 602: +Peters Ekwere j. - Petrophysics

Figure 5.51. Average solvent concentration profiles for Experiment 1. M = 0.84, Ng = 0.3656, Nρ = 0.0899, NPe = 45.5 (Li, 1997)

Figure 5.52 shows the permeability field generated for the Berea

sandstone core and used as the porous medium for the numerical

simulation. The parameters of the log normal permeability field are k =

622 md, φ = 24%, ax = 5.0 cm, ay = 2.5 cm, az = 0.25 cm and V = 0.20.

The experiment was simulated using 0.0975Lα = cm and 0.0031Tα = cm.

Figure 5.53 compares the simulated and the experimental solvent

concentration distributions. The agreement between the simulation and

the experiment is good. Figure 5.54 compares the simulated and

experimental solvent concentration profiles. The agreement between the

5–84

Page 603: +Peters Ekwere j. - Petrophysics

simulation and the experiment is acceptable. Finally, Figure 5.55 shows

a head to head comparison of the simulated and experimental solvent

concentrations at the same cross-sections of the core. The data are fairly

well distributed about the 45 degree line. Therefore, the agreement

between the simulation and the experiment is reasonable.

Figure 5.52. Permeability distribution for homogeneous Berea sandstone core obtained by Sequential Gaussian Simulation. k = 622 md, φ = 24%,

ax = 5.0 cm, ay = 2.5 cm, az = 0.25 cm, V = 0.20 (Shecaira and Peters, 1998).

5-85

Page 604: +Peters Ekwere j. - Petrophysics

Figure 5.53. A comparison of the simulated and experimental solvent concentration distributions for Experiment 1 (Shecaira and Peters,

1998).

5–86

Page 605: +Peters Ekwere j. - Petrophysics

Figure 5.54. A comparison of the simulated and experimental solvent concentration profiles for Experiment 1 (Shecaira and Peters, 1998).

Figure 5.56. A comparison of the simulated and experimental solvent

concentration at the same cross-sections for Experiment 1 (Shecaira and Peters, 1998).

5-87

Page 606: +Peters Ekwere j. - Petrophysics

Experiment 2

This experiment was conducted in the same Berea sandstone core

and at the same mobility ratio as Experiment 1. However, the gravity

number was reduced from 0.3656 to 0.0180 by increasing the Darcy

velocity from 0.00014 cm/s (0.40 ft/day) to 0.00294 cm/s (8.33 ft/day)

in an effort to eliminate the gravity tonguing observed in Experiment 1.

Figures 5.57 and 5.58 show the solvent concentration images and the

solvent concentration profiles for the experiment. The images of Figure

5.57 appear to show more tonguing than in Experiment 1.

5–88

Page 607: +Peters Ekwere j. - Petrophysics

Figure 5.57. Solvent concentration images for Experiment 2. M = 0.84, Ng = 0.0180, Nρ = 0.0899, NPe = 49.5 (Li, 1997)

5-89

Page 608: +Peters Ekwere j. - Petrophysics

Figure 5.58. Average solvent concentration profiles for Experiment 2. M = 0.84, Ng = 0.0180, Nρ = 0.0899, NPe = 49.5 (Li, 1997)

Figures 5.59 to 5.61 compare the initial simulation of Experiment

2 with the experiment. The agreement between the simulation and the

experiment is poor. The simulation model is the same as for Experiment

1, except that the injection rate was increased from 0.40 ft/day to 8.33

ft/day. The images from the simulation clearly show that at the gravity

number of 0.0180, there should be no gravity tonguing. Therefore, it can

be safely concluded that the tonguing in Experiment 2 was not caused by

5–90

Page 609: +Peters Ekwere j. - Petrophysics

gravity segregation but rather by a mechanical problem with the

experiment.

Figure 5.59. A comparison of the preliminary simulated and experimental solvent concentration distributions for Experiment 2 (Shecaira and

Peters, 1998).

5-91

Page 610: +Peters Ekwere j. - Petrophysics

Figure 5.60. A comparison of the preliminary simulated and experimental solvent concentration profiles for Experiment 2 (Shecaira and Peters,

1998).

Figure 5.61. A comparison of the preliminary simulated and experimental

solvent concentration at the same cross-sections for Experiment 2 (Shecaira and Peters, 1998).

5–92

Page 611: +Peters Ekwere j. - Petrophysics

Figure 5.62 shows the end piece of the core holder through which

fluid is injected into the core. Fluid is injected as a point source. Radial

grooves are machined on the face of the end piece to assist in

distributing the injected fluid uniformly over the inlet face of the core. It

would appear that at the injection rate of 0.40 ft/day in Experiment 2,

the grooves were effective in distributing the injected fluid over the inlet

face of the core. However, when the rate was increased to 8.33 ft/day,

the grooves became ineffective in distributing the injected fluid over the

inlet face of the core. As a result, the injected fluid was distributed

unevenly over the inlet face with more fluid being injected at the center of

the core than at the periphery of the core .

Figure 5.62. Schematic diagram of end piece of core holder showing fluid injection hole and grooves.

In order to test the fluid injection hypothesis, the inlet boundary

condition of the simulation model was modified to inject more fluid at the

center than at the periphery of the core as shown in Figure 5.62. Figures

5.63 to 5.65 show comparisons of the results of the simulation with the

modified inlet boundary condition and the experiment. The agreement

between the simulation and the experiment is excellent. The hypothesis

5-93

Page 612: +Peters Ekwere j. - Petrophysics

about the nonuniform fluid injection appears to be supported by the

simulation.

Figure 5.63. A comparison of the simulated and experimental solvent concentration distributions for Experiment 2 with a modified inlet

boundary condition (Shecaira and Peters, 1998).

5–94

Page 613: +Peters Ekwere j. - Petrophysics

Figure 5.64. A comparison of the simulated and experimental solvent concentration profiles for Experiment 2 with a modified inlet boundary

condition (Shecaira and Peters, 1998).

Figure 5.65. A comparison of the simulated and experimental solvent

concentration at the same cross-sections for Experiment 2 with a modified inlet boundary condition (Shecaira and Peters, 1998).

5-95

Page 614: +Peters Ekwere j. - Petrophysics

Experiment 3

This experiment was performed in the same Berea sandstone core

as Experiments 1 and 2 but at an adverse mobility ratio of 98 at the high

Darcy velocity of 0.00294 cm/s (8.33 ft/day). The combination of high

mobility ratio and high injection normally result is viscous fingering as is

evident in Figure 5.66, which shows the solvent concentration images for

this experiment. The solvent concentration profiles are shown in Figure

5.67.

Figures 5.68 to 5.70 compare the results of a preliminary

simulation with the experiment. There is no agreement between the two.

The preliminary simulation was based on a uniform permeability

distribution in the core and a uniform fluid injection at the core inlet.

The simulation was then refined by the using permeability distribution of

Figure 5.52 and the modified inlet boundary condition. More transverse

dispersion ( 0.0122Tα = cm) was included in the numerical model than in

the simulations of Experiments 1 and 2. Figures 5.71 to 5.73 compare

the results of the refined simulation with the experiment. The agreement

between the simulation and the experiment is excellent.

5–96

Page 615: +Peters Ekwere j. - Petrophysics

Figure 5.66. Solvent concentration images for Experiment 3. M = 98, Ng = -0.0002, Nρ = -0.097, NPe = 49.5 (Li, 1997)

5-97

Page 616: +Peters Ekwere j. - Petrophysics

Figure 5.67. Average solvent concentration images for Experiment 3. M = 98, Ng = -0.0002, Nρ = -0.097, NPe = 49.5 (Li, 1997)

5–98

Page 617: +Peters Ekwere j. - Petrophysics

Figure 5.68. A comparison of the preliminary simulated and experimental solvent concentration distributions for Experiment 3 (Shecaira and

Peters, 1998).

5-99

Page 618: +Peters Ekwere j. - Petrophysics

Figure 5.69. A comparison of the preliminary simulated and experimental solvent concentration profiles for Experiment 3 (Shecaira and Peters,

1998).

Figure 5.70. A comparison of the preliminary simulated and experimental

solvent concentration at the same cross-sections for Experiment 3 (Shecaira and Peters, 1998).

5–100

Page 619: +Peters Ekwere j. - Petrophysics

Figure 5.71. A comparison of the simulated and experimental solvent concentration distributions for Experiment 3 after refinement of the

simulation (Shecaira and Peters, 1998).

5-101

Page 620: +Peters Ekwere j. - Petrophysics

Figure 5.72. A comparison of the simulated and experimental solvent concentration profiles for Experiment 3 after refinement of the simulation

model (Shecaira and Peters, 1998).

Figure 5.73. A comparison of the simulated and experimental solvent

concentration at the same cross-sections for Experiment 3 after refinement of the simulation model (Shecaira and Peters, 1998).

5–102

Page 621: +Peters Ekwere j. - Petrophysics

Experiment 4

This experiment was performed in a layered Antolini sandstone

core at a favorable mobility ratio of 0.81 at a Darcy velocity of 0.00014

cm/s (0.40ft/day). Figure 5.74 shows the porosity images for sandstone

obtained by NMR. Figures 5.75 and 5.76 show the solvent concentration

images and the solvent concentration profiles for the experiment.

Using the porosity images as guidance, a three-layer permeability

distribution was generated for the sandstone core as shown in Figure

5.77. This distribution was used in the initial simulation of Experiment

4. Figures 5.78 to 5.80 compare the results of the simulation based on

the three-layer model with the experiment. The agreement between the

simulation and the experiment is fair. The permeability field was further

refined into a five-layer model as shown in Figure 5.81. Figures 5.82 to

5.84 show comparisons of the results of the simulation with the five-

layer model with the experiment. The agreement between the simulation

and the experiment is excellent.

5-103

Page 622: +Peters Ekwere j. - Petrophysics

Figure 5.74. Porosity images of layered Antolini sandstone core of Experiment 4 (Li, 1997).

5–104

Page 623: +Peters Ekwere j. - Petrophysics

Figure 5.75. Solvent concentration images for Experiment 4. M = 0.81, Ng = 0.0649, Nρ = 0.103, NPe = 46.8 (Li, 1997)

5-105

Page 624: +Peters Ekwere j. - Petrophysics

Figure 5.76. Average solvent concentration profiles for Experiment 4. M = 0.81, Ng = 0.0649, Nρ = 0.103, NPe = 46.8 (Li, 1997)

5–106

Page 625: +Peters Ekwere j. - Petrophysics

Figure 5.77. A three-layer model for the Antolini sandstone core of Experiment 4.

5-107

Page 626: +Peters Ekwere j. - Petrophysics

Figure 5.78. A comparison of the simulated and experimental solvent concentration distributions for Experiment 4 based on a three-layer

model (Shecaira and Peters, 1998).

5–108

Page 627: +Peters Ekwere j. - Petrophysics

Figure 5.79. A comparison of the simulated and experimental solvent concentration profiles for Experiment 4 based on a three-layer model

(Shecaira and Peters, 1998).

Figure 5.80. A comparison of the simulated and experimental solvent concentration at the same cross-sections for Experiment 4 based on a

three-layer model (Shecaira and Peters, 1998).

5-109

Page 628: +Peters Ekwere j. - Petrophysics

Figure 5.81. A Five-layer model for the Antolini sandstone core of Experiment 4.

5–110

Page 629: +Peters Ekwere j. - Petrophysics

Figure 5.82. A comparison of the simulated and experimental solvent concentration distributions for Experiment 4 based on a five-layer model

(Shecaira and Peters, 1998).

5-111

Page 630: +Peters Ekwere j. - Petrophysics

Figure 5.83. A comparison of the simulated and experimental solvent concentration profiles for Experiment 4 based on a five-layer model

(Shecaira and Peters, 1998).

Figure 5.84. A comparison of the simulated and experimental solvent concentration at the same cross-sections for Experiment 4 based on a

five-layer model (Shecaira and Peters, 1998).

5–112

Page 631: +Peters Ekwere j. - Petrophysics

Experiment 5

This experiment was performed in the same core as Experiment 4

but at an adverse mobility ratio of 11 and a Darcy velocity of 0.00014

cm/s (0.40 ft/day). Figures 5.85 and 5.86 show the solvent

concentration images and the solvent concentration profiles for the

experiment. The objective of the numerical simulation of this experiment

was to predict the outcome of the experiment instead of history matching

it. Having used the five-layer model to successfully simulate Experiment

4, it was decided to use the same simulation to predict Experiment 5 by

increasing the mobility ratio from 0.81 to 11.

Figures 5.87 to 5.89 show comparisons of the predicted

performance and the experiment. Figure 5.87 shows that predicted

solvent concentration images are in qualitative agreement with the

experimental images. However, Figures 5.88 and 5.89 show that the

quantitative prediction is not as good as might be inferred from Figure

5.87. This lack of accurate quantitative prediction is not surprising

because Experiment 5 is an unstable displacement and as such is

unpredictable.

5-113

Page 632: +Peters Ekwere j. - Petrophysics

Figure 5.85. Solvent concentration images for Experiment 5. M = 11, Ng = -0.0027, Nρ = -0.049, NPe = 46.8 (Li, 1997)

5–114

Page 633: +Peters Ekwere j. - Petrophysics

Figure 5.86. Average solvent concentration profiles for Experiment 5. M = 11, Ng = -0.0027, Nρ = -0.049, NPe = 46.8 (Li, 1997)

5-115

Page 634: +Peters Ekwere j. - Petrophysics

Figure 5.87. A comparison of the predicted and experimental solvent concentration distributions for Experiment 5 based on a five-layer model

(Shecaira and Peters, 1998).

5–116

Page 635: +Peters Ekwere j. - Petrophysics

Figure 5.88. A comparison of the predicted and experimental solvent concentration profiles for Experiment 5 based on a five-layer model

(Shecaira and Peters, 1998).

Figure 5.89. A comparison of the predicted and experimental solvent concentration at the same cross-sections for Experiment 5 based on a

five-layer model (Shecaira and Peters, 1998).

5-117

Page 636: +Peters Ekwere j. - Petrophysics

Experiment 6

The purpose of Experiment 6 and the accompanying numerical

simulation was to verify a new method for mapping the porosity and

permeability distributions in a heterogeneous core by NMR (Zuluaga et

al., 2000). A layered Antolini sandstone core was used for this

experiment. The core properties are L = 10 cm, d = 5 cm, φ = 12.8% and

k = 114 md. A technique was developed to map the 3D porosity and

permeability distributions in the core by NMR imaging. A technique was

developed to map the T1 distribution in a core. Permeability distribution

was calculated using the empirical equation of Kenyon et al. presented in

Chapter 2 as Eq.(2.69) and reproduced here for convenience as

(2.69) 41 NMRk C Tφ= 2

1

2T

The constant C1 was obtained by calibrating the NMR-derived

permeability to be equal to the permeability of the core in the direction of

flow measured by Darcy's law and found to equal to 180 μm2/s2. The

equation used to calculate the permeability of each voxel is

(5.77) 41180 NMRk φ=

Figure 5.90 shows a high resolution porosity image of the core. Dipping

layers are clearly visible in the image. Figure 5.91 shows the porosity

distribution after upscaling the high resolution porosity data

(512x512x512 voxels) to the low resolution numerical simulation grid of

40x40x40. The upscaling makes the porosity distribution somewhat

fuzzy. Figure 5.92 shows the upscaled permeability distribution. A high

permeability streak is clearly visible in the center of the core.

5–118

Page 637: +Peters Ekwere j. - Petrophysics

After measuring the porosity and permeability distributions, a

first-contact miscible displacement was conducted in the Antolini core at

a favorable mobility ratio of 0.81 at a Darcy velocity of 0.000692 cm/s

(1.96 ft/day). In an effort to verify the validity of the porosity and

permeability distributions derived by NMR, the displacement was

simulated using the NMR-derived porosity and permeability distributions

as the input data for the core description. No attempt was made to adjust

the porosity and permeability data to history-match the experiment.

Rather the porosity and permeability distributions were used as is in

order to verify their validity.

Figure 5.90. High resolution NMR porosity image of layered Antolini sandstone core (Zuluaga, 1999).

5-119

Page 638: +Peters Ekwere j. - Petrophysics

Figure 5.91. Upscaled porosity distribution from NMR imaging (Zuluaga et al., 2000)

Figure 5.92. Upscaled permeability distribution from NMR imaging

(Zuluaga et al., 2000)

5–120

Page 639: +Peters Ekwere j. - Petrophysics

Figure 5.93 shows a comparison of the solvent concentration

distribution from the experiment and the simulation. The agreement

between the simulation and the experiment is excellent thereby verifying

the validity of the porosity and permeability distributions obtained by

NMR imaging. Figure 5.94 shows the head to head comparison of the

experimental and simulated solvent concentration distributions at the

same cross-sections. The agreement between the experiment and the

simulation is good. The NMR-derived porosity and permeability

distributions are verified.

5-121

Page 640: +Peters Ekwere j. - Petrophysics

Figure 5.93. A comparison of the experimental and simulated solvent concentration distribution for Experiment 6 using NMR-derived porosity and permeability distributions. M= 0.81, Ng = 0.0160, Nρ = 0.1027, NPe =

133 (Zuluaga et al., 2000).

5–122

Page 641: +Peters Ekwere j. - Petrophysics

Figure 5.94. A comparison of the experimental and simulated solvent concentration at the same cross-sections for Experiment 6 using NMR-

derived porosity and permeability distributions. M= 0.81, Ng = 0.0160, Nρ = 0.1027, NPe = 133 (Zuluaga et al., 2000).

NOMENCLATURE

ax = correlation length in the x direction

ay = correlation length in the y direction

az = correlation length in the z direction

A = cross-sectional area of porous medium

C = solvent concentration

CD = dimensionless solvent concentration

Ci = initial solvent concentration

Cj = injected solvent concentration

D = dispersion coefficient tensor

mD = mechanical dispersion coefficient tensor

5-123

Page 642: +Peters Ekwere j. - Petrophysics

Dd = molecular diffusion coefficient

DL = longitudinal dispersion coefficient

Do = binary diffusion coefficient between the solvent and oil

Dp = mean grain diameter of the porous medium

erf = error function

erfc = complementary error function

F = formation resistivity factor

g = gravitational acceleration

J = total mass flux

aJ = mass flux vector due to advection

dJ = mass flux vector due to dispersion

k = permeability

L = length of porous medium

M = mobility ratio

Ng = gravity number

NPe = Peclet number

Nρ = density number

q = volumetric injection rate

Rf = retardation factor

S = specific surface area

t = time

tD = dimensionless time

u = interstitial velocity vector

v = Darcy velocity vector

V = Dykstra-Parson’s coefficient of permeability variation

xD = dimensionless distance

Lα = longitudinal dispersivity

Tα = transverse dispersivity

μo = viscosity of displaced fluid

μs = viscosity of displacing fluid

5–124

Page 643: +Peters Ekwere j. - Petrophysics

φ = porosity

λ = fluid mobility

τ = tortuosity

ρo = density of displaced fluid

ρs = density of displacing fluid

Δx = mixing zone length

Δρ = density difference

∇ = gradient operator

REFERENCES AND SUGGESTED READINGS

Afzal, N.: Investigation of Heterogeneities and Fluid Displacements in Porous Media Using CT, MS Thesis, The University of Texas at Austin, May 1992.

Aris, R.: “On the Dispersion of a Solute in a Fluid Flowing Through a Tube,” Proc. Roy. London A, 235 (1956) 67-78.

Arya, A., T. A. Hewtt, R.G. Larson and L.W. Lake: “Dispersion and Reservoir Heterogeneity,” SPERE (February 1988) 139-48.

Bachmat, Y.: “Basic Transport Coefficients as Aquifer Characteristics,” IASH (International Association of Scientific Hydrology) Symp. Hydrology of Fractured Rocks, Dubbrovnik 1, Publ.73 (1965) 63-75.

Bachmat, Y.: “On the Similitude of Dispersion Phenomena in Homogeneous and Isotropic Porous Medium,” Water Resources Res., 3, No. 4, (1967) 1079-83.

Bachmat, Y. and J. Bear: “The General Equations of Hydrodynamic Dispersion,” J. Geophys. Res., 69,No. 12 (1964) 2561-67.

Bear, J.: “On the Tensor Form of Dispersion,” J. Geophys. Res., 66, No. 4 (1961) 1185-97.

Bear, J. and Y. Bachmat: “A Generalized Theory on Hydrodynamic Dispersion in Porous Media,” IASH (International Association of Scientific Hydrology) Publ. 72 (1967) 7-16.

Bear, J.: Dynamics of Fluids in Porous Media, Elsevier, New York, 1972.

5-125

Page 644: +Peters Ekwere j. - Petrophysics

Boersma, L., F.T. Lindstrom and M.A. Barlaz: “Dispersion at the Field Scale Resulting from Spatial Variability of the Hydraulic Conductivity Field,” J. Hydrol., 159, No. 3 (1974) 145-67.

Bond, W.J.: “Velocity-Dependent Hydrodynamic Dispersion During Steady, Unsaturated Soil Water Flow: Experiments,” Water Resources Res., 22 , No. 13, (1986) 1881-89.

Bond, W.J. and D.E. Smiles: “Influence of Velocity on Hydrodynamic Dispersion During Unsteady Soil Water Flow,” Soil Sci. Soc. Am. J., 47, No. 3, (1983) 438-41.

Bosma, W.J. and S.E.A.T.M. van der Dee: “Dispersion of Continuously Injected, Non Linearly Adsorbing Solute in Chemically or Physically Heterogeneous Porous Formations,” J. Contam. Hydrol., 18 , No. 3 (1995) 181-98.

Bresler, E. and Dagan, G.: “Solute Transport in Unsaturated Heterogeneous Soils at Field Scale, II, Applications,” Soil Sci. Soc. Am. J., 43, No. 1 (1979) 467- 72.

Broman, W. H., Jr.: Experimental Verification of A Stability Theory for Miscible Displacements, MS Thesis, University of Texas at Austin, December 1985.

Butters, G.L., W.A. Jury and F.F. Ernst: “Field Scale Transport of Bromide in an Unsaturated Soil, 1, Experimental Methodology and Results,” Water Resources Res., 25, No. 7 (1989) 1575-81.

Carslaw, H.S. and Jaeger, J.C. : Conduction of Heat in Solids, 2nd edition, Clarendon Press, Oxford (1959).

Cavalero, S.: Visualization of Viscous Fingering in Miscible and Immiscible Displacements, MS Thesis, University of Texas at Austin, May 1990.

Clark, J.N., Shearin, H.M., Schultz, W.P., Garms, K. and Moore, J.L. : "Miscible Drive - It’s Theory and Application," JPT (June 1958) 11-20.

Collins, R.E.: Flow of Fluids Through Porous Materials, Research & Engineering Consultants Inc. (1990).

Crank, J.: The Mathematics of Diffusion, 2nd ed., Clarendon, Oxford, England (1979).

Cushman, J.H.: “On Diffusion in Fractal Porous Media,” Water Resources Res., 27, No. 4 (1991) 643-44.

Cushman, J.H. and R.T. Ginn: “On Dispersion in Fractal Porous Media,” Water Resources Res., 29, No. 10 (1993) 3513-15.

5–126

Page 645: +Peters Ekwere j. - Petrophysics

Dagan, G.: “Solute Transport in Heterogeneous Porous Formation,” J. Fluid Mech., 145 (1984) 151-77.

Dagan, G. and E. Bresler: “Solute Transport in Unsaturated Heterogeneous Soils at Field Scale, I, Theory,” Soil Sci. Am. J., 43, No. 1 (1979) 461-67.

Domenico, P.A. and Schwartz, F.W. : Physical and Chemical Hydrogeology, John Wiley and Sons, New York, 1990.

Dullien, F.A.L.: Porous Media - Fluid Transport and Pore Structure, Academic Press, New York, (1979).

Ellsworth, T.R. and W.A. Jury: “A Three-Dimensional Field Study of Solute Transport through Unsaturated Layers, Porous Media, 2, Characterization of Vertical Dispersion,” Water Resour. Res., 27 No. 5, 967-81 (1991).

Fetter, C.W. : Contaminant Hydrogeology, Prentice Hall, Upper Saddle River, New Jersey, 1993.

Freeze, R.A. and Cherry, J.A. : Groundwater, Prentice Hall, Englewood Cliffs, New Jersey, 1979.

Freyberg, D.L.: “A Natural Gradient Experiment on Solute Transport in a San Aquifer, 2, Spatial Moments and the Advection and Dispersion of Non reactive Tracers,” Water Resour. Res., 22, No. 1, 2030-46 (1986).

Fried, J.J. and Combarnous, M.A. :”Dispersion in Porous Media,” Advances in Hydroscience, Vol. 7 (1971) 169–282.

Frisch, H.L., T.T. Wang and T.K. Kwei: “Diffusion in Glassy Polymers, II”, J. Polymer Sci., Part A, 2, No. 7, 879-87 (1969).

Furtado, F., J. Glimm, B. Lindquist, F. Pereira and Q. Zhang: “Time Dependent Anomalous Diffusion for Flow in Multi-Fractal Porous Media,” Rep. SUNYSB-AMS-91-06, Dept. of Appl. Math. and Stat., State University of New York at Stony Brook (1991).

Furtado, F., J. Glimm, B. Lindquist, F. Pereira and Q. Zhang: “Flow in Multi-Fractal and Partially Layered Porous Media,” Report, 7 pp, Dept. of Appl. Math. and Stat., State University of New York at Stony Brook (19 91).

Gelhar, L.W. and Collins, M.A.: "General Analysis of Longitudinal Dispersion in Nonuniform Flow," Water Resources Research, 7, No.6 (1971) 1511-1521.

5-127

Page 646: +Peters Ekwere j. - Petrophysics

Gelhar, L.W. : " Stochastic Subsurface Hydrology From Theory to Application," Water Resources Research 22, No. 9 (1986) 135S-145S.

Gelhar, L.W. and C.L. Axness: “Three-Dimensional Stochastic Analysis of Macro dispersion in Aquifers,” Water Resour. Res., 19, No. 1, 161-80 (1983).

Gelhar, L.W., C. Welty and K.R. Rehfeldt: “A Critical Review of Data on Field-Scale Dispersion in Aquifers,” Water Resour. Res., 28, No. 7, 1955-74 (1992).

Gharbi, R., Peters, E.J., and Elkamel, A.: "Scaling Miscible Flood Displacements in Porous Media," Energy & Fuels, Vol. 12, No. 4, 1998, 801-811.

Gharbi, R. and Peters, E.J. : “Scaling Coreflood Experiments to Heterogeneous Reservoirs,” Journal of Petroleum Science and Engineering, 10, (1993) 83-95.

Gharbi, R.: Numerical Modeling of Fluid Displacements in Porous Media Assisted by Computed Tomography Imaging, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1993.

Giona, M.: “First Order Reaction-Diffusion Kinetics in Complex Fractal Media,” Chem Eng. Sci., 47, No. 6, 1503-15 (1992).

Guven, O., F.J. Molz and J.G. Melville: “An Analysis of Dispersion in a Stratified Aquifer,” Water Resour. Res., 20, No. 10, 1337-54 (1984).

Hardham, W. D.: Computerized Tomography Applied to the Visualization of Fluid Displacements, MS Thesis, University of Texas at Austin, December 1988.

Harleman, D.R.F. and R.R. Rumer: “Longitudinal and Lateral Dispersion in an Isotropic Porous Medium,” J. Fluid Mech., 16, No. 3, 385-94 (1963).

Hill, J.M. and J.N. Dewynne: Heat Conduction, Blackwell Sci., Cambridge, Mass. (1987).

Hoopes, J.A. and Harleman, D.R.F. : "Dispersion in Radial Flow From a Recharge Well," Jour. of Geophysical Research, 72, No. 14 (1967) 3595-3607.

Huyakorn, P., Andersen, P.F., Molz, F.J., Guven, O. and Melville, J.G. : “A Curvilinear Finite Element Model for Simulating Two-Well Tracer Tests and Transport in Stratified Aquifers at the Chalk River and Mobile Sites” Water Resources Research, Vol. 22, No. 7 (1986) 1016-1030.

5–128

Page 647: +Peters Ekwere j. - Petrophysics

Javandel, I., C. Doughty and C.F. Tsang: “Groundwater Transport: Handbook of Mathematical Models,” Water Resour. Monogr. Ser., 10, AGU, Washington, DC. (1984).

Jost, W.: Diffusion in Solids, Liquids, Gases, San Diego, Calif., (1952).

Kasap, E.: Numerical Simulation of Unstable Miscible Displacements by a Modified Finite Element Method, MS Thesis, University of Texas at Austin, December 1986.

Khataniar, S.: A Numerical Study of the Performance of Unstable Displacements in Heterogeneous Media, Ph.D. Dissertation, University of Texas at Austin, August 1991.

Khataniar, S. and E. J. Peters: “The Effect of Heterogeneity on the Performance of Unstable Displacements,” Journal of Petroleum Science and Engineering, 7, No. 3/4 (May 1992) 263-81.

Khataniar, S. and E. J. Peters: “A Comparison of the Finite Difference and Finite Element Methods for Simulating Unstable Displacements,” Journal of Petroleum Science and Engineering, 5, (1991) 205-218.

Knight, J. H.: Solute Transport and Dispersion: Commentary, in Flow and Transport in the Natural Environment: Advances and Applications, ed. W.L. Steffen and O.T. Denmead, pp 17-29, Springer-Verlag, New York (1988).

Knight, J. H. and J.R. Philip: “Exact Solution in Nonlinear Diffusion,” J. Eng. Math, 8, 219-27 (1992).

Korvin, G.: Fractal Models in the Earth Sciences, Elsevier, New York (1992).

Lake, L.W. : Enhanced Oil Recovery, Prentice Hall, Englewood Cliffs, New Jersey, 1989.

Lau, L.K., W.J. Kaufman and D.K. Todd: “Dispersion of a Water Tracer in a Radial, Laminar Flow through Homogeneous Porous Media,” Progr. Rep. 5, Canal Seepage Eng. Res. Lab., Univ. of Calif., Berkeley (July 1959).

Li, Ping: Nuclear Magnetic Resonance Imaging of Fluid Displacements in Porous Media, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1997.

Majors, P.D., Li, P. and Peters, E.J. :”NMR Imaging of Immiscible Displacements in Porous Media,” Society of Petroleum Engineers Formation Evaluation (September 1997) 164-169.

5-129

Page 648: +Peters Ekwere j. - Petrophysics

Marle, C.M.: Multiphase Flow in Porous Media, Gulf Publishing Company, Houston, Texas, (1981).

Mercado, A.A.: “A Note on Micro and Macrodispersion,” Ground Water, 22, No. 6, 790-91 (1984).

Molz, F.J., O. Guven and J.G. Melville, “An Examination of Scale-Dependent Dispersion Coefficients,” Ground Water, 21, No. 6, 715-25 (1983).

Muralidhar, R. and D. Ramkrishhna: “Diffusion in Pore Fractals: A Review of Linear Response Models,” Transp. Porous Media, 1, 79-95 (1993).

Neuman, S.P.: “Universal Scaling of Hydraulic Conductivities and Dispersivities in Geologic Media,” Water Resour. Res., 26, No. 8, 1749-58 (1990).

O’ Shaughnessy, B. and I. Procaccia: “Analytical Solutions for Diffusion on Fractal Objects,” Phys. Rev. Lett., 54, No. 5, 455-58 (1995).

Perkins, T.K. and Johnston, O.C. : “A Review of Diffusion and Dispersion in Porous Media,” Soc. Pet. Eng. J. (March 1963) 70-84.

Peters, E. J., J. A. Broman and W. H. Broman, Jr.: “Computer Image Processing: A New Tool for Studying Viscous Fingering in Corefloods,” SPE Reservoir Engineering (November 1987) 720-28.

Peters, E. J. and E. Kasap: “Simulation of Unstable Miscible Displacements by Finite Element Method,” SPE 15597, Proceedings of the 61st Annual Technical Conference of the Society of Petroleum Engineers (October 1986) New Orleans.

Peters, E. J. and W. D. Hardham: “A Comparison of Unstable Miscible and Immiscible Displacements,” SPE 19640, Proceedings of the 64th Annual Technical Conference of the Society of Petroleum Engineers (October 1989) San Antonio.

Peters, E.J., Gharbi, R. and Afzal, N. : “A Look at Dispersion in Porous Media Through Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 15 (1996) 23-31.

Peters, E.J. and Gharbi, R. : “Numerical Modeling of Laboratory Corefloods,” Journal of Petroleum Science and Engineering, 9, (1993) 207-221.

Peters, E.J. and Afzal, N. : “Characterization of Heterogeneities in Permeable Media with Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 7, No. 3/4, (May 1992) 283-296.

5–130

Page 649: +Peters Ekwere j. - Petrophysics

Peters, E.J. and Hardham, W.D. : “Visualization of Fluid Displacements in Porous Media Using Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 4, No. 2, (May 1990) 155-168.

Peters, E.J. and Reid, C.A. : “A Microcomputer-Based Imaging System for the Visualization of Fluid Displacements,” J. Pet. Tech. (May 1990) 558-563.

Philip, J.R.: “Sorption and Infiltration in Heterogeneous Media,” Aust. J. Soil Res., 5, 1-10 (1967).

Philip, J.R.: “Issues in Flow and Transport in Heterogeneous Porous Media,” Transp. Porous Media, 1, 319-38 (1986).

Pickens, J. F. and G.E. Grisak: “Scale-Dependent Dispersion in a Stratified Granular Aquifer,” Water Resour. Res., 17, No. 4, 1191-1211 (1981).

Pickens, J. F. and G.E. Grisak: “Modeling of Scale-Dependent Dispersion in Hydrogeologic Systems,” Water Resour. Res., 17, No. 6, 1701-11 (1981).

Reid, C.: Development of an Image Processing Workstation for the Quantitative Analysis of Fluid Displacements, MS Thesis, University of Texas at Austin, August 1988.

Saffman, P.G.: “A Theory of Dispersion in a Porous Medium,” J. Fluid Mech, 6, 321-49 (1959).

Saffman, P.G.: “Dispersion Due to Molecular Diffusion and Macroscopic Mixing in Flow Through a Network of Capillaries,” J. Fluid Mech., 7, 194-208 (1960).

Sahimi, M.: “Fractal and Superdiffusive Transport and Hydrodynamic Dispersion in Heterogeneous Porous Media, Transp. Porous Media, 13, 3-40 (1993).

Sauty, J.-P.: “An Analysis of Hydrodispersive Transfer in Aquifers,” Water Resour. Res., 16, No. 1, 145-58 (1980).

Scheidegger, A.E.: “General Theory of Dispersion in Porous Media,” J. Geophys. Res., 66, No. 10, 3273-78 (1961).

Scheidegger, A.E.: The Physics of Flow Through Porous Media, University of Toronto Press, Toronto, Canada (1960).

Shecaira, F.S. and Peters, E.J. : "Numerical Modeling of Miscible Displacements in Permeable Media Monitored by Imaging Techniques," Paper submitted to Computational Geosciences, 1998.

5-131

Page 650: +Peters Ekwere j. - Petrophysics

Shecaira, F.S.: Numerical Modeling of Miscible Displacements in Permeable Media Monitored by Imaging Techniques, PhD Dissertation, The University of Texas at Austin, Austin, Texas, May 1998.

Schulin, R., M.T. van Genuchten, H. Fluhler and P. Ferlin: “An Experimental Study of Solute Transport in a Stony Field Soil,” Water Resour. Res., 2, No. 9, 1785-95 (1987).

Schwartz, F.W.: “Macrodispersion in Porous Media: The Controlling Factors,” Water Resour. Res., 1, No. 4, 743-52 (1977).

Shouse, P.J., T.R. Ellsworth and J.A. Jobes: “Steady-State Infiltration as a function of Measurement Scale,” Soil Sci., 157, No. 3, 129-36 (1994).

Tang, D.H., E.O. Frind and E.A. Sudicky: “Contaminant Transport in Fractured Porous Media: Analytical Solutions for a Single Fracture,” Water Resour. Res., 17, No. 3, 555-64 (1981).

Taylor, G.: “Diffusion by Continuous Movements,” Proc. London Math Soc., 20, 196-211 (1981).

Taylor, G.: “Dispersion of Soluble Matter in Solvent Flowing Slowly Through a Tube,” Proc. R. Soc. London A, 219, 186-203 (1953).

Tyler, S.W. and S.W. Wheatcraft: “Reply to Comment by J.R. Philip on ‘An Explanation of Scale-Dependent Dispersivity in Heterogeneous Aquifers Using Concepts of Fractal Geometry,’” by S.W. Wheatcraft and S.W. Tyler, Water Resour. Res., 28, No. 5, 1487-90 (1992).

van Genuchten, M.T. and W.J. Alves: “Analytical Solutions of the One-Dimensional Convective-Dispersive Solute Transport Equation,” Tech. Bull. U.S. Dep. Agric., 1661 (1992).

van Wesenbeeck, J. and R.G. Kachanoski: “Spatial Scale Dependence of In Situ Solute Transport,” Soil Sci. Soc. Am. J., 55, No. 1, 3-7 (1991).

Wang, T.T., T.K. Kwei and H.L. Frisch: “Diffusion in Glassy Polymers, III,” J. Polymer Sci., Part A 2, 7, 2019-28 (1969).

Wheatcraft, S.W. and S.W. Tyler: “An Explanation of Scale-Dependent Dispersivity in Heterogeneous Aquifers Using Concepts of Fractal Geometry,” Water Resour. Res., 24, No. 4, 566-78 (1988).

Yeh, G.T.: “A Lagrangean-Eulerian Method with Zoomable Hidden Fine-Mesh Approach to Solving Advection-Dispersion Equations,” Water Resour. Res., 26, No. 6, 1133-44 (1990).

5–132

Page 651: +Peters Ekwere j. - Petrophysics

Zuluaga, E.: A Simulation Approach to Validate Petrophysical Data from Nuclear Magnetic Resonance Imaging, MS Thesis, University of Texas at Austin, August 1999.

Zuluaga, E., Majors, P.D. and Peters, E.J.: "A Simulation Approach to Validate Petrophysical Data from Nuclear Magnetic Resonance Imaging," SPE Journal, March 2002, 35-39.

5-133

Page 652: +Peters Ekwere j. - Petrophysics

PROBLEMS

5.1. A laboratory core, which is 6 cm in diameter and 40 cm in length, has a porosity of 0.35. The core is initially saturated with water. A tracer is then injected continuously into the core at the rate of 1000 cm3/hour. The effluent data shown in Table 5.4 were recorded at the outlet end of the core. C/Co is the relative tracer concentration. Determine the dispersion coefficient of the core.

Table 5.4. Effluent Data for Tracer Test of Problem 5.1.

Time

(hour)

C/Co

0.35 0.075

0.37 0.215

0.385 0.37

0.396 0.5

0.41 0.65

0.43 0.83

0.44 0.89

0.46 0.96 5.2. Table 5.5 gives the data for a one-dimensional core tracer test.

Determine the Peclet number for the tracer test.

Table 5.5. Data for Problem 5.2.

PV

Injected

(tD)

1 D

D

tt−

CD

0.60 0.516 0.010

0.65 0.434 0.015

0.70 0.359 0.037

0.80 0.224 0.066

0.90 0.105 0.300

1.00 0.000 0.502

5–134

Page 653: +Peters Ekwere j. - Petrophysics

1.10 -0.095 0.685

1.20 -0.183 0.820

1.30 -0.263 0.906

1.40 -0.338 0.988

1.50 -0.408 0.997 5.3. Fresh water at relative concentration C/Co = 0 is injected into a

sandpack saturated with salt water at relative concentration C/Co = 1. As the salt water is displaced the concentration measurements in Table 5.6 were made at a specific time t.

Table 5.6. Tracer Data for Problem 5.3.

Distance from Core Inlet

(cm)

C/Co

(%)

48.2 0.8

49.7 2.2

51.5 3.5

53.6 9.5

55.4 21.7

57.2 50.3

59.3 78.0

61.3 94.5

63.2 94.5

65.4 98.7

68.2 98.7

73.4 100 The average interstitial velocity of flow was 1.6 cm/minute. a. How long after the initiation of flow were these readings taken? b. Determine the dispersion coefficient and the dispersivity of the

sandpack.

5.4. The molecular diffusion coefficient of a porous medium is equal to 5x10-10 m2/s. If a tracer is placed in contact with the inlet end of the porous medium, estimate the relative tracer concentration C/Co at a distance of 5 m from the inlet after 100 years of diffusion. Comment on the effectiveness of molecular diffusion as a transport mechanism in a porous medium.

5-135

Page 654: +Peters Ekwere j. - Petrophysics

5.5. A tracer test is conducted in a relatively homogeneous cylindrical

sandpack using continuous injection of a nonreactive tracer. The injected tracer and the displaced liquid have the same density and

viscosity. The relative concentration C/Co of 0.75 was observed to arrive at the outlet end of the core after 0.9 hour from the start of injection. Other data about the test are:

Length of sandpack = 30 cm Diameter of sandpack = 10 cm

Volumetric injection rate = 1000 cm3/hour Porosity of the sandpack = 35% Hydraulic gradient = 0.1 Fluid viscosity = 1.0 cp

Fluid density = 1.0 gm/cm3

Gravitational acceleration = 981cm/s2

1 atmosphere = 1.0133x106 dynes/cm2 a. Calculate the dispersivity of the porous medium and state its

units. b. Calculate the permeability of the porous medium and state its

units. Problem 3 (15 points) A tracer test was conducted in a long core to determine the dispersivity of a porous medium using a tracer that had the same density and the same viscosity as the displaced liquid. The tracer test was imaged by CT at t = 50 minutes from the beginning of the test. The interstitial velocity for the test was 1.6 cm/minute. Figure 3 shows the 0.9 and 0.1 tracer concentration contours inside the porous medium obtained by imaging. a. Calculate the dispersivity of the porous medium and state its units. b. Is the core homogeneous? If yes, why? If no, why not? c. How would the dispersivity you calculated in part a compare with

the dispersivity obtained from the breakthrough curve in the same experiment?

5–136

Page 655: +Peters Ekwere j. - Petrophysics

Figure 3. Tracer concentration contours at t = 50 minutes.

Problem 3 (20 points)

A solution of nonabsorbing sodium iodide is injected continuously to

displace a brine solution from a core of length L, porosity φ, permeability

k and cross sectional area A at a constant volumetric rate q. The injected

and displaced fluids are miscible in all proportions and have the same

density and viscosity. Sodium iodide solution is used as a solvent in this

experiment because its high x-ray absorption coefficient permits the

solvent concentration profiles to be imaged by x-ray computed

tomography (CT scanning). Probes are installed at the inlet and outlet

5-137

Page 656: +Peters Ekwere j. - Petrophysics

ends of the core to monitor and report the solvent concentrations as

functions of time during the entire experiment.

a. Sketch the solvent concentration profile you would expect to see from the CT scanning after injecting 0.5 pore volume of the solvent. Show the inlet and outlet ends of the core on your sketch as well as the critical values of solvent concentration. Do your sketch on Figure 2.

b. Sketch the graphs of solvent concentration versus dimensionless time you expect from the probes at the inlet and outlet ends of the core. Superimpose the two graphs so that one can make a qualitative comparison of the two. Please label your graphs with inlet and outlet. Do your sketches on Figure 3.

c. Other than the absolute permeability, what transport property of the porous medium and fluids can be estimated from this experiment?

d. Predict the solvent concentration at the outlet end of the core at t = 108 minutes, given:

L = 30 cm q = 50 cm

3/hour

A = 20 cm2

φ = 0.15 DL = 400x10–5 cm

2/s

6. (20 points) A nonabsorbing tracer is pumped through a porous medium of length 30

cm at an interstitial velocity of 1x10-2 cm/s. A relative concentration of 0.42 was

measured in the effluent after 46.6 minutes from the start of the test. Determine the

dispersivity of the porous medium.

4. A tracer test is conducted in a relatively homogeneous cylindrical sandpack using

continuous injection of a nonreactive tracer. The injected tracer and the displaced

liquid have the same density and viscosity. The relative concentration C/Co of 0.75

5–138

Page 657: +Peters Ekwere j. - Petrophysics

was observed to arrive at the outlet end of the core after 0.9 hour from the start of

injection. Other data about the test are:

Length of sandpack = 30 cm

Diameter of sandpack = 10 cm

Volumetric injection rate = 1000 cm3/hour

Porosity of the sandpack = 35%

Hydraulic gradient = 0.1

Fluid viscosity = 1.0 cp

Fluid density = 1.0 gm/cm3

Gravitational acceleration = 981cm/s2

1 atmosphere = 1.0133x106 dynes/cm

2

a. Calculate the dispersivity of the porous medium. Please state the units of your

answer.

5-139

Page 658: +Peters Ekwere j. - Petrophysics

b. Calculate the permeability of the porous medium. Please state the units of

your answer.

6. (10 points) A tracer test was conducted in a long core to determine the dispersivity of a porous medium using a tracer that had the same density and the same viscosity as the displaced liquid. The tracer test was imaged by CT at t = 50 minutes from the beginning of the test. The interstitial velocity for the test was 1.6 cm/minute. Figure 2 shows the 0.9 and 0.1 tracer concentration contours inside the porous medium obtained by imaging.

(a) Calculate the dispersivity of the porous medium and state its

units. (b) Is the core homogeneous? If yes, why? If no, why not? 2. (20 points) A nonabsorbing tracer was injected into a 30 cm long core at a

constant interstitial velocity of 36 cm/hr. The relative concentrations measured at the outlet end of the core were as follows:

Time (Hour) C/Co 0.7 0.25 0.833 0.50 0.992 0.75

Determine the dispersivity of the core.

5–140

Page 659: +Peters Ekwere j. - Petrophysics

5-141

Figure 2. Tracer concentration contours at t = 50 minutes.

Page 660: +Peters Ekwere j. - Petrophysics

CHAPTER 6

INTERFACIAL PHENOMENA AND WETTABILITY

6.1 INTRODUCTION

The pore space in a petroleum reservoir rock is usually occupied by

more than one fluid. In an oil reservoir, water and oil occupy the pore space.

In a gas reservoir, water and gas occupy the pore space. In some oil

reservoirs, at some stage of depletion, water, oil and gas may occupy the pore

space.

When more than one fluid occupies the pore space of a porous medium,

new set of problems arise. Fluid saturations must be tracked. Interfacial

forces (surface forces) between the immiscible fluids and between the fluids

and the rock surface come into play. Because the pores are of capillary

dimensions, capillarity plays a role. Interfaces separate the fluids within the

pores giving rise to differences in fluid pressure between the phases (capillary

pressure) and differences in the flow capacity (relative permeability) of the

rock and fluids. Capillarity also ensures that an immiscible displacement can

never be complete. There is always a residual saturation of the displaced fluid

that is trapped by capillarity. Further, the rock surface can show a marked

affinity for one of the fluids. Such an affinity is characterized by the concept of

wettability. Interfacial phenomena and wettability are presented in this

6-1

Page 661: +Peters Ekwere j. - Petrophysics

chapter. Capillary pressure and relative permeability are presented in

subsequent chapters.

6.2. SURFACE AND INTERFACIAL TENSIONS

6.2.1 Surface Tension

Surface tension is the contractile force that exists at the interface of a

liquid and its vapor (or air). Surface tension makes the surface of a liquid

drop act like a membrane. The force is caused by unequal molecular

attractions of the fluid particles at the surface as shown in Figure 6.1. The

force per unit length (σ = Force/Length) tending to contract the surface of a

liquid is a measure of the surface tension of the liquid. It is a property of the

liquid and is usually expressed in units of dynes/cm.

Figure 6.1: Apparent surface film caused by unequal attraction of

surface molecules of a liquid.

If the forces acting on a molecule at the surface or interface are different

from those acting on a molecule in the body of the liquid, a new interface can

6-2

Page 662: +Peters Ekwere j. - Petrophysics

only be created if work is done. It can be shown (e.g., by soap film

experiment) that the reversible work required to create a new interface is given

by

W daδ σ= − (6.1)

where δW is reversible work required to create a new area, σ is surface

(interfacial) tension and da is the increase in area. The reversible work shown

in Eq.(6.1) corresponds to a free energy quantity. It is apparent from Eq.(6.1)

that surface (interfacial) tension also can be viewed as free surface energy per

unit area with units of erg per square centimeter. Because a system at

equilibrium minimizes its free energy, liquids at equilibrium tend to minimize

their surface area. A liquid jet tends to break up into spherical drops because

a sphere has the smallest surface area per unit volume.

The surface tension of pure water at 70 °F is 72.5 dynes/cm, and at

200 °F is 60.1 dynes/cm. The surface tensions of crude oils at 70 °F range

from 24 to 38 dynes/cm. High temperatures and dissolved gas both tend to

reduce the surface tension of crude oils. Values on the order of 1 dyne/cm

may be expected at temperatures and pressures exceeding 150 °F and 3,000

psig. Table 4.1 shows the surface tensions of some selected liquids.

Factors that affect the surface tension of a liquid include pressure,

temperature and solute concentration. An increase in pressure leads to a

reduction in the surface tension of a liquid. This is because pressure exerts a

compressive force on the surface which reduces the tensile or contractile

tendency of the surface. An increase in temperature leads to a reduction in

the surface tension of a liquid. The increase in temperature causes increased

randomness at the surface which leads to an increase in the surface entropy.

It can be shown from thermodynamics that the surface entropy is given by

6-3

Page 663: +Peters Ekwere j. - Petrophysics

s

P

STσ∂⎛ ⎞= −⎜ ⎟∂⎝ ⎠

(6.2)

where Ss is the surface entropy and T is the temperature. It is clear from

Eq.(6.2) that surface tension decreases with temperature. Figure 6.2 shows

the variation of surface tensions of hydrocarbons with temperature.

Table 4.1. Surface Tension of Pure Liquids

Liquid T (°C) σ (dynes/cm) Water 20 72.8 Water 25 72.0 Benzene 20 28.88 Benzene 25 28.22 Toluene 20 28.43 Carbon tetrachloride

20 26.9

Ethanol 20 22.39 n-Octane 20 21.8 Ethyl ether 20 17.01

The effect of solute concentration on the surface tension of a liquid

depends on the liquid and the nature of the solute. Four general cases may

be identified.

1. Liquids having fairly close values of surface tension. Generally, the

surface tension of the mixture varies approximately linearly with the

composition. For example, for a mixture of acetone and chloroform at

18° C, the surface tension increases linearly with composition (mole %

chloroform) from 22 dynes/cm for pure acetone to 27 dynes/cm for

pure chloroform.

6-4

Page 664: +Peters Ekwere j. - Petrophysics

Figure 6.2. Variation of surface tensions of hydrocarbons with temperature.

2. Liquids having widely different values of surface tension. In general,

the surface tension of a liquid is reduced substantially by addition of a

liquid of lower surface tension, but is only slightly increased by addition

of a liquid of higher surface tension. For example, addition of ethanol to

water causes a rapid reduction in the surface tension of water with

ethanol concentration. However, addition of water to benzene raises its

surface tension from 28.2 to only 29.3 dynes/cm.

3. Solutions of inorganic electrolytes. In general, the surface tension

increases with solute concentration. For example, the surface tension

6-5

Page 665: +Peters Ekwere j. - Petrophysics

of water at 20° C will be increased by addition of sodium chloride from

72.8 dynes/cm to 80 dynes/cm at a concentration of 5 moles of sodium

chloride per liter of solution.

4. Solutions of colloidal (long chain) electrolytes. In general, the surface

tension decreases with solute concentration but is followed by a region

over which the surface tension is virtually unchanged by solute

concentration. For example, the surface tension of water at 25° C will

be reduced by addition of sodium lauryl sulfate from 72 dynes/cm to 40

dynes/cm at a concentration of 0.01 moles per liter of solution. The

surface tension remains constant above this concentration.

A useful empirical relation often used to calculate surface tension is

through the concept of the parachor. The parachor for a pure substance is

defined as

14

L g

Mσρ ρ

Λ =−

(6.3)

where Λ is the parachor, M is the molecular weight of the liquid, σ is the

surface tension of the liquid in dynes/cm, ρL is the saturated liquid density in

g/cm3 and ρg is the saturated vapor density in g/cm3. Parachor has definite

values for specific atoms and structures. Parachors are predicted from the

structure of the molecules or can be calculated for pure substances and

mixtures from surface tension measurements at atmospheric pressure. The

parachors for pure substances are given in Table 6.2. Correlations for

parachors with molecular weight are shown in Figures 6.3 and 6.4. The

saturated liquid and vapor densities for various liquids are given in Figure

6.5. Equation (3.4) can be rearranged to calculate the surface tension as

6-6

Page 666: +Peters Ekwere j. - Petrophysics

4

L g

Mρ ρ

σ⎡ − ⎤⎛ ⎞

= Λ⎢ ⎥⎜⎝ ⎠

⎟⎣ ⎦

(6.4)

Table 4.2. Parachors for Computing Surface and Interfacial Tensions (Katz et

al., 1959)

Constituent Parachor Methane 77.0 Ethane 108.0 Propane 150.3 i-Butane 181.5 n-Butane 181.5 i-Pentane 225.0 n-Pentane 231.5 n-Hexane 271.0 n-Heptane 312.5 n-Octane 351.5 Hydrogen 34 (approx) Nitrogen 41 (approx)

Carbon dioxide 78

6-7

Page 667: +Peters Ekwere j. - Petrophysics

Figure 6.3. Parachors for computing interfacial tension of normal paraffin

hydrocarbons (Katz et al., 1959).

6-8

Page 668: +Peters Ekwere j. - Petrophysics

Figure 6.4. Parachors of heavy fractions for computing interfacial tension of

reservoir liquids (Firoozabadi et al., 1988).

6-9

Page 669: +Peters Ekwere j. - Petrophysics

Figure 6.5: Saturated liquid and vapor densities of various substances.

6-10

Page 670: +Peters Ekwere j. - Petrophysics

6.2.2 Interfacial Tension

Interfacial tension is the contractile force per unit length that exists at

the interface of two immiscible fluids such as oil and water (Figure 6.6). The

forces acting on the surface molecules are similar to those in the liquid-vapor

system, but the mutual attraction of unlike molecules across the interface

becomes important. The free energy required to create a fresh interface is

referred to as the excess interfacial free energy. The specific excess interfacial

free energy is dimensionally equivalent and is numerically equal to the

interfacial tension. Like surface tension, the unit of measurement of

interfacial tension is dynes/cm (or ergs/cm2).

Consider butanol (C4H9OH) with a surface tension of 24 dynes/cm in

contact with water (H2O) with a surface tension of 72 dynes/cm. What will be

the interfacial tension between the two liquids? For this system, the

interfacial tension is 1.8 dynes/cm, a fairly low number considering the

surface tensions of water and butanol. The low interfacial tension indicates

that the molecules of butanol must concentrate at the interface, decreasing

the contractile tendency of the interface. Interfacial orientation of the butanol

is favored because the hydrocarbon chain in the butanol is hydrophobic. At

20 ºC, the surface tension of ethanol is 22.39 dynes/cm and that of water is

72.80 dynes/cm. What is the interfacial tension between ethanol and water?

The answer is zero because ethanol and water are miscible. These examples

show that there is no simple relationship between the surface tensions of

liquids and their interfacial tensions. Table 6.3 lists accurately known

interfacial tensions for various organic liquids against water.

The presence of a third component of the right kind can significantly

reduce the interfacial tension between two liquids. For example, the

interfacial tension of water and iso-pentanol is 4.4 dynes/cm. If ethanol is

added to the system, the ethanol molecules will adsorb at the interface,

6-11

Page 671: +Peters Ekwere j. - Petrophysics

thereby reducing the interfacial tension between the water and the iso-

pentanol. When 25% by weight of ethanol is added, the interfacial tension is

reduced to zero. The system then becomes miscible and forms a single phase.

Figure 6.6. Apparent surface film caused by unequal attraction of molecules at the interface of two liquids.

Table 6.3. Interfacial Tension between Water and Pure Liquids

Liquid T (°C) σ (dynes/cm) n-Hexane 20 51.0 n-Octane 20 50.8 Carbon disulphide 20 48.0 Carbon tetrachloride 20 45.1 Carbon tetrachloride 25 43.7 Bromobenzene 25 38.1 Benzene 20 35.0 Benzene 25 34.71 Nitrobenzene 20 26.0 Ethyl ether 20 10.7

6-12

Page 672: +Peters Ekwere j. - Petrophysics

n-Octanol 20 8.5 n-Hexanol 25 6.8 Aniline 20 5.85 n-Pentanol 25 4.4 Ethyl acetate 30 2.9 Isobutanol 20 2.1 n-Butanol 20 1.8 n-Butanol 25 1.6

Chemicals that adsorb at interfaces are usually referred to as surface active

agents or surfactants. Such chemicals form a monolayer at the interface.

The monolayer is in a state of compression which reduces the contractile

tendency of the interface, thereby reducing the interfacial tension. The

presence of the adsorbed molecules creates a surface or spreading pressure

which reduces the interfacial tension. In fact,

oσ σ= − Π (6.5)

where σ is the reduced interfacial intension, σo is the original interfacial

tension and Π is the spreading pressure. Surfactants are often used to reduce

the interfacial tension between oil and water in order to improve oil recovery.

The interfacial tensions between reservoir water and crude oils have

been measured for a number of reservoirs and found to range from 15 to 35

dynes/cm at 70 °F, 8 to 25 dynes/cm at 100 °F, and 8 to 19 dynes/cm at 130

°F. Table 6.4 presents the results of interfacial tension measurements for

some fluid pairs.

The interfacial tension between reservoir oil and gas can be estimated

using parachors as

6-13

Page 673: +Peters Ekwere j. - Petrophysics

4

1

i NgL

i i ii L g

x yM M

ρρσ=

=

⎡ ⎤⎛ ⎞= Λ −⎢ ⎜⎜⎢ ⎥⎝ ⎠⎣ ⎦

∑ ⎥⎟⎟ (6.6)

where σ is the interfacial tension between oil and gas in dynes/cm, Λi is the

parachor of component i, xi is the mole fraction of component i in the liquid,

ML is the apparent molecular weight of the liquid, ρL is the saturated liquid

density in g/cm3, yi is the mole fraction of component i in the gas, Mg is the

apparent molecular weight of the gas, N is the total number of components in

the mixture and ρg is the saturated gas density in g/cm3.

Table 6.4. Typical Interfacial Tensions and Contact Angles for Fluid Pairs

(Archer and Wall, 1986)

Wetting Phase

Non-Wetting Phase

Conditions Contact Angle (º)

Interfacial Tension

(dynes/cm) Brine Oil Reservoir 30 30 Brine Oil Laboratory 30 48 Brine Gas Laboratory 0 72 Brine Gas Reservoir 0 50 Oil Gas Reservoir 0 4 Gas Mercury Laboratory 140 480

Many reservoir phenomena depend on the interfacial tensions between

the reservoir fluids and between the reservoir fluids and the reservoir rock.

Interfacial tensions of reservoir water and oil can be reduced significantly by

the addition of surface active agents to either the oil or to the water. Some of

these surface active agents occur naturally in crude oils.

6-14

Page 674: +Peters Ekwere j. - Petrophysics

The residual oil saturation for an immiscible displacement in a porous

medium is a function of the interfacial tension between the fluids, the

wettability, the fluid viscosities and the displacement rate. Therefore, we can

write

( )1 cos , , ,or nw wS f vσ θ μ μ= (6.7)

Using the technique of Appendix A for dimensional analysis, it can be shown

that the rank of the dimensional matrix obtained from the four variables

cosσ θ , nwμ , wμ and is two. Therefore, two independent dimensionless groups

can be derived from the four variables. It can be shown that

v

1

23

3

4

cos 0 11 1

1 00 1

nw

w

xx

4x xxxv

σ θμμ

−⎡ ⎤⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥= +⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥

⎣ ⎦ ⎣ ⎦ ⎣ ⎦⎣ ⎦

(6.8)

Let us choose x3 = -1, x4 = 0. The corresponding dimensionless group is given

by

1nw

w

μπμ

= (6.9)

Next, let us choose x3 = x4 = 1. The corresponding dimensionless group is

2 coswvμπ

σ θ= (6.10)

The dimensionless group in Eq.(6.10), cos

wvμσ θ

or wvμσ

in the case of perfect

wetting, is known as the capillary number. It is the ratio of the viscous to the

capillary force. Thus,

6-15

Page 675: +Peters Ekwere j. - Petrophysics

cos

wc

vN or wvμ μσ θ σ

= (6.11)

The functional relationship between the residual oil saturation and the two

dimensionless groups can be written as

2 ,cos

nw wor

w

vS f μ μμ σ θ

⎛ ⎞= ⎜

⎝ ⎠⎟ (6.12)

For a fixed viscosity ratio,

3 cosw

orvS f μ

σ θ⎛= ⎜⎝ ⎠

⎞⎟ (6.13)

The oil recovery is given by

41

1 cor wi w

wi

S S vR fS

μosσ θ

− − ⎛= = ⎜− ⎝ ⎠⎞⎟ (6.14)

Figure 6.7 shows typical correlations for residual saturations versus

capillary number for a wetting fluid displacing a nonwetting fluid and for a

nonwetting fluid displacing a wetting fluid. Such correlations are usually

referred to as capillary desaturation curves (CDC). Consider a wetting fluid

displacing a nonwetting fluid such as a waterflood in a water wet reservoir. It

can be seen that there is a critical capillary number below which the residual

nonwetting fluid saturation is constant and independent of the capillary

number. Normal waterfloods usually fall in this range of capillary number.

Above the critical capillary number, the residual oil saturation decreases as

the capillary number increases. Thus, the residual nonwetting phase can be

mobilized and displaced by increasing the capillary number of the

displacement. Capillary number can be increased by increasing wμ and v .

However, the most effective way to increase the capillary number is by

6-16

Page 676: +Peters Ekwere j. - Petrophysics

lowering the interfacial tension between the wetting and nonwetting phases

with a surfactant. Interfacial tensions less that 0.1 dyne/cm can be achieved.

In the limit, if the interfacial tension could be lowered to zero, the fluids would

become miscible and no residual saturation would be left behind. Of course,

in a practical process, the interfacial tension cannot be reduced to zero except

in a miscible process.

Consider the case of a nonwetting fluid displacing a wetting fluid, such

as a waterflood of an oil wet reservoir. The capillary desaturation curve is

similar to that for the residual nonwetting fluid except that the critical value

of the capillary number is higher. It should be noted that the numerical

values of the residual saturations given in Figure 6.7 are for illustrative

purposes. It should not be assumed that waterfloods always have a residual

oil saturation of 30%, nor should one infer from the figure that it is more

efficient to displace a wetting fluid with a nonwetting fluid. The residual oil

saturation can be greater or less than 30% depending on such factors as the

mobility ratio of the displacement and the properties of the porous medium

such as pore structure, pore size distribution, permeability and wettability to

name a few. Also, in general, it is more difficult for a nonwetting fluid to

displace a wetting fluid than for a wetting fluid to displace a nonwetting fluid.

6-17

Page 677: +Peters Ekwere j. - Petrophysics

Figure 6.7. Typical correlations of residual nonwetting and wetting phase saturations with capillary number (Lake, 1989).

Figure 6.8 shows capillary desaturation experimental data from Abrams

(1975) obtained on the same core sample but at different viscosity ratios.

Clearly, the decrease in residual oil saturation with increasing capillary

number is evident. Figure 6.9 shows the same data plotted against a modified

capillary number that includes the viscosity ratio, 0.4

w w

o

vμ μσ μ

⎛ ⎞⎛ ⎞⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠. The inclusion

of the viscosity ratio appears to improve the correlation.

6-18

Page 678: +Peters Ekwere j. - Petrophysics

Figure 6.8. Capillary desaturation data (Abrams, 1975).

Figure 6.9. Capillary desaturation data with the effect of viscosity ratio included (Abrams, 1975).

6-19

Page 679: +Peters Ekwere j. - Petrophysics

6.2.3 Measurements of Surface and Interfacial Tensions

Several techniques are used to measure surface and interfacial

tensions. These include (1) capillary rise method, (2) sessile drop method, (3)

pendant drop method, (4) ring method and (5) spinning drop method.

Capillary Rise Experiment

When a capillary tube is dipped into a wetting liquid, the liquid will be

spontaneously imbibed (sucked) into the capillary tube as shown Figure 6.10.

Figure 6.10. Capillary rise experiment.

6-20

Page 680: +Peters Ekwere j. - Petrophysics

The equilibrium height is determined by the balance between the capillary

suction force and the pull of gravity. The capillary force acts upward and for a

circular capillary tube is given by

2 cosCapillary Force Force Up rπ σ θ= = (6.15)

The gravitational force acts downward and is given by

( )2 w nwGravitational Force Force Down r h gπ ρ ρ= = − (6.16)

The downward force also can be expressed in terms of the pressures on the

opposite sides of the meniscus as

( )2Pr nw wessure Force Force Down r P Pπ= = − (6.17)

At equilibrium, the force up is equal to the force down. Equating Eqs.(6.15)

and (6.16) gives

(2 cosw nwh

r) gσ θ ρ ρ= − (6.18)

Eq.(6.18) can be rearranged as

( )2cos

w nwrh gρ ρσ

θ−

= (6.19)

The surface tension, σ, can be estimated by measuring the variables on the

right side of Eq.(6.19) in a capillary rise experiment. The experiment can be

simplified by using air as the nonwetting phase and treating the capillary tube

such that it is perfectly wetted by the wetting fluid. In this case, nw wρ ρ<< , θ =

0 and cosθ = 1. Eq.(6.19) becomes

6-21

Page 681: +Peters Ekwere j. - Petrophysics

2

wrh gρσ = (6.20)

By measuring r, h and wρ , an estimate of the surface tension can easily be

obtained from Eq.(6.20).

Eq.(6.20) can be used to define a characteristic capillary length scale as

1

2w

rhg

σκρ

− = = (6.21)

For water at 25 ºC, σ = 72 dynes/cm, 1wρ = g/cm3 and of course, g = 981

cm/s2. Thus, for water,

1 72 0.27091 981x

κ − = = cm or 2.71 mm.

The significance of the capillary length is that when dealing with a system

with the characteristic length scale 1δ κ −< , the effect of gravity is negligible

and capillary effect dominates the process. For example, at the pore scale

where the characteristic length is of the order of microns, capillary effect

dominates gravity effect.

Let us define another constant for the capillary rise experiment as

2 2

w

ag

σρ

= (6.22)

The constant a2 is a property of the wetting fluid only. For water at 25 ºC, a2 =

0.1468 cm2. Eq.(6.20) can be written as

2a rh= (6.23)

6-22

Page 682: +Peters Ekwere j. - Petrophysics

Some famous scientists have studied the capillary rise experiment and have

proposed more elaborate equations to describe the capillary rise than

Eq.(6.23). Jurin (1718) gives the capillary rise equation as

2

3ra r h⎛ ⎞= +⎜

⎝ ⎠⎟ (6.24)

where the 3r corrects for the volume of the liquid in the spherical meniscus.

Hagen and Desains (19xx) proposed

2 3

22 3

0.1111 0.074113r ra rhh h h

⎛ ⎞= + − +⎜

⎝ ⎠

r⎟ (6.25)

where the last two terms on the right hand side correct for deviations of the

meniscus from sphericity. Rayleigh (1915) further refined Eq.(6.25) as

2 3

22 3

0.1288 0.131213r ra rhh h h

⎛ ⎞= + − +⎜

⎝ ⎠

r⎟ (6.26)

For our purpose, we will use the simple version of the capillary rise equation,

Eq.(6.23).

Equating Eqs.(6.15), (6.16) and (6.17) gives

(2 cosnw w w nwP P h

r) gσ θ ρ ρ− = = − (6.27)

Eq.(6.27) gives the excess pressure, (Pnw - Pw), across the curved interface

between the wetting and nonwetting phases in terms of the pertinent

variables of the capillary rise experiment. This excess pressure is known as

the capillary pressure (Pc) and will be the subject of Chapter 7. As shown in

6-23

Page 683: +Peters Ekwere j. - Petrophysics

Eq.(6.27), the excess pressure for the capillary rise experiment is positive,

which means that the pressure in the nonwetting phase is higher than that in

the wetting phase. Eq.(6.27) can now be written as

(2 cosc nw w w nwP P P h g

r)σ θ ρ ρ= − = = − (6.28)

Sessile Drop Method

The sessile drop method of determining the surface tension of a liquid

consists of measuring the number of liquid drops that fall from the capillary

end of the instrument while the surface of the liquid within the bulb is

lowered from the upper to the lower mark as shown in Figure 6.11. The

principle of the method is based on the fact that the size of the liquid drop is

proportional to surface tension of the liquid. The size of the drop is reached

when the surface tension can no longer support its weight. To a first

approximation,

2idealW rπ σ= (6.29)

where Wideal is the weight of the drop that should fall, r is the external

radius of the tube and σ is the surface tension. Figure 6.12 shows the

sequence of shapes for a drop that detaches from a tip. The detached drop

leaves behind some liquid residue. Thus, the actual weight of the drop which

is what is measured is less than the ideal weight. To account for this,

Eq.(6.29) is modified as

2actual idealW W f r fπ σ= = (6.30)

6-24

Page 684: +Peters Ekwere j. - Petrophysics

where f is a correction factor which can be expressed as a function of 13

rV

,

where V is the volume of the drop. Table 6.5 shows the correction factors for

various 13

rV

.

Figure 6.11. Sessile drop method of measuring surface tension.

Figure 6.12. Sequence of shapes for a drop.

6-25

Page 685: +Peters Ekwere j. - Petrophysics

Table 6.5. Correction Factors for Sessile Drop Method (Adamson, 1982).

Pendant Drop Method

The pendant-drop method of measuring surface or interfacial tension

depends only on the density of the fluids and the dimensions of the drop.

Figure 6.13 shows a pendant drop and the relevant dimensions. The surface

or interfacial tension is given by

( )2

e L ggdHρ ρ

σ−

= (6.31)

where σ is the surface tension, de is the maximum diameter of the drop, ρL is

the density of the liquid, ρg is the density of the vapor, H is a constant that is

a function of e

s

dd

and g is the gravitational acceleration. The constant H is

tabulated as a function of de/ds. The pendant drop method can be used to

measure surface tension or interfacial tension. It can also be adapted for

measurements at elevated temperature and pressure.

6-26

Page 686: +Peters Ekwere j. - Petrophysics

Figure 6.13. Surface tension by pendant drop method.

Ring Method

The ring method of determining surface or interfacial tension depends

on measuring the force required to pull the ring free of the interface as shown

in Figure 6.14. Theoretically, the surface or interfacial tension is given by

2FL

σ = (6.32)

where σ is the surface or interfacial tension, F is the force required to pull the

ring free of the interface and L is the circumference of the ring. The factor of 2

accounts for the fact that there are two surfaces around the ring. In practice,

corrections are needed to account for the mass of liquid lifted by the ring in

breaking through the interface as shown in Figure 6.15. Such corrections are

made available with the instrument. Figure 6.16 shows a typical instrument,

known as the du Nouy tensiometer, that employs the ring method for surface

or interfacial tension determination.

6-27

Page 687: +Peters Ekwere j. - Petrophysics

Figure 6.14. Surface tension by ring method.

6-28

Page 688: +Peters Ekwere j. - Petrophysics

Figure 6.15. Condition of liquid surface film at breaking point.

6-29

Page 689: +Peters Ekwere j. - Petrophysics

Figure 6.16. du Nouy tensiometer.

Spinning Drop Method

The spinning drop method of determining surface or interfacial tension

is based on measuring the shape of a drop of liquid or gas bubble in a more

dense liquid contained in a rotating horizontal tube. The drop or bubble is

typically rotated at speeds of 1,200 to 24,000 revolutions per minute (RPM).

Under rotation, the original spherical drop or bubble becomes elongated into a

6-30

Page 690: +Peters Ekwere j. - Petrophysics

cylindrical shape as shown in Figure 6.17. A strobe light is used to visualize

the deformed drop. A microscope is used to measure the diameter of the

drop. The interfacial tension is given by

2 314

rσ ρω= Δ (6.33)

where σ is the surface or interfacial tension, dynes/cm Δρ is the density

difference between the two fluids in g/cm3, ω is the angular velocity in

radians/s and r is the cylindrical radius of the drop in cm.

An instrument based on the spinning drop method has been designed

and patented at the University of Texas at Austin by Schechter and Wade

(Figure 6.18). The instrument is manufactured and sold by the Chemistry

Department at the University of Texas. The instrument is particularly

suitable for measuring low interfacial tensions and is therefore used

extensively in surfactant research. Interfacial tensions as low as 10-6

dyne/cm have been successfully measured with the instrument.

6.3 WETTABILITY

6.3.1 Definition

Wettability is a tendency for one fluid to spread on or adhere to a solid

surface in the presence of other immiscible fluids. The fluid that spreads or

adheres to the surface is known as the wetting fluid. In a petroleum reservoir,

the solid surface is the reservoir rock which may be sandstone, limestone, or

dolomite, together with cementing material. The fluids are water, oil and gas.

Normally, either water or oil is the wetting phase. Gas is always a nonwetting

phase.

6-31

Page 691: +Peters Ekwere j. - Petrophysics

Figure 3.17. Cylindrical liquid drops in a spinning drop apparatus. (A)

benzene-water system at 20,000 RPM, (B) octane-surfactant system at 6,000 RPM (Cayias et al., 1975).

6-32

Page 692: +Peters Ekwere j. - Petrophysics

Figure 6.18. Schematic of spinning drop tensiometer (Cayias et al., 1975).

Consider the water-oil-solid system shown in Figure 6.19. Three

interfacial tensions (specific free surface energies) arise: σos is the solid-oil

interfacial tension, σow is the oil-water interfacial tension and σws is the

water-solid interfacial tension. The angle θ is known as the contact angle and

is measured through the water (the more dense fluid). The contact angle is a

measure of the wettability of the solid.

At equilibrium, the interfacial tensions are related by the Young - Dupre

equation obtained by considering horizontal equilibrium of the point of

contact of the interfacial tensions. For Figure 6.19, this equation is given by

cosos ws owσ σ σ θ− = (6.34)

6-33

Page 693: +Peters Ekwere j. - Petrophysics

The free surface energies for the oil-solid and water-solid interfaces

cannot be measured readily. However, their difference can be determined by

measuring the oil-water interfacial tension and the contact angle. This

difference controls the movement of the interface before equilibrium is

achieved. The following may be deduced from Eq. (6.32):

Figure 6.19. Interfacial tensions in a water-oil-solid system.

1. If the free surface energies for the oil-solid (σos) and the water-solid

(σws) interfaces are equal, the left side of Eq.(6.34) is zero. Since the

oil-water interfacial (σow) tension is nonzero, cosθ on the right side of

Eq.(6.34) must be zero, giving a contact angle of 90°. A contact angle of

90° means that the solid has no preferential wettability for the oil or the

water. This is a situation of neutral or intermediate wettability.

2. If σws < σos , then θ < 90°. The solid is said to be preferentially water

wet. When the oil, water and solid are first brought in contact, water

will advance and spread on the solid surface, displacing the oil until an

equilibrium contact angle is attained according Eq.(6.34). During

6-34

Page 694: +Peters Ekwere j. - Petrophysics

spreading of the water, the free energy of the system is reduced since

σws < σos. The difference (σos − σws) is known as the adhesion tension.

3. If σws > σos , then θ > 90°. The solid is said to be preferentially oil wet.

4. Complete spreading of the oil on the surface takes place if θ = 180° and

complete spreading of water on the surface takes place if θ = 0°.

Complete spreading of crude oil or water on a surface has never been

observed with reservoir fluids.

Figure 6.20 shows the equilibrium contact angles for four wettability states.

In Figure 6.20a, the surface is preferentially oil wet; in Figure 6.20b, the

surface is of neutral wettability; in Figure 6.20c, the surface is water wet; and

in Figure 6.20d, the surface is totally water wet.

6-35

Page 695: +Peters Ekwere j. - Petrophysics

Figure 6.20. Equilibrium contact angles showing four wettability states.

6.3.2 Determination of Wettability

Reservoir wettability is usually determined either by contact angle

measurement using reservoir fluids and a pure mineral surface or by an

imbibition test on a reservoir core sample using refined oil and a synthetic

brine. No wettability determination method involves the simultaneous use of

reservoir fluids and reservoir rock.

6-36

Page 696: +Peters Ekwere j. - Petrophysics

Contact Angle Method

Contact angle is one of the earliest and still most widely used

measurement to evaluate reservoir wettability. The contact angle

measurement essentially seeks to establish whether or not the reservoir oil

contains surface active agents that could make an originally preferentially

water wet mineral surface become preferentially oil wet over time.

Accordingly, the contact angle test uses reservoir oil and brine and a pure,

clean mineral surface which is known to be preferentially water wet at the

outset. In the absence of reservoir brine, synthetic brine is used in the test

since the surface active fluid is the oil and not in the brine. The solids

normally used in the test to represent reservoir rock are pure quartz (silica)

for a sandstone reservoir, pure calcite for a limestone reservoir and pure

dolomite crystal for a dolomite reservoir. These pure minerals are known to

be preferentially water wet initially. If the oil contains surface active agents,

then these will adsorb on the mineral surface over time and increase the

degree of oil wetness. This change in the wettability of the surface can be

observed and quantified by measuring the contact angle over time until an

equilibrium contact angle is obtained. It is reasonable to assume that a

similar wetting equilibrium will be approached in the reservoir.

The contact angle measurement is performed with a contact angle cell

using an instrument known as a goniometer. The mineral surface is

immersed in the brine (or oil) and allowed to equilibrate. A drop of the oil (or

brine) is then introduced on to the surface with a hypodermic syringe as

shown in Figure 6.21. The contact angle is then measured over time. The

test can last several weeks depending on the time required to achieve

adsorption equilibrium. The equipment can be adapted for contact angle

measurements at elevated pressures and temperatures.

6-37

Page 697: +Peters Ekwere j. - Petrophysics

Figure 6.21. Contact angle cell.

Two contact angles are normally measured: the advancing and receding

contact angles. The advancing contact angle (θA) is the contact angle

obtained when water comes into equilibrium with a surface previously in

contact with oil as shown Figure 6.22. The receding contact angle (θR) is the

contact angle obtained when oil comes into equilibrium with a surface

previously in contact with water. The advancing contact angle is always

greater than the receding contact angle. Normally, it is the advancing contact

angle that is reported as the contact angle in a wettability test.

Figure 6.23 shows the results of a contact angle test. Several

interesting observations can be made. The early time contact angle

measurements showed the solid to be preferentially water wet. However, as

6-38

Page 698: +Peters Ekwere j. - Petrophysics

time passed, the degree of water wetness diminished. Eventually, after

adsorption equilibrium was achieved the solid was found to be preferentially

oil wet. Had the contact angle test been terminated prematurely, the

wettability assessment would have been wrong. Note that for this test, over

30 days of aging were needed to establish adsorption equilibrium.

Figure6.22. Advancing and receding contact angles.

Figure 6.23. Approach to equilibrium contact angle (Craig, 1971).

6-39

Page 699: +Peters Ekwere j. - Petrophysics

The major advantages of contact angle measurements are the reliability

of the results and the relative ease of obtaining uncontaminated reservoir

fluid samples compared to uncontaminated reservoir rock samples. The

following disadvantages should be noted. (1) Contact angle is measured on a

flat, clean, homogeneous mineral surface. Such a surface does not exist in the

reservoir. (2) Pure minerals are used in the test to simulate sandstone,

limestone and dolomite reservoir rocks. Pure minerals may not be

representative of actual reservoir mineralogy. (3) The test can be very long and

requires extreme cleanliness and inertness of the test system. (4) There is

evidence that the contact angle is affected by which fluid was first in contact

with the solid.

Amott Wettability Test

The Amott wettability index is obtained by a combined imbibition-

displacement test on a reservoir core sample using refined oil and synthetic

brine. After the reservoir core sample has been flushed with brine to residual

oil saturation and evacuated to remove gas, it is then subjected to the

following tests:

1. The core is immersed in oil (e.g., kerosene) and the volume of brine

displaced by the imbibition of oil is measured after 20 hours in an

imbibition cell as shown in Figure 6.24.

2. The core is centrifuged under kerosene and the additional brine

displaced by centrifuging is measured.

3. The core is immersed in brine and the volume of oil displaced by the

imbibition of brine is measured after 20 hours.

4. The core is centrifuged under brine and the additional oil displaced by

centrifuging is measured.

6-40

Page 700: +Peters Ekwere j. - Petrophysics

Figure 6.24. Imbibition cell.

The wettability indices of water (WIw) and oil (WIo) are calculated as follows:

w

Volume of oil displaced by brine imbibitionWIVolume of oil displaced by brine imbibition forced displacement

=+

(6.35)

o

Volume of brine displaced by oil imbibitionWIVolume of brine displaced by oil imbibition forced displacement

=+

(6.36)

The Amott wettability indices and dimensionless numbers that range from 0

to 1. If the rock is preferentially water-wet, WIo will be 0 and WIw > 0. The

greater the degree of water wetness, the closer will WIw be to 1. Similarly, if

6-41

Page 701: +Peters Ekwere j. - Petrophysics

the rock is preferentially oil wet, WIw will be 0 and WIo > 0. The greater the

degree of oil wetness, the closer will WIo be to 1. For a rock of intermediate or

neutral wettability, WIw and WIo will be 0 or close to 0. Sometimes, the

difference is used to as the wettability measure. In this case, the

wettability index will range from -1 to +1. An index of -1 indicates a strongly

oil wet rock whereas an index of +1 indicates a strongly water wet rock.

( wWI WI− )o

The Amott wettability index is a reliable measure of the wettability of

the core sample. However, the wettability of the core sample may not be

representative of the wettability of the reservoir rock because of the difficulty

of obtaining an unaltered core sample. The wettability of the core sample may

easily be altered by the coring operation.

United States Bureau of Mines (USBM) Wettability Index

The USBM wettability index is obtained by carry out a number of forced

water and oil displacement experiments using a centrifuge. The results of

such experiments are shown in Figure 6.25. The sample is saturated initially

with water. The water is then displaced with oil to irreducible water saturation

using the centrifuge. This is the process labeled I in each figure. Next, the

sample, which contains initial oil saturation and irreducible water saturation

is then centrifuged in water to residual oil saturation. This is the process

labeled II in each figure. The sample, which now contains water and residual

oil saturation is then centrifuged in oil to irreducible water saturation. This is

the process labeled III in each figure. The USBM wettability index is calculated

as

110

2

logwAUSBM Wettability Index IA

⎛ ⎞= = ⎜

⎝ ⎠⎟ (6.37)

6-42

Page 702: +Peters Ekwere j. - Petrophysics

where A1 and A2 are the areas under the capillary pressure curves shown in

each figure.

The area under a capillary pressure curve represents the

thermodynamic work required for the displacement. The displacement of a

nonwetting phase by a wetting phase requires less work than the

displacement of a wetting phase by a nonwetting phase. Therefore, the ratio

of the areas under the capillary pressure curves, 1

2

AA

⎛ ⎞⎜⎝ ⎠

)o

, is a measure of the

degree of wettability of the porous medium. Therefore, the USBM wettability

index for a water wet medium will be positive as shown in Figure 6.25A, that

of an oil wet medium will be negative as shown in Figure 6.25B and that of a

medium of neutral wettability will be 0 as shown in Figure 6.25C. The USBM

wettability index ranges from -1 for a strongly oil wet rock to +1 for a strongly

water wet rock. The absolute value of the index is a measure of the degree of

wettability preference. A wettability index of zero indicates no preferential

wetting by either fluid.

Figure 6.26 compares the USBM wettability index and the Amott

wettability index, , of forty three outcrop rock samples and three

reservoir rock samples. There is a strong correlation between the two

measures of wettability. In particular, both methods show the three reservoir

rock samples to be oil wet as indicated by the negative values for both

wettability indices.

( wWI WI−

6-43

Page 703: +Peters Ekwere j. - Petrophysics

Figure 6.25. Determination of USBM wettability index (Donaldson et al., 1969).

6-44

Page 704: +Peters Ekwere j. - Petrophysics

Figure 6.26. A comparison of the USBM wettability index with Amott wettability index for several core samples (Donaldson et al., 1969).

6.3.3 Wettability of Petroleum Reservoirs

The wettabilities of petroleum reservoirs span the entire spectrum from

preferentially water wet to preferentially oil wet reservoirs. Treibel et al. (1972)

measured the wettabilities of 30 sandstone and 25 carbonate reservoirs by

measuring contact angles at the reservoir temperatures using the reservoir

oils and synthetic brine. Quartz crystal was used to represent the sandstone

reservoirs whereas calcite crystal was used to represent the limestone and

dolomite reservoirs in the contact angle measurements. The wettabilities of

the reservoirs were evaluated using an arbitrary contact angle scale.

Reservoirs with contact angles from 0 to 75° were classified as water wet;

those with contact angles from 75 to 105° were classified as having

6-45

Page 705: +Peters Ekwere j. - Petrophysics

intermediate wettability and those with contact angles from 105 to 180° were

classified as preferentially oil wet.

The results showed that 27% of the reservoirs tested were preferentially

water wet, 66% were preferentially oil wet and the remaining 7% were of

intermediate wettability. It was found that 43% of the sandstones were

preferentially water wet, 50% were preferentially oil wet and 7% were of

intermediate wettability. On the other hand, 84% of the carbonate reservoirs

were preferentially oil wet, 8% were preferentially water wet and 8% were of

intermediate wettability. It would appear from the results of this study that

carbonates are more likely to be preferentially oil wet than preferentially water

wet. However, this assertion cannot be generalized because the 55 reservoirs

used in this study were not obtained by random sampling. A random sample

of reservoirs would be needed if the results of the wettability tests are to be

given statistical significance.

A similar contact angle study by Chiligarian and Chen (1983) on 161

carbonate reservoirs showed 80% of the reservoirs to be preferentially oil wet,

8% to be preferentially water wet and 12% to be of intermediate wettability.

These results are consistent with those of Treibel et al.(1972).

6.3.4 Effect of Wettability on Rock -Fluid Interactions

Wettability has a profound effect on multiphase rock-fluid interactions.

Wettability affects (a) the microscopic fluid distribution at the pore scale in the

porous medium, (b) the magnitude of the irreducible water saturation, (c) the

efficiency of an immiscible displacement in the porous medium, (d) the

residual oil saturation, (e) the capillary pressure curve of the porous medium,

(f) the relative permeability curves of the porous medium and (g) the electrical

properties of the porous medium.

Microscopic Fluid Distribution at the Pore Scale

6-46

Page 706: +Peters Ekwere j. - Petrophysics

Wettability determines the microscopic fluid distribution in a porous

medium at the pore scale. The wetting fluid occupies the small pores, coats

the surface of the solid grains and occupies the corners of the grain contacts.

The wetting phase occupies the small pores, which have high specific surface

areas (S=3(1-φ)/r) in order to minimize the specific surface free energy of the

system. The nonwetting phase occupies the large pores and are located at the

center of the pores. These pore scale fluid distributions are shown

schematically in Figure 6.27 for water wet and oil wet porous media. For the

water wet medium, at the initial state, water being the wetting phase coats the

grain surface and occupy the nooks and crannies of the medium. Oil, being

the nonwetting phase occupies the center of the pores and is surrounded by

water. After waterfooding, the residual oil globules occupy the center of the

pores. For the oil wet medium, oil being the wetting phase coats the grain

surface and occupy the nooks and crannies of the medium. The water, being

the nonwetting phase occupies the center of the pores and is surrounded by

oil. After waterflooding, the water occupies the center of the pores and the

residual oil wets the grain surface and occupies the nooks and crannies of the

medium. These microscopic fluid arrangements have implications for the

nature of the end point relative permeabilities of a water wet rock and an oil

wet rock.

Effect of Wettability on Irreducible Water Saturation

It has been observed that the irreducible water saturation in an oil wet

reservoir rock tends to be less than in a water wet reservoir rock. Craig (1971)

gives the following rule-of-thumb for irreducible water saturation for water

wet and oil wet reservoirs. For water wet reservoirs, the irreducible water

saturation is usually greater than 20 to 25% whereas for oil wet reservoirs it

is generally less than 15% and frequently less than 10% of the pore volume.

6-47

Page 707: +Peters Ekwere j. - Petrophysics

Figure 6.27. Fluid distributions as a function of wettability (adapted from Pirson, 1958).

Effect of Wettability on Electrical Properties of Rocks

Wettability affects the saturation exponent, n, in Archie's resistivity

index equation. For water wet rocks, the exponent is typically around 2.

However, for oil wet rocks, the exponent can increase to rather high values as

the water saturation decreases. Table 6.6 shows the result of laboratory

measurements of n as a function of water saturation in an oil wet sandstone.

It can be observed that below a certain water saturation, the saturation

exponent increases above the usual value of 2. An exponent as high as 9 was

measured in the experiments.

6-48

Page 708: +Peters Ekwere j. - Petrophysics

Table 6.6. Archie Saturation Exponent in Oil Wet Rocks (Mungan and Moore,

1968).

Figure 5.28 shows the effect of wettability on the resistivity index of

carbonate cores. In this study, the cores were rendered preferentially water

wet by heating up to 500 ºF and preferentially oil wet by washing with an

organic acid. Measurements were also made on cores that were of neutral

wettability. The wettability classification was based on imbibition tests. The

equations relating resistivity index to water saturation for neutral and

preferentially water wet cores are and , giving saturation

exponents of 1.92 and 1.61, respectively. The data for the preferentially oil wet

cores separated into two distinct trends described by the equations

for the first trend and for the second trend, giving

saturation exponents of 12.27 and 8.09. The separation of the oil wet data

was attributed to differences in the pore size distributions of the cores.

1.92wI S −=

0.37I =

1.61wI S −=

12.270.000027 wI −= S 8.09wS −

6-49

Page 709: +Peters Ekwere j. - Petrophysics

Figure 6.28. Effect of wettability on the resistivity index of carbonate cores (Sweeney and Jennings, 1960).

It is reasonable to expect that wettability will affect the resistivity and

hence the saturation exponent of a partially saturated porous medium. In an

oil wet medium, water being the nonwetting phase occupies the center of the

large pores. At high water saturations, the water is continuous and therefore

conducts electrical current. As the water saturation is decreased, below a

certain water saturation, the water will breakup into disconnected globules

and can no longer conduct electrical current. As a result, the resistivity of the

system will increase and this increase in resistivity is reflected in the increase

in the water saturation exponent as observed in the experiments.

6-50

Page 710: +Peters Ekwere j. - Petrophysics

Effect of Wettability on the Efficiency of an Immiscible Displacement

Wettability has a significant effect on the efficiency of an immiscible

displacement in a porous medium. Figure 6.29 shows schematically the

microscopic displacement of oil from a water wet medium and an oil wet

medium at the pore scale. In the water wet medium, the injected water is

imbibed into the medium along the pore walls in a manner that enhances the

oil displacement efficiency. The residual oil is trapped at the center of the

large pores. In the oil wet medium, the injected water channels through the

large pores leaving behind considerable residual oil in the small pores, at the

solid contacts and as coatings on the solid grains. From this pore level

picture, it is easy to see that the waterflood efficiency will be higher in the

water wet medium than in the oil wet medium everything else being equal.

The higher waterflood efficiency of the water wet rock compared to the

oil wet rock seen at the pore scale manifests itself at the macroscopic scale

(core scale) as well. Owens and Archer (1971) performed waterflood

experiments in core plugs (1.9 cm diameter and 4.4 cm length) at various

wettability conditions. The core plugs were rendered progressively oil wet by

dissolving a sulfonate in the oil phase. Figure 6.30 shows the oil recovery

curves for the waterfloods as a function of the wettability of the core. The

decline in the oil recovery efficiency with increasing oil wetness is obvious.

6-51

Page 711: +Peters Ekwere j. - Petrophysics

Figure 6.29. Microscopic displacement of oil from a pore during a waterflood: (a) strongly water wet medium, (b) strongly oil wet medium (Raza et al., 1968).

6-52

Page 712: +Peters Ekwere j. - Petrophysics

Figure 6.30. Effect of wettability on waterflood performance at an oil-water

viscosity ratio of 5 (Archer and Owens, 1971).

Peters and Hardham (1989) have conducted similar waterflood

experiments in unconsolidated sandpacks at a larger scale (4.8 cm diameter

and 54 cm length) than core plugs. The results of two such experiments are

presented here. In Experiment 1, the sandpack was first saturated with brine

and the brine was displaced by a viscous silicon-based test oil (103.4 cp) to

establish irreducible water saturation in contact with the sand grains. The

viscous oil was then displaced by brine to simulate a waterflood at an

unfavorable viscosity ratio of 85. In Experiment 2, a second sandpack was

6-53

Page 713: +Peters Ekwere j. - Petrophysics

first saturated with the same viscous silicon-based oil and the oil was then

displaced by the same brine to simulate a waterflood at the same unfavorable

mobility ratio as in Experiment 1. Although the wettabilities of the sandpacks

were not measured directly, it is believed that the first sandpack would

behave as a water wet system whereas the second sandpack would behave as

an oil wet system over the short time scale of the experiments. Both

waterfloods were imaged by X-ray CT to visualize the insitu fluid saturations

in time and space.

The oil recovery curves for the two waterflood experiments are shown in

Figure 6.31. They show the displacement in the water wet sandpack to be

more efficient than in the oil wet sandpack . These results are in agreement

with those of Archer and Owens. The low water breakthrough recoveries in

this study are due to the high oil-water viscosity ratio of 85.

Figures 6.32 and 6.33 show the water saturation images for the two

waterfloods at several pore volumes injected. The images for the water wet

sandpack show a relatively uniform and efficient displacement of the oil by

the water, with relatively high water saturations. In contrast, the images for

the oil wet sandpack show a chaotic, fragmented and inefficient displacement,

with relatively low water saturations. These images clearly show the important

role of wettability in determining the efficiency of waterfloods at the

macroscopic scale. Figures 6.34 and 6.35 show the water saturation profiles

for the experiments. They confirm the higher displacement efficiency of

Experiment 1 compared to Experiment 2.

6-54

Page 714: +Peters Ekwere j. - Petrophysics

Figure 6.31. Effect of wettability on waterflood performance at an oil-water

viscosity ratio of 91 (Peters and Hardham, 1989).

6-55

Page 715: +Peters Ekwere j. - Petrophysics

6-56

Page 716: +Peters Ekwere j. - Petrophysics

6-57

Page 717: +Peters Ekwere j. - Petrophysics

Figure 6.32. Water saturation images for a waterflood in a water wet sandpack at a viscosity ratio of 91. (A) tD = 0.05, (B) tD = 0.10, (C ) tD = 0.25, (D) tD = 0.50, (E) tD = 1.0, (F) tD = 2.0, (G) tD = 3.0 (Peters and Hardham, 1989).

6-58

Page 718: +Peters Ekwere j. - Petrophysics

6-59

Page 719: +Peters Ekwere j. - Petrophysics

6-60

Page 720: +Peters Ekwere j. - Petrophysics

6-61

Page 721: +Peters Ekwere j. - Petrophysics

Figure 6.33. Water saturation images for a waterflood in an oil wet sandpack at a viscosity ratio of 91: (A) tD = 0.05, (B) tD = 0.10, (C ) tD = 0.25, (D) tD = 0.50, (E) tD = 1.0, (F) tD = 2.0, (G) tD = 3.0 (Peters and Hardham, 1989).

6-62

Page 722: +Peters Ekwere j. - Petrophysics

Figure 6.34. Water saturation profiles for a waterflood in a water wet

sandpack.

Figure 6.35. Water saturation profiles for a waterflood in an oil wet sandpack.

6-63

Page 723: +Peters Ekwere j. - Petrophysics

6.4 THERMODYNAMICS OF INTERFACES

6.4.1 Characterization of Interfacial Tension as Specific Surface Free Energy

Consider a closed system consisting of the interface between two

immiscible fluids or between an immiscible fluid and a solid surface. The first

law of thermodynamics applied to this system gives

dU dQ dW= − (6.38)

where dU is the change in total internal energy of the system, dQ is the heat

input into the system and dW is the work done by the system. For an

equilibrium system, infinitesimal changes are reversible so that the reversible

work is given by

dW PdV daσ= − (6.39)

where da is the interfacial area. For a closed system,

dQ TdS= (6.40)

Substitution of Eqs.(6.39) and (6.40) into Eq.(6.38) gives for a closed system,

dU TdS PdV daσ= − + (6.41)

For an open system, Eq.(6.41) becomes

(6.42) 1

i N

i ii

dU TdS PdV da dnσ μ=

=

= − + + ∑

where μi is the chemical potential or molal free energy of component i in the

system, ni is the moles of component i in the system and N is the total

6-64

Page 724: +Peters Ekwere j. - Petrophysics

number of components in the system. At equilibrium, μi is the same in the

bulk fluid and at the interface.

We may obtain the interfacial tension from Eq.(6.42) in terms of the

internal energy of the system as

, , iS V n

Ua

σ ∂⎛ ⎞= ⎜ ⎟∂⎝ ⎠ (6.43)

The interfacial tension also can be expressed in terms of Helmholtz Free

Energy (A) and Gibbs Free Energy (G). The Helmholtz free energy is defined by

A U TS= − (6.44)

Differentiating Eq.(6.44) gives

dA dU TdS SdT= − − (6.45)

Substituting Eq.(6.42) into Eq.(6.45) gives

(6.46) 1

i N

i ii

dA SdT PdV da dnσ μ=

=

= − − + + ∑

We may obtain the interfacial tension from Eq.(6.46) in terms of the Helmholtz

free energy of the system as

, , iT V n

Aa

σ ∂⎛ ⎞= ⎜ ⎟∂⎝ ⎠ (6.47)

Gibbs free energy is defined by

(6.48) G A PV U TS PV H TS= + = − + = −

6-65

Page 725: +Peters Ekwere j. - Petrophysics

where H is the enthalpy of the system. Differentiating Eq.(6.48) gives

(6.49) dG dU VdP PdV TdS SdT= + + − −

Substituting Eq.(6.42) into Eq.(6.49) gives

(6.50) 1

i N

i ii

dG VdP SdT da dnσ μ=

=

= − + + ∑

We may obtain the interfacial tension from Eq.(6.50) in terms of Gibbs free

energy of the system as

, , iT P n

Ga

σ ∂⎛ ⎞= ⎜ ⎟∂⎝ ⎠ (6.51)

In general, changes in P and V accompanying surface changes are small.

6.4.2 Characterization of Microscopic Pore Level Fluid Displacements

We wish to examine the direction of energy change during an

immiscible displacement at the pore scale. The Helmholtz free energy of the

system will change as the displacement progresses. Our objective is to

determine whether the displacement will lead to a decrease or an increase in

the Helmholtz free energy of the system. A system in equilibrium always seeks

to minimize its free energy. Therefore, a displacement that leads to a decrease

in the free energy of the system is favored. Such a displacement will occur

spontaneously (without pumping) given the chance. On the other hand, a

displacement that leads to an increase in the free energy of the system is not

favored and will not occur spontaneously. Such a displacement will have to be

forced by pumping the displacing fluid.

Case 1. Displacement of a Nonwetting Phase by a Wetting Phase

6-66

Page 726: +Peters Ekwere j. - Petrophysics

Consider the displacement of a nonwetting phase by a wetting phase,

such as water displacing oil in a water wet medium, at the pore level as

shown in Figure 6.36. The interfacial forces and interfacial areas at an instant

are also shown in Figure 6.36. The Helmholtz free energy of the system is a

function of the interfacial areas. Thus,

( ), ,sw so woA f a a a= (6.52)

Let the interface move to the right by a small distance dx. The change in the

Helmholtz free energy of the system as the interfacial areas change is given by

sw sosw so wo

A A AdA da da daa a a∂ ∂ ∂

= + +∂ ∂ ∂ wo (6.53)

Figure 6.36. Displacement of a nonwetting phase by a wetting phase at the pore scale.

Suppose the wetting–nonwetting phase interface maintains the same shape

during the displacement. Then

6-67

Page 727: +Peters Ekwere j. - Petrophysics

0woda = (6.54)

Also, from the geometry of Figure 6.36,

2 swda rdx increase in the areaπ= + = (6.55)

2 soda rdx decrease in the areaπ= − = (6.56)

where r is the radius of the pore. The interfacial forces are given in terms of

the Helmholtz free energy by Eq.(6.47). Substituting Eqs.(6.47), (6.54), (6.55)

and (6.56) into Eq.(6.53) gives the change in Helmholtz free energy of the

system as

(2 )so swdA rdx

π σ σ= − − (6.57)

For the wetting–nonwetting interface to move to the right requires an

imbalance in the interfacial forces given by

cosso sw woσ σ σ> + θ (6.58)

Eq.(6.58) can be rewritten as

( ) cosso sw woσ σ σ− > θ (6.59)

The right side of Eq.(5.59) is a positive number. The left side of Eq.(6.59) is

larger than this positive number. When these facts are applied to Eq.(6.57),

we see that the change in the Helmholtz free energy during the displacement

is negative. This means that the Helmholtz free energy will decrease as the

nonwetting phase is displaced by the wetting phase. This is a favored

displacement. In fact, the displacement will occur spontaneously if given the

opportunity to do so. This is the origin of spontaneous imbibition, which

6-68

Page 728: +Peters Ekwere j. - Petrophysics

explains the spontaneous imbibition of a wetting fluid in the capillary tube

experiment. We can calculate an effective displacing force as

0effectivedAFdx

= − > (6.60)

Thus, an effective displacing force develops spontaneously to enable the

wetting phase to displace the nonwetting phase. Such a force will develop

spontaneously to enable water to displace oil in a water wet porous medium.

Case 2. Displacement of a Wetting Phase by a Nonwetting Phase

Consider the displacement of a wetting phase by a nonwetting phase,

such as water displacing oil in an oil wet medium, at the pore level as shown

in Figure 6.37. Let the interface move to the right by a small distance dx.

Because the nonwetting phase never contacts the solid,

0swda = (6.61)

Because the wetting phase is always in contact with the solid,

0soda = (6.62)

The change in the wetting–nonwetting phase interfacial area is given by

(6.63) *2woda r dxπ= +

6-69

Page 729: +Peters Ekwere j. - Petrophysics

Figure 6.37. Displacement of a wetting phase by a nonwetting phase at the pore scale.

where r* is the radius from the center of the pore to the thin wetting phase

film on the surface of the pore. Substituting Eqs.(6.47), (6.61), (6.62) and

(6.63) into Eq.(6.53) gives the change in the Helmholtz free energy of the

system as

2 wodA rdx

π σ= + (6.64)

Eq.(6.64) shows that the Helmholtz free energy of the system increases during

the displacement. This is not a favored displacement. Thus, the displacement

will not occur spontaneously. It must be forced by pumping. The effective

displacement force in this case is given by

0effectivedAFdx

= − < (6.65)

Thus, a negative displacement force arises to oppose the displacement of the

wetting phase by the nonwetting phase. This is why a positive displacement

6-70

Page 730: +Peters Ekwere j. - Petrophysics

pressure is required in order to initiate a drainage capillary pressure

measurement.

To minimize the increase in free energy of the system during the

displacement, r* will be as small as possible. Thus, the injected nonwetting

phase will channel or finger through the wetting phase leaving behind a

significant wetting phase film and residual wetting phase saturation. This

means that the waterflood efficiency of an oil wet medium will be less than the

waterflood efficiency of a water wet medium. This observation is in agreement

with the microscopic picture of the displacements shown in Figure 6.29.

NOMENCLATURE

a = interfacial area

A = Helmholtz free energy

F = force

g = gravitational acceleration

G = Gibbs free energy

h = equilibrium height in a capillary rise experiment

H = enthalpy

M = molecular weight

Mg = apparent molecular weight of gas

ML = apparent molecular weight of liquid

ni = moles of component i

N = total number of components in the mixture

Nc = capillary number

P = pressure

Pc = capillary pressure

Pnw = pressure in the nonwetting phase

Pw = pressure in the wetting phase

r = radius of capillary tube, pore radius, radius of spinning drop

Q = heat

6-71

Page 731: +Peters Ekwere j. - Petrophysics

R = oil recovery

S = entropy

Ss = surface entropy

Sor = residual oil saturation

T = temperature

U = internal energy

v = Darcy velocity

V = volume

W = work

xi = mole fraction of component i in the liquid

yi = mole fraction of component i in the gas

σ = surface or interfacial tension

osσ = oil solid interfacial tension (oil solid specific surface energy)

owσ = oil water interfacial tension (oil water specific surface energy)

wsσ = water solid interfacial tension (water solid specific surface energy)

θ = contact angle

θA = advancing contact angle

θR = receding contact angle

Lρ = saturated liquid density

gρ = saturated gas density

wρ = wetting phase density

nwρ = nonwetting phase density

iμ = chemical potential of component i

wμ = viscosity of wetting phase

nwμ = viscosity of nonwetting phase

Λ = parachor

Π = spreading pressure

6-72

Page 732: +Peters Ekwere j. - Petrophysics

REFERENCES AND SUGGESTED READINGS

Abrams, A. : "The Influence of Fluid Viscosity, Interfacial Tension, and Flow Velocity on Residual Oil Saturation Left by Waterflood," Soc. Pet. Eng. Jour. (Oct., 1975) 437-447.

Adamson, A.W. and Gast, A.P.: Physical Chemistry of Surfaces, Sixth Edition, John Wiley and Sons, Inc., New York, 1997.

Amott, E. : “Observations Relating to Wettability of Porous Rock,” Trans., AIME (1959) 216, 156-162.

Anderson, W.G. : “Wettability Literature Survey - Part 1: Rock/Oil/Brine Interactions and the Effects of Core Handling on Wettability,” J. Pet. Tech. (October 1986) 1125-1144.

Anderson, W.G. : “Wettability Literature Survey - Part 2: Wettability Measurement,” J. Pet. Tech. (November 1986) 1246-1262.

Anderson, W.G. : “Wettability Literature Survey - Part 3: The Effects of Wettability on the Electrical Properties of Porous Media,” J. Pet. Tech. (December 1986) 1371-1378.

Anderson, W.G. : “Wettability Literature Survey - Part 4: Effects of Wettability on Capillary Pressure,” J. Pet. Tech. (October 1987) 1283-1300.

Anderson, W.G. : “Wettability Literature Survey - Part 5: The Effects of Wettability on Relative Permeability,” J. Pet. Tech. (November 1987) 1453-1468.

Anderson, W.G. : “Wettability Literature Survey - Part 6: The Effects of Wettability on Waterflooding,” J. Pet. Tech. (December 1987) 1605-1622.

Archer, J.S. and Wall, C.G. : Petroleum Engineering, Graham & Trotman, London, England, 1986.

Bear, J. : Dynamics of Fluids in Porous Media, Elsevier, New York, 1972.

Benner, F.C. and Bartell, F.E. : “The Effect of Polar Impurities Upon Capillary and Surface Phenomena in Petroleum Production,” API Drilling and Production Practice (1941) 341-348.

Bobek, J.E., Mattax, C.C. and Denekas, M.O. : “Reservoir Rock Wettability - Its Significance and Evaluation,” Trans., AIME (1958) 213, 155-160.

Chatzis, I. and Morrow, N.R. : “Correlation of Capillary Number Relationships for Sandstones,” SPE 10114, Presented at the 56th Annual Fall Technical Conference and Exhibition of the Society of Petroleum Engineers, San Antonio, October 5-7, 1981.

Chilingar, G.V. and Yen, T.F. : “Some Notes on Wettability and Relative Permeabilities of Carbonate Rocks,” Energy Sources , vol. 7, No. 1 (1983) 67-75.

6-73

Page 733: +Peters Ekwere j. - Petrophysics

Choquette, P.W. and Pray, L.C. : “Geologic Nomenclature and Classification of Porosity in Sedimentary Carbonates,” AAPG Bull., Vol. 54, No. 2 (1970) 207-250.

Collins, R.E. : Flow of Fluids Through Porous Materials, Van Nostrand Reinhold Company, 1961. Reprinted by the Petroleum Publishing Company, 1976. Reprinted by Research & Engineering Consultants Inc., 1990.

Craig, F.F., Jr. : The Reservoir Engineering Aspects of Waterflooding, SPE Monograph Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1971.

Cuiec, L.E. : “Study of Problems Related to the Restoration of Natural State of Core Samples,” J. Canadian Pet. Tech (Oct. - Dec. 1977) 68-80.

de Gennes, P.G. and Quere, D. : Capillarity and Wetting Phenomena, Springer Science and Business Media, Inc., 2004.

Denekas, M.O., Mattax, C.C. and Davis, G.T. : “Effects of Crude Oil Components on Rock Wettability,” Trans., AIME (1959) 216, 330-333.

Donaldson, E.C., Thomas, R.D. and Lorenz, P.B. : “Wettability Determination and Its Effect on Recovery Efficiency,” Soc. Pet. Eng. J. (March 1969) 13-20.

Donaldson, E.C., Kendall, R.F., Pavelka, E.A. and Crocker, M.E. : “Equipment and Procedures for Fluid Flow and Wettability Tests of Geological Materials,” DOE/BETC/IC-79/5, Nat. Tech. Info. Sv, Springfield, VA 2216, 1980.

Garnes, J.M., Mathisen, A.M., Scheie, A. and Skauge, A. : "Capillary Number Relations for Some North Sea Reservoir Sandstones," SPE/DOE 20264, Presented at the SPE/DOE Seventh Symposium on Enhanced Oil Recovery, Tulsa, April 22-25, 1990.

Howard, J.J. : "Wettability and Fluid Saturations Determined From NMR T1 Distribution," Magnetic Resonance Imaging, " Vol. 12, No. 2 (1994) 197-200.

Jennings, H.Y. : “Surface Properties of Natural and Synthetic Porous Media,” Producers Monthly (March 1957) 20-24.

Lake, L.W. : Enhanced Oil Recovery, Prentice Hall, Englewood Cliffs, New Jersey, 1989.

Marle, C.M. : Multiphase Flow in Porous Media, Gulf Publishing Company, Houston, Texas, 1981.

Melrose, J.C. : "Interfacial Phenomena as Related to Oil Recovery Mechanisms," Cnd J. Chem. Eng., Vol. 48 (Dec. 1970) 638-644.

Morrow, N.R. : “Wettability and Its Effect on Oil Recovery,” J. Pet. Tech. (December 1990) 1476-1484.

6-74

Page 734: +Peters Ekwere j. - Petrophysics

Morrow, N.R., Cram, P.J. and McCaffery, F.G. : "Displacement Studies in Dolomite With Wettability Control by Octanoic Acid," SPEJ (August 1973) 221-232.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 2 - Interfacial Phenomena and Oil Recovery: Wettability,” World Oil (March 1981) 77-83.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 3 - Interfacial Phenomena and Oil Recovery: Capillarity,” World Oil (May 1981) 149-158.

Mungan, N. and Moore, E.J. : "Certain Wettability Effects on Electrical Resisitivity in Porous Media," J. Cdn. Pet. Tech. (Jan.-March 1968) 7, No.1, 20-25.

Owens, W.W. and Archer, D.L. : “The Effect of Rock Wettability on Oil-Water Relative Permeability Relationships,” J. Pet. Tech. (July 1971) 873-878.

Peters, E. J. and W. D. Hardham: “A Comparison of Unstable Miscible and Immiscible Displacements,” SPE 19640, Proceedings of the 64th Annual Technical Conference of the Society of Petroleum Engineers (October 1989) San Antonio.

Peters, E.J. and Hardham, W.D. : “Visualization of Fluid Displacements in Porous Media Using Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 4, No. 2, (May 1990) 155-168.

Pirson, S.J. : Oil Reservoir Engineering, McGraw-Hill Book Company, Inc., New York, 1958.

Raza, S.H., Treiber, L.E. and Archer, D.L. : “Wettability of Reservoir Rocks and Its Evaluation,” Producers Monthly (April 1968) 2-7.

Rowlinson, J.S. and Widom, B. : Molecular Theory of Capillarity, Dover Publications, Inc., Mineola, New York, 1982.

Sweeny, S.A. and Jennings, H.Y. : "Effect of Wettability on the Electrical Resistivity of Carbonate Rock from a Petroleum Reservoir," J. Phy. Chem. (May 1960) 64, 551-553.

Tiab, D. and Donaldson, E.C. : Petrophysics, Second Edition, Elsevier, New York, 2004.

Treiber, L.E., Archer, D.L. and Owens, W.W. : “A Laboratory Evaluation of the Wettability of Fifty Oil-Producing Reservoirs,” Soc. Pet. Eng. J. (December 1972) 531-540.

Wagner, O.R. and Leach, R.O. : “Improving Oil Displacement Efficiency by Wettability Adjustment,” Trans., AIME (1959) 216, 65-72.

6-75

Page 735: +Peters Ekwere j. - Petrophysics

6-76

Welge, H.J. and Bruce, W.A. : “A Restored-State Method for Determination of Oil in Place and Connate Water,” API Drilling and Production Practice (1947) 161-165.

Willhite, G. P. : Waterflooding, SPE Textbook Series Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1986.

Page 736: +Peters Ekwere j. - Petrophysics

7-1

CHAPTER 7

CAPILLARY PRESSURE

7.1 DEFINITION OF CAPILLARY PRESSURE

When two immiscible fluids are in contact, there is a pressure

discontinuity between the two fluids which depends upon the curvature of the

interface separating the two fluids. This pressure difference or excess

pressure is known as the capillary pressure. The pressure on the concave side

of the interface is higher than that on the convex side of the interface. Figure

7.1 shows a curve interface between to immisicible fluids labeled 1 and 2. The

pressure P2 is greater than P1. The capillary pressure is given by Laplace

equation (sometimes referred to as Young-Laplace equation) as

2 11 2

1 1cP P P

r rσ

⎛ ⎞= − = +⎜ ⎟

⎝ ⎠ (7.1)

In Eq.(7.1), r1 and r2 are referred to as the principal radii of curvature of the

interface. They are mutually perpendicular. The curvature of the interface is

given by

Page 737: +Peters Ekwere j. - Petrophysics

7-2

1 2

1 1Curvaturer r

⎛ ⎞= +⎜ ⎟

⎝ ⎠ (7.2)

Figure 7.1. Equilibrium at a curved interface between two immiscible fluids.

Page 738: +Peters Ekwere j. - Petrophysics

7-3

Laplace equation can be derived by considering the mechanical

equilibrium of the interface or by energy considerations. We derive it here by

energy considerations. Let the interface be expanded by a small amount in the

xyz directions, where z is vertically upward. The increase is surface area of the

interface is given by

( )( )da x dx y dy xy xdy ydx dxdy xdy ydx= + + − = + + = + (7.3)

where dxdy is considered negligibly small. The increase in surface energy is

given by

( )dE xdy ydxσ= + (7.4)

The work done in increasing the interfacial area is given by

( )2 1W P P xydzδ = − (7.5)

At equilibrium, the work done is equal to the increase in surface energy.

Thus,

( ) ( )2 1P P xydz xdy ydxσ− = + (7.6)

From similar triangles (Figure 7.2),

1 1

x dx xr dz r

+=

+ (7.7)

2 2

y dy yr dz r

+=

+ (7.8)

Substituting Eqs.(7.7) and (7.8) into Eq.(7.6) and rearranging gives

Page 739: +Peters Ekwere j. - Petrophysics

7-4

2 11 2

1 1cP P P

r rσ

⎛ ⎞= − = +⎜ ⎟

⎝ ⎠ (7.1)

which is Laplace equation. A mean radius of curvature, rm, may be defined as

1 2

1 1 1 12mr r r

⎛ ⎞= +⎜ ⎟

⎝ ⎠ (7.9)

In terms of the mean radius of curvature, Laplace equation becomes

2c

m

Prσ

= (7.10)

Figure 7.2. Fluid interface in the planes of the principal radii of curvature.

Page 740: +Peters Ekwere j. - Petrophysics

7-5

Laplace equation is the fundamental equation of capillarity. Several

special cases of Laplace equation are of interest. For a spherical liquid drop,

r1 = r2 = r, the radius of the drop. The capillary pressure or the excess

pressure of the drop is given by

2cP

= (7.11)

For a soap bubble in air, r1 = r2 = r, the radius of the bubble. The capillary

pressure or excess pressure is given by

4cP

= (7.12)

where a factor of 2 has been incorporated to account for the two gas-liquid

interfaces of a soap bubble. For a flat interface, r1 = r2 = ∞. In this case, the

capillary pressure is zero. If the two immiscible fluids are in contact with a

solid surface, the interface will intersect the solid at an equilibrium contact

angle θ given by the Young-Dupre equation. Such an interface is shown in

Figure 7.3 for the capillary rise experiment. Laplace equation holds at the

interface. Assuming the interface lies on a sphere as shown in the figure, then

r1 = r2 = r/cosθ, where r is the radius of the capillary tube. The capillary

pressure is given by

1 2

1 1 cos cos 2 coscP

r r r r rθ θ σ θσ σ

⎛ ⎞ ⎛ ⎞= + = + =⎜ ⎟ ⎜ ⎟⎝ ⎠⎝ ⎠

(7.13)

Page 741: +Peters Ekwere j. - Petrophysics

7-6

Figure 7.3. Interface for capillary rise experiment.

For a pendular ring of wetting fluid at the contacts of two spherical

sand grains in an idealized porous medium consisting of a cubic pack of

uniform spheres as shown in Figure 7.4, the capillary pressure is given by

Laplace equation as

1 2

1 1coscPr r

σ θ⎛ ⎞

= −⎜ ⎟⎝ ⎠

(7.14)

Page 742: +Peters Ekwere j. - Petrophysics

7-7

In this case, the principal radii of curvature are on opposite sides of the

interface. By sign convention, one radius will be positive and the other will be

negative.

Figure 7.4. Immiscible fluid interface in an idealized porous medium.

If the wetting fluid saturation in the pendular ring is reduced, r1 and r2 will be

reduced. However, r1 will be reduced more than r2 as the wetting phase

recedes into the corners of the contact of the grains. As a result, the capillary

pressure will increase. If the wetting fluid saturation is increased, r1 and r2

will increase and the capillary pressure will decrease. Therefore, an inverse

relationship exists between the capillary pressure and the wetting phase

saturation for a porous medium. Low wetting phase saturation corresponds

Page 743: +Peters Ekwere j. - Petrophysics

7-8

to high capillary pressure and high wetting phase saturation corresponds to

low capillary pressure.

7.2 CAPILLARY PRESSURE - SATURATION RELATIONSHIP FOR A POROUS MEDIUM

Before considering the capillary pressure versus saturation relationship

for a porous medium, it is instructive to consider the relationship for an

idealized medium consisting of a bundle of capillary tubes of varied radii. In

this case, the capillary pressure versus wetting phase saturation relationship

can be calculated. Let the bundle of capillary tubes medium be dipped into

the wetting phase and allowed to attain capillary equilibrium as shown in

Figure 7.5. The wetting fluid will rise to a different elevation (z) above the free

wetting fluid level in each tube depending on its radius as shown in the figure.

Let the model consist of ten capillary tubes with their radii as shown in Table

7.1. Let the wetting fluid be water with a surface tension of 72 dynes/cm and

a contact angle of 0 with the solid. The non-wetting phase is air. The

equilibrium height of water in each capillary tube can be calculated with

Eq.(6.18) as

2 cos 0.1468

w

z cmr g rσ θ

ρ= = (7.15)

The capillary pressure in each tube is given by Laplace equation as

22 cos 144 /cP dynes cmr r

σ θ= = (7.16)

The wetting phase saturation as a function of elevation z is calculated from

the dimensions of the capillary tubes and is presented in Table 7.1.

Page 744: +Peters Ekwere j. - Petrophysics

7-9

Figure 7.5. Capillary rise experiment for a bundle of capillary tubes medium.

Table 7.1. Capillary Pressure versus Wetting Phase Saturation for Bundle of

Capillary Tubes Model.

Radius Pc Pc Pc z Volume Volume Water

(μm) (dynes/cm2

)

(atm) (psi) cm (cm3) Fraction Saturation

0.00 1.000

100 14400.0 0.014 0.209 14.68 0.0046 0.182 0.818

90 16000.0 0.016 0.232 16.31 0.0042 0.164 0.655

80 18000.0 0.018 0.261 18.35 0.0037 0.145 0.509

70 20571.4 0.020 0.298 20.97 0.0032 0.127 0.382

60 24000.0 0.024 0.348 24.46 0.0028 0.109 0.273

50 28800.0 0.028 0.418 29.36 0.0023 0.091 0.182

40 36000.0 0.036 0.522 36.70 0.0018 0.073 0.109

30 48000.0 0.047 0.696 48.93 0.0014 0.055 0.055

20 72000.0 0.071 1.044 73.39 0.0009 0.036 0.018

10 144000.0 0.142 2.088 146.79 0.0005 0.018 0.000

Page 745: +Peters Ekwere j. - Petrophysics

7-10

0.0254

Figures 7.6 and 7.7 show the capillary pressure versus wetting phase

saturation for this idealized medium. In Figure 7.6, the capillary pressure is

presented as height in cm above the free water level whereas in Figure 7.7,

the capillary pressure is given in psi. Both presentations are valid and can be

used for different purposes. The presentation in Figure 7.6 gives the water

saturation distribution as a function of the height above the free water level.

This type of presentation can be used to determine the water saturation

distribution in a petroleum reservoir starting from the free water level at or

below the oil water contact to the top of the reservoir. The presentation in

Figure 7.7 is useful for calculating pore size distribution.

The capillary pressure versus saturation relationship for the idealized

porous medium shown in Figures 7.6 and 7.7 captures the general trend of

capillary pressure curves for porous media. The features include a non-zero

displacement pressure at a wetting phase saturation of 1 and the inverse

relationship between capillary pressure and wetting phase saturation

mentioned previously. The capillary pressure curves have a stair-case shape

in this case because of the limited number and size of the capillary tubes

used in the experiment. The curve will approach a smooth curve if more tubes

are used and the differences in the tube diameters are made small. The only

limitation of the capillary pressure curves for the bundle of capillary tube

model is the absence of an irreducible wetting phase saturation. There is no

possibility of trapping an irreducible saturation for a model consisting of

straight and isolated capillary tubes.

Page 746: +Peters Ekwere j. - Petrophysics

7-11

Figure 7.6. Capillary pressure expressed as height of water above the free water level versus wetting phase saturation for a bundle of capillary tubes

medium.

Figure 7.7. Capillary pressure in psi versus wetting phase saturation for a bundle of capillary tubes medium.

Page 747: +Peters Ekwere j. - Petrophysics

7-12

Figure 7.8 shows capillary rise experiments for two porous media

having different grain sizes (pore sizes). The wetting phase will rise higher in

the finer grain porous medium than in the courser grain medium. Such

experiments are widely used in soil science to determine the capillary

pressure curve (referred to as water retention curve or matric suction head in

soil science) for unconsolidated soils. The soil, which is packed in a tube is

dipped into water and allowed to sit for several days or weeks to achieve

capillary equilibrium. The tube often is instrumented to measure the

resistivity of the medium in order to calculate the water saturation along the

column by Archie's equation. The capillary pressure is calculated as height of

water above the free water level as was done in the bundle of capillary tube

model.

Figure 7.8. Capillary rise experiments for two porous media of different grain sizes.

Page 748: +Peters Ekwere j. - Petrophysics

7-13

In an actual porous medium, such as reservoir rock, the complexity of

the pore structure and the fluid interface arrangements preclude the use of

Laplace equation to calculate the capillary pressure. Further, this complexity

also precludes the calculation of the wetting phase saturation from the fluid

interface arrangements as was done for the bundle of capillary tubes

experiment. Instead, the capillary pressure versus wetting phase saturation

relationship is measured experimentally. We demonstrate one possible way of

making this measurement for a porous medium using the idealized porous

medium and the experimental set up shown in Figure 7.9. Let water be the

wetting phase and air the non-wetting phase in the experiment. The porous

medium consists of one pore, which is a capillary tube with three radii as

shown in the figure. The radius and length of each segment of the pore are

shown in Table 7.2. The medium, which is strongly water wet (θ = 0º), is

initially saturated with water. The medium rests on a semi-permeable plate at

the bottom of the apparatus. This semi-permeable plate is manufactured such

that it has very fine and uniform pores. It is strongly water wet and is fully

saturated with water. It will permit water to flow through it but because of its

fine pores will prevent the air from flowing through it. Initially, the apparatus

is open to the atmosphere so that the water in the core, the semi-permeable

plate and the connecting vessel below the semi-permeable plate is at

atmospheric pressure. Gas is admitted into the apparatus at a low pressure of

Pg. If Pg is less than 2σcosθ/r1, nothing will happen. The gas pressure is not

high enough for the gas to displace the water from the largest pore. The gas

pressure is then increased to Pg1 until it is equal to 2σcosθ/r1 and the segment

of the pore with radius r1 will be drained of water. The water drainage will

stop after draining the largest pore because Pg1 is not high enough to drain

the pore with radius r2. Next, the gas pressure is increased to Pg2 equal to

2σcosθ/r2 and the segment of the pore with radius r2 will be drained.

Eventually, the pressure of the gas is increased to Pg3 equal to 2σcosθ/r3 and

the segment of the pore with radius r3 will be drained. The volume of water

Page 749: +Peters Ekwere j. - Petrophysics

7-14

drained at each capillary pressure is measured and is used to calculate the

water saturation in the medium corresponding to that capillary pressure. The

graph of Pgi versus water saturation gives the capillary pressure curve of the

porous medium. For this simple porous medium, the capillary pressure curve

can be calculated using Laplace equation and the dimensions of the pore and

is presented in Figure 7.10. Note how the shape of the capillary pressure

curve reflects the pore size distribution of the porous medium. This is the

basis for estimating the pore size distribution of a porous medium from its

drainage capillary pressure curve.

Figure 7.11 shows the effect of wettability on the capillary pressure

curve for the idealized porous medium. It compares the capillary pressure

curve for a contact angle of 0º and a contact angle of 75º. Thus, if the medium

is less water wet, the magnitude of the capillary pressure will be less at each

wetting phase saturation than when it was more water wet. Since the

wettability preference of the rock for the water is less at a contact angle of 75º

than at 0º, less work is required to desaturate the rock. Therefore, the

capillary pressure that needs to be applied to desaturate the rock will be less

at a contact angle of 75º than at 0º at any saturation level.

Page 750: +Peters Ekwere j. - Petrophysics

7-15

Figure 7.9. Capillary pressure measurement for an idealized porous medium.

Table 7.2. Capillary Pressure Curve for Idealized Porous Meium.

Capillary Radius Length Volume Fractiona

l

Pc Pc Pc Water

Tube # (μm) (μm) (cm3) Volume (dynes/cm2

)

(atm) (psi) Saturation

1 10 2 6.283E-

10

0.5747 144000 0.142 2.088 0.4253

2 4 8 4.021E-

10

0.3678 360000 0.355 5.221 0.0575

3 1 20 6.283E-

11

0.0575 1440000 1.421 20.884 0.0000

1.093E-

09

Page 751: +Peters Ekwere j. - Petrophysics

7-16

Figure 7.10. Capillary pressure curve for an idealized porous medium.

Page 752: +Peters Ekwere j. - Petrophysics

7-17

Figure 7.11. Effect of wettability on the capillary pressure curve for an idealized porous medium.

7.3 DRAINAGE CAPILLARY PRESSURE CURVE

If the idealized porous medium of Figure 7.9 were replaced by an actual

porous medium and the experiment repeated, the capillary pressure curve

would look like the one shown in Figure 7.12. This figure shows a typical

drainage capillary pressure curve obtained by displacing the wetting phase

from a porous medium with a non-wetting phase. A process in which the

wetting phase saturation decreases is known as drainage whereas the

converse process in which the wetting phase saturation increases is known as

imbibition. The drainage capillary pressure curve has several characteristic

features. The curve shows that a minimum positive pressure (Pd) must be

applied to the non-wetting phase in order to initiate the drainage. This

minimum pressure, which is known as the displacement pressure, the

threshold pressure or the entry pressure, is determined by the size of the

largest pores connected to the surface of the medium. It can be estimated

with Laplace equation where r is the largest pore radius connected to the

surface. If the rock does not have a strong wettability preference for the

initially saturating fluid, then the displacement pressure will be zero. If the

rock has a strong preference for the displacing fluid, then no pressure is

required to initiate the displacement because it will occur spontaneously. In

this case, the capillary pressure will start at the initial fluid saturation of less

than 1. As the pressure of the non-wetting phase is increased, smaller and

smaller pores are invaded by the non-wetting fluid. Eventually, the wetting

phase becomes discontinuous and can no longer be displaced from the

medium by increasing the capillary pressure. Therefore, an irreducible

wetting phase saturation is achieved for the porous medium at a high

capillary pressure. At the irreducible wetting phase saturation, the capillary

pressure curve becomes nearly vertical. The irreducible wetting phase

Page 753: +Peters Ekwere j. - Petrophysics

7-18

saturation is a function of the grain size (pore size), the wettability of the

medium and the interfacial tension between the wetting and non-wetting

fluids.

Figure 7.12. A typical drainage capillary pressure curve.

What information does the capillary pressure curve for a reservoir rock

provide about the rock? If one reservoir rock has a higher permeability than

another, we know that the higher permeability rock will permit faster fluid

flow through it, everything else being equal, than the lower permeability rock.

Thus, the higher permeability rock is more desirable than the lower

permeability rock as a petroleum reservoir rock. If one reservoir rock has a

higher porosity than another, we know that the higher porosity rock will store

more reserves than the lower porosity rock. Therefore, the higher porosity

Page 754: +Peters Ekwere j. - Petrophysics

7-19

rock is a more desirable reservoir rock than the lower porosity rock. If one

rock has a higher capillary pressure at the same wetting phase saturation

than another, what can we say about the rocks? Is the rock with the higher

capillary pressure curve more desirable or less desirable as a petroleum

reservoir rock than the rock with the lower capillary pressure curve?

Figure 7.13 shows the drainage capillary pressure curves for four rocks:

A, B, C and D. Rock A has the least displacement pressure. Therefore, it has

the largest pores connected to the surface. Its capillary pressure curve

remains essentially flat as the wetting phase saturation is decreased from

100% to 60%. This means that many of the pores are invaded by the non-

wetting fluid at essentially the same capillary pressure. This indicates that A

has uniform pores or is well sorted. Rock A also has the least irreducible

wetting phase saturation, indicating that it has relatively larger grains and

pores than the other rocks. Rock B has a higher displacement pressure than

A. Therefore, it has smaller pores than A. The capillary pressure curve at the

high wetting phase saturations is relatively flat, indicating good sorting. Rock

B has a higher irreducible wetting phase saturation than A, which is

consistent with its finer grains and pores. Rock C is even more fine grained

than B because of its higher displacement pressure. The shape of its capillary

pressure curve shows that a higher capillary pressure is required at each

wetting phase saturation to desaturate the rock. This means that C has a

wider pore size distribution than A and B. Therefore, C is poorly sorted. It

has a higher irreducible water saturation than B, which is consistent with its

finer grains and pores. Rock D is extremely fine grained, extremely poorly

sorted and would be a very poor reservoir rock. This observation is based on

its very high displacement pressure, very steep capillary pressure curve and

very high irreducible wetting phase saturation. Without being told, one can

easily infer that this rock is essentially made of clay.

Page 755: +Peters Ekwere j. - Petrophysics

7-20

Since permeability is proportional to the square of the mean grain size

(pore size), it is easy to see that Rock A has the highest permeability, followed

by B, C and D in that order. From this discussion, we conclude that the rock

with the higher capillary pressure curve is a less desirable reservoir rock than

the one with the lower capillary pressure curve.

In general, the capillary pressure curve for a porous medium is a

function of (1) pore size, (2) pore size distribution, (3) pore structure, (4) fluid

saturation, (5) fluid saturation history, (6) wettability of the rock and (7)

interfacial tension of the fluids involved.

Figure 7.13. Capillary pressure curves for four different rocks.

Page 756: +Peters Ekwere j. - Petrophysics

7-21

7.4 CONVERSION OF LABORATORY CAPILLARY PRESSURE DATA TO RESERVOIR CONDITIONS

Typically, capillary pressure curves are measured in the laboratory

using fluids other than reservoir fluids. It is not uncommon to measure the

capillary pressure curves to be used for analyzing an oil-water reservoir using

air and water or mercury and air in the laboratory. When this is done, it

becomes necessary to convert the laboratory data to reservoir conditions.

This conversion is done using Laplace equation as follows. From Laplace

equation,

( ) ( )2 coslab

c labm

Pr

σ θ= (7.17)

( ) ( )2 cosreservoir

c reservoirm

Pr

σ θ= (7.18)

where rm is the mean radius of curvature of the interface in the rock at a

particular fluid saturation. Eliminating rm from Eqs.(7.17) and (7.18) gives

( ) ( ) ( )( )

coscos

reservoirc creservoir lab

lab

P Pσ θ

σ θ= (7.19)

This ability to scale the laboratory capillary pressure data to reservoir

conditions provides the flexibility for making laboratory capillary pressure

measurements with more convenient fluids than reservoir fluids.

7.5 AVERAGING CAPILLARY PRESSURE DATA

The capillary pressure curves for rock samples from the same reservoir

having different permeabilities will be different. It is often necessary to

average the capillary pressure data for cores from the same reservoir believed

Page 757: +Peters Ekwere j. - Petrophysics

7-22

to have the same pore structure in order to obtain one capillary pressure

curve that can be used for reservoir performance analysis. This averaging can

be done using the Leverett J-function, which is a dimensionless capillary

pressure function (Leverett, 1941).

The Leverett J-function can be derived by dimensional analysis as

follows. The capillary pressure curve of a porous medium is a function of

several variables as shown in Eq.(7.20).

( ), , cos , ,c w w nwkP f S gσ θ ρ ρφ

⎛ ⎞= Γ −⎜ ⎟

⎝ ⎠ (7.20)

where Pc is the capillary pressure, Sw is the wetting phase saturation, Γ is a

dimensionless pore structure function that accounts for such things as pore

size distribution, tortuosity, cementation and dead end pores, φ is the

porosity, σ is the interfacial tension, θ is the contact angle, k is the absolute

permeability of the porous medium, ρw is the wetting phase density, ρnw is the

non-wetting phase density and g is the gravitational acceleration. The wetting

phase saturation (Sw) and the pore structure function (Γ) are dimensionless

and should be set aside from the dimensional analysis until the end. We form

the dimensionless product with the remaining variables as

( ) ( )2

31 4cos dimensionless constantx

xx xw nw c

k g Pσ θ ρ ρφ

⎛ ⎞− =⎡ ⎤⎜ ⎟ ⎣ ⎦

⎝ ⎠ (7.21)

Carrying out the dimensional analysis yields the following solution to the

dimensional analysis problem:

Page 758: +Peters Ekwere j. - Petrophysics

7-23

( )

11

2 23 4

3

4

cos1 1

11 00 1w nw

c

xk

xx x

xg

xP

σ θ

φρ ρ

⎡ ⎤− −⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥

⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥= +⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥− ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎣ ⎦ ⎣ ⎦⎣ ⎦⎢ ⎥⎣ ⎦

(7.22)

Let us choose x3 = 1 and x4 = 0. The corresponding dimensionless group is

given by

( ) ( )1

/cos

w nw g kρ ρ φπ

σ θ−

= (7.23)

Next, let us choose x3 = 0 and x4 = 1. The corresponding dimensionless group

is given by

2/

coscP k φ

πσ θ

= (7.24)

From the dimensional analysis, we can write

( ) ( )1

// , ,cos cos

w nwcw

g kP k f Sρ ρ φφ

σ θ σ θ−⎛ ⎞

= Γ⎜ ⎟⎝ ⎠

(7.25)

The dimensionless group given as π2 in Eq.(7.23) is the ratio of gravity force to

the capillary forces at the pore scale. At the pore scale, capillary forces

dominate the gravity force. Therefore, π2 will be small and can be neglected.

For example, for an air-water capillary pressure curve, σ = 72 dynes/cm, ρw =

1 g/cm3, θ = 0º, g = 981 cm/s2. For a porous medium of 1 darcy permeability

and a porosity of 25%, π2 is of the order of 710− . For a mercury-air capillary

pressure curve, it is of the order of 610− . Eq.(7.25) then becomes

Page 759: +Peters Ekwere j. - Petrophysics

7-24

( ) ( )2/ , ,

cosc

w wP k f S J Sφσ θ

= Γ = Γ (7.26)

where ( ),wJ S Γ is a dimensionless capillary pressure function known as the

Leverett J-function. Eq.(7.26) suggests that porous media that have the same

pore structure but different permeability and porosity will have the same

Leverett J-function. Therefore, if the different capillary pressure curves of the

porous media are rescaled as a Leverett J-function, they should plot as one

curve. This curve provides the means to average capillary pressure data.

Figure 7.14 shows the Leverett J-function for nine unconsolidated

sands with widely different permeabilities ranging from 0.057 to 2160 darcies.

It is remarkable that the data plot as one curve. Figure 7.15 shows the J-

function for a carbonate reservoir. If the porous media have different pore

structures, then the Leverett J-functions for the different rocks will be

different and will not plot as one curve as may be seen in Figure 7.16.

Page 760: +Peters Ekwere j. - Petrophysics

7-25

Figure 7.14. Leverett J function for unconsolidated sands (Leverett, 1941).

Page 761: +Peters Ekwere j. - Petrophysics

7-26

Figure 7.15. Leverett J-functions for a carbonate reservoir; (a) all cores; (b) limestone cores; (c) dolomite cores; (d) microgranular limestone cores; (e)

coarse-grained limestone cores (Brown, 1951).

Page 762: +Peters Ekwere j. - Petrophysics

7-27

Figure 7.16. Leverett J-functions for different rock types (Rose and Bruce,

1949).

Page 763: +Peters Ekwere j. - Petrophysics

7-28

7.6 DETERMINATION OF THE INITIAL STATIC RESERVOIR FLUID SATURATIONS BY USE OF DRAINAGE CAPILLARY PRESSURE CURVE

Initially, the petroleum reservoir was saturated with water before oil

migrated into the reservoir and displaced the water. This displacement of a

wetting phase by a non-wetting phase is simulated in the laboratory

measurement of the drainage capillary pressure curve. The final fluid

distribution in the reservoir is determined by the equilibrium between

capillary and gravitational forces.

Consider the static equilibrium for the water and oil. From

hydrostatics, for the water,

ww

dP gdz

ρ= − (7.27)

where is z is pointed vertically upwards. Similarly, for the oil,

oo

dP gdz

ρ= − (7.28)

Assuming the fluids are incompressible, Eqs.(7.27) and (7.28) can be

integrated to obtain

( ) ( )0w w wP z P gzρ= − (7.29)

and

( ) ( )0o o oP z P gzρ= − (7.30)

Subtracting Eq.(7.29) from (7.30) gives

Page 764: +Peters Ekwere j. - Petrophysics

7-29

( ) ( ) ( )0c c w oP z P gzρ ρ= + − (7.31)

We select as datum the free water level at which the capillary pressure is zero.

With this choice of datum, Eq.(7.31) becomes

( ) ( )c w oP z gz gzρ ρ ρ= − = Δ (7.32)

where Δρ is the density of water minus the density of oil. It is remarkable that

Eq.(7.32) is the same as Eq.(6.26) for the capillary rise experiment. The free

water level occurs at a depth do below the oil-water contact given by

do

Pdgρ

(7.33)

where Pd is the displacement pressure of the capillary pressure curve. Thus,

the elevation above the oil-water contact of any particular saturation is given

by

c dP Phgρ

−=

Δ (7.34)

If the displacement pressure of the capillary pressure curve is 0, then the free

water level and the oil-water contact will be the same. However, this is a

special case. In general, the free water level and the oil-water contact are not

the same.

Figure 7.17 shows (1) a typical static fluid distribution in a

homogeneous reservoir, (2) the oil-water contact, (3) the free water level and

(4) the oil and water pressure profiles. Note the transition zone above the oil-

water contact in which the water saturation decreases from 100% to the

irreducible water saturation. The height of this transition zone is a function

Page 765: +Peters Ekwere j. - Petrophysics

7-30

of the wettability of the rock, the oil-water density contrast, the oil-water

interfacial tension, the grain size (pore size) and sorting, which determine the

permeability of the rock.

Figure 7.17. Initial static fluid distribution in a homogeneous reservoir.

Page 766: +Peters Ekwere j. - Petrophysics

7-31

It should be noted that the capillary pressure in the reservoir is highest

at the top of the reservoir. In order to prevent escape of the hydrocarbon from

the reservoir, the cap rock must have a displacement pressure that is higher

than the maximum capillary pressure labeled Pcap in the figure. Shales

typically form the cap rock in many reservoirs. Shales are fine grained and

have very high displacement pressures. The shale, which is saturated with

water, will prevent the oil from penetrating it because its displacement

pressure is higher than the maximum capillary pressure in the reservoir. Of

course, some may think that the oil does not penetrate the shale because it

has a low permeability. However, the correct analysis is that the shale

prevents the oil from penetrating it because its displacement pressure is

much higher than the pressure in the oil phase. This is a capillary

phenomenon not a Darcy law phenomenon.

Figure 7.18 shows the initial fluid distribution in an actual petroleum

reservoir based on log analysis. The track labeled "Bulk Volume Analysis"

shows the water and oil distributions in the pay zone as a percent of the bulk

volume of reservoir rock. The dark area gives the oil content and the light area

to its left gives the water content. A careful examination of this section of the

plot shows a water saturation versus depth graph that is similar to the water

saturation versus depth graph sketched in Figure 7.17.

Page 767: +Peters Ekwere j. - Petrophysics

7-32

Figure 7.18. Initial static fluid distribution in an actual petroleum reservoir.

Page 768: +Peters Ekwere j. - Petrophysics

7-33

In a layered reservoir in which the layers have different capillary

pressure curves, the layers are in capillary equilibrium. As a result,

saturation discontinuities will occur. However, there will be only one free

water level. Figure 7.19 shows the water saturation distribution for a well

that has penetrated a layered reservoir. Given the capillary pressure curve for

each layer, it is a simple matter to apply Eqs.(7.32) and (7.33) to calculate the

water saturation distribution from the free water level to the top of the

reservoir. The steps for calculating the water saturation distribution in such a

heterogeneous reservoir is as follows.

1. Using the displacement pressure of the bottom layer, calculate the free

water level using Eq.(7.33).

2. Take a small value of z measured from the free water level.

3. Calculate the capillary pressure at that level using Eq.(7.32).

4. Determine the layer in which z occurs.

5. Using the capillary pressure curve for the layer in which z occurs, read

or calculate the water saturation for the value of capillary pressure from

step 3.

6. If z is at the boundary of two layers, there will be a saturation

discontinuity at that value of z. Two saturation values should be

calculated one from each of the capillary pressure curves of the two

layers involved.

7 Increase the value of z and repeat steps 3 through 6 until z reaches the

top of the reservoir.

Page 769: +Peters Ekwere j. - Petrophysics

7-34

This is how the saturation distribution in Figure 7.18c was calculated. If you

look closely at Figure 7.18b in which the four layers have been identified and

their capillary pressure curves have been plotted as height above the free

water level, you can mentally sketch the water saturation distribution over the

entire column of the well.

Figure 7.18. Fluid distribution for a layered reservoir; (a) well penetrating a

layered reservoir; (b) capillary pressure curves for the layers; (c) water staturation profile observed at the well (Archer and Wall, 1986).

Page 770: +Peters Ekwere j. - Petrophysics

7-35

Example 7.1

Table 7.3 gives the properties of an idealized oil reservoir consisting of four

layers with distinct petrophysical properties. The top of the reservoir is at

8000 ft below the surface and the oil water contact is at 8185 ft. Table 7.4

gives the drainage oil-water capillary pressure curve for Layer 1. All the layers

have the same pore structure but different permeabilities and porosities.

Table 7.3. Petrophysical Properties of Idealized Layered Reservoir.

Layer 1 Layer 2 Layer 3 Layer 4 Depth

(ft) 8000-8050

8050-8070

8070-8125

8125-8185

h (ft) 50 20 55 60 k (md) 144 50 10 200

φ %) 23.5 20 18 24

Table 7.4. Drainage Capillary Pressure Curve for Layer 1

Pc1

Sw (psi)

1.000 1.973

0.950 2.377

0.900 2.840

0.850 3.377

0.800 4.008

0.750 4.757

0.700 5.663

0.650 6.781

0.600 8.195

0.550 10.039

0.500 12.547

0.450 16.154

0.400 21.787

0.350 31.817

0.300 54.691

0.278 78.408

Page 771: +Peters Ekwere j. - Petrophysics

7-36

Other properties for the reservoir are as follows:

1.036wρ = g/cm3

0.822oρ = g/cm3

35σ = dynes/cm

0θ =

1. Calculate and plot the graph of the Leverett J-function for the reservoir.

2. Calculate and plot the capillary pressure curves for Layers 2, 3 and 4,

together with that of Layer 1.

3. Calculate the depth of the free water level for the reservoir.

4. Calculate and plot graphs of the initial water and oil saturations in the

reservoir from 8000 ft to the free water level assuming the reservoir is in

capillary equilibrium.

5. Calculate and plot graphs of the water and oil pressures at the initial

reservoir conditions.

6. A well drilled into the reservoir has been perforated from 8090 to 8110

ft. Determine the type of reservoir fluid that will be produced initially.

Solution to Example 7.1

1. The Leverett J-function is calculated with Eq.(7.26) using the capillary

pressure curve for Layer 1. Consistent units are required to make the

function dimensionless. For example, a set of consistent units is cP in

Page 772: +Peters Ekwere j. - Petrophysics

7-37

dynes/cm2, k in cm2 and σ in dynes/cm. For example, at 0.278wS = ,

78.408cP = psi, the J-function is calculated as

( )( )( ) ( )( ) ( )6 978.408 /14.696 1.0133 10 144 /1000 9.689 10 / 0.235

0.278, 11.90235cos0

x xJ

Γ = =

The calculated Leverett J-function is presented in Table 7.5 and Figure

7.19. It is the characteristic underlying dimensionless capillary

pressure curve for the reservoir.

Table 7.5. Summary of Calculated Leverett J-function and the Capillary Pressure Curves for Layers 2, 3 and 4 for Example 7.1.

Layer 1 Layer 2 Layer 3 Layer 4

Pc1 Pc2 Pc3 Pc4

Sw (psi) J(Sw) (psi) (psi) (psi)

1.000 1.973 0.299 3.089 6.553 1.692

0.950 2.377 0.361 3.721 7.894 2.038

0.900 2.840 0.431 4.446 9.432 2.435

0.850 3.377 0.513 5.287 11.215 2.896

0.800 4.008 0.608 6.275 13.311 3.437

0.750 4.757 0.722 7.447 15.799 4.079

0.700 5.663 0.860 8.866 18.807 4.856

0.650 6.781 1.029 10.616 22.520 5.815

0.600 8.195 1.244 12.830 27.217 7.027

0.550 10.039 1.524 15.717 33.341 8.609

0.500 12.547 1.905 19.643 41.670 10.759

0.450 16.154 2.452 25.290 53.649 13.852

0.400 21.787 3.307 34.109 72.357 18.683

0.350 31.817 4.830 49.812 105.668 27.283

0.300 54.691 8.302 85.624 181.635 46.898

0.278 78.408 11.902 122.755 260.402 67.235

Page 773: +Peters Ekwere j. - Petrophysics

7-38

Figure 7.19. Leverett J-function for the reservoir of Example 7.1.

2. Since all the layers have the same pore structure, they share the same

Leverett J-function. Thus, Eq.(7.26) can be solved for Pc using the

known J-function from Layer 1. However, for this example, the capillary

pressure curve for Layer j can be calculated from the data for Layer 1 as

11

1

jcj c

j

kP Pkφ

φ⎛ ⎞⎛ ⎞

= ⎜ ⎟⎜ ⎟⎜ ⎟⎝ ⎠⎝ ⎠

For example, at Sw = 0.278,

2144 2078.408 122.75523.5 50cP ⎛ ⎞⎛ ⎞= =⎜ ⎟⎜ ⎟

⎝ ⎠⎝ ⎠ psi

Page 774: +Peters Ekwere j. - Petrophysics

7-39

The calculated capillary pressure curves are presented in Table 7.5 and

Figure 7.20.

Figure 7.20. Capillary pressure curves for all the layers for the reservoir of Example 7.1.

3. The free water level occurs in Layer 4. Therefore, the displacement

pressure for Layer 4 is used to calculate the depth of the free water level

below the oil water contact from Eq.(7.33) as

( )( )

( )( )

61.6919 /14.696 1.0133 10555.69

1.036 0.822 981o

xd = =

− cm

555.69 / 30.48 18.23od = = ft

Page 775: +Peters Ekwere j. - Petrophysics

7-40

Eq.(7.33) can also be written in oilfield units as

[ ] [ ] [ ]3 3

144/ /

144

c co

P psi P psid ft

lb ft lb ftρ ρ= =

⎡ ⎤ ⎡ ⎤Δ Δ⎣ ⎦ ⎣ ⎦

( )( )( )( ) ( )( )

144 1.691918.24

1.036 62.4 0.822 62.4od = =−

ft

The free water level is at 8185+18.24 = 8203.24 ft.

4. The capillary pressure is zero at the free water level and increases as a

linear function of height above the free water level for incompressible

liquids. The height of any point in the reservoir above the free water

level is designated as z in Eq.(7.32) and can be used to calculate the

capillary pressure at that point in the reservoir. Clearly, at any depth D,

z is given by

z FWL D= −

For example at D = 8100 ft, which is in Layer 3,

8203.24 8100 103.24z = − = ft

The capillary pressure is calculated with Eq.(7.32) in oilfield units as

( )( ) ( )( ) ( )1.036 62.4 0.822 62.4

103.24 9.574144cP−⎡ ⎤⎣ ⎦= = psi

The water saturation at 8100 ft is calculated from the capillary pressure

curve for Layer 3 by linear interpolation as

Page 776: +Peters Ekwere j. - Petrophysics

7-41

( )9.574 9.4320.900 0.850 0.900 0.89611.215 9.432wS −⎛ ⎞= + − =⎜ ⎟−⎝ ⎠

The oil saturation is 1 1 0.896 0.104o wS S= − = − = . Table 7.6 shows the

calculated water and oil saturations from 8000 ft to 8203.24 ft (FWL).

Figure 7.21 shows the calculated saturation distributions. It can be

observed that none of the layers is at irreducible water saturation.

Therefore, each layer will produce some water if perforated.

Table 7.6. Calculated Saturations and Pressures for Example 7.1.

Depth z Pc Sw So Pw Po

ft ft psi

8000 203.24 18.847 0.426 0.574 3688.95 3670.10

8005 198.24 18.383 0.430 0.570 3689.15 3670.77

8010 193.24 17.920 0.434 0.566 3689.36 3671.44

8015 188.24 17.456 0.438 0.562 3689.57 3672.12

8020 183.24 16.992 0.443 0.557 3689.78 3672.79

Layer 1 8025 178.24 16.529 0.447 0.553 3689.99 3673.46

8030 173.24 16.065 0.451 0.549 3690.20 3674.13

8035 168.24 15.601 0.458 0.542 3690.40 3674.80

8040 163.24 15.138 0.464 0.536 3690.61 3675.47

8045 158.24 14.674 0.471 0.529 3690.82 3676.15

8050 153.24 14.210 0.477 0.523 3691.03 3676.82

8050 153.24 14.210 0.576 0.424 3691.03 3676.82

8055 148.24 13.747 0.584 0.416 3691.24 3677.49

Layer 2 8060 143.24 13.283 0.592 0.408 3691.44 3678.16

8065 138.24 12.819 0.600 0.400 3691.65 3678.83

8070 133.24 12.356 0.611 0.389 3691.86 3679.51

8070 133.24 12.356 0.823 0.177 3691.86 3679.51

8075 128.24 11.892 0.834 0.166 3692.07 3680.18

8080 123.24 11.428 0.845 0.155 3692.28 3680.85

8085 118.24 10.965 0.857 0.143 3692.49 3681.52

8090 113.24 10.501 0.870 0.130 3692.69 3682.19

Page 777: +Peters Ekwere j. - Petrophysics

7-42

Layer 3 8095 108.24 10.037 0.883 0.117 3692.90 3682.86

8100 103.24 9.574 0.896 0.104 3693.11 3683.54

8105 98.24 9.110 0.910 0.090 3693.32 3684.21

8110 93.24 8.646 0.926 0.074 3693.53 3684.88

8115 88.24 8.183 0.941 0.059 3693.73 3685.55

8120 83.24 7.719 0.957 0.043 3693.94 3686.22

8125 78.24 7.255 0.974 0.026 3694.15 3686.90

8125 78.24 7.255 0.593 0.407 3694.15 3686.90

8130 73.24 6.792 0.610 0.390 3694.36 3687.57

8135 68.24 6.328 0.629 0.371 3694.57 3688.24

8140 63.24 5.864 0.648 0.352 3694.78 3688.91

8145 58.24 5.401 0.672 0.328 3694.98 3689.58

8150 53.24 4.937 0.696 0.304 3695.19 3690.25

Layer 4 8155 48.24 4.473 0.725 0.275 3695.40 3690.93

8160 43.24 4.010 0.755 0.245 3695.61 3691.60

8165 38.24 3.546 0.791 0.209 3695.82 3692.27

8170 33.24 3.082 0.833 0.167 3696.02 3692.94

8175 28.24 2.619 0.880 0.120 3696.23 3693.61

8180 23.24 2.155 0.935 0.065 3696.44 3694.29

WOC 8185 18.24 1.691 1.000 0.000 3696.65 3694.96

Aquifer 8185 18.24 1.691 1.000 0.000 3696.65 3694.96

8190 13.24 1.228 1.000 0.000 3696.86 3695.63

8195 8.24 0.764 1.000 0.000 3697.06 3696.30

FWL 8203.24 0.00 0.000 1.000 0.000 3697.41 3697.41

Page 778: +Peters Ekwere j. - Petrophysics

7-43

Figure 7.21. Fluid saturation distribution for the reservoir of Example 7.1.

5. The water pressure is given by Eq.(7.29), which can be written in oilfield

units as

( ) ( )0144

ww w

zP z P ρ= −

Page 779: +Peters Ekwere j. - Petrophysics

7-44

where ( )0wP is the water pressure at the free water level. The water

pressure at the free water level is given by

( ) ( )( ) ( )( )( )1.036 62.4 8203.240 14.7 3697.41

144 144w

w atm

FWLP P

ρ= + = + = psia

At D = 8100 ft,

( ) ( )( ) ( )1.036 62.4103.24 3697.40 103.24 3651.06

144wP = − = psia

At the free water level, the pressure in the oil phase is equal to the

pressure in the water phase because the free water level is the reference

depth at which the capillary is zero. Thus,

( ) ( )0 0 3697.41o wP P= = psia

Of course, there is no oil in the reservoir below the water oil contact.

Therefore, there can be no oil pressure below the water oil contact. The

oil pressure starts at the water contact. However, if the oil pressure is

extrapolated to the free water level, its value will be equal to the water

pressure of 3697.41 psia. It should be noted that the difference between

the water oil pressure and the water pressure at the water oil contact is

equal to the displacement pressure of Layer 4 of 1.692 psi.

At D = 8100 ft, the oil pressure can be calculated as

( ) ( ) ( )103.24 103.24 103.24 3651.06 9.574 3641.87o w cP P P= − = − = psia

The calculated phase pressures are presented in Table 7.6 and Figure

7.22.

Page 780: +Peters Ekwere j. - Petrophysics

7-45

Figure 7.22. Water and oil phase pressures for Example 7.1.

6. The well is perforated in Layer 3, the lowest permeability layer, where

the oil saturation ranges from only 2.6% to 17.7%. This is in the

saturation range of residual oil for most reservoirs. Therefore, only

water will be produced from the well. The well should have been

perforated in Layer 1, where it would have produced a mixture of oil

and water.

7.7 CAPILLARY PRESSURE HYSTERESIS

Capillary pressure curves show a marked hysteresis depending on

whether the curve is determined under a drainage process or an imbibition

process. Figure 7.23 shows typical drainage and spontaneous imbibition

Page 781: +Peters Ekwere j. - Petrophysics

7-46

capillary pressure curves for the same porous medium. At any wetting phase

saturation, the drainage capillary pressure is higher than the imbibition

capillary pressure. At a capillary pressure of zero, the spontaneous imbibition

curve terminates at a wetting phase saturation that may or may not

correspond to the true residual non-wetting phase saturation depending on

the wettability of the rock. If the rock has a strong preference for the wetting

phase, then the wetting phase saturation at which the imbibition curve

terminates will be close to the true residual non-wetting phase saturation, Sor,

which is equal to (1-Swro). This is the case shown in Figure 7.23. If the rock

does not have a strong preference for the wetting phase, then the wetting

phase saturation at zero capillary pressure on the imbibition curve will not

correspond to the true residual non-wetting phase saturation. This means

that (1-Swro) will be larger than Sor. Additional oil can be displaced from the

rock, say be centrifuging the sample in water. This is the case shown in

Figure 7.24. The branch of the imbibition curve labeled 3 on the figure is the

forced imbition capillary pressure curve of the rock. Note that this branch

constitutes a negative capillary pressure. Note also that the true residual non-

wetting saturation (Sor) in this case can only by determined by forced

displacement not by spontaneous imbibition.

Figure 7.25 shows several cycles of capillary pressure

measurements on the same rock. The primary drainage curve labeled 1 was

performed first, followed by the spontaneous imbibition curve labeled 2. The

secondary drainage curve labeled 3 was performed after the spontaneous

imbibition measurement. It should be noted that the secondary drainage

curve will be less than the primary drainage curve at any given wetting phase

saturation. This is another aspect of capillary pressure hysteresis. If the

spontaneous imbibitition experiment is interupted and the measurement

reversed, then a different drainage curve will be followed as shown in curve 4.

If the drainage experiment is interupted and reversed, then a different

imbibition curve will be followed as shown in curve 5. Curves 4 and 5 form a

Page 782: +Peters Ekwere j. - Petrophysics

7-47

loop known as a scanning curve. Note that the area under the secondary

drainage curve was one of the areas used to define the USBM wettability

index.

Figure 7.23. Drainage and imbibition capillary pressure curves. (1) drainage

curve, (2) spontaneous imbibition curve (Killins et al., 1953).

Capillary pressure hysteresis can be explained in a variety of ways. In

Section 6.4.2, it was shown from energy considerations that more work is

required for a non-wetting phase to displace a wetting phase than for a

wetting phase to displace a non-wetting phase. This means that at any level of

saturation, more work is required during the drainage capillary pressure

measurement than during the imbibition measurement. Since work during

the capillary pressure measurement is PcΔV, where ΔV is the volume of fluid

Page 783: +Peters Ekwere j. - Petrophysics

7-48

displaced at that capillary pressure, the capillary pressure on the drainage

cycle will be greater than on the imbibition cycle to displace the same volume

of fluid.

Figure 7.24. Drainage and imbibition capillary pressure curves. (1) drainage curve, (2) spontaneous imbibition curve, (3) forced imbibition curve (Killins et

al., 1953).

Page 784: +Peters Ekwere j. - Petrophysics

7-49

Contact angle hysteresis plays a part in the capillary pressure

hysteresis. During drainage, the wetting phase recedes from the porous

medium and the contact angle is the receding contact angle, θR. During

imbibition, the wetting phase advances into the porous medium and the

contact angle is the advancing contact angle, θA. Since θR is less than θA,

2σcosθR /rm, the drainage capillary pressure, is larger than 2σcosθA /rm, the

imbibition capillary pressure at the same saturation state.

Figure 7.25. Cycles of capillary pressure measurements. (1) primary drainage, (2) spontaneous imbibition, (3) secondary drainage, (4-5) scanning curve.

Page 785: +Peters Ekwere j. - Petrophysics

7-50

The very nature of immiscible displacement plays a role in the capillary

pressure hysteresis. When the capillary pressure experiment is reversed to

measure the spontaneous imbibition curve, the pressure in the non-wetting

phase is reduced to allow the wetting phase to be imbibed. As the wetting

phase is imbibed into the rock, some non-wetting phase will be trapped in

certain pores. This trapping causes the wetting phase saturation on the

imbibition curve to be less than on the drainage curve at the same capillary

pressure.

The pore structure also plays a role in the capillary pressure hysteresis.

Consider the capillary pressure versus wetting phase saturation relationship

for the idealized pore shown in Figure 7.26 during drainage and imbibition.

During drainage, the pore is initially full of the wetting fluid at a capillary

pressure given by Laplace equation as shown in Figure 7.26a. Next, the

capillary pressure is increased to a higher value to drain some of the wetting

fluid as shown in Figure 7.26b. The higher capillary pressure versus the

wetting phase saturation is a point on the drainage capillary pressure curve.

Next, we consider the imbibition process as shown in Figures 7.26c and d. At

c, the capillary pressure is high at a wetting phase saturation of nearly zero.

After the wetting fluid has been imbibed to the equilibrium level shown in

Figure 7.26c, the imbibition capillary pressure will be approximately the same

as the drainage capillary pressure of Figure 7.26b because the mean

curvature of the interfaces at c and d are about the same. However, the

wetting phase saturation at d is considerably lower than at b. Thus, at the

same capillary pressure, the wetting phase saturation for imbibition is less

than for drainage. This is hysteresis.

Page 786: +Peters Ekwere j. - Petrophysics

7-51

Figure 7.26. Drainage and imbibition capillary pressures versus saturation for an idealized pore.

Another effect of pore structure is shown in Figure 7.27. The pore

(Figure 7.27a) is initially saturated with the wetting phase. The drainage

capillary pressure that must be applied to force the non-wetting phase into

the pore is given by

11

4 coscdrP

Deσ θ

= (7.35)

However, this capillary pressure is not sufficient to drain the entire pore

because neck De2 is smaller than De1. The interface will stop at De2. To

continue with the drainage, the applied capillary pressure must be increased

to Pcdr2 given by

Page 787: +Peters Ekwere j. - Petrophysics

7-52

22

4 coscdrP

Deσ θ

= (7.36)

Looking at the remaining necks of the pore, we see that Pcdr2 is large enough

to drain the remaining portion of the pore. Now, let us reduce the capillary

pressure to start the imbibition process. When the capillary pressure is

reduced to Pcibm1 given by

11

4 coscimbP

Dσ θ

= (7.37)

the first pore will be emptied of the non-wetting phase and the interface will

come to equilibrium at the location marked Imb1. To drain the non-wetting

phase further, the capillary pressure must be further reduced to Pcimb2 given

by

22

4 coscimbP

Dσ θ

= (7.38)

Looking at the remaining necks of the pore, we see that Pcimb2 is low enough to

empty the remaining portion of the pore of the non-wetting phase. Figure

7.27b shows the drainage and imbibition capillary pressure curves for the

experiment just described. We see capillary pressure hysteresis.

Figure 7.28 shows how the imbibition capillary pressure curve can be

used along with the drainage curve to determine the type of fluid that will be

produced at various depths in a reservoir. If the well is perforated above the

transition zone, only clean oil (water free oil) will be produced initially. If the

well is perforated in the upper part of the transition zone, both oil and water

will be produced from day one. If the well is perforated in the bottom part of

the transition zone, only water will be produced even though the zone has oil

saturation. The oil saturation in this zone is residual oil saturation.

Page 788: +Peters Ekwere j. - Petrophysics

7-53

Figure 7.27. Capillary pressure hysteresis in for an idealized pore (Dullien, 1992).

Capillary pressure hysteresis presents no problem in reservoir

engineering analysis as it is usually clear which curve should be used for a

particular analysis. The drainage curve should be used for estimating the

initial fluid saturation distribution in the reservoir whereas the imbibition

curve should be used for analyzing a waterflood performance in a water-wet

reservoir.

Page 789: +Peters Ekwere j. - Petrophysics

7-54

Figure 7.28. Drainage and imbibition capillary pressure curves showing the

depth of water free oil production (Archer and Wall, 1986).

7.8 CAPILLARY IMBIBITION

Consider a reservoir consisting of two layers with different

permeabilities and capillary pressure curves as shown in Figure 7.29 (a) and

(b). Initially, both layers are in capillary equilibrium at their respective

irreducible water saturations. Let this equilibrium be disturbed by

Page 790: +Peters Ekwere j. - Petrophysics

7-55

waterflooding the two layers. The injected water will advance further into the

more permeable layer (Figure 7.29 (c)). The oil and water pressures are

continuous across the boundary between the two layers. Thus, at the

boundary,

1 2o oP P= (7.39)

and

1 2w wP P= (7.40)

Subtracting Eq.(7.40) from (7.39) gives the condition for equilibrium as

1 2c cP P= (7.41)

Thus, at equilibrium, the capillary pressures in the two porous media will be

equal at their boundary.

In Figure 7.29c, sections A and D and C and F are in capillary

equilibrium, so no fluid exchanges will occur between these sections.

Sections B and E are not in capillary equilibrium, so fluid exchanges will

occur in an effort to achieve capillary equilibrium. Section E will loose water

to section B and gain oil from B while section B will gain water from E and

loose oil to E until a new capillary equilibrium is achieved. Thus, water will

be imbibed into the less permeable layer from the more permeable layer and

oil will be expelled from the less permeable layer into the more permeable

layer for subsequent displacement. This fluid exchange is beneficial to the oil

recovery process. However, the imbibition process is very slow. Therefore, the

water injection rate must be sufficiently slow for imbibition to assist in

waterflooding the low permeability layer.

Page 791: +Peters Ekwere j. - Petrophysics

7-56

Figure 7.29. Capillary imbibition; (a) reservoir before waterflooding; (b)

capillary pressure curves for the layers; (c) reservoir after waterflooding.

Naturally fractured reservoirs (fissured reservoirs) present another

example of capillary imbibition. The fractures have zero capillary pressure

whereas the matrix blocks have normal capillary pressure curves. When the

fractures become 100% saturated with water which comes in contact with the

oil saturated matrix blocks, the capillary equilibrium will be disturbed. Water

will be imbibed into the matrix blocks, expelling oil from the matrix blocks

into the fractures. Ultimately, the oil saturation in the matrix will be reduced

to the residual oil saturation over time.

Page 792: +Peters Ekwere j. - Petrophysics

7-57

7.9 CAPILLARY END EFFECT IN A LABORATORY CORE

7.9.1 Capillary End Effect

Another capillary phenomenon of interest is the capillary end effect

often experienced in laboratory coreflooding experiments. The end of the core

is in contact with the outside which could be viewed as a second medium with

zero capillary pressure. The condition for capillary equilibrium (Eq. 7.41)

requires that the capillary pressure inside the core at the outlet end be equal

to zero.

Consider a porous medium initially saturated with a non-wetting phase

(say oil) and irreducible wetting phase saturation (say water). The outlet end

of the core is at a higher capillary pressure than the outside. If the medium is

flooded with the wetting phase (waterflooded), initially, only the non-wetting

phase (oil) will be expelled from the outlet end at a higher capillary pressure

than the outside (Figure 7.30a). When the wetting phase arrives at the outlet

end, however, the system now has a chance to seek capillary equilibrium.

This equilibrium will be achieved by the accumulation of the wetting phase at

the outlet end of the core until the wetting phase saturation equals the

wetting phase saturation at zero capillary pressure on the imbibition capillary

pressure curve (Figure 7.30b). This saturation is marked Swro in Figure 7.30b.

Thus, the production of the wetting phase is delayed until well after the

arrival of the wetting phase at the outlet end of the core. This phenomenon is

known as capillary end effect.

This phenomenon has several undesirable consequences. The observed

breakthrough recovery of the non-wetting phase will be falsely high and the

wetting phase saturation distribution in the core will be opposite what would

normally be expected, with the wetting phase saturation being higher towards

the core outlet than in the rest of the core (Figure 7.30c). Most corefloods are

blind tests because one cannot see the fluid distribution inside the core.

Page 793: +Peters Ekwere j. - Petrophysics

7-58

Therefore, the breakthrough recovery is usually taken to be a good measure of

the displacement efficiency. In the presence of capillary end effect, the

breakthrough recovery will be too large and will give a false sense of the

displacement efficiency. Also, in the unsteady state method for relative

permeability measurement described in Chapter 8, the fractional flow of the

wetting phase versus the wetting phase saturation at the outlet end of the

core is used to calculate relative permeabilities on the assumption that there

is no capillary end effect. Therefore, if there is capillary end effect in the

experiment, the calculated relative permeabilities will be wrong.

Figure 7.30. Capillary end effect; (a) coreflood; (b) spontaneous imbibition capillary pressure curve; (c) wetting phase saturation profiles; (d) relative

permeabililty curves.

Page 794: +Peters Ekwere j. - Petrophysics

7-59

7.9.2 Mathematical Analysis of Capillary End Effect

We can derive the mathematical model for the immiscible displacement

shown in Figure 7.30a and use it to explain the capillary end effect

phenomenon. Darcy's law for the wetting and non-wetting phases is given by

w ww

w

k A Pqxμ

∂= −

∂ (7.42)

and

nw nwnw

nw

k A Pqxμ

∂= −

∂ (7.43)

where wk and nwk are the effective permeabilities to the wetting and non-

wetting phases. Let us define the relative permeabilities of the wetting and

non-wetting phases as

wrw

kkk

= (7.44)

and

nwrnw

kkk

= (7.45)

Eqs.(7.42) and (7.43) can be written in terms of the relative permeabilities as

rw ww

w

kk A Pqxμ

∂= −

∂ (7.46)

and

Page 795: +Peters Ekwere j. - Petrophysics

7-60

rnw nwnw

nw

kk A Pqxμ

∂= −

∂ (7.47)

Capillary equilibrium gives

( )nw w c wP P P S− = (7.48)

Assuming incompressible fluids, then

w nwq q q= + (7.49)

The continuity equation for the wetting phase is

0w wS qAt x

φ ∂ ∂+ =

∂ ∂ (7.50)

Finally, the saturation constraint gives

1w nwS S+ = (7.51)

Eqs.(7.46) through (7.51) constitute the complete mathematical description of

two-phase immiscible displacement in the absence of the effect of gravity.

Subtracting Eq.(7.46) from (7.47) and rearranging gives

w w nw nw nw w

rw rnw

q q P Pkk A kk A x x

μ μ ∂ ∂− = −

∂ ∂ (7.52)

Substituting Eqs.(7.48) and (7.49) into (7.52) gives upon rearrangement

Page 796: +Peters Ekwere j. - Petrophysics

7-61

1

1

rnw c

w nw

rnw w

rw nw

kk A Pq q x

kqk

μμ

μ

∂+

∂=

+ (7.53)

Let the true fractional flow of the wetting phase be defined as

ww

qfq

= (7.54)

Let an approximate fractional flow of the wetting phase be defined as

1

1w

rnw w

rw nw

F kk

μμ

=+

(7.55)

The approximate fractional flow of the wetting phase also can be defined as a

function of the mobility ratio as

111

wF

M

=+

(7.56)

Both wf and wF are functions of saturation. Substituting Eqs.(7.54) and (7.55)

into (5.53) gives the true fractional flow of the wetting phase as

1 rnw cw w

nw

kk A Pf Fq xμ

⎛ ⎞∂= +⎜ ⎟∂⎝ ⎠

(7.57)

Let the dimensionless distance from the inlet end be defined as

DxxL

= (7.58)

Page 797: +Peters Ekwere j. - Petrophysics

7-62

Let the spontaneous imbibition capillary pressure curve be given in terms of

its Leverett J-function as

( ) ( )cos ,/c w wP S J S

kσ θ

φ= Γ (7.59)

Substituting Eqs.(7.58) and (7.59) into (7.57) gives the true fractional flow of

the wetting phase as

cos1w w rnwnw D

A k Jf F kq L x

σ θ φμ

⎡ ⎤⎛ ⎞ ∂= +⎢ ⎥⎜ ⎟⎜ ⎟ ∂⎢ ⎥⎝ ⎠⎣ ⎦

(7.60)

The term in the inner bracket on the right side of Eq.(7.60) is a dimensionless

number that gives the ratio of the capillary to viscous forces in the

displacement. Let this dimensionless number be defined as

cos

capnw

A kNq L

σ θ φμ

= (7.61)

Substituting Eq.(7.61) into (7.60) gives

1w w cap rnwD

Jf F N kx

⎡ ⎤∂= +⎢ ⎥∂⎣ ⎦

(7.62)

Let the dimensionless time be defined as

Dqtt

A Lφ= (7.63)

Substituting Eq.(7.63) into (7.50) gives the continuity equation for the wetting

phase as

Page 798: +Peters Ekwere j. - Petrophysics

7-63

0w w

D D

S ft x

∂ ∂+ =

∂ ∂ (7.64)

Substituting Eq.(7.62) into (7.64) gives

0w w wcap w rnw

D w D D D

S dF S JN F kt dS x x x

⎛ ⎞ ⎛ ⎞∂ ∂ ∂ ∂+ + =⎜ ⎟ ⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠⎝ ⎠

(7.65)

Because J is a function of wS , Eq.(6.65) can be written as

0w w w wcap w rnw

D w D D w D

S dF S SdJN F kt dS x x dS x

⎛ ⎞ ⎛ ⎞∂ ∂ ∂∂+ + =⎜ ⎟ ⎜ ⎟∂ ∂ ∂ ∂⎝ ⎠ ⎝ ⎠

(7.66)

Eq.(7.66) is the final form of the partial differential equation for the

wetting phase saturation for two phase immiscible displacement in a linear

porous medium. When supplemented with appropriate initial and boundary

conditions together with the rock and fluid properties, its solution can be

used to calculate the performance of the immiscible displacement such as a

waterflood or a gas flood. The solution of this equation is deferred to Chapter

8. Our concern here is the capillary end effect, not the prediction of the overall

displacement performance.

Let us examine in detail the fractional flow of the wetting phase at the

outlet end of the core. Applied to the outlet end of the core, Eq.(7.62) can be

written as

1w w cap rnwD

J Jf F N kxδ

− +⎡ ⎤⎛ ⎞−= +⎢ ⎥⎜ ⎟

⎝ ⎠⎣ ⎦ (7.67)

where the derivative of the J-function with respect to Dx is given by the

following finite difference approximation:

Page 799: +Peters Ekwere j. - Petrophysics

7-64

D D

J J Jx xδ

− +⎛ ⎞∂ −⎜ ⎟∂ ⎝ ⎠

(7.68)

In Eq.(7.68), J + is the J-function inside the porous medium, J − is the J-

function outside the porous medium and Dxδ is a small distance in the

neighborhood of the outlet end of the porous medium as shown in Figure

7.31. Of course, J − is equal to zero. Therefore, the fractional flow of the

wetting phase at the outlet end of the porous medium becomes

1w w cap rnwD

Jf F N kxδ

+⎡ ⎤= −⎢ ⎥

⎣ ⎦ (7.69)

Depending on the values of capN , rnwk , and J + , it is possible for the following

inequality to prevail during the displacement:

1 0cap rnwD

JN kxδ

+

− ≤ (7.70)

Figure 7.31. Initial capillary barrier at the outlet end of the core.

Page 800: +Peters Ekwere j. - Petrophysics

7-65

If

1 0cap rnwD

JN kxδ

+

− = (7.71)

then

0wf = (7.72)

at the outlet end of the core. Because the fractional flow of the wetting phase

is zero at the outlet end of the core, the wetting phase cannot flow out of the

core but instead will accumulate there raising the wetting phase saturation to

an abnormal level. This is the capillary end effect phenomenon at work. If

1 0cap rnwD

JN kxδ

+

− < (7.73)

the wetting phase cannot flow out of the core either but will accumulate at the

end as before, raising the wetting saturation there to an abnormal level.

Although the inequality given by Eq.(7.73) appears to indicate that the wetting

phase will flow backwards into the core, this will not happen because there is

no supply of the wetting phase outside the outlet end of the core for it to be

imbibed into the core. Therefore, Eq.(7.70) gives the condition for capillary end

effect to occur in an immiscible displacement.

Let us now examine the physics of the displacement before and after

wetting phase breakthrough. Before the arrival of the wetting phase at the

outlet end of the core, only the non-wetting phase will be produced. The

wetting phase saturation profiles will be as shown in Figure 7.30a at t1 and t2.

At time at , the wetting phase arrives at the outlet end of the core. At this time,

J + has its maximum possible value at the irreducible wetting phase

Page 801: +Peters Ekwere j. - Petrophysics

7-66

saturation. If the values of capN , rnwk and J + are such that Eq.(7.70) holds,

then capillary end effect will occur and the wetting phase begins to

accumulate at the outlet end of the core resulting in the increase of the

wetting saturation, ( ),wS L t . As the wetting phase saturation increases, J +

decreases dramatically as shown in Figure 7.30b. When the wetting phase

saturation has increased to wroS , J + becomes zero and capillary equilibrium is

achieved between the porous medium and the outside. At this time, Eq.(7.69)

gives

0w wf F= > (7.74)

and the wetting phase flows out of the core and is produced along with the

non-wetting phase. The wetting phase saturation profile at breakthrough is

shown in Figure 7.30c at btt . After breakthrough, both phases will be

produced until residual non-wetting phase is achieved in the core after many

pore volumes of wetting phase injection. The wetting phase saturation profile

will be as shown in Figure 7.30c at t∞ . Beyond this time, only the wetting

phase will be produced.

How can capillary end effect be eliminated from the experiment? The

condition for eliminating the capillary end effect is obtained from Eq.(7.69) as

1 0cap rnwD

JN kxδ

+

− > (7.75)

or

Dcap

rnw

xNk Jδ

+< (7.76)

Page 802: +Peters Ekwere j. - Petrophysics

7-67

Thus, capN should be as small as possible in the experiment to eliminate

capillary end effect. A critical value of capN can be defined as

Dcapcritical

rnw

xNk Jδ

+= (7.77)

For capN below the critical value, capillary end effect will be eliminated. Above

the critical value, capillary end effect will occur in the displacement.

How can the value of capN be controlled in the experiment? The only

means to control capN in the experiment is through the injection rate, q.

Examination of Eq.(7.61) shows that capN can be made small by the use of a

high injection rate in the experiment. Substituting Eq.(7.61) into (7.76) gives

the condition for the injection rate to eliminate capillary end effect as

cosrnw

nw D

Ak J kqL x

σ θ φμ δ

+

> (7.78)

A critical injection rate can be defined as

cosrnw

criticalnw D

Ak J kqL x

σ θ φμ δ

+

= (7.79)

For injection rates below the critical value, capillary end effect will occur. For

rates above the critical, capillary end effect will be eliminated. In terms of

Darcy velocity, Eqs.(7.78) and (7.79) become

cosrnw

nw D

k J kqvA L x

σ θ φμ δ

+

= > (7.80)

Page 803: +Peters Ekwere j. - Petrophysics

7-68

cosrnwcritical

nw D

k J kvL x

σ θ φμ δ

+

= (7.81)

Is capillary end effect a problem in field displacements? The answer is no

because the large value of L at field scale ensures that capN is always smaller

than is required to eliminate capillary end effect at field rates.

7.9.3 Mathematical Model of Capillary End Effect During Steady State Relative Permeability Measurement.

The steady state method for relative permeability measurements

involves the simultaneous injection into the core of the wetting and non-

wetting fluids as shown in Figure 7.32. The fluids are injected at various

ratios of qw/qnw until steady state is achieved. If the total injection rate, q,

satisfies the condition of Eq.(7.76), then the injection rate is high enough to

eliminate capillary end effect and the wetting phase saturation will be uniform

throughout the core. However, if the total rate is not high enough to eliminate

capillary end effect, the wetting phase saturation will be non uniform in the

core. The wetting phase saturation will be higher at the outlet end of the core

than at the inlet end. The steady state saturation profile can be derived from

Eq.(7.62) as follows. At steady state conditions, Eq.(7.62) can be written as

1 ww w cap rnw

w D

dSdJf F N kdS dx

⎡ ⎤= +⎢ ⎥

⎣ ⎦ (7.82)

Rearranging Eq.(7.82) gives

1w

ww

Dcap rnw

w

fFdS

dJdx N kdS

⎛ ⎞−⎜ ⎟

⎝ ⎠= (7.83)

Page 804: +Peters Ekwere j. - Petrophysics

7-69

Figure 7.32. Steady state experiment.

Eq.(7.83) is the partial differential equation for the steady state wetting phase

saturation. It is a first order nonlinear equation that can easily be integrated

to obtain the steady state saturation profile. The appropriate boundary

condition for the equation is

( )1w wroS S= (7.84)

Figure 7.33 shows the steady state saturation profile obtained by solving

Eqs.(7.83) with the boundary condition given by Eq.(7.84).

Page 805: +Peters Ekwere j. - Petrophysics

7-70

Figure 7.33. Steady state wetting phase saturation profile in the presence of capillary end effect.

7.9.4 Experimental Evidence of Capillary End Effect

Perkins (1957) has presented experimental data that show capillary end

effect at work. He conducted waterfloods in laboratory cores at two rates, one

below the critical rate for capillary end effect and one above the critical rate.

The core was 12 inches in length and 1.25 inches in diameter. The oil and

water viscosities were 1.8 and 0.9 cp. The low injection rate was 2.4 ft/day

whereas the high injection rate was 36 ft/day. The injected water was 0.1

normal sodium chloride solution. The core was instrumented with two current

electrodes and nineteen potential electrodes distributed along its length.

These electrodes enabled the resistivity of the core to be measured along the

length of the core, from which the water saturation profiles were calculated

using Archie's equation.

Page 806: +Peters Ekwere j. - Petrophysics

7-71

Figure 7.34 shows the water saturation profiles for the low rate flood.

The water arrived at the outlet end of the core at tD = 0.41 pore volume

injected. However, no water was produced from the core until tD = 0.60 pore

volume injected. Thus, the flood was affected by capillary end effect. The water

saturation profile at tD = 0.60 clearly shows the capillary end effect.

Figure 7.35 shows the water saturation profiles for the high rate flood.

The water arrived at the outlet end of the core at tD = 0.60 pore volume

injected and got produced shortly thereafter at tD = 0.65 pore volume injected.

Thus, the capillary end effect was significantly reduced in the high rate

displacement compared to the low rate displacement.

Figure 7.34. Wetting phase saturation profiles at low injection rate (Perkins,

1957).

Page 807: +Peters Ekwere j. - Petrophysics

7-72

Figure 7.35. Wetting phase saturation profiles at high injection rate (Perkins,

1957).

The second example from Richardson et al. (1952) investigated capillary

end effect in steady state experiments. The core was a Berea sandstone of

length 30 cm and diameter 6.85 cm. The core was cut into 8 segments as

shown in Figure 7.36. Steps were taken to ensure capillary continuity

between the segments. The core was initially saturated with oil, which was the

wetting phase. Then helium, the non-wetting phase, and oil were injected

simultaneously at various ratios. After steady state was achieved, the core

segments were weighed to determine the wetting phase saturation in each

segment.

Page 808: +Peters Ekwere j. - Petrophysics

7-73

Figure 7.36. Segmented core used in steady state experiments (Richardson et

al., 1953).

Figures 7.37 to 7.39 show the measured and calculated wetting phase

saturation profiles at steady state for successively higher total injection rates.

The theoretical saturation profiles were calculated with Eqs.(7.83) and (7.84)

using the drainage relative permeability curves and drainage capillary

pressure curves shown in Figure 7.40. Capillary end effect occurred in the

three experiments shown in Figures 7.37 to 7.39. However, it can be seen that

the capillary end effect was reduced as the total injection rate was increased.

The lesson from these examples is that capillary end effect could

dominate laboratory scale immiscible displacements. Since such

displacements are normally used to determine relative permeability curves in

Page 809: +Peters Ekwere j. - Petrophysics

7-74

the laboratory, it is imperative that steps be taken to eliminate or at least

minimize capillary end effect in these experiments. Failure to do so will result

in the relative permeability curves so derived being wrong.

Figure 7.36. Capillary end effect in a steady state experiment. qw = 0.15

cm3/s, qnw = 0.000336 cm3/s (Richardson et al., 1953).

Page 810: +Peters Ekwere j. - Petrophysics

7-75

Figure 7.37. Capillary end effect in a steady state experiment. qw = 0.264

cm3/s, qnw = 0.0022 cm3/s (Richardson et al., 1953).

Figure 7.38. Capillary end effect in a steady state experiment. qw = 0.80

cm3/s, qnw = 0.00288 cm3/s (Richardson et al., 1953).

Page 811: +Peters Ekwere j. - Petrophysics

7-76

Figure 7.39. Drainage relative permeability curves and capillary pressure

curve for Berea sandstone (Richardson et al., 1953).

7.10 CAPILLARY PRESSURE MEASUREMENTS

Three methods are commonly used to measure capillary pressure

curves in the petroleum industry: the restored state method (porous plate

method), mercury injection method and centrifuge method.

7.10.1 Restored State Method (Porous Plate Method).

In this method, capillary pressure is measured by placing the sample,

initially saturated with a wetting fluid, in a vessel filled with the non-wetting

fluid. The bottom of the vessel consists of a semi-permeable plate, which

allows the wetting phase displaced from the sample to pass through while

blocking the passage of the non-wetting phase. Extending from the porous

Page 812: +Peters Ekwere j. - Petrophysics

7-77

plate is a graduated tube which allows the volume of the wetting phase

displaced to be measured as shown in Figure 7.41. With the sample in place,

the pressure of the non-wetting fluid is increased in steps and the system is

allowed to achieve equilibrium after each pressure change. The volume of

wetting phase displaced at each pressure is measured. The capillary pressure

is the non-wetting phase pressure minus the wetting phase pressure at each

step. The wetting phase saturation of the sample is determined from the

volume of wetting phase displaced at each pressure to obtain the capillary

pressure versus saturation relationship.

The porous plate is typically made of porcelain or fritted glass. It must

have a displacement pressure that is higher than the largest capillary

pressure to be measured. This limits the maximum capillary pressure that

can be measured with the method to about 200 psi.

The porous plate apparatus can be used to measure the imbibition

capillary pressure curve as well as the drainage curve. The method gives a

reliable estimate of the irreducible wetting phase saturation. The major

disadvantage of the porous plate method is that it takes too long to obtain the

entire capillary pressure curve. It is not unusual for the capillary pressure

experiment to take several weeks to complete.

7.10.2 Mercury Injection Method

In this method, capillary pressure is measured by injecting mercury,

which is a non-wetting phase, into the sample. The apparatus used in the

measurement is shown in Figure 7.42. It consists of a sample cell and a

mercury injection pump. A dry sample is placed in the cell and the cell is

evacuated. Mercury is injected into the cell until the mercury is level with a

graduation on the high-pressure glass capillary above the sample chamber.

Nitrogen pressure is then applied in successive increments and at each step,

Page 813: +Peters Ekwere j. - Petrophysics

7-78

Figure 7.41. Porous plate capillary pressure apparatus (Welge and Bruce,

1947).

mercury is injected to maintain the mercury level with the graduation on the

capillary. From the volume of the cell and the volume of mercury required to

fill the cell with the sample before mercury injection into the sample, the bulk

volume of the sample can be determined. The mercury-air capillary pressure

versus saturation relationship is calculated from the volume of mercury

forced into the sample pore space as a function of the applied nitrogen

pressure.

Page 814: +Peters Ekwere j. - Petrophysics

7-79

Figure 7.38. Capillary pressure cell for mercury injection (Purcell, 1949).

The mercury injection method is very fast. The capillary pressure curve

can be obtained in a matter of hours. The imbibition curve can be obtained

very easily by decreasing the nitrogen pressure and withdrawing mercury

from the system. Figure 7.43 shows typical capillary pressure curves

obtained by mercury injection, mercury withdrawal and mercury re-injection.

Page 815: +Peters Ekwere j. - Petrophysics

7-80

Figure 7.43. Mercury-air capillary pressure curves.

Brown (1951) has shown that the mercury injection method can give

essentially the same capillary pressure curve as the restored state method

except for a scaling factor. Figures 7.44 and 7.45 compare capillary pressure

curves obtained by mercury injection and the restored state method for a

sandstone and a limestone core, respectively. The results show good

agreement between the two methods. The scaling factors for the sandstone

and limestone were 7.5 and 5.5, respectively. These are different from the

Page 816: +Peters Ekwere j. - Petrophysics

7-81

scaling factor of 5.2 suggested by Purcell (1949) based on the ratio of σcosθ of

mercury-air and water-air systems.

The major disadvantage of the mercury injection method is that the core

can no longer be used for other tests after mercury injection. The method

also cannot be used to determine the irreducible wetting phase saturation.

7.10.3 Centrifuge Method.

In this method, the sample saturated with a wetting fluid is placed in a

centrifuge cup containing the non-wetting fluid as shown in Figures 7.46 and

7.47. The sample is rotated at a series of constant angular velocities and the

amount of wetting fluid displaced at equilibrium at each velocity is measured

with the aid of a stroboscopic light. The only data measured directly in this

method are the volume of wetting fluid displaced and the corresponding

rotational speed of the centrifuge. These data can be used to derive the

capillary pressure versus saturation relationship of the porous medium.

The theory underlying the method is that the centrifuge imposes a

centrifugal force (typically, over 1000 times the force of gravity) on the sample.

This causes the denser wetting fluid to be displaced outward away from the

center of rotation and the non-wetting fluid to flow into the sample through

the inlet face of the sample. Consider an oil-water system in which water is

the wetting phase and oil is the non-wetting phase. At equilibrium, the

pressure gradients in the water and oil are given by

2ww

dP rdr

ρ ω= − (7.85)

2oo

dP rdr

ρ ω= − (7.86)

Page 817: +Peters Ekwere j. - Petrophysics

7-82

Figure 7.44. A comparison of water-nitrogen and mercury-air capillary pressure curves for a sandstone core (Brown, 1951). Note the different

capillary pressure scales for the two sets of data.

Page 818: +Peters Ekwere j. - Petrophysics

7-83

Figure 7.45. A comparison of water-nitrogen and mercury-air capillary pressure curves for a limestone core (Brown, 1951). Note the different

capillary pressure scales for the two sets of data.

Page 819: +Peters Ekwere j. - Petrophysics

7-84

Figure 7.46. Positions of core and graduated tube in a centrifuge for

measurement of oil-displacing-water capillary pressure curve (Donaldson et al., 1980).

Page 820: +Peters Ekwere j. - Petrophysics

7-85

Figure 7.47. Positions of core and graduated tube in a centrifuge for

measurement of water-displacing-oil capillary pressure curve (Donaldson et al., 1980).

where ω is the angular velocity of the centrifuge in radians/s. Subtracting

Eqs. 7.85 from 7.86 gives

( ) 2 2cw o

dP r rdr

ρ ρ ω ρω= − − = −Δ (7.87)

Page 821: +Peters Ekwere j. - Petrophysics

7-86

Integration of Eq. 7.87 gives the capillary pressure at radius r as

2

2crP CρωΔ

= − + (7.88)

where C is an integration constant. Application of the Hassler-Brunner

boundary condition that Pc = 0 at r = r2 gives the capillary pressure at any r

as

( )2

2 222cP r rρωΔ

= − (7.89)

where r2 is the outlet face of the core measured from the center of rotation. At

any stage of the centrifuge experiment, the highest capillary pressure occurs

at r1, the inlet face of the core measured from the center of rotation. At the

inlet face, the capillary pressure is given by

( )2

2 21 2 12cP r rρωΔ

= − (7.90)

At any stage of the centrifuge experiment, the water saturation in the

core varies from a minimum at the core inlet, r1, to a value of 1.0 at the core

outlet, r2. Since the capillary pressure at the core inlet can be calculated

from Eq.(7.90), it is only necessary to estimate the water saturation at the

core inlet in order to obtain the required capillary pressure versus saturation

relationship. The water saturation at the core inlet can be estimated from the

average water saturation determined from the amount of water displaced at

each speed. The average water saturation in the core is given by

2

1

2 1

r

wrwav

S drS

r r=

−∫

(7.91)

Page 822: +Peters Ekwere j. - Petrophysics

7-87

Multiplying Eq.(7.91) by Δρω2r1 and rearranging gives

2

1

2 21 1

r

wav wrLr S S rdrρω ρωΔ = Δ∫ (7.92)

where L is the length of the core. Noting that Δρω2Lr1 is approximately equal

to Pc1 since L is small relative to r1 and r2 and Δρω2r1dr is equal to 1cdP− ,

Eq.(7.92) can be written as

( )1

1 0

cP

c wav w c cP S S P dP= ∫ (7.93)

Differentiating Eq.(7.93) gives the water saturation at the core inlet as

( )11

1

c wavw

c

d P SS

dP= (7.94)

or

1 11

wavw wav c

c

dSS S PdP

⎛ ⎞= + ⎜ ⎟

⎝ ⎠ (7.95)

The approximate water saturation equation, Eq.(7.94) or (7.95), is applicable

for r1/r2 equal to or greater than 0.8. When applied to experimental data,

experience shows that Eq.(7.94) gives a smoother capillary pressure curve

than Eq.(7.95).

The centrifuge method is fast and allows the capillary pressure

measurement to be completed in a day or less. The method is good for

determining the irreducible water saturation. Figure 7.48 shows a

comparison of the capillary pressure curves of the same sample from

centrifuge measurement and by the restored state method obtained by

Page 823: +Peters Ekwere j. - Petrophysics

7-88

Hassler and Brunner (1945). Although the capillary pressure curve from

centrifuge measurement is somewhat higher than from the restored state

method, the agreement between the two sets of data is reasonable. In this

figure, the open circles are data obtained by the use of Eq.(7.94) and the open

squares are data obtained by the use of Eq.(7.95) to compute the inlet water

saturation. It should be observed that both equations give somewhat different

results for the water saturations. Figure 7.49 shows comparisons of the

capillary pressure curves for the same cores measured by the restored state

method (diaphram), mercury injection and centrifuge experiments. The

agreement between the three methods is excellent.

The disadvantages of the method include (1) inability to measure the

displacement pressure since the water saturation at the core inlet is always

less than the average water saturation of the core, (2) the Hassler-Brunner

boundary condition at the core outlet may be violated at high centrifuge

speeds, (3) the calculated water saturation at the core inlet is an

approximation, and (4) inability to obtain spontaneous imbibition capillary

pressure curve. Melrose (1988) has investigated the Hassler-Brunner

boundary condition and concluded that it was unlikely to be violated under

normal core analysis conditions. Rajan (1986) has presented a more accurate

method for calculating the water saturation at the core inlet.

Page 824: +Peters Ekwere j. - Petrophysics

7-89

Figure 7.48. A comparison of capillary pressure obtained by centrifuge and by

the restored state method (Hassler and Brunner, 1945).

Page 825: +Peters Ekwere j. - Petrophysics

7-90

Figure 7.49. A comparison of capillary pressure curves obtained by centrifuge, mercury injection and by the restored state method (Hermansen et al., 1991).

Example 7.2

Table 7.7 shows the data obtained in a centrifuge experiment for determining

the air-water capillary pressure curve of a core sample. Other data from the

experiment are as follows:

Core length = 2.0 cm

Core diameter = 2.53 cm

Core pore volume = 1.73 cm3

Core permeability = 144 md

Page 826: +Peters Ekwere j. - Petrophysics

7-91

Centrifuge arm (r2) = 8.6 cm

Water-air density difference = 0.9988 gm/cm3

2 8.6r = cm

Table 7.7. Centrifuge Experimental Data Centrifuge Speed (RPM)

Volume of Water Displaced

(cc) 1300 0.30 1410 0.40 1550 0.50 1700 0.60 1840 0.70 2010 0.75 2200 0.80 2500 0.90 2740 1.00 3120 1.05 3810 1.10 4510 1.20 5690 1.25

1. Calculate the capillary pressure curves for the sample using Eq.(7.94)

(method 1) and Eq. (7.95) (method 2) for the inlet water saturation.

2. Compare the capillary pressure curves obtained from the two methods.

3. Calculate the acceleration imposed on the inlet of the core at the

centrifuge speed of 5690 RPM and compare to the acceleration due to

gravity.

Page 827: +Peters Ekwere j. - Petrophysics

7-92

Solution to Example 7.2

The results of the calculations are summarized in Table 7.8.

1. The entries in the table were computed as follows at the centrifuge

speed of 1300 revolutions per minute (RPM).

( ) ( )/ 1.73 0.30 /1.73 0.827wav p w pS V V V= − = − =

( )2 / 60 2 1300 / 60 136.14Nω π π= = = radians/s

( ) ( )( )( )2 2 2 2 2 22 1

1

0.9988 136.14 8.6 6.6281362.40

2 2c

r rP

ρωΔ − −= = = dynes/cm2

61 281362.40 /1.0133 10 0.28cP x= = atm

( )( )1 6

281362.40 14.6964.08

1.0133 10cPx

= = psi

Figure 7.50 shows the graph of 1wav cS P versus 1cP . The regression

equation is

0.62171 11.3563wav c cS P P=

Application of Eq.(7.94) gives the inlet water saturation as

( ) ( )( ) ( )( ) 0.37831

1 61

281362.40 14.6960.6217 1.3563 0.495

1.0133 10wav c

wc

d S PS

dP x

−⎡ ⎤

= = =⎢ ⎥⎣ ⎦

Page 828: +Peters Ekwere j. - Petrophysics

7-93

Table 7.8. Calculated Results for Example 7.2.

1 2 3 4 5 6 7 8 9 10 11

Centrifug

e

Method

1(Eq.7.94)

Method 2

(Eq.7.95)

Speed Vw Swav ω Pc1 Pc1 Pc1 SwavPc1 Sw1 1

wav

c

dSdP

Sw1

RPM cc radians/

s

dynes/cm2 atm psi psi psi-1 psi

1300 0.30 0.827 136.14 281362.40 0.28 4.08 3.37 0.495 -0.0739 0.525

1410 0.40 0.769 147.65 330992.07 0.33 4.80 3.69 0.466 -0.0590 0.485

1550 0.50 0.711 162.32 399984.12 0.39 5.80 4.12 0.434 -0.0455 0.447

1700 0.60 0.653 178.02 481146.36 0.47 6.98 4.56 0.404 -0.0353 0.407

1840 0.70 0.595 192.68 563657.13 0.56 8.17 4.87 0.381 -0.0283 0.364

2010 0.75 0.566 210.49 672622.63 0.66 9.76 5.53 0.356 -0.0222 0.350

2200 0.80 0.538 230.38 805795.28 0.80 11.69 6.28 0.333 -0.0173 0.335

2500 0.90 0.480 261.80 1040541.4

3

1.03 15.09 7.24 0.302 -0.0122 0.296

2740 1.00 0.422 286.93 1249915.0

1

1.23 18.13 7.65 0.282 -0.0095 0.251

3120 1.05 0.393 326.73 1620647.4

3

1.60 23.50 9.24 0.255 -0.0066 0.238

3810 1.10 0.364 398.98 2416736.5

4

2.39 35.05 12.76 0.220 -0.0038 0.231

4510 1.20 0.306 472.29 3386354.6

6

3.34 49.11 15.05 0.193 -0.0024 0.189

5690 1.25 0.277 595.86 5390187.7

2

5.32 78.17 21.69 0.162 -0.0013 0.179

Figure 7.51 shows the graph of wavS versus 1cP . The regression equation

is

0.378311.3563wav cS P−=

Page 829: +Peters Ekwere j. - Petrophysics

7-94

( )( ) ( )( ) 1.3783

61

281362.40 14.6960.3783 1.3563 0.0739

1.0133 10wav

c

dSdP x

−⎡ ⎤

= − = −⎢ ⎥⎣ ⎦

Application of Eq.(7.95) gives the inlet water saturation as

( ) ( )( ) ( )1 1 61

1.73 0.30 281362.40 14.6960.0739 0.525

1.73 1.0133 10wav

w wav cc

dSS S PdP x

− ⎡ ⎤⎛ ⎞= + = − =⎜ ⎟ ⎢ ⎥

⎝ ⎠ ⎣ ⎦

Figure 7.50. Graph of 1wav cS P versus 1cP for Example 7.2.

Page 830: +Peters Ekwere j. - Petrophysics

7-95

Figure 7.51. Graph of wavS versus 1cP for Example 7.2.

2. Figure 7.52 shows a comparison of the capillary pressure curves from

Eqs.(7.94) and (7.95). Both equations give essentially the same capillary

pressure curves. However, Eq.(7.95) shows more scatter in the

computed water saturations than Eq.(7.94).

3. The acceleration imposed at the inlet end of the core at the centrifuge

speed of 5690 RPM is given by

( ) ( )22

21 1

2 56902 6.6 2343288.1960 60

Nacceleration r rππω

⎡ ⎤⎛ ⎞= = = =⎢ ⎥⎜ ⎟⎝ ⎠ ⎣ ⎦

cm/s2

This acceleration is 2388 times the acceleration due to gravity.

Page 831: +Peters Ekwere j. - Petrophysics

7-96

Figure 7.52. Comparison of the capillary pressure curves derived from Eqs.7.94 and 7.95 for Example 7.1.

7.11 PORE SIZE DISTRIBUTION

7.11.1 Introduction

We saw in Section 7.2 that the capillary pressure versus saturation

relationship of an idealized porous medium captured the pore size

distribution of the medium. It should come as no surprise that one of the

principal applications of capillary pressure curve is for estimating the pore

size distribution of porous media. Because mercury does not wet most solids,

the capillary pressure curve derived from mercury injection is particularly well

suited for probing the pore structure of porous media. Thus, mercury

Page 832: +Peters Ekwere j. - Petrophysics

7-97

injection porosimetry is widely used in the petroleum and material science

industries to determine the pore size distribution of porous materials.

When mercury is injected into the porous medium at a low pressure,

mercury will invade those pores having pore throat radii equal to or greater

than the radius given by Laplace equation:

2 cos

cPR

σ θ= (7.96)

where R is the pore throat radius. As the capillary pressure is increased,

pores with smaller pore throat sizes are invaded by mercury. If the mercury

pressure is high enough, all the pores in the porous medium will be invaded

by mercury. Therefore, the cumulative volume of mercury injected versus the

capillary pressure can be used to determine the pore size distribution of the

medium.

Two different pore size distribution functions can be derived from

mercury porosimetry. The first is the pore volume distribution, which does not

assume a model for the porous medium such as the bundle of capillary tubes

model. The only assumption is that the pore throat is circular. The second

pore size distribution function is based on the assumption of a bundle of

capillary tubes model of the porous medium and represents the distribution of

the pore throat size assuming the pores are capillary tubes. Because they

represent different things, the two pore size distribution functions are not the

same.

7.11.2 Pore Volume Distribution

Consider the result of the mercury injection experiment as shown in

Figure 7.53. The figure shows the cumulative volume of mercury injected

expressed as a fraction of the total pore volume plotted against the capillary

Page 833: +Peters Ekwere j. - Petrophysics

7-98

pressure. This is the raw data obtained from the experiment. Note that the

cumulative volume of mercury injected expressed as a fraction of the pore

volume is the non-wetting phase saturation, Snw. Shown on the capillary

pressure axis is the corresponding pore throat size obtained from Eq.(7.96) as

2 cos

c

RP

σ θ= (7.97)

Figure 7.53. Cumulative pore volume of mercury injected versus capillary pressure.

Also shown on the figure is the wetting phase saturation versus capillary

pressure obtained from

1w nwS S= − (7.98)

Page 834: +Peters Ekwere j. - Petrophysics

7-99

It should be noted that the pore throat radius, R, decreases from left to right

in the figure. Let us replot the saturations versus pore throat size such that

the pore throat size increases from left to right as shown in Figure 7.54. It

should be observed that the Sw versus R curve now represents the cumulative

probability distribution function for the pore volume distribution whereas the

Snw versus R curve represents the expectation curve for the pore volume

distribution. At any value of R, Sw is the fraction of the pore volume occupied

by pores having pore throat size equal to R or less. If the probability density

function for the pore volume distribution is f (R), then

( )final

R

w RS f R dR= ∫ (7.99)

where

( )0 1final

R

Rf R dR =∫ (7.100)

Differentiating Eq.(7.99) using Leibnitz's rule for differentiating an integral

gives the probability density function as

( ) wdSf RdR

= (7.101)

Because of the relationship between Sw and Snw, the probability density

function for the pore volume distribution also is given by

( ) nwdSf RdR

= − (7.102)

Various alternative expressions can be derived for the probability density

function by use of Eq.(7.96). From Eq.(7.96),

Page 835: +Peters Ekwere j. - Petrophysics

7-100

2 cos a constantcP R σ θ= = (7.103)

Differentiating Eq.(7.103) gives

0c cP dR RdP+ = (7.104)

Figure 7.54. Saturations versus pore throat size from mercury injection.

Substituting Eqs.(7.96) and (7.104) into Eqs.(7.101) and (7.102) gives the

following alternative expressions for the probability density function for the

pore volume distribution:

( ) 2

2 cosw c w w

c c

dS P dS dSf RdR R dP R dP

σ θ= = − = − (7.105)

Page 836: +Peters Ekwere j. - Petrophysics

7-101

( ) 2

2 cosnw c w nw

c c

dS P dS dSf RdR R dP R dP

σ θ= − = = (7.106)

Thus, differentiating the saturation versus the pore throat radius curve from

the experiment leads directly to the probability density function for the pore

volume distribution of the rock. Further, a plot of the incremental pore

volume of mercury injected versus pore throat size also is a measure of the

pore volume distribution.

Differentiation of experimental data can be a noisy affair. The three-

point central difference approximation given previously for the calculation of

the welltest derivative function (Eq.(3.65)) can be used to differentiate the

experimental data. It should be noted that the welltest derivative function

should be divided by ti to obtain the first derivative of the function. There are

other differentiation schemes that may be less noisy than the three-point

central difference formula. The five-point central difference formula proposed

by Akima (1970) may be less noisy than the three-point central difference

formula. The computational template for the five-point central difference

formula is shown in Figure 7.55. The first derivative at x3 is given by

Page 837: +Peters Ekwere j. - Petrophysics

7-102

Figure 7.55. Computational template for calculating the first derivative using Akima's method (Akima, 1970).

3

4 3 2 2 1 3

4 3 2 1x

m m m m m mdydx m m m m

− + −⎛ ⎞ =⎜ ⎟ − + −⎝ ⎠ (7.107)

where

5 44

5 4

y ymx x

−=

− (7.108)

4 33

4 3

y ymx x

−=

− (7.109)

3 22

3 2

y ymx x

−=

− (7.110)

Page 838: +Peters Ekwere j. - Petrophysics

7-103

2 11

2 1

y ymx x

−=

− (7.111)

7.11.3 Pore Size Distribution Based on Bundle of Capillary Tubes Model

The objective is to estimate the pore size distribution of a porous

medium from the drainage capillary pressure curve based on a bundle of

capillary tubes model of the porous medium. Further, the pore size

distribution should be scaled into a pore size probability density function.

Let the probability density function for the pore size distribution be

( )Rδ , where R is the pore throat radius. Then, for the porous medium,

0

( ) 1R dRδ∞

=∫ (7.112)

The fractional number of pores with radii between R and R+dR is ( )R dRδ . The

number of pores with radii between R and R+dR is ( )N R dRδ , where N is the

total number of pores (capillary tubes) making up the porous medium. The

cross-sectional area of the pores with radii between R and R+dR is given by

2 ( )cdA R N R dRπ δ= (7.113)

The cross-sectional area occupied by all the pores can be obtained by

integrating Eq.(7.113) to obtain

2

0

( )c TA A N R R dRφ π δ∞

= = ∫ (7.114)

where AT is the total cross-sectional area of the porous medium and φ is the

porosity of the porous medium. Let

Page 839: +Peters Ekwere j. - Petrophysics

7-104

2 2

0

( ) a constant R R R dRδ∞

= =∫ (7.115)

Eq.(7.114) can be rewritten as

2c TA A NRφ π= = (7.116)

The pore volume of the porous medium is given by

2 2

0

( )p TV A L NL R R dR NLRφ π δ π∞

= = =∫ (7.117)

At capillary equilibrium, the wetting phase occupies all the pores with radii

less than R corresponding to the capillary pressure given by

2 cos

( )c wP SR

σ θ= (7.118)

The wetting phase volume is given by

2

0

( )R

wV NL R R dRπ δ= ∫ (7.119)

The wetting phase saturation is given by

2 2

0 02

2

0

( ) ( )

( )

R R

ww

p

R R dR R R dRVSV R

R R dR

δ δ

δ∞= = =∫ ∫

∫ (7.120)

Eq.(7.120) can be rearranged as

Page 840: +Peters Ekwere j. - Petrophysics

7-105

2 2 2

0 0

( ) ( )R R

wR S R R dR r r drδ δ= =∫ ∫ (7.121)

where r is a dummy integration variable, which is not to be confused with the

radius R, the upper limit of the integral. Differentiating Eq.(7.121) with

respect to R gives

2 2

0

( )R

wdS dR r r drdR dR

δ= ∫ (7.122)

The right side of Eq.(7.122) can be evaluated using Leibnitz’s rule for

differentiating a definite integral as follows:

2

2 2

0 0

( )0( ) ( ) 0 (0)R R r rd dR dr r dr R R dr

dR dR dR Rδ

δ δ δ⎡ ⎤∂ ⎣ ⎦= − +

∂∫ ∫ (7.123)

The integrand of the integral on the right side of Eq.(7.123) is zero. Therefore,

Eq.(7.123) simplifies to

2 2

0

( ) ( )Rd r r dr R R

dRδ δ=∫ (7.124)

Substituting Eq.(7.124) into Eq.(7.122) gives

2 2 ( )wdSR R RdR

δ= (7.125)

Differentiating Eq.(7.118) with respect to R gives

2

2 cos( )c wdP SdR R

σ θ= − (7.126)

Page 841: +Peters Ekwere j. - Petrophysics

7-106

Dividing Eq.(7.126) by Eq.(7.125) gives

2

4

2 cos( )( )

c w

w

RdP SdS R R

σ θδ

= − (7.127)

The probability density function for the pore size distribution can be obtained

from Eq.(7.127) as

24

2 cos( )( )c w

w

RR dP SR

dS

σ θδ= −

⎡ ⎤⎢ ⎥⎣ ⎦

(7.128)

Using Eq.(7.96), Eq.(7.128) can be rewritten in the following alternative form:

23

( )( )( )

c w

c w

w

P SRR dP SR

dS

δ= −

⎡ ⎤⎢ ⎥⎣ ⎦

(7.129)

Eqs.(7.118) and (7.129) can be used to calculate the pore size distribution

from a drainage capillary pressure curve as follows:

1. Pick a high Pc(Sw) value corresponding to a low wetting phase

saturation, Sw, and a small pore size, R.

2. Calculate the pore radius, R, using Eq.(7.118).

3. Calculate the derivative of the capillary pressure curve with respect to

the wetting phase saturation at the value of the Pc(Sw) in step 1. Note

that this is a negative quantity.

4. Calculate 2( ) /R Rδ using Eq.(7.129).

Page 842: +Peters Ekwere j. - Petrophysics

7-107

5. Pick lower values of Pc(Sw) and repeat steps 2 through 4 until the entire

capillary pressure curve has been used in the pore size distribution

calculation.

6. Plot the graph of 2( ) /R Rδ versus R. Calculate the area under the graph,

Ag. Using Ag, calculate the constant 2R so as to satisfy Eq.(7.112),

which is the requirement for a probability density function.

7. Using the value of 2R , calculate and plot the graph of ( )Rδ versus R,

which is the required probability density function for the pore size

distribution.

Eq.(7.129) was derived for a general capillary pressure curve obtained

by any method. In the case of the capillary pressure curve obtained by

mercury porosimetry, Eq.(7.129) can be transformed by substitution of

Eq.(7.104) to obtain the probability density function as

( )2

2wdSRR

R dRδ = (7.130)

or

( )2

2nwdSRR

R dRδ = − (7.131)

Using Eq.(7.130) or (7.131), the probability density function can easily be

calculated from the mercury injection data.

Figure 7.56 shows the pore size distributions of various sedimentary

rocks determined from drainage capillary pressure curves.

Page 843: +Peters Ekwere j. - Petrophysics

7-108

Figure 7.56. Pore size distributions of sedimentary rocks based on the bundle of capillary tubes model of the rock (Crocker, 1983).

Page 844: +Peters Ekwere j. - Petrophysics

7-109

Example 7.3

The first three columns of Table 7.9 give the mercury injection data for a low

permeability sandstone sample with k = 0.048 md and φ = 5.6%. Calculate

and plot the following graphs:

1. Snw and Sw versus pore throat radius, R.

2. ΔSnw versus pore throat radius, R.

3. Probability density function for pore volume distribution.

4. Probability density function for pore radius distribution assuming a

bundle of capillary tubes model of the porous medium.

Table 7.9. Mercury Injection Data and Calculated Pore Size Distributions for

Example 7.3.

1 2 3 4 5 6 7 8 9

Pore Volume Pore Radius

Distribution Distribution

Pc Pc ΔSnw R R ( ) wdSf RdR

= ( )2

2wdSRR

R dRδ =

Snw Sw (psi) (dynes/cm2

)

(cm) (μm) (1/μm) (1/μm)

0.000 1.000 124.92 8.613E+06 0.0000

0

8.538E-

05

0.853

8

0.000 0.000

0.031 0.969 150.75 1.039E+07 0.0312

1

7.075E-

05

0.707

5

0.323 0.002

0.068 0.932 175.43 1.210E+07 0.0370

6

6.080E-

05

0.608

0

0.424 0.004

0.104 0.896 200.40 1.382E+07 0.0355

3

5.322E-

05

0.532

2

0.508 0.006

0.168 0.832 249.77 1.722E+07 0.0640

4

4.270E-

05

0.427

0

0.691 0.012

Page 845: +Peters Ekwere j. - Petrophysics

7-110

0.229 0.771 300.12 2.069E+07 0.0613

1

3.554E-

05

0.355

4

0.928 0.024

0.330 0.670 400.20 2.759E+07 0.1011

0

2.665E-

05

0.266

5

1.385 0.064

0.428 0.572 499.51 3.444E+07 0.0981

5

2.135E-

05

0.213

5

2.310 0.166

0.513 0.487 599.28 4.132E+07 0.0847

8

1.780E-

05

0.178

0

2.499 0.259

0.577 0.423 699.36 4.822E+07 0.0637

2

1.525E-

05

0.152

5

2.507 0.354

0.625 0.375 799.13 5.510E+07 0.0478

7

1.335E-

05

0.133

5

2.501 0.461

0.660 0.340 899.29 6.201E+07 0.0351

3

1.186E-

05

0.118

6

2.289 0.534

0.686 0.314 999.18 6.889E+07 0.0264

7

1.067E-

05

0.106

7

2.268 0.654

0.709 0.291 1098.89 7.577E+07 0.0227

2

9.706E-

06

0.097

1

2.636 0.919

0.730 0.270 1198.44 8.263E+07 0.0212

7

8.900E-

06

0.089

0

2.637 1.094

0.748 0.252 1298.50 8.953E+07 0.0180

9

8.214E-

06

0.082

1

2.638 1.284

0.765 0.235 1398.24 9.641E+07 0.0165

3

7.628E-

06

0.076

3

2.875 1.623

0.780 0.220 1498.24 1.033E+08 0.0149

6

7.119E-

06

0.071

2

3.042 1.971

0.794 0.206 1598.66 1.102E+08 0.0141

0

6.672E-

06

0.066

7

3.245 2.394

0.806 0.194 1695.09 1.169E+08 0.0124

5

6.292E-

06

0.062

9

3.255 2.700

0.818 0.182 1797.32 1.239E+08 0.0114

5

5.934E-

06

0.059

3

3.268 3.048

0.829 0.171 1895.87 1.307E+08 0.0107

6

5.626E-

06

0.056

3

3.390 3.518

Page 846: +Peters Ekwere j. - Petrophysics

7-111

0.838 0.162 2000.98 1.380E+08 0.0095

6

5.330E-

06

0.053

3

3.539 4.091

0.856 0.144 2196.94 1.515E+08 0.0175

7

4.855E-

06

0.048

5

3.800 5.295

0.871 0.129 2396.98 1.653E+08 0.0154

9

4.450E-

06

0.044

5

3.866 6.413

0.885 0.115 2597.89 1.791E+08 0.0135

1

4.105E-

06

0.041

1

4.010 7.812

0.897 0.103 2799.03 1.930E+08 0.0121

2

3.810E-

06

0.038

1

4.038 9.133

0.907 0.093 2997.38 2.067E+08 0.0100

5

3.558E-

06

0.035

6

4.054 10.514

0.918 0.082 3248.15 2.240E+08 0.0113

2

3.284E-

06

0.032

8

4.196 12.778

0.928 0.072 3495.86 2.410E+08 0.0098

7

3.051E-

06

0.030

5

4.253 15.004

0.937 0.063 3744.60 2.582E+08 0.0087

5

2.848E-

06

0.028

5

4.213 17.054

0.944 0.056 3996.64 2.756E+08 0.0065

9

2.669E-

06

0.026

7

4.068 18.757

0.950 0.050 4246.84 2.928E+08 0.0064

0

2.511E-

06

0.025

1

4.068 21.180

0.956 0.044 4494.10 3.099E+08 0.0056

2

2.373E-

06

0.023

7

4.041 23.561

0.960 0.040 4745.57 3.272E+08 0.0047

7

2.247E-

06

0.022

5

3.725 24.214

0.965 0.035 4997.24 3.446E+08 0.0041

9

2.134E-

06

0.021

3

3.675 26.495

0.968 0.032 5245.84 3.617E+08 0.0036

7

2.033E-

06

0.020

3

3.645 28.956

0.972 0.028 5496.45 3.790E+08 0.0035

1

1.940E-

06

0.019

4

3.958 34.517

0.975 0.025 5746.20 3.962E+08 0.0036

5

1.856E-

06

0.018

6

4.128 39.346

Page 847: +Peters Ekwere j. - Petrophysics

7-112

0.978 0.022 5994.05 4.133E+08 0.0030

6

1.779E-

06

0.017

8

3.692 38.296

0.981 0.019 6246.10 4.307E+08 0.0025

8

1.708E-

06

0.017

1

3.509 39.516

0.983 0.017 6497.47 4.480E+08 0.0022

9

1.642E-

06

0.016

4

3.491 42.548

0.985 0.015 6744.53 4.650E+08 0.0022

2

1.581E-

06

0.015

8

3.308 43.446

0.987 0.013 6996.48 4.824E+08 0.0016

0

1.524E-

06

0.015

2

2.964 41.879

0.990 0.010 7497.19 5.169E+08 0.0031

6

1.423E-

06

0.014

2

2.913 47.272

0.992 0.008 7997.18 5.514E+08 0.0020

4

1.334E-

06

0.013

3

2.862 52.840

0.995 0.005 8494.95 5.857E+08 0.0025

4

1.256E-

06

0.012

6

2.734 56.951

0.997 0.003 8995.38 6.202E+08 0.0018

8

1.186E-

06

0.011

9

2.661 62.153

0.998 0.002 9495.55 6.547E+08 0.0016

3

1.123E-

06

0.011

2

2.452 63.825

0.999 0.001 9996.48 6.893E+08 0.0009

4

1.067E-

06

0.010

7

1.491 43.010

1.000 0.000 10496.0

0

7.237E+08 0.0006

5

1.016E-

06

0.010

2

0.524 16.671

1.000 0.000 10997.0

3

7.583E+08 0.0000

8

9.699E-

07

0.009

7

0.000 0.000

1.000 0.000 11495.5

6

7.926E+08 0.0000

0

9.278E-

07

0.009

3

0.000 0.000

1.000 0.000 11996.4

7

8.272E+08 0.0000

0

8.891E-

07

0.008

9

0.000

1.000 0.000 12495.5

4

8.616E+08 0.0000

0

8.536E-

07

0.008

5

0.000

1.000 0.000 12996.0

7

8.961E+08 0.0000

0

8.207E-

07

0.008

2

0.000

Page 848: +Peters Ekwere j. - Petrophysics

7-113

1.000 0.000 13495.1

1

9.305E+08 0.0000

0

7.903E-

07

0.007

9

0.000

1.000 0.000 13995.9

0

9.650E+08 0.0000

0

7.621E-

07

0.007

6

0.000

1.000 0.000 14495.6

8

9.995E+08 0.0000

0

7.358E-

07

0.007

4

0.000

1.000 0.000 14996.2

0

1.034E+09 0.0000

0

7.112E-

07

0.007

1

0.000

1.000 0.000 15495.9

8

1.068E+09 0.0000

0

6.883E-

07

0.006

9

0.000

1.000 0.000 15995.1

3

1.103E+09 0.0000

0

6.668E-

07

0.006

7

0.000

1.000 0.000 16495.5

2

1.137E+09 0.0000

0

6.466E-

07

0.006

5

0.000

1.000 0.000 16995.2

8

1.172E+09 0.0000

0

6.276E-

07

0.006

3

0.000

1.000 0.000 17495.3

3

1.206E+09 0.0000

0

6.096E-

07

0.006

1

0.000

1.000 0.000 17995.5

0

1.241E+09 0.0000

0

5.927E-

07

0.005

9

0.000

1.000 0.000 18495.9

5

1.275E+09 0.0000

0

5.766E-

07

0.005

8

0.000

1.000 0.000 18996.3

7

1.310E+09 0.0000

0

5.615E-

07

0.005

6

0.000

1.000 0.000 19495.3

3

1.344E+09 0.0000

0

5.471E-

07

0.005

5

0.000

1.000 0.000 19995.7

3

1.379E+09 0.0000

0

5.334E-

07

0.005

3

0.000

1.000 0.000 20995.3

1

1.448E+09 0.0000

0

5.080E-

07

0.005

1

0.000

1.000 0.000 21995.7

9

1.517E+09 0.0000

0

4.849E-

07

0.004

8

0.000

1.000 0.000 22995.6

4

1.586E+09 0.0000

0

4.638E-

07

0.004

6

0.000

Page 849: +Peters Ekwere j. - Petrophysics

7-114

1.000 0.000 23995.7

2

1.655E+09 0.0000

0

4.445E-

07

0.004

4

0.000

1.000 0.000 24996.3

4

1.724E+09 0.0000

0

4.267E-

07

0.004

3

0.000

1.000 0.000 25994.9

6

1.792E+09 0.0000

0

4.103E-

07

0.004

1

0.000

1.000 0.000 26995.6

5

1.861E+09 0.0000

0

3.951E-

07

0.004

0

0.000

1.000 0.000 27995.6

6

1.930E+09 0.0000

0

3.810E-

07

0.003

8

0.000

1.000 0.000 28996.0

8

1.999E+09 0.0000

0

3.678E-

07

0.003

7

0.000

1.000 0.000 29995.4

9

2.068E+09 0.0000

0

3.556E-

07

0.003

6

0.000

1.000 0.000 30996.2

1

2.137E+09 0.0000

0

3.441E-

07

0.003

4

0.000

1.000 0.000 31995.5

3

2.206E+09 0.0000

0

3.333E-

07

0.003

3

0.000

1.000 0.000 32996.2

0

2.275E+09 0.0000

0

3.232E-

07

0.003

2

0.000

1.000 0.000 33996.5

7

2.344E+09 0.0000

0

3.137E-

07

0.003

1

0.000

1.000 0.000 34996.1

7

2.413E+09 0.0000

0

3.048E-

07

0.003

0

0.000

1.000 0.000 35996.3

2

2.482E+09 0.0000

0

2.963E-

07

0.003

0

0.000

1.000 0.000 36995.5

3

2.551E+09 0.0000

0

2.883E-

07

0.002

9

0.000

1.000 0.000 37996.3

9

2.620E+09 0.0000

0

2.807E-

07

0.002

8

0.000

1.000 0.000 38996.3

8

2.689E+09 0.0000

0

2.735E-

07

0.002

7

0.000

1.000 0.000 39995.7

5

2.758E+09 0.0000

0

2.667E-

07

0.002

7

0.000

Page 850: +Peters Ekwere j. - Petrophysics

7-115

1.000 0.000 41995.3

9

2.896E+09 0.0000

0

2.540E-

07

0.002

5

0.000

1.000 0.000 43995.2

2

3.034E+09 0.0000

0

2.424E-

07

0.002

4

0.000

1.000 0.000 45993.7

9

3.171E+09 0.0000

0

2.319E-

07

0.002

3

0.000

1.000 0.000 47991.1

8

3.309E+09 0.0000

0

2.222E-

07

0.002

2

0.000

1.000 0.000 49990.5

5

3.447E+09 0.0000

0

2.134E-

07

0.002

1

0.000

1.000 0.000 51989.8

4

3.585E+09 0.0000

0

2.051E-

07

0.002

1

0.000

1.000 0.000 53989.7

5

3.723E+09 0.0000

0

1.975E-

07

0.002

0

0.000

1.000 0.000 55989.4

0

3.861E+09 0.0000

0

1.905E-

07

0.001

9

0.000

1.000 0.000 57988.7

6

3.998E+09 0.0000

0

1.839E-

07

0.001

8

0.000

Solution to Example 7.3

The calculated results are shown in columns 4 to 9 of Table 7.9. For

mercury, σ = 480 dynes/cm and θ = 140º. Figure 7.57 shows the graphs of

non-wetting phase saturation, wetting phase saturation and incremental pore

volume of mercury injected versus pore throat size. It should be observed that

the incremental pore volume of mercury injected provides a rough estimate of

the pore volume distribution for the porous medium.

Page 851: +Peters Ekwere j. - Petrophysics

7-116

Figure 7.57. Graphs of Snw, Sw and ΔSnw versus pore throat radius for Example 7.3.

Columns 8 and 9 of Table 7.9 show the probability density functions for the

pore volume and pore throat radius distributions calculated with the 5-point

central difference formula of Akima. For the pore radius distribution, 2R was

found to be 0.003284. Figures 7.58 and 7.59 show the graphs of the

probability density functions for pore volume and pore radius distributions.

Page 852: +Peters Ekwere j. - Petrophysics

7-117

Figure 7.58. Probability density function for pore volume distribution for Example 7.3.

Figure 7.59. Probability density function for pore throat radius distribution assuming a bundle of capillary tubes model for the sample of Example 7.3.

Page 853: +Peters Ekwere j. - Petrophysics

7-118

7.11.4 Mercury Injection Porosimeter

Several different types of equipment are used for mercury injection

porosimetry. The mercury injection equipment used by Purcell (1949) was

described in Section 7.9.2. However, more modern mercury porosimeters are

now available. One such porosimeter is the Autopore IV 9520, manufactured

by Micromeritics. Figure 7.60 shows this equipment. It is a compact, high

precision, computer-controlled instrument capable of injecting mercury into a

sample at capillary pressures up to 60,000 psi. Pressure is measured with

transducers that produce an electrical signal that is proportional to the

applied pressure. This analog signal is converted into digital code for

processing by the monitoring computer.

The transducer that detects the volume of mercury injected is

integrated into the sample holder assembly as shown in Figure 7.61. The

sample cup has a capillary stem, which serves both as the mercury reservoir

and as an element of the mercury volume transducer. The capillary stem is

made of glass, which is an insulator but the outer face of the capillary is

plated with metal, which is a conductor. The sample is placed in the sample

cup and the cup along with the capillary stem is filled with mercury, which is

a conductor. The combination of two concentric electrical conductors

separated by an insulator produces a co-axial capacitor, whose capacitance is

a function of the length of mercury in the capillary. As mercury is injected

into the sample, the length of mercury in the capillary decreases and this in

turn changes the capacitance of the penetrometer assembly. The capacitance

of the assembly is measured at each step of mercury injection and used to

calculate the volume of mercury injected into the sample.

Page 854: +Peters Ekwere j. - Petrophysics

7-119

Figure 7.60. Autopore IV 9520 mercury injection porosimeter (from Micromeritics).

Page 855: +Peters Ekwere j. - Petrophysics

7-120

Figure 7.61. Penetrometer for the Autopore IV 9520 mercury injection

porosimeter (from Micromeritics).

Page 856: +Peters Ekwere j. - Petrophysics

7-121

7.12 CALCULATION OF PERMEABILITY FROM DRAINAGE CAPILLARY PRESSURE CURVE

7.12.1 Calculation of Absolute Permeability from Drainage Capillary Pressure Curve

The objective is to estimate the permeability of a porous medium from

the drainage capillary pressure curve based on the bundle of capillary tubes

model of the porous medium. This problem was addressed in Section 3.7.3

using the probability density function for the pore throat diameter. Here, we

present the estimate of the absolute permeability of the porous medium based

on the probability density function for the pore throat radius obtained from

mercury injection porosimetry.

Let the probability density function for the pore throat radius be ( )Rδ ,

where R is the pore throat radius. Then, for the porous medium,

0

( ) 1R dRδ∞

=∫ (7.132)

The fractional number of pores with radii between R and R+dR is ( )R dRδ . The

number of pores with radii between R and R+dR is ( )N R dRδ , where N is the

total number of pores making up the porous medium. The cross-sectional

area of the pores with radii between R and R+dR is given by

2 ( )cdA R N R dRπ δ= (7.133)

The cross-sectional area occupied by all the pores can be obtained by

integrating Eq.(7.133) to obtain

2

0

( )c TA A N R R dRφ π δ∞

= = ∫ (7.134)

Page 857: +Peters Ekwere j. - Petrophysics

7-122

where AT is the total cross-sectional area of the porous medium and φ is the

porosity of the porous medium. Let

2 2

0

( ) a constant R R R dRδ∞

= =∫ (7.135)

Eq.(7.135) can be written as

2c TA A NRφ π= = (7.136)

The volumetric flow rate for pores with radii between R and R+dR is

given by Hagen-Poiseuille’s law as

( )2

21 ( )8 ( )

R Pdq R N R dRR L

π δμ τ

⎡ ⎤ Δ= ⎢ ⎥

⎣ ⎦ (7.137)

where ( )Rτ is the tortuosity of the porous medium and L is the length of the

medium. Note that the tortuosity is a function of the pore throat size, R. The

wetting phase occupies the pores with radii less than R. The volumetric flow

rate of the wetting phase is obtained by integrating Eq.(7.137) as

4

0

( )8 ( )

R

ww

N P R Rq dRL R

π δμ τ

Δ= ∫ (7.138)

The effective permeability to the wetting phase is given by Darcy’s law as

4

0

( )8 ( )

Rw w

wT T

q L N R Rk dRA P A R

μ π δτ

= =Δ ∫ (7.139)

The pore volume of the porous medium is given by

Page 858: +Peters Ekwere j. - Petrophysics

7-123

2 2

0

( )p TV A L NL R R dR NLRφ π δ π∞

= = =∫ (7.140)

At capillary equilibrium, the wetting phase occupies all the pores with radii

less than R corresponding to the capillary pressure given by

2 cos

( )c wP SR

σ θ= (7.141)

The wetting phase volume is given by

2

0

( )R

wV NL R R dRπ δ= ∫ (7.142)

The wetting phase saturation is given by

2 2

0 02

2

0

( ) ( )

( )

R R

ww

p

R R dR R R dRVSV R

R R dR

δ δ

δ∞= = =∫ ∫

∫ (7.143)

Eq.(7.143) can be rearranged as

2 2 2

0 0

( ) ( )R R

wR S R R dR r r drδ δ= =∫ ∫ (7.144)

where r is a dummy integration variable, which is not to be confused with the

pore throat radius R, the upper limit of the integral. Differentiating Eq.(7.144)

with respect to R gives

2 2

0

( )R

wdS dR r r drdR dR

δ= ∫ (7.145)

Page 859: +Peters Ekwere j. - Petrophysics

7-124

The right side of Eq.(7.145) can be evaluated using Leibnitz’s rule for

differentiating a definite integral as follows:

2

2 2

0 0

( )0( ) ( ) 0 (0)R R r rd dR dr r dr R R dr

dR dR dR Rδ

δ δ δ⎡ ⎤∂ ⎣ ⎦= − +

∂∫ ∫ (7.146)

The integrand of the integral on the right side of Eq.(7.146) is zero. Therefore,

Eq.(7.146) simplifies to

2 2

0

( ) ( )Rd r r dr R R

dRδ δ=∫ (7.147)

Substituting Eq.(7.147) into Eq.(7.145) gives

2

2

( )w

R R dRdSR

δ= (7.148)

Substituting Eqs.(7.136) and (7.148) into Eq.(7.139) gives the effective

permeability to the wetting phase as

2

08 ( )

wS

w wRk dSR

φτ

= ∫ (7.149)

The R in Eq.(7.149) is the largest pore size filled with the wetting phase. This

can be obtained from the capillary pressure curve as

2 cos

( )c w

RP Sσ θ

= (7.150)

Let us approximate the tortuosity by a function of the form

Page 860: +Peters Ekwere j. - Petrophysics

7-125

( ) aRRατ = (7.151)

where a and α are numerical constants. This function satisfies the

requirement that the tortuosity increases as the pore size decreases.

Substituting Eqs.(7.150) and (7.151) into Eq.(7.149) gives the effective

permeability to the wetting phase as

( )2

20

2 cos8

wSw

wc

dSka P

α

α

σ θφ

+

+= ∫ (7.152)

The absolute permeability of the porous medium can be obtained from

Eq.(7.152) for the medium fully saturated by the wetting phase as

( )2 1

20

2 cos8

w

c

dSka P

α

α

σ θφ

+

+= ∫ (7.153)

Let α = 0 and 1/a = F1. Substituting these values into Eq.(7.153) gives the

absolute permeability of the porous medium as

( )2 1

1 20

2 cos8

w

c

dSk FP

σ θφ= ∫ (7.154)

If the capillary pressure is in psi, the permeability is in millidarcy and the

surface tension is in dynes/cm, Eq.(7.154) becomes, after the units

conversion,

( )1

2

1 20

10.6566 cos w

c

dSk FP

σ θ φ= ∫ (7.155)

Page 861: +Peters Ekwere j. - Petrophysics

7-126

In the case of mercury injection, σ = 480 dynes/cm and θ = 140º. Substituting

these values into Eq.(7.155) gives the permeability as

1

61 2

0

1.441 10 w

c

dSk x FP

φ= ∫ (7.156)

Eq.(7.156) is Purcell's equation in which F1 is a lithology factor to

account for tortuosity of the porous medium. Purcell (1949) performed

measurements on core samples and computed the lithology factors that made

the measured permeabilities to be equal to the calculated permeabilities from

the drainage capillary pressure curves of the samples. His results are shown

in Table 7.10. The permeabilities of samples 17 and 18 were too low to

measure but were estimated to be less than 0.1 md. The calculated lithology

factors ranged from 0.085 to 0.363, with an average value of 0.216. He

suggested the use of the average value of 0.216 for estimating the

permeability of a rock from its drainage capillary pressure curve. However, as

shown in Figure 7.62, there is a positive correlation between the lithology

factor and permeability. Clearly, the higher the permeability, the higher the

lithology factor. Similarly, the lower the permeability, the lower the lithology

factor. The lithology factor is related to the reciprocal of the tortuosity of the

porous medium. Therefore, the positive correlation between the lithology

factor and permeability is to be expected. Because of this correlation, the use

of the average lithology factor of 0.216 to calculate the permeability of all

porous media can lead to poor estimates of permeability. Figure 7.63

compares the measured permeability with the calculated permeability using

the average lithology factor of 0.216 plotted on log scales. The correlation

between the two sets of data appears very strong. However, when the same

data are plotted on linear scales as shown in Figure 7.64, the correlation is

weaker than in Figure 7.63.

Table 7.10. Lithology factors for various core samples (Purcell, 1949).

Page 862: +Peters Ekwere j. - Petrophysics

7-127

Lithology Calculated

Permeability

Sample Permeability Factor Using F1 = 0.216

No (md) (md)

1 1.2 0.085 3.04

2 12 0.122 21.2

3 13.4 0.168 17.3

4 36.9 0.149 53.5

5 57.4 0.200 61.9

6 70.3 0.165 91.6

7 110 0.257 92.3

8 116 0.256 97.5

9 144 0.191 163

10 336 0.107 680

11 430 0.216 430

12 439 0.273 348

13 496 0.276 388

14 772 0.185 902

15 1070 0.282 816

16 1459 0.363 865

17 0.003

18 0.1

19 35.7 0.182 42.2

20 40.2 0.158 54.9

21 184 0.231 172

22 235 0.276 183

23 307 0.215 308

24 320 0.163 422

25 506 0.284 383

26 634 0.272 502

27 1150 0.338 734

Page 863: +Peters Ekwere j. - Petrophysics

7-128

Figure 7.62. Correlation between lithology factor and permeability (Purcell, 1949).

Figure 7.63. A comparison of measured permeability and calculated permeability using the average lithology factor of 0.216 plotted on log scales

(Purcell, 1949).

Page 864: +Peters Ekwere j. - Petrophysics

7-129

Figure 7.64. A comparison of measured permeability and calculated permeability using the average lithology factor of 0.216 plotted on linear

scales (Purcell, 1949).

Example 7.4

Estimate the permeability of the sample of Example 7.3 using the mercury

injection capillary pressure data of Table 7.9.

Solution to Example 7.4

Table 7.11 presents the calculated results. Figure 7.65 shows the graphs of cP

and 2

1

cP versus wS . The integration called for in Eq.(7.156) was performed with

the trapezoidal rule to obtain

1 6 220

8.355 10 1/w

c

dS x psiP

−=∫

Page 865: +Peters Ekwere j. - Petrophysics

7-130

Using the average lithology factor of 0.216 in Eq.(7.156) gives

( )( )( )6 61.441 10 0.216 0.056 8.355 10 0.146k x x −= = md, compared to the measured

permeability of 0.048 md. However, if the correlation of the lithology factor

with permeability is recognized and taken into account, the lithology factor for

the very low permeability sample should be much less than 0.216. Using the

lowest lithology factor of 0.085 in Table 7.10 gives the permeability as

( )( )( )6 61.441 10 0.085 0.056 8.355 10 0.057k x x −= = md, compared to the measured

value of 0.048 md. Thus, the low lithology factor of 0.085 gives a better

estimate of the permeability than the average value of 0.216 for this tight

sample.

Figure 3.30 gives a correlation between tortuosity and porosity based on

the experimental data of Winsauer et al. (1952). The equation of the

regression line is

27.35 10.987τ φ= − +

For the sample with a porosity of 0.056, the tortuosity predicted by the

regression line is 9.4554. Therefore, the lithology factor (1/τ) predicted by the

correlation is 0.1058. Using the this value of the lithology factor in Eq.(7.156)

gives the estimated permeability as

( )( )( )6 61.441 10 0.1058 0.056 8.355 10 0.071k x x −= = md, compared to the measured

permeability of 0.048 md. This estimate also is better than that based on the

average lithology factor of 0.216.

Table 7.11. Calculated Results for Example 7.4.

Page 866: +Peters Ekwere j. - Petrophysics

7-131

cP 2

1

cP 21

wS w

c

dSP∫

wS psi ( )21/ psi ( )21/ psi

1.000 124.919 6.408E-

05

0.000E+0

0

0.969 150.747 4.401E-

05

1.687E-06

0.932 175.428 3.249E-

05

3.104E-06

0.896 200.401 2.490E-

05

4.124E-06

0.832 249.768 1.603E-

05

5.434E-06

0.771 300.121 1.110E-

05

6.266E-06

0.670 400.203 6.244E-

06

7.143E-06

0.572 499.515 4.008E-

06

7.646E-06

0.487 599.278 2.784E-

06

7.934E-06

0.423 699.359 2.045E-

06

8.088E-06

0.375 799.133 1.566E-

06

8.174E-06

0.340 899.290 1.237E-

06

8.223E-06

0.314 999.184 1.002E-

06

8.253E-06

0.291 1098.894 8.281E-

07

8.274E-06

0.270 1198.443 6.962E-

07

8.290E-06

0.252 1298.503 5.931E- 8.302E-06

Page 867: +Peters Ekwere j. - Petrophysics

7-132

07

0.235 1398.241 5.115E-

07

8.311E-06

0.220 1498.239 4.455E-

07

8.318E-06

0.206 1598.665 3.913E-

07

8.324E-06

0.194 1695.092 3.480E-

07

8.328E-06

0.182 1797.317 3.096E-

07

8.332E-06

0.171 1895.866 2.782E-

07

8.335E-06

0.162 2000.983 2.498E-

07

8.338E-06

0.144 2196.945 2.072E-

07

8.342E-06

0.129 2396.980 1.740E-

07

8.345E-06

0.115 2597.892 1.482E-

07

8.347E-06

0.103 2799.027 1.276E-

07

8.349E-06

0.093 2997.379 1.113E-

07

8.350E-06

0.082 3248.154 9.478E-

08

8.351E-06

0.072 3495.865 8.183E-

08

8.352E-06

0.063 3744.603 7.132E-

08

8.353E-06

0.056 3996.642 6.261E-

08

8.353E-06

0.050 4246.843 5.545E- 8.353E-06

Page 868: +Peters Ekwere j. - Petrophysics

7-133

08

0.044 4494.100 4.951E-

08

8.354E-06

0.040 4745.567 4.440E-

08

8.354E-06

0.035 4997.241 4.004E-

08

8.354E-06

0.032 5245.841 3.634E-

08

8.354E-06

0.028 5496.453 3.310E-

08

8.354E-06

0.025 5746.203 3.029E-

08

8.354E-06

0.022 5994.055 2.783E-

08

8.355E-06

0.019 6246.104 2.563E-

08

8.355E-06

0.017 6497.474 2.369E-

08

8.355E-06

0.015 6744.532 2.198E-

08

8.355E-06

0.013 6996.476 2.043E-

08

8.355E-06

0.010 7497.188 1.779E-

08

8.355E-06

0.008 7997.178 1.564E-

08

8.355E-06

0.005 8494.954 1.386E-

08

8.355E-06

0.003 8995.378 1.236E-

08

8.355E-06

0.002 9495.550 1.109E-

08

8.355E-06

0.001 9996.479 1.001E- 8.355E-06

Page 869: +Peters Ekwere j. - Petrophysics

7-134

08

0.000 10496.00

0

9.077E-

09

8.355E-06

0.000 10997.02

9

8.269E-

09

8.355E-06

0.000 11495.55

9

7.567E-

09

8.355E-06

0.000 11996.47

0

6.949E-

09

8.355E-06

0.000 12495.53

8

6.405E-

09

8.355E-06

0.000 12996.07

4

5.921E-

09

8.355E-06

0.000 13495.10

8

5.491E-

09

8.355E-06

0.000 13995.90

2

5.105E-

09

8.355E-06

0.000 14495.68

0

4.759E-

09

8.355E-06

0.000 14996.19

5

4.447E-

09

8.355E-06

0.000 15495.97

6

4.164E-

09

8.355E-06

0.000 15995.13

0

3.909E-

09

8.355E-06

0.000 16495.52

0

3.675E-

09

8.355E-06

0.000 16995.27

7

3.462E-

09

8.355E-06

0.000 17495.33

2

3.267E-

09

8.355E-06

0.000 17995.50

4

3.088E-

09

8.355E-06

0.000 18495.94 2.923E- 8.355E-06

Page 870: +Peters Ekwere j. - Petrophysics

7-135

7 09

0.000 18996.36

5

2.771E-

09

8.355E-06

0.000 19495.32

6

2.631E-

09

8.355E-06

0.000 19995.73

2

2.501E-

09

8.355E-06

0.000 20995.31

3

2.269E-

09

8.355E-06

0.000 21995.79

1

2.067E-

09

8.355E-06

0.000 22995.64

3

1.891E-

09

8.355E-06

0.000 23995.72

1

1.737E-

09

8.355E-06

0.000 24996.34

4

1.600E-

09

8.355E-06

0.000 25994.96

3

1.480E-

09

8.355E-06

0.000 26995.65

4

1.372E-

09

8.355E-06

0.000 27995.66

0

1.276E-

09

8.355E-06

0.000 28996.07

8

1.189E-

09

8.355E-06

0.000 29995.49

4

1.111E-

09

8.355E-06

0.000 30996.21

1

1.041E-

09

8.355E-06

0.000 31995.52

7

9.768E-

10

8.355E-06

0.000 32996.19

9

9.185E-

10

8.355E-06

0.000 33996.57 8.652E- 8.355E-06

Page 871: +Peters Ekwere j. - Petrophysics

7-136

4 10

0.000 34996.16

8

8.165E-

10

8.355E-06

0.000 35996.32

0

7.718E-

10

8.355E-06

0.000 36995.53

1

7.306E-

10

8.355E-06

0.000 37996.39

1

6.927E-

10

8.355E-06

0.000 38996.37

9

6.576E-

10

8.355E-06

0.000 39995.75

4

6.251E-

10

8.355E-06

0.000 41995.39

1

5.670E-

10

8.355E-06

0.000 43995.21

9

5.166E-

10

8.355E-06

0.000 45993.78

5

4.727E-

10

8.355E-06

0.000 47991.17

6

4.342E-

10

8.355E-06

0.000 49990.55

5

4.002E-

10

8.355E-06

0.000 51989.83

6

3.700E-

10

8.355E-06

0.000 53989.75

4

3.431E-

10

8.355E-06

0.000 55989.39

8

3.190E-

10

8.355E-06

0.000 57988.76

2

2.974E-

10

8.355E-06

0.000 59988.26

6

2.779E-

10

8.355E-06

Page 872: +Peters Ekwere j. - Petrophysics

7-137

Figure 7.65. cP and 2

1

cP versus wS for Example 7.4.

7.12.2 Calculation of Relative Permeabilities from Drainage Capillary Pressure Curve

The relative permeability to the wetting phase is given by

Page 873: +Peters Ekwere j. - Petrophysics

7-138

2

01

20

( )

wSw

cwrw w

w

c

dSPkk S

k dSP

α

α

+

+

= =∫

∫ (7.157)

The relative permeability to the non-wetting phase is given by

1

2

1

20

( ) w

w

cSnwrnw w

w

c

dSPkk S

k dSP

α

α

+

+

= =∫

∫ (7.158)

If α = 0, Eqs.(7.157) and (7.158) become

2

01

20

( )

wSw

cwrw w

w

c

dSPkk S

k dSP

= =∫

∫ (7.159)

and

1

2

1

20

( ) w

w

cSnwrnw w

w

c

dSPkk S

k dSP

= =∫

∫ (7.160)

The weakness of the above relative permeability models is that (krw+krnw) = 1,

which does not agree with experimental observations of relative permeability

functions. Experiments show that, in general, (krw+krnw) < 1. This lack of

agreement of the models with experiments is because the tortuosity of the

Page 874: +Peters Ekwere j. - Petrophysics

7-139

porous medium has been neglected. In fact, the tortuosity of the medium in

the presence of multiphase fluids is a function of saturation. Furthermore, the

models do not allow for residual saturations of the wetting and non-wetting

phases. These defects of the models will be corrected in Chapter 8 to derive

more realistic relative permeability curves from drainage capillary pressure

curves.

7.13 EMPIRICAL CAPILLARY PRESSURE MODELS

Often, it is desirable to fit analytical models to capillary pressure curves

to simply reservoir performance calculations, especially in numerical

simulations. Capillary pressure appears in the immiscible displacement model

as a derivative, either as c

D

Px

∂∂

or c

w

dPdS

. Therefore, an analytical model will allow

the derivatives to be calculated without numerical noise. Several empirical

analytical models are available for this purpose. We present two such models

from Brooks and Corey (1966) and van Genuchten (1980).

7.13.1 Brooks-Corey Capillary Pressure Models

A popular capillary pressure model in the petroleum industry and soil

physics is the Brooks-Corey model (Brooks and Corey, 1966). Based on

evaluations of several drainage capillary pressure curves for consolidated

porous media, Brooks and Corey observed that all the drainage capillary

pressure curves they examined could be represented by linear functions of the

form

*ln ln lnw c eS P Pλ λ= − + (7.161)

or

Page 875: +Peters Ekwere j. - Petrophysics

7-140

*1ln ln lnc w eP S Pλ

= − + (7.162)

by an appropriate choice of irreducible wetting phase saturation, where *wS is

the reduced wetting phase saturation defined by

*

1w wirr

wwirr

S SSS

−=

− (7.163)

In Eqs.(7.161) and (7.162), Pe is a constant given by the value of Pc on the

straight lines at *wS = 1, and λ is the pore size distribution index obtained from

the slopes of the straight lines. It should be observed from Eq.(7.162) that λ

controls the slope of the linear capillary pressure plot. A large value of λ gives

a small slope, which corresponds to the capillary pressure curve with a

narrow pore size distribution whereas a small value of λ gives a large slope,

which corresponds to the capillary pressure curve for a wide pore size

distribution. Thus, small values of λ indicate a wide pore size distribution

whereas large values of λ indicate a narrow pore size distribution. A porous

medium with a uniform pore size corresponds to λ = ∞ . In view of these

observations, Brooks and Corey called λ the pore size distribution index.

Eqs.(7.161) and (7.162) give a drainage capillary pressure model of the form

( )1

*c e wP P S λ

−= (7.164)

Brooks and Corey also proposed an imbibition capillary pressure model of the

form

( )1

1c e eP P S λ−⎡ ⎤

= −⎢ ⎥⎣ ⎦

(7.165)

Page 876: +Peters Ekwere j. - Petrophysics

7-141

where Se is the effective wetting phase saturation defined by

1

w wirre

wirr nwr

S SSS S

−=

− − (7.166)

where Snwr is the residual non-wetting phase saturation. To fit this model to

measured drainage capillary pressure data, a log-log plot of the drainage

capillary pressure data is made either as *ln wS versus ln cP or as ln cP versus

*ln wS . If the plot is nonlinear, then wirrS is adjusted until the plot is linear. Pe is

determined from the linear log-log plot at * 1wS = and λ is determined from the

slope of the straight line. The two parameters are then substituted into

Eqs.(7.164) and (7.165) to calculate the drainage and imbibition capillary

pressure curves.

Example 7.5

1. Fit the Brooks-Corey model to the air-water drainage capillary pressure

data given in Table 7.12 and compare the result of the model to the

original data.

2. Derive the spontaneous imbibition capillary pressure curve for the

sample using the Brooks-Corey spontaneous imbibition capillary

pressure model.

Table 7.12. Air-Water Drainage Capillary Pressure Data for Example 7.5.

Drainag

e

Page 877: +Peters Ekwere j. - Petrophysics

7-142

Pc

Sw psi

1.000 1.973

0.950 2.377

0.900 2.840

0.850 3.377

0.800 4.008

0.750 4.757

0.700 5.663

0.650 6.781

0.600 8.195

0.550 10.039

0.500 12.547

0.450 16.154

0.400 21.787

0.350 31.817

0.300 54.691

0.278 78.408

Solution to Example 7.4

1. Using Swirr = 0.278 as indicated by the drainage capillary pressure data

of Table 7.12, a log-log graph of *wS versus cP was plotted as shown in

Figure 7.66. Clearly, the graph is nonlinear. The curve fitting procedure

calls for the adjustment of Swirr until the graph is linear. Figure 7.67

shows the log-log plot with Swirr = 0.100. It is linear. Shown in Figure

7.68 is the corresponding log-log plot of cP versus *wS for Swirr = 0.100.

As expected, it too is linear.

Page 878: +Peters Ekwere j. - Petrophysics

7-143

Figure 7.66.Log-log plot of *wS versus cP for 0.278wirrS = for Example 7.5.

Figure 7.67. Log-log plot of *wS versus cP for 0.100wirrS = for Example 7.5.

Page 879: +Peters Ekwere j. - Petrophysics

7-144

Figure 7.68. Log-log plot of cP versus *wS for 0.100wirrS = for Example 7.5.

From Figure 7.68, the resulting Brooks-Corey drainage capillary

pressure straight line is

*ln 2.1443ln ln 2.2238c wP S= − +

Therefore,

1 2.1443λ

− = −

0.4664λ =

2.2238eP =

Page 880: +Peters Ekwere j. - Petrophysics

7-145

The Brooks-Corey drainage capillary pressure equation for this example

is

( ) ( )1 2.1443* *2.2238c e w wP P S Sλ

− −= =

Figure 7.69 shows a comparison of the Brooks-Corey capillary pressure

model with the measured capillary pressure curve. The agreement is

good. The Brooks-Corey log-log plot extended the capillary pressure

data to an irreducible wetting phase saturation of 0.10. It also changed

the displacement pressure from 1.973 to 2.224 psi. Figure 7.70 shows

the same comparison as in Figure 7.68 but on an enlarged capillary

pressure scale. The overall agreement between the model fit and the

original data is good.

Figure 7.69. A comparison of the Brooks-Corey drainage capillary pressure curve with the measured data for Example 7.5.

Page 881: +Peters Ekwere j. - Petrophysics

7-146

Figure 7.70. A comparison of the Brooks-Corey drainage capillary pressure curve with the measured data on an enlarged capillary pressure scale for

Example 7.5.

2. Figure 7.71 shows the predicted Brooks-Corey drainage and imbibition

capillary pressure curves for Snwr = 0.25 for the sample of Example 7.5.

Table 7.13 shows the original capillary pressure data along with the

calculated results with the Brooks-Corey model.

Page 882: +Peters Ekwere j. - Petrophysics

7-147

Figure 7.71. A comparison of the Brooks-Corey drainage and imbibition capillary pressure curves for the sample of Example 7.5.

Table 7.13. Calculated Results with the Brooks-Corey Model for Example 7.5.

Original

Data Brooks-Corey Model

Drainage Drainage Imbibition

Pc Pc Pc

Sw psi *wS psi Se psi

1.000 1.973 1.000 2.224

0.950 2.377 0.944 2.514

0.900 2.840 0.889 2.863

0.850 3.377 0.833 3.288

0.800 4.008 0.778 3.812

0.750 4.757 0.722 4.468 1.000 0.000

0.700 5.663 0.667 5.305 0.923 0.416

Page 883: +Peters Ekwere j. - Petrophysics

7-148

0.650 6.781 0.611 6.393 0.846 0.958

0.600 8.195 0.556 7.843 0.769 1.679

0.550 10.039 0.500 9.831 0.692 2.669

0.500 12.547 0.444 12.656 0.615 4.075

0.450 16.154 0.389 16.851 0.538 6.163

0.400 21.787 0.333 23.452 0.462 9.448

0.350 31.817 0.278 34.672 0.385 15.032

0.300 54.691 0.222 55.947 0.308 25.620

0.278 78.408 0.198 71.829 0.274 33.524

0.250 0.167 103.678 0.231 49.374

0.200 0.111 247.331 0.154 120.867

0.150 0.056 1093.394 0.077 541.933

0.100 0.000 0.000

The Brooks-Corey capillary pressure model cannot adequately fit a

capillary pressure curve with an inflection (S-shaped curved) such as shown

in Figure 7.13 for sample C. In fact, fitting the Brooks-Corey model to such a

capillary pressure curve will eliminate the inflection and turn the capillary

curve into a hyperbola. The shape of the drainage capillary pressure curve of

Example 7.5 was that of a hyperbola. As a result, the Brooks-Corey model

gave a good fit of the data. If the original data had an inflection as typically

observed in poorly sorted porous media, the Brooks-Corey model would have

given a poor fit, especially at high wetting phase saturations. Since the shape

of the drainage capillary pressure curve is a reflection of the pore throat size

distribution, other capillary pressure models have been devised that preserve

the shape of the drainage capillary pressure curve, especially at high wetting

phase saturation.

Page 884: +Peters Ekwere j. - Petrophysics

7-149

7.13.2 van Genuchten Capillary Pressure Model

An empirical capillary pressure model that preserves the shape of the

capillary pressure curve at high wetting phase saturations was proposed by

van Genuchten (1980). This model is widely used in soil physics and in

hydrology. The model is given by

( )

* 11

m

w nc

SPα

⎡ ⎤= ⎢ ⎥

+⎢ ⎥⎣ ⎦ (7.167)

where α, n and m are the fitting parameters.

Example 7.6

Fit the van Genuchten model to the air-water drainage capillary pressure data

given in Table 7.14 and compare the result of the model to the original data.

Solution to Example 7.6

Table 7.14 shows the calculated results for the van Genuchten model using

the following parameters:

α = 0.018

m = 45

n = 6

Figure 7.72 compares the van Genuchten model with the measured data. The

agreement is good especially at high values of the wetting phase saturation. It

should be observed that this model can match capillary pressure curves with

inflections, unlike the Brooks-Corey model. It can also match a capillary

pressure curve with a hyperbolic shape.

Page 885: +Peters Ekwere j. - Petrophysics

7-150

Table 7.14. Calculated Results for van Genuchten model.

Measured Data

van Genuchten

Model

Pc

Sw (psi) *wS *

wS Sw

0.20 50.0 0.000 0.000 0.200

0.24 39.0 0.050 0.006 0.205

0.25 38.0 0.063 0.012 0.210

0.28 35.0 0.100 0.065 0.252

0.30 34.0 0.125 0.100 0.280

0.35 32.0 0.188 0.199 0.359

0.40 30.0 0.250 0.332 0.466

0.50 28.5 0.375 0.444 0.555

0.56 28.0 0.450 0.481 0.585

0.60 27.8 0.500 0.496 0.597

0.68 27.0 0.600 0.555 0.644

0.70 26.8 0.625 0.569 0.655

0.70 26.8 0.625 0.569 0.655

0.75 26.0 0.688 0.625 0.700

0.80 24.8 0.750 0.701 0.761

0.85 23.0 0.813 0.798 0.838

0.90 21.0 0.875 0.877 0.902

0.95 18.0 0.938 0.949 0.959

1.00 13.5 1.000 0.991 0.993

Page 886: +Peters Ekwere j. - Petrophysics

7-151

Figure 7.72. A comparison of van Genuchten model with measured capillary pressure data for Example 7.6.

7.14 CAPILLARY TRAPPING IN POROUS MEDIA

Capillary trapping ensures that an immiscible displacement at normal

interfacial tensions and rates is never complete. There is always a residual

phase that is trapped. Several models have been proposed to explain capillary

trapping. We examine two such models here: the pore doublet model and the

snap-off model.

7.14.1 Pore Doublet Model of Capillary Trapping

Figure 7.73 shows the pore doublet model, which consists of two pores

that are joined at the inlet and outlet ends, with one pore larger than the

other. The pore doublet is initially filled with a non-wetting phase. A wetting

phase is then injected at a rate q to displace the non-wetting phase from both

Page 887: +Peters Ekwere j. - Petrophysics

7-152

branches of the pore doublet. The problem is to determine which of the two

interfaces in the capillary tubes will arrive at the outlet first (Point B). We will

assume that once the interface in one of the capillary tubes has arrived at B,

the non-wetting phase in the other tube will be trapped. To determine which

interface will arrive at B first, we need to derive the expressions for the

velocities of the interfaces as a function of the relevant parameters of the

model. Although it is not necessary to do so, let us assume that the wetting

and non-wetting phases have the same viscosity to simplify the analysis.

Figure 7.73. Pore doublet model. (a) in a porous medium; (b) capillary tubes

approximation.

The pressure drop across each capillary tube is given by

Page 888: +Peters Ekwere j. - Petrophysics

7-153

A B A w w nw nw BP P P P P P P P− = − + − + − (7.168)

where Pw and Pnw are the pressures on either side of the interface. From

Hagen-Poisseuille's law,

( )41

118

A wP PrqL

πμ

−= (7.169)

( )( )

41

118

nw BP PrqL L

πμ

−=

− (7.170)

where L1 is the distance of the interface from the inlet end and L is the total

length of the pore doublet from A to B. Substituting Eqs.(7.169) and (7.170)

into Eq.(7.168) and noting that (Pnw - Pw) is the capillary pressure gives

114

1

8A B c

q LP P Prμ

π− = − (7.171)

Similarly, for the second capillary tube,

224

2

8A B c

q LP P Prμ

π− = − (7.172)

Equating Eqs.(7.171) and (7.172) and rearranging gives

( )1 2 2 14 41 2 2 1

8 8 1 12 cosc cL Lq q P P

r r r rμ μ σ θ

π π⎛ ⎞ ⎛ ⎞ ⎛ ⎞

− + = − = −⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠ ⎝ ⎠

(7.173)

where the Laplace equation has been used to replace the capillary pressures.

Assuming incompressible fluids,

1 2q q q+ = (7.174)

Page 889: +Peters Ekwere j. - Petrophysics

7-154

Eqs.(7.173) and (7.174) are two linear simultaneous equations in q1 in q2,

which can easily solved to obtain

4

2 2 11

4 41 2

8 1 12 cos

8 8

L qr r r

qL L

r r

μ σ θπ

μ μπ π

⎛ ⎞ ⎛ ⎞− −⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠=⎛ ⎞ ⎛ ⎞

+⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠

(7.175)

4

1 2 12

4 41 2

8 1 12 cos

8 8

L qr r r

qL L

r r

μ σ θπ

μ μπ π

⎛ ⎞ ⎛ ⎞+ −⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠=⎛ ⎞ ⎛ ⎞

+⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠

(7.176)

Diving Eq.(7.176) by (7.175) and simplifying gives

4 42 2

1 2 124

1 2

2 1

cos 1 14cos 1 1

4

r rqr L r rq

q rqL r r

π σ θμ

π σ θμ

⎛ ⎞ ⎛ ⎞+ −⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠=⎛ ⎞

− −⎜ ⎟⎝ ⎠

(7.177)

The interface velocities are given by

11 2

1

qvrπ

= (7.178)

22 2

2

qvrπ

= (7.179)

Let

2

1

rr

β⎛ ⎞

= ⎜ ⎟⎝ ⎠

(7.180)

Page 890: +Peters Ekwere j. - Petrophysics

7-155

31 cosvcapq LN

π σ θ= (7.181)

Substituting Eqs.(7.178) through (7.181) into (7.177) gives the ratio of the

interface velocities as

2

212

14 1

4 1 1

vcap

vcap

Nv

Nvβ

ββ β

⎛ ⎞+ −⎜ ⎟

⎝ ⎠=⎛ ⎞

− −⎜ ⎟⎝ ⎠

(7.182)

Eq.(7.182) can be used to determine the conditions under which the

non-wetting phase will be trapped in the smaller pore or in the larger pore. If

2

1

1vv

> (7.183)

the non-wetting phase will be trapped in the smaller pore. Substituting

Eq.(7.182) into (7.183) gives the condition for trapping in the smaller pore as

( )( )

2 14 1vcapN

β β

β

+>

+ (7.184)

If

2

1

1vv

< (7.185)

the non-wetting phase will be trapped in the larger pore. Substituting

Eq.(7.182) into (7.185) gives the condition for the non-wetting phase to be

trapped in the larger pore as

Page 891: +Peters Ekwere j. - Petrophysics

7-156

( )( )

2 14 1vcapN

β β

β

+<

+ (7.186)

The critical value of vcapN for trapping in either pore is given by

( )( )

2 14 1vcapcriticalN

β β

β

+=

+ (7.187)

For fixed values of r1, r2, σcosθ, μ and L, vcapN will depend on the rate q. If q is

low, the displacement will be dominated by capillary forces and the non-

wetting phase will be trapped in the larger pore resulting in a low

displacement efficiency. If q is high, the displacement will be dominated by

the viscous forces and the non-wetting phase will be trapped in the smaller

pore resulting in a high displacement efficiency. These observations are in

qualitative agreement with macroscopic observations in corefloods. Figure

7.74 shows the breakthrough oil recovery as a function of vμL obtained by

Rapoport and Leas (1953). The breakthrough oil recovery here is the oil

recovery at the time of water arrival at the outlet end of the core. As such, it is

a measure of the displacement efficiency. It should be observed that vμL is

directly proportional to the macroscopic version of vcapN . Clearly, the

displacement efficiency increases with an increase in vμL or vcapN in agreement

with the prediction of the pore doublet model.

Page 892: +Peters Ekwere j. - Petrophysics

7-157

Figure 7.74. Breakthrough oil recovery versus Rapaport and Leas scaling coefficient, vμL (Rapoport and Leas, 1953).

If

2

1

0vv

< (7.188)

the domination of the displacement by the capillary forces will be strong that

the interface in the larger pore will retract resulting in the trapping of the

non-wetting phase in the larger pore and low displacement efficiency.

Substituting Eq.(7.182) into (7.188) gives the condition for this to happen as

Page 893: +Peters Ekwere j. - Petrophysics

7-158

1 114vcapN

β⎛ ⎞

< −⎜ ⎟⎝ ⎠

(7.189)

It is interesting to compare vcapN with capN , which is the relevant

dimensionless number for diagnosing the problem of capillary end effect.

From Eq.(7.61),

1coscap

q LN A k

μσ θ φ

= (7.190)

A comparison of Eqs.(7.181) and (7.190) shows that both dimensionless

numbers give the ratio of viscous to capillary forces or vice versa. capN gives the

ratio of capillary to viscous forces at the macroscopic scale whereas vcapN gives

the ratio of the viscous to capillary forces at the pore scale. Once again, we

see the competition between viscous and capillary forces that pervades

immiscible displacements in porous media.

Chatzis and Dullien (1983) and Laidlaw and Wardlaw (1983) provide

more detailed theoretical analyses of the pore doublet model along with

experimental verifications.

7.14.2 Snap-Off Model of Capillary Trapping

When a non-wetting phase is forced through a pore constriction, it can

suffer from capillary instability and snap off (breakup) after exiting the

constriction. This phenomenon is controlled by the ratio of the pore body to

the pore throat size. Figure 7.75 shows oil being displaced in two pores, one

with a low aspect ratio and the other with a high aspect ratio. The aspect ratio

is define as

Page 894: +Peters Ekwere j. - Petrophysics

7-159

1

2

DAspect RatioD

= (7.191)

where 1D and 2D are the pore body diameter and the pore throat diameter,

respectively. In Figure 7.75a, the aspect ratio is low and the oil is displaced

through the pore without trapping. In Figure 7.75b, the aspect ratio is high

and the oil suffers capillary instability and snaps off at the pore throat and be

trapped.

Figure 7.75. Capillary trapping by snap-off mechanism in a single pore.(a) low

aspect ratio; (b) high aspect ratio (Chatzis et al., 1983).

Page 895: +Peters Ekwere j. - Petrophysics

7-160

Figure 7.76 shows a drop of non-wetting fluid passing through a pore

constriction is a uniform pack of spheres. The drop will become unstable and

breakup when the capillary pressure at the pore neck exceeds the capillary

pressure at the leading edge of the drop. The condition for snap-off is given by

1 1 2cn

n t f

Pr r r

σσ⎛ ⎞

= − >⎜ ⎟⎝ ⎠

(7.192)

Figure 7.76. Snap-off in a porous medium (Stegemeier, 1976).

Page 896: +Peters Ekwere j. - Petrophysics

7-161

Figure 7.77 shows a sequence of mercury injection and withdrawal into

a pore system. Figure 7.77A shows the capillary pressure hysteresis loops

whereas Figure 7.77B shows mercury trapping by snap-off corresponding to

the hysteresis loops.

7.14.3 Mobilization of Residual Non-Wetting Phase

Once a phase has been trapped, the pressure gradient required to

mobilize it can be significantly higher than can be generated under the

flooding conditions. Let us calculate the pressure gradient required to

mobilize a trapped oil blob in a waterflood as shown in Figure 7.78. For the

blob to pass through the pore throat, the pressure drop across the leading

edge must exceed the entry pressure or the displacement pressure of the pore

throat. Thus, the condition for the blob to pass through the pore throat is

given by

'

1

2 cosB BP P

rσ θ

− ≥ (7.193)

From Laplace equation,

'

2

2 cosA AP P

rσ θ

− = (7.194)

Subtracting Eq.(7.194) from (7.193) gives

( ) ( )' '

1 2

1 12 cosB B A AP P P Pr r

σ θ⎛ ⎞

− − − ≥ −⎜ ⎟⎝ ⎠

(7.195)

Page 897: +Peters Ekwere j. - Petrophysics

7-162

Figure 7.77. Snap-off in mercury injection-withdrawal experiment. A: capillary pressure scanning curves; B: corresponding mercury trapping by snap-off

(Stegemeier, 1976).

Page 898: +Peters Ekwere j. - Petrophysics

7-163

or

( ) ( )' '

1 2

1 12 cosA B B AP P P Pr r

σ θ⎛ ⎞

− + − ≥ −⎜ ⎟⎝ ⎠

(7.196)

Because ' 'B AP P= , Eq.(7.196) becomes

( )1 2

1 12 cosA BP Pr r

σ θ⎛ ⎞

− ≥ −⎜ ⎟⎝ ⎠

(7.197)

The pressure gradient required to mobilize the blob is given by

( )1 2

2 cos 1 1A BP PL L r r

σ θ− ⎛ ⎞≥ −⎜ ⎟

⎝ ⎠ (7.198)

Figure 7.78. Trapped oil blob.

Let us estimate the pressure gradient required to mobilize a trapped oil

droplet in a normal waterflood in a reservoir rock using typical values of the

relevant parameters. Let

1 10r mμ=

Page 899: +Peters Ekwere j. - Petrophysics

7-164

2 50r mμ=

50L mμ=

30 /dynes cmσ =

0θ =

500 wk md=

1 w cpμ =

1 /wv ft day=

The pressure gradient required to mobilize the droplet is given by

( )( )( )( ) ( )( ) ( )( )

( )( )66 6 6

2 30 1 1 1 14.696 30.48 /1.0133 1050 10 100 10 10 100 50 10 100

A BP P psi ftL xx x x− − −

⎡ ⎤− ⎛ ⎞⎢ ⎥≥ − ⎜ ⎟⎝ ⎠⎢ ⎥⎣ ⎦

or

4243.73 /A BP P psi ftL−

This is an enormous pressure gradient requirement. The pressure gradient

generated by the waterflood is given by Darcy's law as

( )( )( )( )( )

1 10.32 /

0.001127 5.615 500P psi ftL

Δ= =

We see that the pressure gradient generated by the waterflood is not sufficient

to mobilize the oil droplet. Therefore, it will remain trapped.

Page 900: +Peters Ekwere j. - Petrophysics

7-165

7.14.4 Oil Migration

Figure 7.79 shows an upward migrating oil bubble from a source rock

into a reservoir initially fully saturated with water. The migrating bubble has

encountered a restriction at a pore throat of radius rH. Obviously, in order for

migration to continue, the leading end of the bubble (A) must squeeze through

the pore throat. Assuming that ends A and B of the bubble are hemispherical

with B having a radius rB, one can calculate the length of the oil blob required

for the blob to pass through the restriction and continue its upward

migration. To pass through the restriction, the capillary pressure at the

leading edge of the blob must exceed the displacement pressure of the

restriction. Therefore, the condition for upward migration through the

restriction is

( ) 2oA wA

H

P Prσ

− > (7.199)

Laplace equation gives

( ) 2oB wB

B

P Prσ

− = (7.200)

From hydrostatics through the water,

wB wA wP P ghρ= + (7.201)

Page 901: +Peters Ekwere j. - Petrophysics

7-166

Figure 7.79. Migrating oil filament.

Page 902: +Peters Ekwere j. - Petrophysics

7-167

From hydrostatics through the oil,

oB oA oP P ghρ= + (7.202)

Subtracting Eq.(7.201) from (7.202) gives

( ) ( ) ( )oB wB oA wA w oP P P P ghρ ρ− = − − − (7.203)

Substituting Eq.(7.200) into (7.203) and rearranging gives

( ) ( ) 2oA wA w o

B

P P ghrσρ ρ− = − + (7.204)

Substituting Eq.(7.204) into (7.199) gives the condition for upward migration

as

( ) 2 2w o

B H

ghr rσ σρ ρ− + > (7.205)

or

( )

2 1 1

w o H B

hg r r

σρ ρ

⎛ ⎞> −⎜ ⎟− ⎝ ⎠

(7.206)

Let

10Hr mμ=

50Br mμ=

30 /dynes cmσ =

ρw = 1.00 g/cm3

Page 903: +Peters Ekwere j. - Petrophysics

7-168

ρo = 0.70 g/cm3

The minimum length of the blob to continue its upward migration is given by

( )( )( )( ) ( )( ) ( )( )6 6

2 30 1 1 163.1 1 0.70 981 10 10 100 50 10 100

h cmx x− −

⎡ ⎤⎢ ⎥> − =

− ⎢ ⎥⎣ ⎦

7.15 EFFECTS OF WETTABILITY AND INTERFACIAL TENSION ON CAPILLARY PRESSURE CURVES

The effects of wettability and interfacial tension on capillary pressure

curves can easily be deduced from Laplace equation. The effect of wettability

on capillary pressure curve is shown qualitatively in Figure 7.80. As the

degree of preferential wettability is reduced, the capillary pressure curve will

decrease. The effect of interfacial tension on capillary pressure is shown

qualitatively in Figure 7.81. Capillary pressure decreases as the interfacial

tension of the fluids decrease.

Page 904: +Peters Ekwere j. - Petrophysics

7-169

Figure 7.80. Effect of wettability on capillary pressure curve.

Page 905: +Peters Ekwere j. - Petrophysics

7-170

Figure 7.81. Effect of interfacial tension on capillary pressure curve.

NOMENCLATURE

A = cross sectional area in the flow direction

Bo = oil formation volume factor

Bw = water formation volume factor

do = depth of free water level below water oil contact

fw = fractional flow of wetting phase

fw = fractional flow of water

fnw = fractional flow of non-wetting phase

fnw2 = fractional flow of non-wetting phase at the outlet end of porous medium

Page 906: +Peters Ekwere j. - Petrophysics

7-171

fo = fractional flow of oil

f(R) = probability density function for pore volume distribution

Fw = approximate fractional flow of wetting phase

1F = lithology factor

g = gravitational acceleration

h = height above water oil contact

J = Leverett J-function

k = absolute permeability of the medium

ko = effective permeability to oil

kw = effective permeability to wetting phase

knw = effective permeability to non-wetting phase

kwr = end-point relative permeability to wetting phase

kg = effective permeability to gas

kro = relative permeability to oil

krw = relative permeability to water

krg = relative permeability to gas

krw = relative permeability to wetting phase

krnw= relative permeability to non-wetting phase

knwr= end-point relative permeability to non-wetting phase

L = length

M = mobility ratio

ME = end-point mobility ratio

N = centrifuge speed in revolutions per minute

NpD = dimensionless cumulative production

Ncap = dimensionless capillary to viscous force ratio

P = pressure

Pc = capillary pressure

Pc1 = capillary pressure at the inlet end of core in a centrifuge

Pd = displacement pressure

Pe = displacement pressure for Brooks-Corey model

Page 907: +Peters Ekwere j. - Petrophysics

7-172

Pg = pressure in the gas phase

Pnw= pressure in the non-wetting phase

Po = pressure in the oil phase

Pw = pressure in the water phase

Pw = pressure in the wetting phase

q = total volumetric injection rate

qo = volumetric flow rate of oil

qg = volumetric flow rate of gas

qnw= volumetric flow rate of non-wetting phase

qw = volumetric flow rate of water

qw = volumetric flow rate of wetting phase

r = radius of capillary tube

mr = mean radius of curvature of an interface

1 2,r r = radii of curvature of an interface

1 2,r r = distance of inlet end and outlet end of core from the center of rotation in a centrifuge

R = pore throat radius

Se = effective wetting phase saturation

Sg = gas saturation

So = oil saturation

Sor = residual oil saturation

Sw = water saturation

Sw = wetting phase saturation

Swirr = irreducible wetting phase saturation

Swro = wetting phase saturation at which the imbibition capillary pressure is zero

Sw1 = wetting phase saturation at the inlet end of porous medium in a centrifuge

Snw= non-wetting phase saturation

Snwr = residual non-wetting phase saturation

Page 908: +Peters Ekwere j. - Petrophysics

7-173

Swav = average wetting phase saturation

Swav = average water saturation

Swf = frontal saturation

*wS = normalized wetting phase saturation

t = time

Dt = dimensionless time

v = flux vector, Darcy velocity vector

vw = Darcy velocity for the wetting phase

vnw = Darcy velocity for the non-wetting phase

x = distance in the direction of flow

Dx = dimensionless distance

z = height above free water level

wρ = density of water

wρ = density of wetting phase

nwρ = density of non-wetting phase

σ = interfacial tension

θ = contact angle

λ = pore size distribution index

μ = viscosity

μg = gas viscosity

μο = oil viscosity

μw = water viscosity

μw = wetting phase viscosity

μnw= non-wetting phase viscosity

δ(R) = probability density function for pore throat size distribution based on bundle of capillary tubes model of porous medium

φ = porosity, fraction

τ = tortuosity

γ = liquid specific gravity

Page 909: +Peters Ekwere j. - Petrophysics

7-174

ω = angular velocity of centrifuge

ΔP = pressure drop

ΔPw = pressure drop in the wetting phase

ΔPnw = pressure drop in the non-wetting phase

Γ = pore structure

REFERENCES AND SUGGESTED READINGS

Abrams, A. : "The Influence of Fluid Viscosity, Interfacial Tension, and Flow Velocity on Residual Oil Saturation Left by Waterflood," Soc. Pet. Eng. Jour. (Oct., 1975) 437-447.

Adamson, A.W. and Gast, A.P.: Physical Chemistry of Surfaces, Sixth Edition, John Wiley and Sons, Inc., New York, 1997.

Afzal, N.: Investigation of Heterogeneities and Fluid Displacements in Porous Media Using CT, MS Thesis, The University of Texas at Austin, May 1992.

Akima, H.: "A New Method of Interpolation and Smooth Curve Fitting Based on Local Procedures," J. ACM, Vol.17, No. 4 (Oct. 1970) 589-602.

Anderson, W.G. : “Wettability Literature Survey - Part 4: Effects of Wettability on Capillary Pressure,” J. Pet. Tech. (October 1987) 1283-1300.

Anderson, W.G. : “Wettability Literature Survey - Part 6: The Effects of Wettability on Waterflooding,” J. Pet. Tech. (December 1987) 1605-1622.

Archer, J.S. and Wall, C.G. : Petroleum Engineering, Graham & Trotman, London, England, 1986.

Arriola, A., Willhite, G.P. and Green, D.W. : "Trapping of Oil Drops in a Noncircular Pore Throat and Mobilization Upon Contact With a Surfactant," SPEJ (Feb. 1983) 99-114.

Bear, J. : Dynamics of Fluids in Porous Media, Elsevier, New York, 1972.

Benner, F.C. and Bartell, F.E. : “The Effect of Polar Impurities Upon Capillary and Surface Phenomena in Petroleum Production,” API Drilling and Production Practice (1941) 341-348.

Bethel, F.T. and Calhoun, J.C. : "Capillary Desaturation in Unconsolidated Beads," Trans. AIME (1953) Vol. 198, 197-202.

Brooks, R.H. and Corey, A.T. : “Properties of Porous Media Affecting Fluid Flow,” Jour. Irrigation and Drainage Div., Proc. Amer. Soc. Of Civil Engr. (June, 1966) 61-88.

Page 910: +Peters Ekwere j. - Petrophysics

7-175

Brown, H.W : “Capillary Pressure Investigations,” Trans., AIME (1951) 192, 67-74.

Buckley, S.E. and Leverett, M.C. : “Mechanism of Fluid Displacement in Sands,” J. Pet. Tech. (May 1941) 107-116.

Burdine, N.T., Gournay, L.S. and Reichertz, P.P. : “Pore Size Distribution of Petroleum Reservoir Rocks,” Trans., AIME (1950) 198, 195-204.

Campbell, G.S. : "A Simple Method for Determining Unsaturated Conductivity from Moisture Retention Data," Soil Sci., Vol. 117, No. 6 (1974) 311-314.

Campbell, G.S. and Gee, G.W. : "Water Potential: Miscellaneous Methods," in American Society of Agronomy, Madison, WI (1986) 619-633.

Charbeneau, R.J. and Daniel, D.E. : "Contaminant Transport in Unsaturated Flow," Chapter 15, in Handbook of Hydrology, D.R. Maidment (Ed.), McGraw-Hill Inc., New York, 1993.

Chatzis, I., Morrow, N.R. and Lim, H.T. : “Magnitude and Detailed Structure of Residual Oil Saturation,” Soc. Pet. Eng. J. (April 1983) 311-326.

Chatzis, I. and Dullien, F.A.L. : “Dynamic Immiscible Displacement Mechanisms in Pore Doublets: Theory versus Experiment,” Jour. Colloid and Interface Science , vol. 91, No. 1 (Jan. 1983) 199-222.

Chatzis, I. and Morrow, N.R. : “Correlation of Capillary Number Relationships for Sandstones,” SPEJ (Oct. 1984) 555-562.

Collins, R.E. and Cooke, C.E.: "Fundamental Basis for the Contact Angle and Capillary Pressure," Trans. Faraday Soc., 55 (1959) 1602-1606.

Collins, R.E. : Flow of Fluids Through Porous Materials, Research & Engineering Consultants Inc., 1990.

Core Laboratories Inc. : Special Core Analysis, 1976.

Corey, A.T. : Mechanics of Heterogeneous Fluids in Porous Media, Water Resources Publications, Fort Collins, Colorado, 1977.

Cosse, R. : Basics of Reservoir Engineering, Editions Technip, Paris, 1993.

Craig, F.F., Jr. : The Reservoir Engineering Aspects of Waterflooding, SPE Monograph Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1971.

Cuiec, L.E. : “Study of Problems Related to the Restoration of Natural State of Core Samples,” J. Canadian Pet. Tech (Oct. - Dec. 1977) 68-80.

de Gennes, P.G. and Quere, D. : Capillarity and Wetting Phenomena, Springer Science and Business Media, Inc., 2004.

Donaldson, E.C., Kendall, R.F., Pavelka, E.A. and Crocker, M.E. : “Equipment and Procedures for Fluid Flow and Wettability Tests of Geological

Page 911: +Peters Ekwere j. - Petrophysics

7-176

Materials,” DOE/BETC/IC-79/5, Nat. Tech. Info. Sv, Springfield, VA 2216, 1980.

Douglas, J., Jr., Blair, P.M. and Wagner, R.J. : “Calculation of Linear Waterflood Behavior Including the Effects of Capillary Pressure,” Trans., AIME (1958) 213, 96-102.

Dullien, F.A.L. : “Effects of Pore Structure on Capillary and Flow Phenomena in Sandstones,” J. Cnd. Pet. Tech. (July-Sept. 1975) 48-55.

Dullien, F.A.L. : Porous Media - Fluid Transport and Pore Structure, Academic Press, New York, 1979.

Forbes, P., Chardaire-Riviere, C., Deflandre, F. and Flcury, M. : “Local Saturation Measurement While Centrifuging (MWC) for Improving Centrifuge Capillary-Pressure Curve Determination,” SCA Conference Paper Number 9215 (1992).

Gardescu, I.I. : “Behavior of Gas Bubbles in Capillary Spaces,” Trans., AIME (1930) 351-370.

Gharbi, R. and Peters, E.J. : “Scaling Coreflood Experiments to Heterogeneous Reservoirs,” Journal of Petroleum Science and Engineering, 10, (1993) 83-95.

Garnes, J.M., Mathisen, A.M., Scheie, A. and Skauge, A. : "Capillary Number Relations for Some North Sea Reservoir Sandstones," SPE/DOE 20264, Presented at the SPE/DOE Seventh Symposium on Enhanced Oil Recovery, Tulsa, April 22-25, 1990.

Graue, A., Kolltvelt, K., Lien, J.R. and Skauge, A. : "Imaging Fluid Saturation Development in Long-Core Flood Displacements," SPE Formation Evaluation (December 1990) 406-412.

Hassler, G.L. and Brunner, E. : “Measurement of Capillary Pressures in Small Core Samples,” Trans., AIME (1945) 160, 114-123.

Haughey, D.P. and Beveridge, G.S.G. : "Structural Properties of Packed Beds - A Review,” Can. J. Chem. Eng., Vol. 47 (April 1969) 130-149.

Hermansen, H., Eliassen, O., Guo, Y. and Skjaeveland, S.M. : "Capillary Pressure From Centrifuge - A New Direct Method," Advance in Core Evaluation II, EuroCAS, London, May 20-22, 1991.

Jennings, H.Y. : “Surface Properties of Natural and Synthetic Porous Media,” Producers Monthly (March 1957) 20-24.

Jennings, H.J. : “How to Handle and Process Soft and Unconsolidated Cores,” World Oil (June 1965) 116-119.

Keelan, D.K. : "A Critical Review of Core Analysis Techniques” The Jour. Can. Pet. Tech. (April-June 1972) 42-55.

Page 912: +Peters Ekwere j. - Petrophysics

7-177

Killins, C.R., Nielsen, R.F. and Calhoun, J.C., Jr.: “Capillary Desaturation and Imbibition in Rocks,” Producers Monthly (February 1953) 18, No. 2, 30-39.

Klute, A. : "Water Retention: Laboratory Methods," Methods of Soil Analysis, Part 1, A. Klute (Ed.), American Society of Agronomy, Madison, WI (1986) 635-686.

Klute, A. : "Hydraulic Conductivity and Diffusivity: Laboratory Methods," Methods of Soil Analysis, Part 1, A. Klute (Ed.), American Society of Agronomy, Madison, WI (1986) 687-734.

Kyte, J.R. and Rapoport, L.A. : “Linear Waterflood Behavior and End Effects in Water-Wet Porous Media,” Trans., AIME (1958) 213, 423-426.

Laidlaw, W.G. and Wardlaw, N.C. : "A Theoretical and Experimental Investigation of Trapping in Pore Doublets," Canadian J. Chemical Engineering, Vol. 61 (October 1983) 719-727.

Lake, L.W. : Enhanced Oil Recovery, Prentice Hall, Englewood Cliffs, New Jersey, 1989.

Land, C.S. : "Calculation of Imbibition Relative Permeability for Two- and Three-Phase Flow From Rock Properties,” SPEJ (June 1968) 149-156.

Leij, F.J., Russell, W.B. and Lesch, S.M. : “Closed-Form Expressions for Water Retention and Conductivity Data,” Ground Water Vol. 35, No. 5 (1997) 848-858.

Leva, M., Weintraub, M., Grummer, M. Pollchick, M. and Storch, H.H. : "Fluid Flow Through Packed and Packed and Fluidized Systems," US Bureau of Mines Bull. No. 504, 1951.

Leverett, M.C. : “Flow of Oil-Water Mixtures through Unconsolidated Sands,” Trans., AIME (1939) 140, xxx-xxx.

Leverett, M.C. : “Capillary Behavior in Porous Solids,” Trans., AIME (1941) 142, 152-169.

Majors, P.D., Li, P. and Peters, E.J. : "NMR Imaging of Immiscible Displacements in Porous Media,” Society of Petroleum Engineers Formation Evaluation (September 1997) 164-169.

Marle, C.M. : Multiphase Flow in Porous Media, Gulf Publishing Company, Houston, Texas, 1981.

Melrose, J.C. : “Role of Capillary Forces in Determining Microscopic Displacement Efficiency for Oil Recovery by Waterflooding,” J. Cnd Pet. Tech. (Oct.-Dec. 1974) 54-62.

Melrose, J.C. : “Interpretation of Centrifuge Capillary Pressure Data,” The Log Analyst (January-February 1988) 40-47.

Page 913: +Peters Ekwere j. - Petrophysics

7-178

Melrose, J.C. : "Interfacial Phenomena as Related to Oil Recovery Mechanisms," Cnd J. Chem. Eng., Vol. 48 (Dec. 1970) 638-644.

Mohanty, K.K., Davis, H.T. and Scriven, L.E. : "Physics of Oil Entrapment in Water-Wet Rock," SPE 9406 Presented at the 55th Annual Technical Conference and Exhibition of the Society of Petroleum Engineers of AIME, Dallas, September 21-24, 1980.

Monicard, R.P. : Properties of Reservoir Rocks, Gulf Publishing Company, Houston, TX, 1980.

Morrow, N.R. : "Capillary Equilibrium in Porous Media," Soc. Pet. Eng. J. (March 1965) 15-24.

Morrow, N.R. : "Capillary Pressure Correlations For Uniformly Wetted Porous Media," J. Cnd Pet. Tech. (Oct.-Dec. 1976)) 49-69.

Morrow, N.R. : "Physics and Thermodynamics of Capillary," Ind. Eng. Chem. Vol. 62 (1970) 32.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 2 - Interfacial Phenomena and Oil Recovery: Wettability,” World Oil (March 1981) 77-83.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 3 - Interfacial Phenomena and Oil Recovery: Capillarity,” World Oil (May 1981) 149-158.

Perkins, F.M. : "An Investigation of the Role of Capillary Forces in Laboratory Water Floods," Trans. AIME (1957) Vol. 210, 409-411.

Purcell, W.R. : “Capillary Pressures - Their Measurement Using Mercury and the Calculation of Permeability There From,” Trans., AIME (1949) 186, 39-48.

Purcell, W.R. : “Interpretation of Capillary Pressure Data” Trans., AIME (1950) 189, 369-371.

Rajan, R.R. : “Theoretically Correct Analytical Solution for Calculating Capillary Pressure-Saturation from Centrifuge Experiments,” SPWLA Trans., 27th Annual Logging Symposium, Paper J, Houston, Tx, June 9-13, 1986.

Rapoport, L.A. and Leas, W.J. : “Properties of Linear Waterfloods,” Trans., AIME (1953) 198, 139-148.

Ritter, L.C. and Drake, R.L. : “Pore Size Distribution in Porous Materials,” Ind. Eng. Chem., An. Ed. (1945) 17, 782-786.

Ritter, L.C. and Drake, R.L. : “Macropore-Size Distributions in Some Typical Porous Substances,” Ind. Eng. Chem., An. Ed. (1945) 17, 787-791.

Rose, H.E. : “ An Investigation into the the Laws of Flow of Fluids Through Granular Material,” Proc. Inst. Mech. Eng. (1945) 153, 141-148.

Page 914: +Peters Ekwere j. - Petrophysics

7-179

Rowlinson, J.S. and Widom, B. : Molecular Theory of Capillarity, Dover Publications, Inc., Mineola, New York, 1982.

Scheidegger, A.E. : The Physics of Flow Through Porous Media, University of Toronto Press, Toronto, 1960.

Slobod, R.L., Chambers, A. and Prehn, W.L., Jr. : “Use of Centrifuge for Determining Connate Water, Residual Oil, and Capillary Pressure Curves of Small Core Samples,” Trans., AIME (1951) 192, 127-134.

Stegemeier, G.L. : “Mechanisms of Entrapment and Mobilization of Oil in Porous Media,” Presented at the 81st National Meeting of AIChE, Kansas City, April 12-14, 1976.

Stegemeier, G.L. : “Relationship of Trapped Oil Saturation to Petrophysical Properties of Porous Media,” SPE 4754 Presented at the Improved Oil Recovery Symposium of the Society of Petroleum Engineers of AIME , Tulsa, April 22-24, 1974.

Taber, J.J. : "Dynamic and Static Forces Required to Remove a Discontinuous Oil Phase From Porous Media Containing Both Oil and Water," SPEJ (March 1969) 3-12.

Tiab, D. and Donaldson, E.C. : Petrophysics, Second Edition, Elsevier, New York, 2004.

Topp, G.C. and Miller, E.E. : "Hysterisis Moisture Characteristics and Hydraulic Conductivities for Glass-Bead Media," Soil Sci. Soc. Am. Proc., Vol. 30 (1966) 156-162.

Topp, G.C., Klute, A. and Peters, D.B. : "Comparison of Water Content-Pressure Head Data Obtained by Equilibrium, Steady State, and Unsteady State Methods," Soil Sci. Soc. Am. Proc., Vol. 31 (1967) 312-314.

Topp, G.C. and Davis, J.L. : "Measurement of Soil Water Using Time-Domain Reflectometry (TDR): Field Evaluation," Soil Sci. Soc. Am. Proc., Vol. 49 (1985) 19-24.

Topp, G.C., Davis, J.L. and Annan, A.P. : "Electromagnetic Determination of Soil Water Content: Measurements In Coaxial Transmission Lines," Water Resources Research, Vol. 16 (19805) 574-582.

van Genuchten, M.T. : "A Closed-Form Equation for Predicting the Hydraulic Conductivity of Unsaturated Soils," Soil Sci. Soc. Am. J., Vol. 44 (1980) 892-898.

Welge, H.J. and Bruce, W.A. : “A Restored-State Method for Determination of Oil in Place and Connate Water,” API Drilling and Production Practice (1947) 161-165.

Willhite, G. P. : Waterflooding, SPE Textbook Series Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1986.

Page 915: +Peters Ekwere j. - Petrophysics

7-180

Winsauer, W.O., Shearin, H.M., Jr., Masson, P.H. and Williams, M. : "Resistivity of Brine-Saturated Sands in Relation to Pore Geometry," AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 253-277.

Wyllie, M.R.J. and Rose, W.D. : "Some Theoretical Considerations Related to the Quantitative Evaluation of the Physical Characteristics of Reservoir Rock from Electrical Log Data,” Trans., AIME (1950) 189, 105-118.

Wyllie, M.R.J. and Spangler, M.B. : "Application of Electrical Resistivity Measurements to Problems of Fluid Flow in Porous Media,” AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 359-403.

Page 916: +Peters Ekwere j. - Petrophysics

7-181

Page 917: +Peters Ekwere j. - Petrophysics

8-1

CHAPTER 8

RELATIVE PERMEABILITY

8.1 DEFINITION OF RELATIVE PERMEABILITY

In a petroleum reservoir, it is possible for two or three fluids to flow

simultaneously. Examples are (a) the flow of gas and water in a gas reservoir,

(b) the flow of oil and water in an oil reservoir, (c) the flow of oil and gas in an

oil reservoir and (d) the flow of oil, water and gas in an oil reservoir. In

multiphase flow situations, the absolute permeability of the porous medium is

no longer sufficient to calculate the flow rate of each fluid type or to calculate

the total flow rate of all the fluids.

In order to make quantitative predictions for multiphase flow, we need

to know the permeability to each fluid in the presence of the other fluids in

the rock. The permeability of one fluid in the presence of the other immiscible

fluids is known as the effective permeability to that fluid. To calculate the

flow rate of each fluid in multiphase flow, we extend Darcy’s Law to

multiphase flow. For example, for simultaneous flow of oil, water and gas in

an inclined linear system, Darcy’s Law is applied to each phase as follows:

sino oo o

o

k A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.1)

sinw ww w

w

k A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.2)

Page 918: +Peters Ekwere j. - Petrophysics

8-2

sing gg g

g

k A Pq g

xρ α

μ∂⎛ ⎞

= − +⎜ ⎟∂⎝ ⎠ (8.3)

where α is the angle of inclination with the horizontal. Eqs.(8.1) to (8.3) show

that using the concept of an effective permeability, Darcy’s Law is applied to

each phase as if the other phases did not exist. Capillary equilibrium between

the phases gives

( )/o w c ow wP P P S− = (8.4)

( )/g o c go oP P P S− = (8.5)

/ / /c ow c go c gwP P P+ = (8.6)

It is often more convenient to work with a dimensionless effective

permeability known as the relative permeability obtained by dividing the

effective permeability by a base permeability such as the absolute

permeability of the porous medium. Thus, for the three phase example, using

the absolute permeability of the porous medium as the base permeability, the

relative permeabilities to oil, water and gas are given by

oro

kkk

= (8.7)

wrw

kkk

= (8.8)

grg

kk

k= (8.9)

In terms of relative permeabilities, Eqs.(8.1) through (8.3) become

sinro oo o

o

kk A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.10)

Page 919: +Peters Ekwere j. - Petrophysics

8-3

sinrw ww w

w

kk A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.11)

sinrg gg g

g

kk A Pq g

xρ α

μ∂⎛ ⎞

= − +⎜ ⎟∂⎝ ⎠ (8.12)

Sometimes, the effective permeability to the non-wetting phase at the

irreducible wetting phase saturation is used as the base permeability for

defining the relative permeability. In this case, the end point relative

permeability to the non-wetting phase will be 1.0. As the base pressure

appears in Darcy's law as shown in Eqs.(8.10) to (8.12), it is necessary to

ascertain the base permeability used to define a relative permeability curve

before such a curve is used in performance calculations. Failure to do so will

lead to wrong results.

Figure 8.1 shows typical imbibition relative permeability curves for a

two-phase system. The following observations can be made about the key

features of the relative permeability curves.

1. The relative permeability curves are nonlinear functions of fluid

saturation.

2. The sum of the relative permeabilities at each saturation is always less

than 1.0.

3. There is an irreducible wetting phase saturation (Swirr) at which the

relative permeability to the wetting phase is zero and the relative

permeability to the non-wetting phase attains a maximum end point

value (knwr)

4. There is a residual non-wetting phase saturation (Snwr) at which the

relative permeability to the non-wetting phase is zero and the relative

Page 920: +Peters Ekwere j. - Petrophysics

8-4

permeability to the wetting phase attains a maximum end point value

(kwr).

Figure 8.1. Typical imbibition relative permeability curves.

5. The relative permeability curves are not defined in the saturation ranges

given by ( )1 1nwr wS S− < < and 0 w wirrS S< < .

6. Two phase flow occurs over the saturation range ( )1wirr w nwrS S S< < − .

Imbibition relative permeability curves typically are used to perform the

following reservoir performance calculations:

• Waterflood calculations in a water wet reservoir in which water

displaces oil and/or gas.

Page 921: +Peters Ekwere j. - Petrophysics

8-5

• Natural water influx calculations in a water wet reservoir in which water

displaces oil and/or gas.

• Oil displaces gas, which occurs when oil is forced into a gas cap.

Figure 8.2 shows typical drainage relative permeability curves with

features that are very similar to the imbibition curves of Figure 8.1. The

obvious differences between the drainage and imbibition curves are that the

drainage curve for the wetting phase starts at a wetting phase saturation of

1.0 and that of the non-wetting phase is zero at the wetting phase saturation

of 1.0. This is because at the start of the drainage relative permeability

measurements, the porous medium was fully saturated with the wetting

phase. The permeability of the wetting phase is then equal to the absolute

permeability of the porous medium. Of course, at the start of the experiment,

there was no non-wetting phase in the medium. Therefore, the relative

permeability to the non-wetting phase must be zero. This is the only occasion

in which the sum of relative permeabilities is equal to 1.0 because there was

only one phase present. As will be shown later, because of capillary pressure

hysteresis, the drainage and imbibition relative permeability curves will be

different.

Drainage relative permeability curves typically are used to perform the

following reservoir performance calculations:

• Solution gas drive calculations in which gas displaces oil.

• Gravity drainage calculations in which gas replaces drained oil.

• Gas drive calculations in which gas displaces oil and/or water.

• Oil or gas displacing water in tertiary recovery processes.

Page 922: +Peters Ekwere j. - Petrophysics

8-6

Figure 8.2. Typical drainage relative permeability curves.

8.2 LABORATORY MEASUREMENT OF TWO-PHASE RELATIVE PERMEABILITIES BY THE STEADY STATE METHOD

The most straight-forward laboratory measurement technique for

relative permeabilities is the steady state method. For imbibition relative

permeability measurement, the test starts with the core initially saturated

with an irreducible wetting phase saturation (Swirr) and a non-wetting phase

saturation of (1-Swirr). Then a mixture of the two phases is injected into the

inlet face of the core at a fixed ratio of nw

w

qq

⎛ ⎞⎜ ⎟⎝ ⎠

until steady state is achieved.

Steady state is achieved when the pressure drop across the core no longer

changes with time and the ratio of the produced fluids is the same and the

ratio of the injected fluids. The steady state pressure drop across the core and

Page 923: +Peters Ekwere j. - Petrophysics

8-7

the injection rate of each phase are measured. The relative permeabilities are

calculated with the integrated forms of Darcy's law for two phase flow as

shown later. The saturations are usually calculated by material balance.

At steady state, the continuity equations for the wetting and non-

wetting phases for horizontal flow are

0wvx

∂=

∂ (8.13)

0nwvx

∂=

∂ (8.14)

Darcy's law applied to each phase gives

a constantw ww

w

k Pvxμ

∂= − =

∂ (8.15)

a constantnw nwnw

nw

k Pvxμ

∂= − =

∂ (8.16)

From capillary equilibrium,

( )nw w c wP P P S− = (8.17)

where Pc(Sw) in this case is the imbibition capillary pressure. If the

saturations, Sw and Snw, are uniform throughout the porous medium, then

kw, knw and Pc are independent of x. Then, Eqs.(8.15) and (8.16) can be

integrated to give

w w w ww

w w

v L q LkP A P

μ μ= =

Δ Δ (8.18)

nw nw nw nwnw

nw nw

v L q LkP A P

μ μ= =

Δ Δ (8.19)

Since ( )c wP S is uniform, wPΔ and nwPΔ are equal and the pressure drop across

the core can be measured in either phase and used to calculate the effective

Page 924: +Peters Ekwere j. - Petrophysics

8-8

permeabilities with Eqs.(8.18) and (8.19). The steady state saturation

distribution in the core can be calculated with Eq.(7.83), which is reproduced

here for convenience:

1w

ww

Dcap rnw

w

fFdS

dJdx N kdS

⎛ ⎞−⎜ ⎟

⎝ ⎠= (7.83)

with a specified inlet boundary condition. Of course, to do so, rwk , rnwk and

( )c wP S must be known.

If the saturation distribution in the core is not uniform because of

capillary end effect, then Eqs.(8.18) and (8.19) are not valid and cannot be

used to calculate the effective permeabilities. As there are no alternative

equations to use, the relative permeability experiment will be a failure.

Therefore, in the steady state experiment, the total injection rate ( )w nwq q q= +

should be sufficiently high to minimize capillary end effect as outlined in

Chapter 7.

Figure 8.3 shows an apparatus that can be used for the steady state

experiment. A typical sequence of steps for obtaining the imbibition relative

permeability curves might be as follows:

1. Install the clean, dry core sample in the Hassler apparatus as shown in

Figure 8.3. Evacuate the core and saturate with the wetting phase.

Determine the absolute permeability of the core by wetting phase flow.

2. Displace the wetting phase with the non-wetting phase until no more

wetting phase flows from the core. Calculate the irreducible wetting

phase saturation and the initial non-wetting phase saturation.

Measure the steady state pressure drop and the non-wetting phase

Page 925: +Peters Ekwere j. - Petrophysics

8-9

injection rate and calculate the relative permeability to the non-wetting

phase at the irreducible wetting phase saturation by use of Eq.(8.19) as

nw nwrnw

nw

q LkkA Pμ

(8.20)

3. Inject a mixture of the wetting and non-wetting phases at rates qnw and

qw such that the ratio, qw/qnw, is very much less than 1 until steady

state is achieved. Steady state is achieved when the injected and

produced qw/qnw ratios are equal and the pressure drop no longer

changes with time.

4. Measure the pressure drop and calculate the wetting phase saturation

by material balance. Calculate the relative permeabilities to the non-

wetting and wetting phases at the latest wetting phase saturation using

Eq.(8.20) and (8.18) as

w wrw

w

q LkkA Pμ

(8.21)

5. Increase the ratio qw/qnw and repeat steps 3 and 4 to calculate the

relative permeabilities at higher and higher wetting phase saturations.

6. Finally, inject only the wetting phase until no more non-wetting phase

flows from the core. Calculate the residual non-wetting phase

saturation. Measure the steady state pressure drop and the wetting

phase injection rate and calculate the relative permeability to the

wetting phase at residual non-wetting phase saturation. This completes

the relative permeability measurements.

In the steady state relative permeability experiment, it is necessary to

minimize capillary end effect. This can be accomplished by injecting at a

sufficiently high total rate or by other means as discussed by Richardson et

al. (1952). Figures 8.4 and 8.5 show the pressure profiles in the gas and oil

Page 926: +Peters Ekwere j. - Petrophysics

8-10

Figure 8.3. Hassler’s apparatus for relative permeability measurement (Osoba

et al., 1951).

phases for a gas-oil steady state relative permeability experiment conducted at

two rates by Richardson et al. (1952). At the lower injection rate (Figure 8.4),

capillary end effect is apparent whereas at the higher rate (Figure 8.5), there

is little or no capillary end effect. Note also, that when the capillary end effect

has been eliminated, the pressure drop in each phase is the same. Therefore,

the pressure drop measured in either phase is sufficient for calculating both

relative permeabilities.

Page 927: +Peters Ekwere j. - Petrophysics

8-11

The various steady state methods such as the Penn State method,

single core dynamic method, dispersed feed method, Hafford method and

Hassler method differ primarily in the techniques used to minimize or

eliminate capillary end effect (Richardson et al., 1952). When capillary end

effect has been eliminated, all the steady state methods give the same results

as shown in Figures 8.6 and 8.7.

Figure 8.4. Steady state oil and gas pressure profiles at a relatively low injection rate (Richardson et al., 1952).

Page 928: +Peters Ekwere j. - Petrophysics

8-12

Figure 8.5. Steady state oil and gas pressure profiles at a relatively high injection rate (Richardson et al., 1952).

Page 929: +Peters Ekwere j. - Petrophysics

8-13

Figure 8.6. Relative permeability curves from six steady state methods, short core section (Richardson et al., 1952).

Page 930: +Peters Ekwere j. - Petrophysics

8-14

Figure 8.7. Relative permeability curves from six steady state methods, long core section (Richardson et al., 1952).

The major problem with the steady state method for relative

permeability measurements is that it takes too long to complete. It is not

unusual for a steady state experiment to take several weeks to complete. An

alternative and much faster technique is the unsteady state method or the

dynamic displacement method based on immiscible displacement theory.

Because the calculation of relative permeabilities from unsteady state

experiment is based on the solution of two-phase immiscible displacement

equation, we must first solve the two-phase immiscible displacement problem

before we can discuss the unsteady state relative permeability measurements.

Page 931: +Peters Ekwere j. - Petrophysics

8-15

8.3 THEORY OF ONE DIMENSIONAL IMMISCIBLE DISPLACEMENT IN A POROUS MEDIUM

8.3.1 Mathematical Model of Two-Phase Immiscible Displacement

Consider the displacement of a non-wetting phase by a wetting phase in

a linear inclined core as shown in Figure 8.8. Darcy’s Law applied to each

phase gives

sinrnw nwnw nw

nw

kk A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.22)

sinrw ww w

w

kk A Pq gx

ρ αμ

∂⎛ ⎞= − +⎜ ⎟∂⎝ ⎠ (8.23)

Capillary equilibrium gives

( )nw w c wP P P S− = (8.17)

Assuming incompressible fluids, mass conservation requires that

w nwq q q= + (8.24)

The true fractional flows of the wetting and non-wetting phases are defined as

follows:

w ww

w nw

q qfq q q

= =+

(8.25)

1nw nwnw w

w nw

q qf fq q q

= = = −+

(8.26)

The continuity equation for the wetting phase is

0w wS qAt x

φ ∂ ∂+ =

∂ ∂ (8.27)

Page 932: +Peters Ekwere j. - Petrophysics

8-16

Figure 8.8. Displacement of a non-wetting phase by a wetting phase in an inclined core.

The saturation constraint gives

1w nwS S+ = (8.28)

Subtracting Eq.(8.22) from (8.23) and rearranging gives

( ) sinw w nw nw nw ww nw

rw rnw

q q P P gkk A kk A x x

μ μ ρ ρ α∂ ∂− = − − −

∂ ∂ (8.29)

Page 933: +Peters Ekwere j. - Petrophysics

8-17

Substituting Eqs.(8.17) and (8.24) into (8.29) gives upon rearrangement

( )1 sin

1

rnw cw nw

w nw

rnw w

rw nw

kk A P gq q x

kqk

ρ ρ αμ

μμ

∂⎡ ⎤+ − −⎢ ⎥∂⎣ ⎦=+

(8.30)

Let an approximate fractional flow of the wetting phase be defined as

1

1w

rnw w

rw nw

F kk

μμ

=+

(8.31)

Substituting Eqs.(8.30) and (8.31) into (8.25) gives the true fractional flow of

the wetting phase as

( )1 sinrnw cw w w nw

nw

kk A Pf F gq x

ρ ρ αμ

⎧ ⎫∂⎡ ⎤= + − −⎨ ⎬⎢ ⎥∂⎣ ⎦⎩ ⎭ (8.32)

Let the dimensionless distance from the inlet end be defined as

DxxL

= (8.33)

Let the spontaneous imbibition capillary pressure curve be given in terms of

its Leverett J-function as

( ) ( )cos ,/c w wP S J S

kσ θ

φ= Γ (8.34)

Substituting Eqs.(8.83) and (8.34) into (8.32) gives the true fractional flow of

the wetting phase as

( ) sincos1 w nww w rnw rnw

nw D nw

kA gA k Jf F k kq L x q

ρ ρ ασ θ φμ μ

⎡ ⎤⎛ ⎞ −⎛ ⎞∂= + −⎢ ⎥⎜ ⎟ ⎜ ⎟⎜ ⎟ ∂⎢ ⎥⎝ ⎠⎝ ⎠⎣ ⎦

(8.35)

Eq.(8.35) can be written as

Page 934: +Peters Ekwere j. - Petrophysics

8-18

1w w cap rnw rnw gD

Jf F N k k Nx

⎡ ⎤∂= + −⎢ ⎥∂⎣ ⎦

(8.36)

where capN is given by

coscap

nw

A kNq L

σ θ φμ

= (7.61)

and gN is given by

( ) sinw nwg

nw

kA gN

qρ ρ α

μ−

= (8.37)

capN is the same dimensionless number we encountered in the analysis of

capillary end effect. It represents the ratio of capillary to viscous forces in the

displacement. gN is a new dimensionless number, which represents the ratio

of gravity to viscous forces in the displacement. The mobility ratio of the

displacement is given by

( ) rw nww

rnw w

kM Sk

μμ

= (8.38)

The mobility ratio as defined in Eq(8.38) is a function of saturation and will be

different at each point in the porous medium depending on the saturation. A

characteristic mobility ratio for the displacement can be defined in terms of

the end-point relative permeabilities as

wr nwE

nwr w

kMk

μμ

= (8.39)

where wrk and nwrk are the end-point relative permeabilities for the wetting and

non-wetting phases. The mobility ratio given in Eq.(8.39) is a characteristic

dimensionless number for the displacement that is independent of saturation.

Eq.(8.36) can also be written as

Page 935: +Peters Ekwere j. - Petrophysics

8-19

1

11

cap rnw rnw gD

w

JN k k Nxf

M

∂+ −

∂=

⎛ ⎞+⎜ ⎟⎝ ⎠

(8.40)

where the approximate fractional flow of the wetting phase is given by

111

wF

M

=⎛ ⎞+⎜ ⎟⎝ ⎠

(8.41)

In order to maximize the displacement efficiency of the non-wetting

phase, we need to minimize the fractional flow of the wetting phase at each

point in the porous medium. Much can be deduced about the immiscible

displacement from the fractional flow equation, Eq.(8.40). Examination of this

equation leads to the following qualitative deductions about immiscible

displacements in porous media:

1. The fractional flow of the wetting phase is a strong function of

saturation.

2. The displacement behavior can be rate-sensitive if the effect of

capillarity or gravity is significant.

3. Capillarity is detrimental to the displacement efficiency as it increases

the fractional flow of the wetting phase at a given saturation.

4. Gravity is beneficial to the displacement efficiency for up-dip

displacement of the lighter non-wetting phase by the heavier wetting

phase as it reduces the fractional flow of the wetting phase at a given

saturation. Conversely, gravity will be detrimental to the displacement

efficiency for down-dip displacement of the lighter non-wetting phase by

the heavier wetting phase.

5. The displacement efficiency can be increased by reducing the mobility

ratio. This can be accomplished in practice by increasing the viscosity

Page 936: +Peters Ekwere j. - Petrophysics

8-20

of the wetting phase (the injected fluid) by use of a polymer. This is the

basis for polymer flooding as an improved oil recovery technique.

6. The effects of gravity and capillarity on the displacement can be reduced

by increasing the injection rate.

There are additional facts about the immiscible displacement that are

not apparent from the fractional flow equation. The fractional flow equation

indicates that the displacement efficiency can be improved by injecting the

wetting phase at a high enough rate to minimize capillary smearing of the

displacement front. This is generally true for a favorable mobility ratio

displacement. If the mobility ratio is unfavorable, an increase in rate can

result in viscous instability which reduces the displacement efficiency. The

fractional flow equation suggests that the effect of gravity will be eliminated if

the porous medium is horizontal. This is misleading because, in practice, if

there is a density contrast between the fluids, the injection rate is sufficiently

low and the core has a vertical dimension (which it does), gravity segregation

will occur even in a horizontal medium. In this case, the one dimensional

displacement model is inadequate to describe the displacement. A

multidimensional model is needed to correctly describe the gravity-dominated

displacement. The only fail proof way to eliminate the effect of gravity is to

eliminate the density contrast between the fluids or perform the displacement

in outer space. One can create a gravity number for displacement in a

horizontal core by replacing sinα in Eq.(8.37) by the aspect ratio (d/L).

The partial differential equation for the wetting phase saturation can be

derived as follows. Let

( )w w rnw gS F k NΨ = − (8.42)

( )w rnww

dJS kdS

Ω = (8.43)

Eq.(8.36) then becomes

Page 937: +Peters Ekwere j. - Petrophysics

8-21

( ) ( ) ( ) ww w w cap w

D

Sf S S N Sx

∂= Ψ + Ω

∂ (8.44)

Eq.(8.27) can be written in dimensionless form as

0w w

D D

S ft x

∂ ∂+ =

∂ ∂ (8.45)

where Dt is given by

Dqtt

A Lφ= (7.63)

Substituting Eq.(8.44) into (8.45) gives the partial differential equation for the

wetting phase saturation as

( ) 0w w wcap w

D w D D D

S S Sd N St dS x x x

⎡ ⎤∂ ∂ ∂Ψ ∂+ + Ω =⎢ ⎥∂ ∂ ∂ ∂⎣ ⎦

(8.46)

We have reduced the immiscible displacement problem to the solution of a

second order, nonlinear, parabolic partial differential equation for the wetting

phase saturation. When supplemented with appropriate initial and boundary

conditions, Eq.(8.46) can be solved, usually numerically, to obtain the wetting

phase saturation in time and space.

8.3.2 Buckley-Leverett Approximate Solution of the Immiscible Displacement Equation

Eq.(8.46) cannot be solved analytically for the saturation profiles. Here,

we examine the approximate solution obtained by Buckley and Leverett

(1941). The continuity equation, Eq.(8.45), can be written as

0w w w

D w D

S df St dS x

∂ ∂+ =

∂ ∂ (8.47)

where the true fractional flow of the wetting phase for horizontal

displacement is given by

Page 938: +Peters Ekwere j. - Petrophysics

8-22

1

11

wcap rnw

w Dw

SdJN kdS xf

M

∂+

∂=

⎛ ⎞+⎜ ⎟⎝ ⎠

(8.48)

It should be observed that the true fractional flow function contains w

D

Sx

∂∂

,

which is unknown. Buckley and Leverett (1941) obtained an approximate

solution to Eq.(8.47) by making a key simplifying assumption. They dropped

the capillary pressure term from Eq.(8.48) and as a result, they approximated

the fractional flow of the wetting phase as

w wf F (8.49)

Substituting Eq.(8.49) into (8.47) gives the partial differential equation for the

wetting phase saturation as

0w w w

D w D

S dF St dS x

∂ ∂+ =

∂ ∂ (8.50)

Eq.(8.50) is known as the Buckley-Leverett equation in the petroleum

industry. The Buckley-Leverett approximation changes the original second

order parabolic partial differential equation for the wetting phase saturation to

a first order, hyperbolic partial differential equation. This is a radical change

in the structure of the mathematical problem. However, the change allows an

approximate analytical solution to be obtained for the wetting phase

saturation profiles that is adequate for making gross performance calculations

for the immiscible displacement.

Eq.(8.50) is a nonlinear, first order, hyperbolic partial differential

equation that can be solved by the method of characteristics. From calculus,

the total time derivative of ( ),D DS x t is given by

w w wD

D D D D

S S dSdxt dt x dt

⎛ ⎞∂ ∂+ =⎜ ⎟∂ ∂⎝ ⎠

(8.51)

Page 939: +Peters Ekwere j. - Petrophysics

8-23

Subtracting Eq.(8.50) from (8.51) gives

w w wD

D w D D

dF S dSdxdt dS x dt

⎛ ⎞ ∂− =⎜ ⎟ ∂⎝ ⎠

(8.52)

Eq.(8.52) can be decomposed into the following two simultaneous equations:

0wD

D w

dFdxdt dS

− = (8.53)

0w

D

dSdt

= (8.54)

Eq.(8.53) gives the characteristic path for the hyperbolic partial differential

equation given by Eq(8.54). Eq.(8.54) shows that along the characteristic

path given by Eq.(8.53), the saturation is a constant.

Eq.(8.53) can be integrated to determine the distance traveled by a

constant saturation at a given time as

( ) ( ) ( )00w D w

wD DDS t DS

w

dFx x t tdS

− = − (8.55)

If there was no prior injection, 0Dt will be zero and all the saturations from Swi

to (1 - Snwr) will be located at the inlet end of the system, making ( )0wDSx equal

to zero. In this case, Eq.(8.55) becomes

( )w D

wDDS t

w

dFx tdS

= (8.56)

Eq.(8.56) can be written as

wD D

w

dFx tdS

= (8.57)

Page 940: +Peters Ekwere j. - Petrophysics

8-24

where Dx is the dimensionless distance traveled by a given saturation at time

Dt . Eq.(8.57) can be written in dimensional form as

( )i w

w

Q t dFxA dSφ

= (8.58)

Eq.(8.58) is usually referred to in the petroleum industry as the Buckley-

Leverett frontal advance equation. It should be emphasized that Eq.(8.57) or

(8.58) applies to a particular wetting phase saturation. To determine the

dimensionless distance traveled by a particular saturation 1wS at time Dt , we

use Eq.(8.57) to compute the distance as

1w

wD D

w S

dFx tdS

⎛ ⎞= ⎜ ⎟

⎝ ⎠ (8.59)

where the derivative of the approximate fractional flow curve is evaluated at

1wS . Eq.(8.57) can be used to derive a similarity transformation for an

immiscible displacement. The similarity transformation is given by

w wD

D w w

dF dfxzt dS dS

= = = (8.60)

If the saturation profiles for an immiscible displacement are plotted as wS

versus D

D

xt

, all the saturation profiles will collapse into one curve. If the

saturation profiles in an immiscible displacement are imaged say by CT or by

NMR, then Eq.(8.60) can be used to calculate the true fractional flow curve,

including the effect of capillarity, as

( ) w

wirr

SD

w w wSD

xf S dSt

= ∫ (8.61)

Given the relative permeability curves and the viscosity ratio, the

approximate fractional flow function and its derivative can be computed and

Page 941: +Peters Ekwere j. - Petrophysics

8-25

plotted as shown in Figure 8.9. In this figure, the S-shaped curve ADBC is

the approximate fractional flow curve (Fw) obtained from the relative

permeability curves and the viscosity ratio. The curve AFE is the derivative of

this function w

w

dFdS

⎛ ⎞⎜ ⎟⎝ ⎠

. Using Eq.(8.59) and this derivative function, the distance

traveled by each wetting phase saturation between Swirr and (1-Snwr) at a

given time Dt can be computed. Figure 8.10 shows the saturation profile that

will be obtained before wetting phase breakthrough by use of Eq.(8.59) and

the approximate derivative function. We see that the Buckley-Leverett

approximation gives rise to multiple-valued saturations at various Dx which is

physically impossible. This multiple-valued solution is caused by neglecting

the capillary term in the fractional flow equation. It is no accident that the

multi-valued solution occurs in the saturation range wirr w wfS S S< < where the

capillary pressure gradient is high and should not have been neglected.

To eliminate the multiple-valued solution, we appeal to physical reality

as follows. At time t, Qi(t) of wetting phase has been injected and the flood

front has traveled a distance fx into the medium. A volumetric balance of the

injected wetting phase can be used to calculate fx as follows:

( ) ( )0

fx

i w wirrQ t A S S dxφ= −∫ (8.62)

Integration of Eq.(8.62) by parts and substitution of Eq.(8.58) gives

( ) ( )1

wf

nwr

S wi f wf wirr i wS

w

dFQ t Ax S S Q dSdS

φ−

= − − ∫ (8.63)

Upon performing the integration in Eq.(8.63) and rearranging, one obtains

Page 942: +Peters Ekwere j. - Petrophysics

8-26

Figure 8.9. Approximate fractional flow function and its first derivative. Note the tangent construction.

Figure 8.10. Calculated water saturation distribution based on the Buckley-Leverett approximation showing the discontinuity in saturation as required by

a material balance.

Page 943: +Peters Ekwere j. - Petrophysics

8-27

( ) ( )w wff i

wf wirr

F SAx Q t

S Sφ =

− (8.64)

From the Buckley-Leverett frontal advance equation, Eq.(8.58), one can also

obtain

( )wf

wf i

w S

dFAx Q tdS

φ⎛ ⎞

= ⎜ ⎟⎝ ⎠

(8.65)

A comparison of Eqs.(8.64) and (8.65) gives

( )

wf

w wfw

w wf wirrS

F SdFdS S S

⎛ ⎞=⎜ ⎟ −⎝ ⎠

(8.66)

The saturation distribution in Figure 8.10 will be single valued if all the

saturations between Swirr and the frontal saturation, Swf, are eliminated.

Eq.(8.66) shows that the frontal saturation (Swf) is the saturation at which the

straight line passing through the point Sw = Swirr and Fw = 0 is tangent to the

approximate fractional flow curve, Fw. This line is shown in Figure 8.9 as AB.

This tangent construction was first suggested by Welge (1952). The effect of

the tangent construction is to correct the approximate fractional flow curve Fw

for the capillary term that was neglected to obtain the true fractional flow

curve fw. Such a correction is needed at the front (low wetting phase

saturation) where the capillary pressure gradient is high and should not have

been neglected. With the tangent construction correction in place, the true

fractional flow curve, fw, is now given by the curve ABC (Fig. 8.9) thereby

eliminating the S-shaped lower portion of Fw, which led to the tripple-valued

saturation solution of Figure 8.10. With this correction, the derivative of the

true fractional flow curve w

w

dfdS

⎛ ⎞⎜ ⎟⎝ ⎠

used in the solution is given by the curve

EFG (Fig. 8.9). After the tangent construction, the true fractional flow curve

and its derivative are given by

Page 944: +Peters Ekwere j. - Petrophysics

8-28

Figure 8.11. Similarity transformation for an immiscible displacement.

Figure 8.12. Integration of the transformed saturation data to calculate the true fractional flow curve including capillarity.

Page 945: +Peters Ekwere j. - Petrophysics

8-29

( )( ) ( )

( )

for

for 1.0 wf

w wirr ww wf w wf wirr w wf

wf wirr w Sw w

w w wf w

S S dFF S S S S S SS S dSf S

F S S S

⎧⎛ ⎞ ⎛ ⎞−= − ≤ ≤⎪⎜ ⎟ ⎜ ⎟⎪⎜ ⎟−= ⎝ ⎠⎨⎝ ⎠

⎪≤ ≤⎪⎩

(8.67)

and

= a constant for

for 1.0

wf

w

wwirr w wf

w Sw

w wwf w

w S

dF S S SdSdf

dS dF S SdS

⎧⎛ ⎞≤ ≤⎪⎜ ⎟

⎝ ⎠⎪= ⎨

⎛ ⎞⎪ ≤ ≤⎜ ⎟⎪⎝ ⎠⎩

(8.68)

The similarity transformation for the immiscible displacement is given

by the curve EFGH (Fig. 8.9) and is shown in Figure 8.11. Figure 8.12 shows

how the transformed saturation data can be integrated to obtain the true

fractional flow curve that includes the effect of capillarity.

We now show that the intersection of the tangent line with the line Fw =

1 (point J in Fig. 8.9) gives the constant average wetting phase saturation

behind the front before and at wetting phase breakthrough. The slope of the

tangent line can be written as

( )1

wf

w wfw

w wav wfS

F SdFdS S S

−⎛ ⎞=⎜ ⎟ −⎝ ⎠

(8.69)

which can be rearranged as

( )1

wf

w wfwav wf

w

w S

F SS S

dFdS

−= +

⎛ ⎞⎜ ⎟⎝ ⎠

(8.70)

Before breakthrough, the average wetting phase saturation behind the front is

given by

Page 946: +Peters Ekwere j. - Petrophysics

8-30

1

0

nwrS

wwav

f

AxdSS

Ax

φ

φ

= ∫ (8.71)

Substituting Eq.(8.58) into (8.71) and integrating gives the average wetting

phase saturation behind the front as

( )1

wf

w wfwav wf

w

w S

F SS S

dFdS

−= +

⎛ ⎞⎜ ⎟⎝ ⎠

(8.72)

Eq.(8.72) is identical to Eq.(8.70), which confirms that the intersection of the

tangent line and the line Fw = 1 gives the wetting phase saturation (Swav)

corresponding to point J in Figure 8.9. Thus, the average wetting phase

saturation at breakthrough can be obtained graphically from the tangent

construction.

The average wetting phase saturation after breakthrough can be

obtained by a tangent construction at the outlet wetting phase saturation

between Swf and 1 - Snwr. The intersection of the tangent line and the line

Fw = 1 gives the average wetting phase saturation in the porous medium

corresponding to the outlet wetting phase saturation. By this tangent

construction, the Buckley-Leverett approximation can be used to predict the

performance of the one-dimensional immiscible displacement after wetting

phase breakthrough. Before breakthrough, the amount of non-wetting phase

recovered is equal to the amount of fluid injected. Thus, the entire

displacement performance can be predicted for a given set of wetting and non-

wetting relative permeability curves and wetting and non-wetting viscosity

ratio.

Page 947: +Peters Ekwere j. - Petrophysics

8-31

8.3.3 Waterflood Performance Calculations from Buckley–Leverett Theory

We now apply Buckley-Leverett theory to calculate a waterflood

performance from beginning to end. It is assumed that the true fractional flow

curve and its derivative have been computed using the relative permeability

curves, the viscosity ratio and the Welge tangent construction. Therefore, the

equations in this section are written in terms of the true fractional flow curve.

The methodology presented also applies to the calculation of the performance

of a gas flood using gas-oil drainage relative permeability curves.

Oil Recovery at any Time

The oil recovery at any time after the initiation of water injection is

given by

1wav wirr

wirr

S SRS−

=−

(8.73)

where R is the oil recovery as a fraction of the initial oil in place, Swav is the

average water saturation in the porous medium at the time of interest and

Swirr is the initial water saturation in the porous medium before water

injection which is assumed to be the irreducible water saturation. Thus, in

principle, the oil recovery can be calculated at any time by first calculating the

average water saturation in the porous medium at that time and applying

Eq.(8.73). However, depending on the stage of water injection, Eq.(8.73) may

not offer the most direct way to calculate the oil recovery. Let us examine the

waterflood performance at various stages of the flood.

Oil Recovery Before Water Breakthrough

Consider a constant rate water injection project. Assuming

incompressible fluids, the amount of oil recovered before water breakthrough

must equal the amount of water injected. Thus, at reservoir conditions,

Page 948: +Peters Ekwere j. - Petrophysics

8-32

( ) ( )w i oqB t Q t Q t= = (8.74)

where q is the constant water injection rate, in surface units, Bw is the water

formation volume factor, t is the time of interest before water breakthrough,

Qi is the cumulative water injected at time t in reservoir units and Qo is

the cumulative oil produced at time t in reservoir units. The cumulative oil

produced at surface conditions is

( ) ( )Cumulative Oil Produced i ow

o o o

Q t Q tqB tB B B

= = = (8.75)

where Bo is the current oil formation volume factor. The oil recovery as a

fraction of the initial oil in place is given by

( )

( )( ) ( )1 1 1

iw i

wirr wirr wirr

Q tqB t WRAL S AL S Sφ φ

= = =− − −

(8.76)

where Wi is the pore volume of water injected.

Oil Recovery at Water Breakthrough

From the Buckley–Leverett frontal advance equation, Eq.(8.58), the

distance traveled by a given saturation is given by

( )w w

iw w w

w wS S

Q tqB t df dfxA dS A dSφ φ

⎛ ⎞ ⎛ ⎞= =⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ (8.77)

Let us apply Eq.(8.77) to the frontal water saturation Swf to get

( )wf wf

iw w wf

w wS S

Q tqB t df dfxA dS A dSφ φ

⎛ ⎞ ⎛ ⎞= =⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ (8.78)

At the moment of water breakthrough, the frontal saturation arrives at the

outlet end of the porous medium and fx equals L. At the moment of water

breakthrough, Eq.(8.78) then becomes

Page 949: +Peters Ekwere j. - Petrophysics

8-33

( )wf wf

iw w w

w wS S

Q tqB t df dfLA dS A dSφ φ

⎛ ⎞ ⎛ ⎞= =⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ (8.79)

Eq.(8.75) can be rearranged as

( ) 1

wf

iwi

w

w S

Q tqB tWAL AL df

dSφ φ

= = =⎛ ⎞⎜ ⎟⎝ ⎠

(8.80)

where Wi is the pore volume of water injected. The cumulative oil recovery at

water breakthrough is equal to the cumulative water injected in reservoir

volumes. The fractional oil recovery at water breakthrough is obtained from

Eq.(8.80) as

( )

( )( ) ( )

11 1 1

1wf

ii w

wi wi wi wwi

w S

Q tW qB tRS AL S AL S dfS

dSφ φ

= = = =− − − ⎛ ⎞

− ⎜ ⎟⎝ ⎠

(8.81)

The breakthrough time, tbt, can be obtained from Eq.(8.80) as

wf

btw

ww S

ALtdfqBdS

φ=

⎛ ⎞⎜ ⎟⎝ ⎠

(8.82)

or in dimensionless form as

1

wf

Dbtw

w S

tdfdS

=⎛ ⎞⎜ ⎟⎝ ⎠

(8.83)

The average water saturation in the porous medium behind the displacement

front before and at water breakthrough is given by

1

0

orS

wwav

f

AxdSS

Ax

φ

φ

= ∫ (8.71)

Page 950: +Peters Ekwere j. - Petrophysics

8-34

Figure 8.13 shows a typical water saturation distribution at time t before

breakthrough. From Figure 8.13, we see that the integral (area under the

curve) in Eq.(8.71) can be split into two parts as follows:

1 or

wf

S

wf f wSwav

f

AS x AxdSS

Ax

φ φ

φ

−+

=∫

(8.84)

Substituting Eq.(8.58) into (8.84) gives the average water saturation as

( )

1 or

wf

S

i wSwav wf

f

Q t dFS S

Axφ

= +∫

(8.85)

Performing the integration in Eq.(8.85) gives

( ) ( ) ( )1i w or w wf

wav wff

Q t f S f SS S

Axφ

⎡ ⎤− −⎣ ⎦= + (8.86)

But Fw at Sw = (1 – Sor) is equal to 1.0. Thus, Eq.(8.81) can rewritten as

( ) ( )1i w wf

wav wff

Q t f SS S

Axφ

⎡ ⎤−⎣ ⎦= + (8.87)

Substituting Eq.(8.78) into (8.87) gives the average water saturation behind

the front as

( )1

wf

w wfwav wf

w

w S

f SS S

dfdS

⎡ ⎤−⎣ ⎦= +⎛ ⎞⎜ ⎟⎝ ⎠

(8.72)

It should be observed in Figure 8.9 that the average water saturation behind

the front up until water breakthrough as given in Eq.(8.72) is the same as the

water saturation at which the tangent to the fractional flow curve intersects

the Fw = 1 axis. Thus, the average water saturation in the porous medium at

water breakthrough can easily be determined graphically. The average water

Page 951: +Peters Ekwere j. - Petrophysics

8-35

saturation can then be substituted into Eq.(8.73) to calculate the oil recovery

at water breakthrough. We can easily show that the result obtained by this

approach will be the same as that obtained by Eq.(8.81). Substituting

Eq.(8.72) into Eq.(8.73) gives

Figure 8.13. Typical water saturation profile at time t before water breakthrough.

Page 952: +Peters Ekwere j. - Petrophysics

8-36

( )1

1wf

w wfwf wirr

w

w S

wirr

f SS S

dfdS

RS

⎡ ⎤−⎣ ⎦+ −⎛ ⎞⎜ ⎟⎝ ⎠

=−

(8.88)

From the equation of the tangent line in Figure 8.9, we find that

( )

wf

w wfw

w wf wirrS

f SdfdS S S

⎛ ⎞=⎜ ⎟ −⎝ ⎠

(8.66)

Substituting Eq.(8.66) into (8.88) gives the oil recovery at water breakthrough

as

( )

1

1wf

wwi

w S

RdfSdS

=⎛ ⎞

− ⎜ ⎟⎝ ⎠

(8.89)

which is identical to Eq.(8.81).

Oil Recovery After Water Breakthrough

After water breakthrough, Eq.(8.77) applied to the outlet end of the

porous medium gives

( )2 2w w

iw w w

w wS S

Q tqB t df dfLA dS A dSφ φ

⎛ ⎞ ⎛ ⎞= =⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ (8.90)

where Sw2 is the water saturation at the outlet end of the porous medium

which now lies between Swf and (1 – Sor). Rearrangement of Eq.(8.90) gives

the pore volumes of water injected as

( )

2

1

w

iwi

w

w S

Q tqB tWAL AL df

dSφ φ

= = =⎛ ⎞⎜ ⎟⎝ ⎠

(8.91)

Page 953: +Peters Ekwere j. - Petrophysics

8-37

where Wi is the pore volumes of water injected since the initiation of water

injection. Eq.(8.91) is analogous to Eq.(8.80) before breakthrough.

A material balance for the water after water breakthrough gives

( ) ( ) ( )0

L

i w wirr wQ t A S S dx Q tφ− − =∫ (8.92)

The integral in Eq.(8.92) can be performed using integration by parts.

Performing integration by parts, Eq.(8.92) can be written as

( ) ( ){ } ( )2

1

0

or

w

SLi w wirr w wS

Q t A S S x AxdS Q tφ φ−

− − − =⎡ ⎤⎣ ⎦ ∫ (8.93)

Substituting the limits for the first integral gives

( ) ( ){ } ( )2

1

2or

w

S

i w wirr w wSQ t AL S S AxdS Q tφ φ

−− − − =∫ (8.94)

Substituting Eq.(8.77) into (8.94) gives

( ) ( ) ( ){ } ( )2

1

2or

w

S

i w wirr i w wSQ t AL S S Q t df Q tφ

−− − − =∫ (8.95)

Performing the integration in Eq.(8.95) gives

( ) ( ) ( ) ( ) ( ){ } ( )2 21i w wirr i w or w w wQ t AL S S Q t f S f S Q tφ− − − − − =⎡ ⎤⎣ ⎦ (8.96)

or

( ) ( ) ( ) ( ){ } ( )2 21i w wirr i w w wQ t AL S S Q t f S Q tφ− − − − =⎡ ⎤⎣ ⎦ (8.97)

since fw at Sw = 1–Sor is equal to 1. Eq.(8.97) can be rearranged as

( ) ( ) ( ) ( )2 21i w iw wirr w w

Q t Q t Q tS S f S

AL ALφ φ−

= + − −⎡ ⎤⎣ ⎦ (8.98)

Page 954: +Peters Ekwere j. - Petrophysics

8-38

( )2 21w wirr pD i w wS S N W f S= + − −⎡ ⎤⎣ ⎦ (8.99)

( )2 21pD w wi i w wN S S W f S= − + −⎡ ⎤⎣ ⎦ (8.100)

where NpD is the oil recovery as a fraction of the total pore volume. We observe

that the sum of the first two terms on the right hand side of Eq.(8.99) is the

average water saturation in the porous medium after water breakthrough.

Thus, Eq.(8.99) can be rewritten as

( )2 21w wav i w wS S W f S= − −⎡ ⎤⎣ ⎦ (8.101)

Substituting Eq.(8.91) into (8.101) and rearranging gives the average water

saturation in the porous medium after water breakthrough as

( )

2

22

1

w

w wwav w

w

w S

f SS S

dfdS

−⎡ ⎤⎣ ⎦= +⎛ ⎞⎜ ⎟⎝ ⎠

(8.102)

which is analogous to Eq.(8.72) at water breakthrough. Figure 8.14 shows

that the average water saturation after water breakthrough as given by

Eq.(8.102) is equal to the water saturation where the tangent line to the

fractional flow curve at the outlet water saturation intersects the Fw = 1 axis.

The average water saturation in the porous medium after water

breakthrough could also have been derived using Eq.(8.71) and the water

saturation profile shown in Figure 8.15. The average water saturation is then

given by

2

1

2or

w

S

w wSwav

AS L AxdSS

AL

φ φ

φ

−+

=∫

(8.103)

Substituting Eq.(8.77) into (8.103) gives the average water saturation as

Page 955: +Peters Ekwere j. - Petrophysics

8-39

( )

2

1

2

or

w

S

i wSwav w

Q t dfS S

ALφ

= +∫

(8.104)

Figure 8.14. Average water saturation after water breakthrough.

Performing the integration in Eq.(8.104) gives

( ) ( ) ( )2

2

1i w or w wwav w

Q t f S f SS S

ALφ− −⎡ ⎤⎣ ⎦= + (8.105)

But fw at Sw = 1 – Sor is equal to 1.0. Thus, Eq.(8.105) can be rewritten as

Page 956: +Peters Ekwere j. - Petrophysics

8-40

Figure 8.15. Typical water saturation profile at time t after water breakthrough.

( ) ( )2

2

1i w wwav w

Q t f SS S

ALφ−⎡ ⎤⎣ ⎦= + (8.106)

Substituting Eq.(8.90) into (8.106) gives the average water saturation after

water breakthrough as

( )

2

22

1

w

w wwav w

w

w S

F SS S

dFdS

−⎡ ⎤⎣ ⎦= +⎛ ⎞⎜ ⎟⎝ ⎠

(8.107)

Page 957: +Peters Ekwere j. - Petrophysics

8-41

which is identical to Eq.(8.102).

Water Production

There is no water production before water breakthrough. After water

breakthrough, the water oil ratio is given by

1

w

ww o w

o w wo

o

fBq B FWOR

q B FfB

⎛ ⎞⎜ ⎟ ⎛ ⎞⎝ ⎠= = = ⎜ ⎟−⎛ ⎞ ⎝ ⎠⎜ ⎟⎝ ⎠

(8.108)

The pore volumes of water produced is given by material balance on the water

as

Water produced Cumulative water injected Water stored= − (8.109)

Substituting appropriate symbols into Eq.(8.109) gives

( )p i wav wirrW W S S= − − (8.110)

Substituting Eq.(8.91) into (8.110) gives the pore volumes of water produced

as

( )

2

1

w

p wav wiw

w S

W S SdfdS

= − −⎛ ⎞⎜ ⎟⎝ ⎠

(8.111)

Example 8.1

A waterflood is to be performed in a linear reservoir. The relative permeability

curves for the reservoir are adequately described by the following analytical

models:

3rw wr ek k S= (8.112)

Page 958: +Peters Ekwere j. - Petrophysics

8-42

( )21rnw nwr ek k S= − (8.113)

where Se is defined as

1

w wirre

wirr nwr

S SSS S

−=

− − (8.114)

The other pertinent data are as follows:

0.20wirrS =

0.30nwrS =

0.95nwrk =

0.35wrk =

10nw oμ μ= = cp

1wμ = cp

1.20oB = RB/STB

1.0wB = RB/STB

1. Calculate and plot graphs of the relative permeability curves.

2. Calculate and plot graphs of the approximate fractional flow curve

( )wF and its derivative w

w

dFdS

⎛ ⎞⎜ ⎟⎝ ⎠

.

3. Perform the Welge tangent construction and from it determine the

frontal water saturation ( )wfS , the average water saturation at water

breakthrough ( )wavS and the true fractional flow curve ( )wf and its

derivative w

w

dfdS

⎛ ⎞⎜ ⎟⎝ ⎠

.

Page 959: +Peters Ekwere j. - Petrophysics

8-43

4. Plot the graphs of the true fractional flow curve and its derivative.

5. Calculate the end point mobility ratio for the waterflood.

6. Calculate and plot graphs of the water saturation profiles at tD = 0.20,

0.30 and 1.0.

7. Calculate the dimensionless breakthrough time.

8. Calculate the breakthrough oil recovery as a fraction of the initial oil in

place.

9. Calculate and plot the graph of oil recovery versus pore volume of water

injected before and after water breakthrough.

10. Calculate and plot the graph of water oil ratio versus oil recovery.

Solution to Example 8.1

The results of the calculations are summarized in Table 8.1.

1. The relative permeability curves calculated with Eqs.(8.112) and (8.113)

are shown in Figure 8.16.

2. Figure 8.17 shows the approximate fractional flow curve calculated with

Eq.(8.41) and its derivative calculated by differentiating Fw with respect

to Sw analytically.

Table 8.1. Calculated Results for Example 8.1.

tD tD tD

0.20 0.30 1.00

Sw krw krnw Fw w

w

dFdS

fw w

w

dfdS

xD xD xD Wi R WOR

0.200 0.00000 0.950 0.00000 0.000 0.000 2.775 0.555 0.833 2.775 0.000 0.000 0.000

0.210 0.00000 0.912 0.00003 0.009 0.028 2.775 0.555 0.833 2.775 0.008 0.023 0.000

0.220 0.00002 0.876 0.00026 0.039 0.056 2.775 0.555 0.833 2.775 0.016 0.045 0.000

0.230 0.00008 0.839 0.00090 0.094 0.083 2.775 0.555 0.833 2.775 0.025 0.068 0.000

0.240 0.00018 0.804 0.00222 0.176 0.111 2.775 0.555 0.833 2.775 0.033 0.091 0.000

Page 960: +Peters Ekwere j. - Petrophysics

8-44

0.250 0.00035 0.770 0.00453 0.290 0.139 2.775 0.555 0.833 2.775 0.041 0.113 0.000

0.260 0.00060 0.736 0.00815 0.441 0.167 2.775 0.555 0.833 2.775 0.049 0.136 0.000

0.270 0.00096 0.703 0.01348 0.632 0.194 2.775 0.555 0.833 2.775 0.057 0.158 0.000

0.280 0.00143 0.670 0.02094 0.866 0.222 2.775 0.555 0.833 2.775 0.066 0.180 0.000

0.290 0.00204 0.639 0.03097 1.147 0.250 2.775 0.555 0.833 2.775 0.074 0.202 0.000

0.300 0.00280 0.608 0.04403 1.473 0.278 2.775 0.555 0.833 2.775 0.082 0.223 0.000

0.310 0.00373 0.578 0.06057 1.844 0.305 2.775 0.555 0.833 2.775 0.090 0.243 0.000

0.320 0.00484 0.549 0.08103 2.254 0.333 2.775 0.555 0.833 2.775 0.098 0.263 0.000

0.330 0.00615 0.520 0.10575 2.693 0.361 2.775 0.555 0.833 2.775 0.106 0.281 0.000

0.340 0.00768 0.492 0.13496 3.150 0.389 2.775 0.555 0.833 2.775 0.115 0.299 0.000

0.350 0.00945 0.466 0.16875 3.607 0.416 2.775 0.555 0.833 2.775 0.123 0.315 0.000

0.360 0.01147 0.439 0.20703 4.044 0.444 2.775 0.555 0.833 2.775 0.131 0.330 0.000

0.370 0.01376 0.414 0.24949 4.439 0.472 2.775 0.555 0.833 2.775 0.139 0.343 0.000

0.380 0.01633 0.389 0.29560 4.772 0.500 2.775 0.555 0.833 2.775 0.147 0.355 0.000

0.390 0.01921 0.365 0.34465 5.024 0.527 2.775 0.555 0.833 2.775 0.156 0.365 0.000

0.400 0.02240 0.342 0.39576 5.181 0.555 2.775 0.555 0.833 2.775 0.164 0.374 0.000

0.410 0.02593 0.320 0.44794 5.238 0.583 2.775 0.555 0.833 2.775 0.172 0.381 0.000

0.420 0.02981 0.298 0.50019 5.195 0.611 2.775 0.555 0.833 2.775 0.180 0.388 0.000

0.430 0.03407 0.277 0.55153 5.058 0.638 2.775 0.555 0.833 2.775 0.188 0.393 0.000

0.440 0.03871 0.257 0.60109 4.842 0.666 2.775 0.555 0.833 2.775 0.197 0.398 0.000

0.450 0.04375 0.238 0.64815 4.561 0.694 2.775 0.555 0.833 2.775 0.205 0.403 0.000

0.460 0.04921 0.219 0.69216 4.234 0.722 2.775 0.555 0.833 2.775 0.213 0.407 0.000

0.470 0.05511 0.201 0.73274 3.879 0.749 2.775 0.555 0.833 2.775 0.221 0.411 0.000

0.480 0.06147 0.184 0.76969 3.511 0.777 2.775 0.555 0.833 2.775 0.229 0.416 0.000

0.490 0.06829 0.168 0.80296 3.144 0.805 2.775 0.555 0.833 2.775 0.237 0.421 0.000

0.491 0.06900 0.166 0.80608 3.107 0.808 2.775 0.555 0.833 2.775 0.246 0.423 0.000

0.492 0.06971 0.164 0.80917 3.071 0.810 2.775 0.555 0.833 2.775 0.254 0.426 0.000

0.493 0.07043 0.163 0.81222 3.035 0.813 2.775 0.555 0.833 2.775 0.262 0.428 0.000

0.494 0.07115 0.161 0.81524 2.999 0.816 2.775 0.555 0.833 2.775 0.270 0.430 0.000

0.495 0.07188 0.160 0.81822 2.964 0.819 2.775 0.555 0.833 2.775 0.278 0.432 0.000

0.496 0.07262 0.158 0.82117 2.928 0.822 2.775 0.555 0.833 2.775 0.287 0.434 0.000

0.497 0.07335 0.157 0.82408 2.893 0.824 2.775 0.555 0.833 2.775 0.295 0.436 0.000

0.498 0.07410 0.155 0.82695 2.857 0.827 2.775 0.555 0.833 2.775 0.303 0.438 0.000

0.499 0.07485 0.154 0.82979 2.822 0.830 2.775 0.555 0.833 2.775 0.311 0.440 0.000

0.500 0.07560 0.152 0.83260 2.788 0.833 2.775 0.555 0.833 2.775 0.319 0.442 0.000

Page 961: +Peters Ekwere j. - Petrophysics

8-45

0.500 0.07568 0.152 0.83288 2.784 0.833 2.775 0.555 0.833 2.775 0.328 0.444 0.000

0.500 0.07575 0.152 0.83316 2.781 0.833 2.775 0.555 0.833 2.775 0.336 0.445 0.000

0.500 0.07583 0.152 0.83343 2.777 0.833 2.775 0.555 0.833 2.775 0.344 0.447 0.000

0.500 0.07583 0.152 0.83343 2.777 0.833 2.775 0.555 0.833 2.775 0.352 0.449 0.000

0.500 0.07586 0.151 0.83357 2.775 0.834 2.775 0.555 0.833 2.775 0.360 0.450 6.010

0.501 0.07636 0.150 0.83537 2.753 0.835 2.753 0.551 0.826 2.753 0.363 0.451 6.089

0.502 0.07712 0.149 0.83810 2.718 0.838 2.718 0.544 0.816 2.718 0.368 0.452 6.212

0.503 0.07789 0.147 0.84081 2.684 0.841 2.684 0.537 0.805 2.684 0.373 0.453 6.338

0.504 0.07866 0.146 0.84347 2.650 0.843 2.650 0.530 0.795 2.650 0.377 0.454 6.466

0.505 0.07944 0.144 0.84611 2.616 0.846 2.616 0.523 0.785 2.616 0.382 0.455 6.598

0.506 0.08023 0.143 0.84871 2.583 0.849 2.583 0.517 0.775 2.583 0.387 0.456 6.732

0.507 0.08102 0.142 0.85127 2.549 0.851 2.549 0.510 0.765 2.549 0.392 0.457 6.868

0.508 0.08181 0.140 0.85380 2.516 0.854 2.516 0.503 0.755 2.516 0.397 0.458 7.008

0.509 0.08261 0.139 0.85630 2.483 0.856 2.483 0.497 0.745 2.483 0.403 0.459 7.151

0.510 0.08341 0.137 0.85877 2.450 0.859 2.450 0.490 0.735 2.450 0.408 0.460 7.297

0.520 0.09175 0.123 0.88169 2.137 0.882 2.137 0.427 0.641 2.137 0.468 0.469 8.943

0.530 0.10062 0.110 0.90160 1.850 0.902 1.850 0.370 0.555 1.850 0.540 0.479 10.995

0.540 0.11005 0.097 0.91878 1.591 0.919 1.591 0.318 0.477 1.591 0.628 0.489 13.575

0.550 0.12005 0.086 0.93351 1.360 0.934 1.360 0.272 0.408 1.360 0.736 0.499 16.849

0.560 0.13064 0.074 0.94606 1.154 0.946 1.154 0.231 0.346 1.154 0.866 0.508 21.048

0.570 0.14183 0.064 0.95668 0.974 0.957 0.974 0.195 0.292 0.974 1.027 0.518 26.502

0.580 0.15364 0.055 0.96561 0.816 0.966 0.816 0.163 0.245 0.816 1.226 0.528 33.693

0.590 0.16609 0.046 0.97306 0.678 0.973 0.678 0.136 0.203 0.678 1.474 0.537 43.348

0.600 0.17920 0.038 0.97923 0.559 0.979 0.559 0.112 0.168 0.559 1.788 0.546 56.589

0.610 0.19298 0.031 0.98430 0.456 0.984 0.456 0.091 0.137 0.456 2.191 0.555 75.235

0.620 0.20745 0.024 0.98841 0.368 0.988 0.368 0.074 0.110 0.368 2.716 0.564 102.358

0.630 0.22262 0.019 0.99171 0.292 0.992 0.292 0.058 0.088 0.292 3.420 0.573 143.471

0.640 0.23852 0.014 0.99430 0.228 0.994 0.228 0.046 0.068 0.228 4.392 0.581 209.224

0.650 0.25515 0.009 0.99629 0.172 0.996 0.172 0.034 0.052 0.172 5.798 0.589 322.295

0.660 0.27254 0.006 0.99777 0.126 0.998 0.126 0.025 0.038 0.126 7.966 0.597 537.909

0.670 0.29070 0.003 0.99882 0.086 0.999 0.086 0.017 0.026 0.086 11.663 0.605 1020.015

0.680 0.30966 0.002 0.99951 0.052 1.000 0.052 0.010 0.016 0.052 19.193 0.612 2444.665

0.690 0.32942 0.000 0.99988 0.024 1.000 0.024 0.005 0.007 0.024 42.067 0.619 10402.648

0.700 0.35000 0.000 1.00000 0.000 1.000 0.000 0.000 0.000 0.000

Page 962: +Peters Ekwere j. - Petrophysics

8-46

Figure 8.16. Relative permeability curves for Example 8.1.

Page 963: +Peters Ekwere j. - Petrophysics

8-47

Figure 8.17. Approximate fractional flow curve and its derivative for Example

8.1.

3. The Welge tangent construction is shown in Figure 8.17. From the

tangent construction,

0.500035wfS =

2.775wf

w

w S

dfdS

⎛ ⎞=⎜ ⎟

⎝ ⎠

0.5603wavS =

4. The true fractional flow curve and its derivative obtained from the

tangent construction are shown in Figure 8.18.

Page 964: +Peters Ekwere j. - Petrophysics

8-48

Figure 8.18. True fractional flow curve and its derivative for Example 8.1.

5. The end point mobility ratio for the waterflood is given by

( ) ( ) ( ) ( )/ / / 0.35 /1 / 0.95 /10 3.68E rw w rnw nwM k kμ μ= = =

6. The water saturation profiles calculated with Eq.(8.57) are shown in

Figure 8.19.

Page 965: +Peters Ekwere j. - Petrophysics

8-49

Figure 8.19. Water saturation profiles for Example 8.1.

7. The dimensionless breakthrough time is calculated with Eq.(8.83) as

1 1 0.4602.775

wf

Dbtw

w S

tdfdS

= = =⎛ ⎞⎜ ⎟⎝ ⎠

pore volume injected.

8. The breakthrough oil recovery as a fraction of the initial oil in place is

calculated with Eq.(8.81) as

( ) ( )( )

1 1 0.4501 0.20 2.775

1wf

btw

wirrw S

RdfSdS

= = =−⎛ ⎞

− ⎜ ⎟⎝ ⎠

9. Before water breakthrough, the oil recovery is a linear function of the

pore volume injected and can be calculated with Eq.(8.86). After water

Page 966: +Peters Ekwere j. - Petrophysics

8-50

breakthrough, the oil recovery is calculated with Eq.(8.100) as

1pD

wirr

NR

S=

−.

Figure 8.20 shows the calculated oil recovery curve.

Figure 8.20. Oil recovery curve for Example 8.1.

10. The producing water oil ratio is zero before water breakthrough. After

water breakthrough, the producing water oil water ratio is calculated

with Eq.(8.108). After breakthrough, the producing water oil ratio

increases rapidly as shown in Figure 8.21.

Page 967: +Peters Ekwere j. - Petrophysics

8-51

Figure 8.21. Producing water oil ratio for Example 8.1.

8.4 LABORATORY MEASUREMENT OF TWO-PHASE RELATIVE PERMEABILITIES BY THE UNSTEADY STATE METHOD

The major problem with the steady state method for relative

permeability measurements is that it is too slow. An alternative and much

faster technique is the unsteady state method or the dynamic displacement

method (Welge, 1952; Johnson et al., 1959; Jones and Roszelle, 1978). In

this method, for an imbibition test, the core is first saturated with the non-

wetting phase at irreducible wetting phase saturation as in the steady state

method. However, only the wetting phase is injected into the core to displace

the non-wetting phase. As the experiment progresses, the wetting phase

breaks through at the outlet end of the core and over time a higher and higher

fraction of the total produced fluid is the wetting phase.

Page 968: +Peters Ekwere j. - Petrophysics

8-52

By measuring the produced fractions of the wetting and non-wetting

phases at the outlet end of the core and the pressure drop across the core

versus time, the relative permeability curves can be calculated from the

production and pressure data using the theory of immiscible displacement in

porous media. This method is much faster than the steady state method,

usually requiring a few hours to complete compared to several weeks for the

steady state method. If adequate precautions are taken, the dynamic

displacement method will give relative permeability curves that are

comparable to those obtained by the steady state method.

Figure 8.22 shows the experimental setup and the measured data.

Because the point of observation is the outlet end of the core, it is necessary

that capillary end effect be minimized otherwise the calculated relative

permeability-saturation relationship will be wrong. It should be noted that

relative permeability curves can only be obtained over the saturation range Swf

to 1-Snwr. Therefore, it is necessary to choose the fluid viscosities that will give

the widest possible saturation window. This is obtained by using performing

and adverse mobility ratio displacement. A favorable mobility ratio

displacement will be unsuitable because for such a displacement, Swf is equal

to (1-Snwr) and there is no saturation window for calculating the relative

permeability curves. The relative permeability to the wetting phase below Swf

can only be obtained by extrapolating the data above Swf.

The technique for calculating relative permeability curves from

unsteady state measurements was developed by Welge (1952) and Johnson,

Bossler and Neumann (JBN, 1959). The fractional flow of the non-wetting

phase at the outlet end of the core is given by

21

1nw

rw nw

rnw w

f kk

μμ

=+

(8.115)

Page 969: +Peters Ekwere j. - Petrophysics

8-53

Figure 8.22. Unsteady state method for determining two-phase relative permeability curves; (a) coreflood; (b) measured data.

It should be noted that for saturations above Swf, Eq.(8.115) gives the true

fractional flow of the non-wetting phase because above Swf, the true fractional

flow and the approximate fractional curves are equal. Eq.(8.115) can be

rearranged to calculate the wetting-non-wetting phase relative permeability

ratio as

2

1 1rw w

rnw nw nw

kk f

μμ

⎛ ⎞= −⎜ ⎟

⎝ ⎠ (8.116)

The fractional flow of the non-wetting phase at the outlet end of the core is

also given by

Page 970: +Peters Ekwere j. - Petrophysics

8-54

( )( )2

pDnwnwnw

i i

dNdQ tqfq dQ t dW

= = = (8.117)

where Qnw(t) and Qi(t) are the cumulative non-wetting phase produced and

the cumulative wetting phase injected and pDN and iW are their dimensionless

counterparts as fractions of the total pore volume. Eqs.(8.116) and (8.117)

were first presented by Welge (1952). It should be noted that these equations

give no useful information before breakthrough because the fractional flow of

the non-wetting phase at the outlet end of the core is 1 and the relative

permeability to wetting phase is zero. This is why the unsteady state relative

permeability method is limited to only post breakthrough wetting phase

saturations between Swf and 1-Snwr.

After wetting phase breakthrough, we need to associate the computed

relative permeability ratio with the wetting phase saturation at the outlet end

of the core, the point of observation. To determine the wetting phase

saturation at the outlet end of the core, we perform a material balance for the

wetting phase after breakthrough to obtain

( )2 21w wirr pD i w wS S N W F S= + − −⎡ ⎤⎣ ⎦ (8.99)

Eq.(8.99) can be written in terms of the fractional flow of the non-wetting

phase as

2 2w wirr pD i nwS S N W f= + − (8.118)

Using Eqs.(8.116) and (8.118), rw

rnw

kk

versus 2wS can be computed.

Johnson, Bossler and Neumann (JBN, 1959) presented equations for

calculating the individual relative permeability curves by the unsteady state

method by incorporating the pressure drop into the computations. The

pressure drop across the porous medium at time t is given by

Page 971: +Peters Ekwere j. - Petrophysics

8-55

0

L PP dxx

∂Δ = −

∂∫ (8.119)

Darcy’s Law for the non-wetting phase gives

rnwnw

nw

kk A Pqxμ

∂= −

∂ (8.120)

Dividing Eq.(8.120) by q and rearranging gives the pressure gradient as

nwnw

rnw

qP fx kk A

μ⎛ ⎞∂= −⎜ ⎟∂ ⎝ ⎠

(8.121)

Substituting Eq.(8.121) into (8.119) gives

0

Lnw nw

rnw

q fP dxkA kμ⎛ ⎞Δ = ⎜ ⎟

⎝ ⎠ ∫ (8.122)

Applying the Buckley-Leverett frontal advance equation, Eq.(8.77), at the

outlet end of the core after breakthrough gives

( )2w

i w

w S

Q t dfLA dSφ

⎛ ⎞= ⎜ ⎟

⎝ ⎠ (8.90)

Dividing Eq.(8.77) by (8.90) gives

'

'2

w

w

fxL f

= (8.123)

where 'wf and '

2wf are the derivatives of the fractional flow functions at any

distance and at the core outlet, respectively. Differentiating Eq.(8.123) with

respect to 'wf gives

''2

ww

Ldx dff

= (8.124)

Page 972: +Peters Ekwere j. - Petrophysics

8-56

Substituting Eq.(8.124) into (8.122) and rearranging gives

'

2'

' 20

wf nw ww

rnw nw

f PkAfdfk q Lμ

Δ=∫ (8.125)

Let

a constants nw

q kAP Lμ

⎛ ⎞ = =⎜ ⎟Δ⎝ ⎠ (8.126)

Substituting Eq.(8.126) into (8.125) gives

'

2

'2

'

0

wwf nw s

wrnw

q ff Pdf

qkP

⎛ ⎞⎜ ⎟Δ⎝ ⎠=

⎛ ⎞⎜ ⎟Δ⎝ ⎠

∫ (8.127)

Let a relative injectivity ratio be defined as

r

s

qPI

qP

⎛ ⎞⎜ ⎟Δ⎝ ⎠=⎛ ⎞⎜ ⎟Δ⎝ ⎠

(8.128)

Substituting Eq.(8.128) into (8.127) gives

'

2'

' 20

wf nw ww

rnw r

f fdfk I

=∫ (8.129)

Differentiating Eq.(8.129) with respect to '2wf gives

'

2 2'2

nw w

rnw w r

f fdk df I

⎛ ⎞= ⎜ ⎟

⎝ ⎠ (8.130)

Substituting Eq.(8.91) into (8.130) gives

2 11

nw

rnw i r

i

f dk W I

dW

⎛ ⎞= ⎜ ⎟⎛ ⎞ ⎝ ⎠

⎜ ⎟⎝ ⎠

(8.131)

Page 973: +Peters Ekwere j. - Petrophysics

8-57

Eq.(8.131) can be used to calculate the relative permeability of the non-

wetting phase as

2

11

nwrnw

i r

i

fkd

W Id

W

=⎛ ⎞⎜ ⎟⎛ ⎞ ⎝ ⎠

⎜ ⎟⎝ ⎠

(8.132)

Knowing the relative permeability of non-wetting phase, the relative

permeability of the wetting phase can be calculated from Eq.(8.116) as

2

1 1wrw rnw

nw nw

k kf

μμ

⎛ ⎞= −⎜ ⎟

⎝ ⎠ (8.133)

The advantage of the unsteady method over the steady state method of

relative permeability measurement is that it is considerably faster. Because

the method is based on the Buckley-Leverett displacement model, the

unsteady state method can only be used to calculate relative permeability

curves between Swf and the wetting phase saturation at the residual

nonwetting phase saturation (1-Snwr) as previously noted. If Swf is high as in

the case of a favorable mobility ratio displacement, then much of the relative

permeability curves cannot be obtained because one is limited to a very small

saturation observation window. To solve this problem, unfavorable mobility

ratio displacements are typically used to determine relative permeability

curves by the unsteady state method. Further, in order to minimize capillary

end effect, high displacement rates are also typically used. The combination

of high rate and adverse mobility ratio can lead to viscous instability that will

make the displacement performance to be rate sensitive. If this happens, the

relative permeability curves obtained by the unsteady state method will be

rate sensitive and can be quite different from the relative permeability curves

of the same porous medium obtained by the steady state method (Peters and

Khataniar, 1987).

Page 974: +Peters Ekwere j. - Petrophysics

8-58

Eqs.(8.117) and (8.132) call for differentiating the measured

experimental data. The challenge in calculating the relative permeability

curves from these equations is to ensure that the curves are smooth. Any type

of finite difference approximation of the derivatives will result in numerical

noise leading to noisy relative permeability curves. The best way to process

the experimental data is by fitting well behaved functions to the experimental

data and then differentiating the functions. Peters and Khataniar (1987) have

suggested the following curve fits, which they have shown to work well.

( ) ( )21 2 3ln lnpD i iN A A W A W= + + (8.134)

2

1 2 31 1 1ln ln lni r i i

B B BW I W W

⎡ ⎤⎛ ⎞ ⎛ ⎞ ⎛ ⎞= + + ⎢ ⎥⎜ ⎟ ⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ ⎝ ⎠⎣ ⎦ (8.135)

Example 8.2

Table 8.2 gives the experimental data for an unsteady state relative

permeability measurement for a sandpack. In the experiment, water was used

to displace a viscous oil at a constant injection rate. The pore volume of water

injected (Wi), the cumulative oil produced (Qo) and the pressure drop across

the sandpack (ΔP) were measured as functions of time.

Table 8.2. Experimental Data for Unsteady State Relative Permeability

Measurements.

Wi Qo ΔP

PV %IOIP psi

0.339 38.28 9.02

0.351 38.95 8.30

0.395 40.10 6.91

0.439 40.91 6.07

0.502 41.92 5.42

0.587 42.95 4.87

Page 975: +Peters Ekwere j. - Petrophysics

8-59

0.670 43.77 4.55

0.840 45.11 4.00

1.137 46.55 3.32

1.604 47.96 2.78

2.029 48.96 2.52

2.624 50.08 2.42

3.225 50.78 2.30

4.346 51.78 2.13

5.719 52.67 1.99

7.092 53.23 1.90

8.464 53.67 1.83

10.516 54.16 1.79

11.203 54.34 1.75

12.578 54.60 1.74

13.271 54.71 1.70

14.644 54.82 1.70

16.016 54.90 1.70

Other data for the experiment are as follows:

Injection rate = 100 cc/hr

Irreducible water saturation = 11.90%

Length of porous medium = 54.7 cm

Diameter of porous medium = 4.8 cm

Average porosity of porous medium = 30.58%

Absolute permeability of porous medium = 3.42 Darcies

Oil viscosity = 108.37 cp

Oil density = 0.959 gm/cm3

Water viscosity = 1.01 cp

Water density = 0.996 gm/cm3

Oil-water interfacial tension = 26.7 dynes/cm

Page 976: +Peters Ekwere j. - Petrophysics

8-60

Effective permeability to oil at irreducible water saturation = 3.16 Darcies

Oil recovery at water breakthrough = 38.28 % IOIP

Final oil recovery at termination of experiment = 54.9 % IOIP

1. Plot graphs of the raw experimental data.

2. Perform the curve fits suggested in Eqs.(8.134) and (8.135) and display

the results graphically.

3. Calculate the oil-water relative permeability curves for the porous

medium using the Johnson-Bossler-Neumann (JBN) method.

4. Plot graphs of the relative permeability curves.

5. Plot the graph of the true fractional flow curve measured in the

experiment.

6. How long was this test?

Solution to Example 8.2

The results of the calculations are summarized in Table 8.3.

1. Figure 8.23 shows the graphs of the raw experimental data.

2. Figures 8.24 and 8.25 show the curve fits of pDN versus ln iW and

1lni rW I

⎛ ⎞⎜ ⎟⎝ ⎠

versus 1lniW

⎛ ⎞⎜ ⎟⎝ ⎠

. The curve fit equations are

( )20.4026 0.0474ln 0.0066 lnpD i iN W W= + +

2

1 1 1ln 2.3600 1.5798ln 0.1130 lni r i iW I W W

⎡ ⎤⎛ ⎞ ⎛ ⎞ ⎛ ⎞= − + + ⎢ ⎥⎜ ⎟ ⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠ ⎝ ⎠⎣ ⎦

These equations can be differentiated analytically to obtain

Page 977: +Peters Ekwere j. - Petrophysics

8-61

( )( )2

0.0474 2 0.0066 lnpD inw

i i

dN Wf

dW W−

= =

( )( ) 2

1 12.3600 1.5798ln 0.1130 ln2

1 12 0.1130 ln1.5798

1 1 1i iW Wi r inw

rnw

i i i

dW I Wf e

kd

W W W

⎛ ⎞⎡ ⎤⎛ ⎞ ⎛ ⎞⎜ ⎟− + + ⎢ ⎥⎜ ⎟ ⎜ ⎟⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠⎣ ⎦⎝ ⎠

⎡ ⎤⎛ ⎞ ⎛ ⎞⎢ ⎥⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠⎢ ⎥= = +⎢ ⎥⎛ ⎞ ⎛ ⎞ ⎛ ⎞⎢ ⎥⎜ ⎟ ⎜ ⎟ ⎜ ⎟⎢ ⎥⎝ ⎠ ⎝ ⎠ ⎝ ⎠⎣ ⎦

Table 8.3. Calculated Results for Example 8.2.

Wi NpD ΔP

PV PV psi ln(Wi) 1ln

iW

⎛ ⎞⎜ ⎟⎝ ⎠

fnw2 Sw2 Ir 1

i rW I 1

lni rW I

⎛ ⎞⎜ ⎟⎝ ⎠

2nw

rnw

f

k krnw krw

0.119 0.924 0.000

0.339 0.337 9.02 -1.082 1.082 0.182 0.395 4.335 0.680 -0.385 0.368 0.494 0.021

0.351 0.343 8.30 -1.047 1.047 0.174 0.401 4.711 0.605 -0.503 0.356 0.490 0.022

0.395 0.353 6.91 -0.929 0.929 0.151 0.413 5.659 0.447 -0.804 0.319 0.473 0.025

0.439 0.360 6.07 -0.823 0.823 0.133 0.421 6.442 0.354 -1.040 0.290 0.457 0.028

0.502 0.369 5.42 -0.689 0.689 0.113 0.432 7.214 0.276 -1.287 0.258 0.436 0.032

0.587 0.378 4.87 -0.533 0.533 0.093 0.443 8.029 0.212 -1.550 0.226 0.411 0.037

0.67 0.386 4.55 -0.400 0.400 0.079 0.452 8.594 0.174 -1.751 0.203 0.388 0.042

0.84 0.397 4.00 -0.174 0.174 0.059 0.467 9.775 0.122 -2.106 0.170 0.349 0.052

1.137 0.410 3.32 0.128 -0.128 0.040 0.483 11.778 0.075 -2.595 0.136 0.295 0.066

1.604 0.423 2.78 0.473 -0.473 0.026 0.500 14.065 0.044 -3.116 0.108 0.237 0.084

2.029 0.431 2.52 0.708 -0.708 0.019 0.512 15.517 0.032 -3.449 0.094 0.199 0.097

2.624 0.441 2.42 0.965 -0.965 0.013 0.526 16.158 0.024 -3.747 0.082 0.162 0.113

3.225 0.447 2.30 1.171 -1.171 0.010 0.534 17.001 0.018 -4.004 0.074 0.135 0.125

4.346 0.456 2.13 1.469 -1.469 0.006 0.547 18.358 0.013 -4.379 0.064 0.100 0.144

5.719 0.464 1.99 1.744 -1.744 0.004 0.559 19.649 0.009 -4.722 0.057 0.074 0.162

7.092 0.469 1.90 1.959 -1.959 0.003 0.566 20.580 0.007 -4.983 0.053 0.057 0.175

8.464 0.473 1.83 2.136 -2.136 0.002 0.573 21.367 0.006 -5.198 0.050 0.045 0.185

10.516 0.477 1.79 2.353 -2.353 0.002 0.580 21.844 0.004 -5.437 0.047 0.033 0.197

11.203 0.479 1.75 2.416 -2.416 0.001 0.582 22.344 0.004 -5.523 0.047 0.030 0.200

12.578 0.481 1.74 2.532 -2.532 0.001 0.586 22.472 0.004 -5.644 0.045 0.025 0.206

Page 978: +Peters Ekwere j. - Petrophysics

8-62

13.271 0.482 1.70 2.586 -2.586 0.001 0.588 23.001 0.003 -5.721 0.045 0.022 0.208

14.644 0.483 1.70 2.684 -2.684 0.001 0.590 23.001 0.003 -5.820 0.044 0.019 0.213

16.016 0.484 1.70 2.774 -2.774 0.001 0.592 23.001 0.003 -5.909 0.043 0.016 0.217

Figure 8.23. Raw experimental data for the unsteady state relative permeability measurements of Example 8.2.

Page 979: +Peters Ekwere j. - Petrophysics

8-63

Figure 8.24. Curve fit of pDN versus ln iW for Example 8.2.

Figure 8.25. Curve fit of 1lni rW I

⎛ ⎞⎜ ⎟⎝ ⎠

versus 1lniW

⎛ ⎞⎜ ⎟⎝ ⎠

for Example 8.2.

3. The oil-water relative permeability data calculated with Eqs.(8.132) and

(8.133) are presented in Table 8.2.

4. Figure 8.26 shows the oil-water relative permeability curves from the

unsteady state experiment. It should be noted that the relative

permeability curves are obtained over the limited saturation range of

0.395 0.592wS≤ ≤ . The relative permeability curves between 0.119wirrS =

and 0.395wfS = cannot be obtained from the experiment. They can only

be obtained by extrapolation of the computed data to the conditions at

the irreducible water saturation where 0.000rwk = and 0.924nwrk = . The

Page 980: +Peters Ekwere j. - Petrophysics

8-64

experiment predicts a residual oil saturation of 40% in this

homogeneous high permeability sand.

Figure 8.26. Computed relative permeability curves for Example 8.2.

5. The true fractional flow curve measured in the experiment is shown in

Figure 8.27. It is interesting to note that it is only that portion of the

true fractional flow curve that is equal to the approximate fractional

flow curve that can be measured in the experiment. If the saturation

profiles in the experiment could be imaged, then it would possible to

calculate the true fractional flow curve between wirrS and wS by the

similarity transformation and the integration outlined in Figures 8.11

and 8.12.

6. The unsteady state experiment lasted 48.48 hours compared to several

weeks for the steady state experiment.

Page 981: +Peters Ekwere j. - Petrophysics

8-65

Figure 8.27. True fractional flow curve measured in the unsteady state experiment of Example 8.2.

8.5 FACTORS AFFECTING RELATIVE PERMEABILITIES

The factors that affect or could affect relative permeability curves

include (1) fluid saturation, (2) fluid saturation history, (3) Wettability, (4)

injection rate, (5) viscosity ratio, (6) interfacial tension, (7) pore structure, (8)

temperature and (9) heterogeneity.

8.5.1 Fluid Saturation

Relative permeabilities are strongly dependent on fluid saturations. The

higher the fluid saturation, the higher the relative permeability to that fluid.

In general, relative permeabilities are nonlinear functions of fluid saturation

as shown in Figures 8.6 and 8.7.

Page 982: +Peters Ekwere j. - Petrophysics

8-66

8.5.2 Saturation History

Like capillary pressure curves, relative permeability curves show

saturation hysteresis. Figure 8.28 shows typical relative permeability curves

for drainage and imbibition. The imbibition non-wetting phase relative

permeability curve is generally lower than the drainage curve at the same

saturations. The imbibition wetting phase relative permeability curve is

slightly greater than the drainage curve. These differences can easily be

explained. During drainage, the non-wetting phase flows through the large

pores displacing the wetting phase along the way. The thin film of wetting

phase that coats the grain surface acts as a lubricant for the flow of the non-

wetting phase. Therefore, the relative permeability to the non-wetting phase

will be high during drainage. That of the wetting phase also will be high

because it starts from 1 and decreases as the non-wetting phase begins to

occupy some of the pores that were previously occupied by the wetting phase.

During imbibition, some of the non-wetting phase will be trapped in the large

pores. This capillary trapping reduces the amount of non-wetting phase

available to flow during imbibition compared to during drainage. It also

reduces the cross-sectional area of the medium occupied by the connected

non-wetting phase. As a result, the imbibition relative permeability to the

non-wetting phase is reduced compared to that during drainage. Because of

capillary trapping of the non-wetting phase during imbibition, the wetting

phase is forced to occupy and flow through pore sizes that are larger than it

would otherwise have flowed if there was no trapping of the non-wetting

phase. This forcing of the wetting phase to flow through larger pores than it

would otherwise have done in the absence of trapping enhances the relative

permeability of the wetting phase on the imbibition cycle compared to the

drainage cycle. These observations are in accord with the experimental results

shown in Figure 8.28.

Page 983: +Peters Ekwere j. - Petrophysics

8-67

Figure 8.28. Relative permeability hysteresis (Osoba et al., 1951).

8.5.3 Wettability

Relative permeability curves are markedly affected by the wettability of

the medium. Jennings (1957) measured steady state oil water relative

permeability curves on a core that was initially strongly water wet. He then

treated the core with a surface active agent (organo chlorosilane) that

rendered the core oil wet and repeated the relative permeability

measurements. The results are shown Figure 8.29. In general, the relative

permeability to oil decreases while the relative permeability to water increases

as the medium changes from a strongly water wet to a strongly oil wet

Page 984: +Peters Ekwere j. - Petrophysics

8-68

medium. It is interesting to replot the relative permeability curves of Figure

8.29 as functions of wetting phase saturation instead of water saturation. The

replotted curves are shown in Figure 8.30. We see that when plotted against

the wetting phase saturation, the relative permeability curves for the oil wet

core and the water wet core are close to each other. They are not identical

because the degree of wettability preference in the two experiments may be

different. However, the relative permeability curves for the wetting phase and

the non-wetting phase from the two experiments are essentially the same.

Based on experimental observations, Craig (1971) gives the following

rules of thumb about the relative permeabilities for water wet and oil wet

media. (1) The irreducible water saturation for a water wet medium is usually

greater than 20% to 25% whereas that of an oil wet medium it is generally

less than 15%, and frequently less than 10%. (2) The water saturation at

which the oil and water relative permeabilities are equal is greater than 50%

for a strongly water wet medium whereas it is less than 50% for a strongly oil

wet medium. (3) The end-point relative permeability to water is generally less

than 30% for a strongly water wet medium and greater than 50% and

approaches the oil end point for a strongly oil wet medium. These

observations are consistent with the effect of wettability on the fluid

distribution and displacement discussed in Section 6.3.4.

It should be emphasized that the above rules of thumb are applicable

only to systems that show a strong preferential wettability to either water or

oil. In general, one cannot infer the wettability of a porous medium based

solely on the relative permeability curves. For example, relative permeability

curves that intersect at a water saturation of 50% does not mean that the

medium is of “neutral” wettability.

Page 985: +Peters Ekwere j. - Petrophysics

8-69

Figure 8.29. Effect of strong preferential wettability on steady state relative permeability curves (Jennings, 1957).

Figure 8.30. Relative permeability curves from Figure 8.28 replotted as functions of wetting phase saturation (adapted from Jennings, 1957).

Page 986: +Peters Ekwere j. - Petrophysics

8-70

At a given saturation, the relative permeability for a phase is higher

when that phase is the non-wetting phase than when it is the wetting phase.

This is observation can be seen in Figure 8.29. At any water saturation, the

relative permeability to water is higher when the water was the non-wetting

phase than when it was the wetting phase. Similarly, at any water saturation,

the relative permeability to oil is higher when the oil was the non-wetting

phase than when it was the wetting phase.

Owens and Archer (1971) measured relative permeability curves of

sandstones that were rendered progressively oil wet with a surface active

agent. Figure 8.31 shows their results for contact angles ranging from 0 to

180°. Note the general decrease in the oil relative permeabilities and increase

in the water relative permeabilities as the system was made progressively

more oil wet. Note also, that the strongly preferentially wet systems with

contact angles of 0 and 180° generally obey Craig’s rules of thumb regarding

the end-point water relative permeability and the water saturation at which

the water and oil relative permeabilities are equal. The rules of thumb do not

strictly apply to the intervening degrees of wettability.

8.5.4 Injection Rate

Injection rate usually does not affect relative permeabilities obtained by

the steady state method provided the rate is sufficiently high to minimize

capillary end effect. However, Peters and Khataniar (1987) have shown that

relative permeabilities obtained by the unsteady state displacement method

can show rate sensitivity due to viscous instability. Figures 8.32 and 8.33

show the effects of rate and viscosity ratio on relative permeability curves for

oil wet and water wet sandpacks. Both the curves for the oil wet medium and

the water wet medium shift to lower water saturations as the injection rate

(stability number) is increased. It should be noted that for the water wet

system, the relative permeability curves obtained by the unsteady state

method deviate from the steady state curves as the degree of instability is

Page 987: +Peters Ekwere j. - Petrophysics

8-71

increased. The water curve increases and the oil curve decreases away from

the steady state curves as the degree of instability of the displacement

experiment increases.

Figure 8.31. Relative permeabilities for range of wetting conditions (Owens

and Archer, 1971).

Page 988: +Peters Ekwere j. - Petrophysics

8-72

Figure 8.32. Effect of stability number on unsteady state relative permeability curves for oil wet sandpacks (Peters and Khataniar, 1987).

Page 989: +Peters Ekwere j. - Petrophysics

8-73

Figure 8.33. Effect of stability number on unsteady state relative permeability

curves for water wet sandpacks (Peters and Khataniar, 1987).

8.5.5 Viscosity Ratio

Viscosity ratio usually does not affect relative permeabilities obtained by

the steady state method since there is no displacement involved. Figure 8.34

shows the relative permeability curves obtained with the steady state method

at various viscosity ratios. Clearly, no viscosity ratio effect is apparent.

However, Peters and Khataniar (1987) have shown that relative permeabilities

obtained by the unsteady state displacement method at adverse viscosity

ratios can show sensitivity to injection rate and viscosity ratio due to viscous

instability (Figures 8.32 and 8.33).

Page 990: +Peters Ekwere j. - Petrophysics

8-74

Figure 8.34. Effect of viscosity ratio on relative permeability curves obtained

by the steady state method (Leverett, 1939).

8.5.6 Interfacial Tension

Relative permeability curves are affected by interfacial tension only at

interfacial tensions lower than 0.1 dyne/cm. Above this value, relative

permeabilities are unaffected by interfacial tension.

Bardon and Longeron (1978) studied the effect of interfacial tension on

gas-oil relative permeabilities using methane and normal heptane

displacement experiments. In their study, interfacial tensions were calculated

using parachors and the gas-oil relative permeabilities were calculated using

relative permeability models and a numerical simulator to history match the

displacement data. Figure 8.35 shows the relative permeability curves that

gave the best fit to the displacement recovery data at the calculated interfacial

tensions. The results show that the relative permeabilities to gas and oil

Page 991: +Peters Ekwere j. - Petrophysics

8-75

increased as the interfacial tension decreased. The residual fluid saturations

decreased as the interfacial tension decreased as expected from the effect of

capillary number on residual fluid saturations. In the limit, at ultra-low

interfacial tensions, the relative permeability curves were approximately

straight lines. These general trends in the effect of interfacial tensions on

relative permeability curves have been confirmed by Amaefule and Handy

(1981).

Figure 8.35. Effect of interfacial tension on gas-oil relative permeability curves

(Bardon and Longeron, 1978).

8.5.7 Pore Structure

Morgan and Gordon (1970) have presented results that show that rocks

with large pores and correspondingly small specific surface areas have low

irreducible water saturations that leave a relatively large amount of pore

Page 992: +Peters Ekwere j. - Petrophysics

8-76

space available for multiphase flow. Therefore, for such rocks, end point

relative permeabilities are high and a large saturation change may occur

during two phase flow. By contrast, rocks with small pores have larger

specific surface areas and larger irreducible water saturations that leave less

room for multiphase flow. As a result, the end point relative permeabilities

are lower and the saturation range for two phase flow is smaller than in rocks

with large pores. Finally, rocks having some relatively large pores connected

by small pores have a large surface area, resulting in high irreducible water

saturation and relative permeability behavior that is similar to rocks with

small pores only. These observations are summarized in Figure 8.36.

8.5.8 Temperature

There are data in the literature that suggest that relative permeability

curves are affected by temperature. Poston et al., (1970) found that

temperature causes residual oil saturation to decrease and irreducible water

saturation to increase, with corresponding increases in relative permeability

curves (Figure 8.37). On the other hand, there are data in the literature that

also show that relative permeabilities are not temperature dependent (Miller

and Ramey, 1985). Apparently, the effect of temperature on relative

permeabilities is still and open question. This situation is understandable

because temperature can affect rock and fluid properties which in turn can

affect relative permeability curves. For example, high temperature can change

the wettability of the rock which affects relative permeabilities. It can also

reduce interfacial tensions, which can affect relative permeabilities and the

irreducible saturations. Because of the effect of temperature on the other

properties of the system that can affect relative permeabilities, it is difficult to

categorically determine the effect of temperature on relative permeabilities.

Page 993: +Peters Ekwere j. - Petrophysics

8-77

Figure 8.36. Effect of pore structure on relative permeability curves; (a)

sandstone with large, well-connected pores with k = 1314 md; (b) sandstone with small, well-connected pores with k = 20 md; (c) sandstone with a few

large pores connected with small pores with k = 36 md (Morgan and Gordon, 1970).

Page 994: +Peters Ekwere j. - Petrophysics

8-78

Figure 8.37. Effect of temperature on relative permeability curves (Poston et

al., 1970).

8.5.9 Heterogeneity

Relative permeabilities are typically measured on homogeneous core

samples. These curves are then used in numerical simulators to model the

performance of heterogeneous reservoirs. It is often necessary to adjust the

laboratory measured relative permeability curves in order to successfully

history match the performance of the heterogeneous reservoirs. Gharbi and

Peters (1993) simulated the waterflood performance of a heterogeneous

reservoir using a set of input relative permeability curves and then used the

simulated oil recovery versus pore volumes of water injected and the

simulated pressure drop to calculate the equivalent relative permeability

curves for the heterogeneous medium by the JBN method. Figure 8.38

Page 995: +Peters Ekwere j. - Petrophysics

8-79

compares the input relative permeabilities with the computed equivalent

relative permeabilities for the heterogeneous medium. The effect of the

heterogeneity is to shift the oil and water relative permeabilities to low water

saturations thereby increasing the water relative permeability curve and

decreasing the oil relative permeability curve. Thus, the relative

permeabilities for the heterogeneous medium are similar to the relative

permeabilities for a strongly oil wet medium.

Figure 8.38. Effect of heterogeneity on relative permeability curves (Gharbi

and Peters, 1993).

8.6 THREE-PHASE RELATIVE PERMEABILITIES

Three phase relative permeabilities are required to predict the

performance of three phase flow of oil, water and gas. There are considerably

less experimental data in the literature on three phase relative permeabilities

than two phase relative permeabilities. Figure 8.39 shows, on a ternary

Page 996: +Peters Ekwere j. - Petrophysics

8-80

saturation diagram, the approximate regions of single phase flow, two phase

flow and three phase flow in an oil, water and gas system ( Leverett and Lewis,

1941). It can be seen that the three phase flow region is small compared to

single phase and two phase flow regions. Figures 8.40, 8.41 and 8.42 show

the three phase water, oil and gas relative permeabilities measured by

Leverett and Lewis (1941). They found that the relative permeability to water

was only a function of the water saturation. However, the relative

permeabilities to oil and gas were functions of all three fluid saturations.

Figure 8.39. Approximate limits of saturations giving 5 per cent or more of all

components in flow stream for the flow of nitrogen, kerosene and brine. Arrows point to increasing fraction of respective components in stream

(Leverett and Lewis, 1941).

Page 997: +Peters Ekwere j. - Petrophysics

8-81

Figure 8.40. Three phase relative permeability to water (Leverett and Lewis,

1941).

Figure 8.41. Three phase relative permeability to oil (Leverett and Lewis,

1941).

Page 998: +Peters Ekwere j. - Petrophysics

8-82

Figure 8.42. Three phase relative permeability to gas (Leverett and Lewis,

1941).

Three phase relative permeabilities are not routinely measured in the

laboratory as two phase relative permeabilities. Instead, three phase relative

permeabilities are usually calculated from two phase relative permeability

data using various relative permeability models. Delshad and Pope (1989)

have reviewed the various three phase relative permeability models and found

that some of them do not always agree with the available experimental three

phase relative permeability data.

8.7 CALCULATION OF RELATIVE PERMEABILITIES FROM DRAINAGE CAPILLARY PRESSURE CURVE

In Section 7.12.2, we derived the following approximate drainage

relative permeability curves for wetting and non-wetting phases from the

drainage capillary pressure curve:

Page 999: +Peters Ekwere j. - Petrophysics

8-83

2

01

20

( )

wSw

cwrw w

w

c

dSPkk S

k dSP

= =∫

∫ (7.159)

and

1

2

1

20

( ) w

w

cSnwrnw w

w

c

dSPkk S

k dSP

= =∫

∫ (7.160)

We found that these models were defective in two respects: (1) they do not

include trapped residual saturations and (2) the sum of the relative

permeabilities is equal to 1, which is contrary to experimental observations.

These deficiencies result from the fact that the models neglect certain facts

about the nature of two phase flow in porous media. First, the cross-sectional

area open to the flow of the wetting phase is not a constant as assumed in the

models but is a function of the wetting phase saturation. Second, the

tortuosity for the flow of the wetting phase, which was neglected in the

models, is also a function of the wetting phase saturation. Burdine (1953)

proposed the following normalized drainage relative permeability models,

which account for these saturation dependencies in the cross-sectional area

and the tortuosity for two phase flow:

( )

*

*2* 2* * 0

1**

20

1( )( )

( 1) 1

wS

wcw w

rw w ww w

wc

dSPk Sk S S

k SdS

P

= ==

∫ (8.136)

( ) *

1*

2* 2* *1*

*2

0

1

( )( ) 1( 0) 1

w

wcSnw w

rnw w wnw w

wc

dSPk Sk S S

k SdS

P

= = −=

∫ (8.137)

Page 1000: +Peters Ekwere j. - Petrophysics

8-84

where *wS is the normalized wetting phase saturation given by

*

1w wirr

wwirr

S SSS

−=

− (8.138)

In Eqs.(8.136) and (8.137), the ratios of the integrals on the right side

account for the cross-sectional area changes with saturations and the terms

( )2*wS and ( )2*1 wS− account for the tortuosity changes with saturations. It

should be noted that the base permeability used in Eq.(8.136) to define the

normalized relative permeability to the wetting phase is equal to the absolute

permeability of the medium, whereas the base permeability used to define the

normalized relative permeability to the non-wetting phase in Eq.(8.137) is

equal to the effective permeability to the non-wetting phase at the irreducible

wetting phase saturation. Thus, the normalized wetting phase relative

permeability given by Eq.(8.136) is also the true relative permeability to the

wetting phase. However, the normalized non-wetting phase relative

permeability given by Eq.(8.137) must be multiplied by the end point relative

permeability to the non-wetting phase (knwr) in order to obtain the true non-

wetting phase relative permeability. Furthermore, the normalized non-wetting

phase relative permeability of Eq.(8.137) starts at a wetting phase saturation

of 1.0 or a non-wetting phase saturation of zero. Normally, a critical non-

wetting phase saturation is required before the non-wetting phase can flow.

Thus, the end-point non-wetting phase relative permeability and a critical

non-wetting phase saturation must be introduced into Eq.(8.137) to obtain

the true relative permeability for the non-wetting phase. Given the drainage

capillary pressure curve, the integrals in Eqs.(8.136) and (8.137) can easily be

calculated numerically to obtain the normalized drainage relative permeability

curves.

An alternative approach to evaluating the integrals in Eqs.(8.136) and

(8.137) is to fit the Brooks-Corey (1966) model to the drainage capillary

Page 1001: +Peters Ekwere j. - Petrophysics

8-85

pressure curve and then integrate the resulting linear function. As discussed

in Section 7.13.1, the Brooks-Corey drainage capillary pressure model is given

by

*ln ln lnw c eS P Pλ λ= − + (7.161)

or

*1ln ln lnc w eP S Pλ

= − + (7.162)

and

( )1

*c e wP P S λ

−= (7.164)

where λ is the pore size distribution index obtained from the straight line

given by Eq.(7.161) or (7.162). Substituting Eq.(7.164) into Eqs.(8.136) and

(8.137) and performing the integrations gives the normalized drainage relative

permeability curves as

( )2 3

*( )rw w wk S Sλ

λ+

= (8.139)

and

( ) ( )22* *( ) 1 1rnw w w wk S S S

λλ+⎡ ⎤

= − −⎢ ⎥⎣ ⎦

(8.140)

A critical saturation can be introduced into the relative permeability model for

the non-wetting phase as

( )2 2

*( ) 1 1w wirrrnw w w

m wirr

S Sk S SS S

λλ+⎛ ⎞ ⎡ ⎤−

= − −⎜ ⎟ ⎢ ⎥− ⎣ ⎦⎝ ⎠ (8.141)

Page 1002: +Peters Ekwere j. - Petrophysics

8-86

where Sm is the wetting phase saturation corresponding to the critical non-

wetting phase saturation. Finally, the true relative permeability curve for the

wetting and non-wetting phases are given by

( )2 3

*( )rw w wk S Sλ

λ+

= (8.142)

( )2 2

*( ) 1 1w wirnw w nwr w

m wi

S Sk S k SS S

λλ+⎛ ⎞ ⎡ ⎤−

= − −⎜ ⎟ ⎢ ⎥− ⎣ ⎦⎝ ⎠ (8.143)

where nwrk is the non-wetting phase relative permeability at the irreducible

wetting phase saturation.

Example 8.3

Use the air-water capillary pressure data of Table 8.4 to calculate the

drainage relative permeability curves by the method of Brooks and Corey for a

core sample.

Table 8.4. Drainage Capillary Pressure Curves for Example 8.3.

Saturation Capillary Pressure

(psi)

1.000 1.973

0.950 2.377

0.900 2.840

0.850 3.377

0.800 4.008

0.750 4.757

0.700 5.663

0.650 6.781

0.600 8.195

0.550 10.039

0.500 12.547

0.450 16.154

0.400 21.787

0.350 31.817

Page 1003: +Peters Ekwere j. - Petrophysics

8-87

0.300 54.691

0.278 78.408

Solution to Example 8.3

Figure 8.43 shows the graph of lnPc versus *ln wS for Swirr = 0.10. The

equation of the resulting straight line is given by

*ln 2.1443ln ln 2.2238c wP S= − +

Therefore,

1 2.1443λ

− = −

0.4664λ =

2.2238eP =

Figure 8.44 shows the graph of *ln wS versus lnPc for Swi = 0.10. It also is

linear and could have been used for the subsequent calculations. The Brooks-

Corey drainage capillary pressure equation is given by

( ) ( )1 2.1443* *2.2238c e w wP P S Sλ

− −= =

Figure 8.45 shows the normalized drainage relative permeability curves

calculated with Eqs.(8.139) and (8.140). Figure 8.46 shows the true drainage

relative permeability curves for Sm = 0.95 and knwr = 0.961. The results of the

calculations are summarized in Table 8.5.

Page 1004: +Peters Ekwere j. - Petrophysics

8-88

Figure 8.43. Log-log graph of Pc versus *wS for Example 8.3.

Figure 8.44. Log-log graph of *wS versus Pc for Example 8.3.

Page 1005: +Peters Ekwere j. - Petrophysics

8-89

Figure 8.45. Normalized drainage relative permeability curves for Example 8.3.

Figure 8.46. True drainage relative permeability curves for Example 8.3.

Page 1006: +Peters Ekwere j. - Petrophysics

8-90

Table 8.5. Results of Drainage Relative Permeability Calculations for Example

8.3

Original Data

Brooks-Corey

Model Drainage Relative Permeability Curves

Pc Pc

Sw psi *wS psi ( )rw wk S ( )rnw wk S rwk rnwk

1.000 1.973 1.000 2.224 1.000 0.000 1.000 0.000

0.950 2.377 0.944 2.514 0.659 0.001 0.659 0.000

0.900 2.840 0.889 2.863 0.424 0.006 0.424 0.002

0.850 3.377 0.833 3.288 0.265 0.017 0.265 0.008

0.800 4.008 0.778 3.812 0.160 0.036 0.160 0.022

0.750 4.757 0.722 4.468 0.093 0.063 0.093 0.044

0.700 5.663 0.667 5.305 0.052 0.098 0.052 0.073

0.650 6.781 0.611 6.393 0.028 0.140 0.028 0.111

0.600 8.195 0.556 7.843 0.014 0.189 0.014 0.156

0.550 10.039 0.500 9.831 0.006 0.244 0.006 0.208

0.500 12.547 0.444 12.656 0.003 0.304 0.003 0.266

0.450 16.154 0.389 16.851 0.001 0.371 0.001 0.331

0.400 21.787 0.333 23.452 0.000 0.443 0.000 0.401

0.350 31.817 0.278 34.672 0.000 0.521 0.000 0.479

0.300 54.691 0.222 55.947 0.000 0.605 0.000 0.562

0.278 78.408 0.198 71.829 0.000 0.643 0.000 0.601

0.250 0.167 103.678 0.000 0.694 0.000 0.652

0.200 0.111 247.331 0.000 0.790 0.000 0.749

0.150 0.056 1093.394 0.000 0.892 0.000 0.852

0.100 0.000 0.000 1.000 0.000 0.962

Page 1007: +Peters Ekwere j. - Petrophysics

8-91

NOMENCLATURE

A = cross sectional area in the flow direction

Bo = oil formation volume factor

Bw = water formation volume factor

fw = fractional flow of wetting phase

fw = fractional flow of water

fnw = fractional flow of non-wetting phase

fnw2 = fractional flow of non-wetting phase at the outlet end of porous medium

fo = fractional flow of oil

Fw = approximate fractional flow of wetting phase

g = gravitational acceleration

Ir = relative injectivity

J = Leverett J-function

k = absolute permeability of the medium

ko = effective permeability to oil

kw = effective permeability to water

kwr = end-point relative permeability to wetting phase

kg = effective permeability to gas

kro = relative permeability to oil

krw = relative permeability to water

krg = relative permeability to gas

krw = relative permeability to wetting phase

krnw= relative permeability to non-wetting phase

knwr= end-point relative permeability to non-wetting phase

L = length

M = mobility ratio

ME = end-point mobility ratio

Ncap = dimensionless capillary to viscous force ratio

Ng = gravity number

NpD = dimensionless cumulative production

Page 1008: +Peters Ekwere j. - Petrophysics

8-92

Nvcap = capillary number

P = pressure

Pc = capillary pressure

Pe = displacement pressure for Brooks-Corey model

Pg = pressure in the gas phase

Pnw= pressure in the non-wetting phase

Po = pressure in the oil phase

Pc/ow = oil-water capillary pressure curve

Pc/go = gas-oil capillary pressure curve

Pc/gw = gas-water capillary pressure curve

Pw = pressure in the water phase

Pw = pressure in the wetting phase

q = total volumetric injection rate

qo = volumetric flow rate of oil

qg = volumetric flow rate of gas

qnw= volumetric flow rate of non-wetting phase

qw = volumetric flow rate of water

qw = volumetric flow rate of wetting phase

Qi = cumulative injection

Qnw = cumulative non-wetting phase produced

Qo = cumulative oil produced

R = oil recovery as a fraction of initial oil in place

Se = effective wetting phase saturation

Sg = gas saturation

So = oil saturation

Sor = residual oil saturation

Sw = water saturation

Sw = wetting phase saturation

Swirr = irreducible wetting phase saturation

Page 1009: +Peters Ekwere j. - Petrophysics

8-93

Swro = wetting phase saturation at which imbibition capillary pressure is zero

Sw2 = wetting phase saturation at the outlet end of porous medium

Snw= non-wetting phase saturation

Snwr = residual non-wetting phase saturation

Swav = average wetting phase saturation

Swav = average water saturation

Swf = frontal saturation

*wS = normalized wetting phase saturation

t = time

tbt = breakthrough time

Dt = dimensionless time

v = flux vector, Darcy velocity vector

vw = Darcy velocity for the wetting phase

vnw = Darcy velocity for the non-wetting phase

x = distance in the direction of flow

fx = distance to the displacement front

Dx = dimensionless distance

Dfx = dimensionless distance to the displacement front

iW = dimensionless pore volume injected

pW = cumulative water produced

WOR = water oil ratio

δx = small length in the neighborhood of the outlet end of porous medium

gρ = density of gas

oρ = density of oil

wρ = density of water

wρ = density of wetting phase

nwρ = density of non-wetting phase

σ = interfacial tension

Page 1010: +Peters Ekwere j. - Petrophysics

8-94

θ = contact angle

λ = pore size distribution index

μ = viscosity

μg = gas viscosity

μο = oil viscosity

μw = water viscosity

μw = wetting phase viscosity

μnw= non-wetting phase viscosity

φ = porosity, fraction

τ = tortuosity

γ = liquid specific gravity

ω = angular velocity of centrifuge

ΔP = pressure drop

ΔPw = pressure drop in the wetting phase

ΔPnw = pressure drop in the non-wetting phase

Γ = pore structure

REFERENCES AND SUGGESTED READINGS

Amaefule, J.O. and Handy, L.L. : “The Effect of Interfacial Tensions on Relative Oil-Water Permeabilities of Consolidated Porous Media,” SPE/DOE 9783, presented at the SPE/DOE 2nd Joint Symposium on Enhanced Oil Recovery of the Society of Petroleum Engineers, Tulsa, OK, April 5-8, 1981.

Anderson, W.G. : “Wettability Literature Survey - Part 5: The Effects of Wettability on Relative Permeability,” J. Pet. Tech. (November 1987) 1453-1468.

Bardon, C. and Longeron, D. : “Influence of Low Interfacial Tensions on Relative Permeability,” SPE 7609, presented at the 53rd Annual Fall Technical Conference and Exhibition of the Society of Petroleum Engineers, Houston, Tx, October 1-3, 1978.

Brooks, R.H. and Corey, A.T. : “Properties of Porous Media Affecting Fluid Flow,” Jour. Irrigation and Drainage Div., Proc. Amer. Soc. Of Civil Engr. (June, 1966) 61-88.

Page 1011: +Peters Ekwere j. - Petrophysics

8-95

Brutsaert, W. : "Some Methods of Calculating Unsaturated Permeability," Trans. of the Amer. Soc. of Agricultural Engineers, Vol. 10 (1967) 400-404.

Buckley, S.E. and Leverett, M.C. : “Mechanism of Fluid Displacement in Sands,” J. Pet. Tech. (May 1941) 107-116.

Burdine, N.T. : “Relative Permeability Calculations From Pore Size Distribution Data,” Trans., AIME (1953) 71-78.

Campbell, G.S. : "A Simple Method for Determining Unsaturated Conductivity from Moisture Retention Data," Soil Sci., Vol. 117, No. 6 (1974) 311-314.

Charbeneau, R.J. and Daniel, D.E. : "Contaminant Transport in Unsaturated Flow," Chapter 15, in Handbook of Hydrology, D.R. Maidment (Ed.), McGraw-Hill Inc., New York, 1993.

Collins, R.E. : Flow of Fluids Through Porous Materials, Research & Engineering Consultants Inc., 1990.

Corey, A.T. : Mechanics of Heterogeneous Fluids in Porous Media, Water Resources Publications, Fort Collins, Colorado, 1977.

Craig, F.F., Jr. : The Reservoir Engineering Aspects of Waterflooding, SPE Monograph Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1971.

Daniel, D.E. : "Permeability Test for Unconsolidated Soil," Geotechnical Testing Journal., Vol. 6, No. 2 (1983) 81-86.

Daniel, D.E., Trautwein, S.J., Boyton, S.S. and Foreman, D.E. : "Permeability Testing with Flexible-Wall Permeabilities for Unconsolidated Soil," Geotechnical Testing Journal., Vol. 7, No. 3 (1984) 113-122.

Delshad, M. and Pope, G.A. : “Comparison of the Three-Phase Oil Relative Permeability Models,” Transport in Porous Media 4 (1989) 59-83.

Douglas, J., Jr., Blair, P.M. and Wagner, R.J. : “Calculation of Linear Waterflood Behavior Including the Effects of Capillary Pressure,” Trans., AIME (1958) 213, 96-102.

Dullien, F.A.L. : Porous Media - Fluid Transport and Pore Structure, Academic Press, New York, 1979.

Dykstra, H. and Parsons, R.L. : “The Prediction of Oil Recovery by Waterflood,” Secondary Recovery of Oil in the United States, American Petroleum Institute (1950) 160-175.

Gharbi, R. and Peters, E.J. : “Scaling Coreflood Experiments to Heterogeneous Reservoirs,” Journal of Petroleum Science and Engineering, 10, (1993) 83-95.

Page 1012: +Peters Ekwere j. - Petrophysics

8-96

Gharbi, R.: Numerical Modeling of Fluid Displacements in Porous Media Assisted by Computed Tomography Imaging, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1993.

Graue, A., Kolltvelt, K., Lien, J.R. and Skauge, A. : "Imaging Fluid Saturation Development in Long-Core Flood Displacements," SPE Formation Evaluation (December 1990) 406-412.

Hardham, W. D.: Computerized Tomography Applied to the Visualization of Fluid Displacements, MS Thesis, University of Texas at Austin, December 1988.

Honarpour, M., Koederitz, L. and Harvey, A.H. : Relative Permeability of Petroleum Reservoirs, CRC Press, Inc., Boca Raton, Florida, 1986.

Jennings, H.Y. : “Surface Properties of Natural and Synthetic Porous Media,” Producers Monthly (March 1957) 20-24.

Jennings, H.J. : “How to Handle and Process Soft and Unconsolidated Cores,” World Oil (June 1965) 116-119.

Johnson, E.F., Bossler, D.P. and Naumann, V.O. : “Calculation of Relative Permeability from Displacement Experiments,” Trans., AIME (1959) 216, 370-372.

Jones, S.C. and Roszelle, W.O. : “Graphical Techniques for Determining Relative Permeability from Displacement Experiments,” J. Pet. Tech. (May 1978) 807-817.

Keelan, D.K. : "A Critical Review of Core Analysis Techniques” The Jour. Can. Pet. Tech. (April-June 1972) 42-55.

Khataniar, S. and E. J. Peters: “The Effect of Heterogeneity on the Performance of Unstable Displacements,” Journal of Petroleum Science and Engineering, 7, No. 3/4 (May 1992) 263-81.

Khataniar, S. and E. J. Peters: “A Comparison of the Finite Difference and Finite Element Methods for Simulating Unstable Displacements,” Journal of Petroleum Science and Engineering, 5, (1991) 205-218.

Khataniar, S. : An Experimental Study of the Effect of Instability on Dynamic Displacement Relative Permeability Measurements, MS Thesis, University of Texas, Austin, Tx, August 1985.

Khataniar, S.: A Numerical Study of the Performance of Unstable Displacements in Heterogeneous Media, Ph.D. Dissertation, University of Texas at Austin, August 1991.

Killins, C.R., Nielsen, R.F. and Calhoun, J.C., Jr.: “Capillary Desaturation and Imbibition in Rocks,” Producers Monthly (February 1953) 18, No. 2, 30-39.

Page 1013: +Peters Ekwere j. - Petrophysics

8-97

Klute, A. : "Water Retention: Laboratory Methods," Methods of Soil Analysis, Part 1, A. Klute (Ed.), American Society of Agronomy, Madison, WI (1986) 635-686.

Klute, A. : "Hydraulic Conductivity and Diffusivity: Laboratory Methods," Methods of Soil Analysis, Part 1, A. Klute (Ed.), American Society of Agronomy, Madison, WI (1986) 687-734.

Kyte, J.R. and Rapoport, L.A. : “Linear Waterflood Behavior and End Effects in Water-Wet Porous Media,” Trans., AIME (1958) 213, 423-426.

Lake, L.W. : Enhanced Oil Recovery, Prentice Hall, Englewood Cliffs, New Jersey, 1989.

Land, C.S. : "Calculation of Imbibition Relative Permeability for Two- and Three-Phase Flow From Rock Properties,” SPEJ (June 1968) 149-156.

Land, C.S. : "Comparison of Calculated with Experimental Imbibition Relative Permeability,” SPEJ (Dec. 1971) 419-425.

Lefebvre du Prey, E.J. : “Factors Affecting Liquid-Liquid Relative Permeabilities of a Consolidated Porous Medium,” Soc. Pet. Eng. J. (Feb. 1973) 39-47.

Leva, M., Weintraub, M., Grummer, M. Pollchick, M. and Storch, H.H. : "Fluid Flow Through Packed and Packed and Fluidized Systems," US Bureau of Mines Bull. No. 504, 1951.

Leverett, M.C. : “Flow of Oil-Water Mixtures through Unconsolidated Sands,” Trans., AIME (1939) 140, xxx-xxx.

Leverett, M.C. : “Capillary Behavior in Porous Solids,” Trans., AIME (1941) 142, 152-169.

Li, Ping: Nuclear Magnetic Resonance Imaging of Fluid Displacements in Porous Media, PhD Dissertation, The University of Texas at Austin, Austin, Texas, August 1997.

Majors, P.D., Li, P. and Peters, E.J. :”NMR Imaging of Immiscible Displacements in Porous Media,” Society of Petroleum Engineers Formation Evaluation (September 1997) 164-169.

Marle, C.M. : Multiphase Flow in Porous Media, Gulf Publishing Company, Houston, Texas, 1981.

Miller, M.A. and Ramey, H.J., Jr. : “Effect of Temperature on Oil/Water Relative Permeability of Unconsolidated and Consolidated Sands,” Soc. Pet. Eng. J. (Dec. 1985) 945-953.

Morrow, N.R., Cram, P.J. and McCaffery, F.G. : "Displacement Studies in Dolomite With Wettability Control by Octanoic Acid," SPEJ (August 1973) 221-232.

Page 1014: +Peters Ekwere j. - Petrophysics

8-98

Morgan, J.T. and Gordon, D.T. : “Influence of Pore Geometry on Water-Oil Relative Permeability,” J. Pet. Tech. (October 1970) 1199-1208.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 2 - Interfacial Phenomena and Oil Recovery: Wettability,” World Oil (March 1981) 77-83.

Mungan, N. : “Enhanced Oil Recovery Using Water as a Driving Fluid; Part 3 - Interfacial Phenomena and Oil Recovery: Capillarity,” World Oil (May 1981) 149-158.

Mungan, N. and Moore, E.J. : "Certain Wettability Effects on Electrical Resisitivity in Porous Media," J. Cdn. Pet. Tech. (Jan.-March 1968) 7, No.1, 20-25.

Muqeem, M.A., Bentsen, R.G. and Maini, B.B. : “Effect of Temperature on Three-phase Water-oil-gas Relative Permeabilities of Unconsolidated Sand,” J. Canadian Pet. Tech. (March 1995) 34-41.

Osoba, J.S., Richardson, J.G., Kerver, J.K., Hafford, J.A. and Blair, P.M. : “Laboratory Measurements of Relative Permeability,” Trans., AIME (1951) 192, 47-56.

Perkins, F.M. : "An Investigation of the Role of Capillary Forces in Laboratory Water Floods," Trans. AIME (1957) Vol. 210, 409-411.

Peters, E.J. : Stability Theory and Viscous Fingering in Porous Media, PhD Dissertation, University of Alberta, Edmonton, Alberta, Canada, January 1979.

Peters, E.J. and Flock, D.L. : “The Onset of Instability During Two-Phase Immiscible Displacement in Porous Media,” Soc. Pet. Eng. J. (April 1981) 249-258; Trans., AIME (1981) 271.

Peters, E. J., J. A. Broman and W. H. Broman, Jr.: “Computer Image Processing: A New Tool for Studying Viscous Fingering in Corefloods,” SPE Reservoir Engineering (November 1987) 720-28

Peters, E. J. and W. D. Hardham: “A Comparison of Unstable Miscible and Immiscible Displacements,” SPE 19640, Proceedings of the 64th Annual Technical Conference of the Society of Petroleum Engineers (October 1989) San Antonio.

Peters, E.J., Afzal, N. and Gharbi, R. : “On Scaling Immiscible Displacements in Permeable Media,” Journal of Petroleum Science and Engineering, 9, (1993) 183-205.

Peters, E.J. and Gharbi, R. : “Numerical Modeling of Laboratory Corefloods,” Journal of Petroleum Science and Engineering, 9, (1993) 207-221.

Page 1015: +Peters Ekwere j. - Petrophysics

8-99

Peters, E.J. and Hardham, W.D. : “Visualization of Fluid Displacements in Porous Media Using Computed Tomography Imaging,” Journal of Petroleum Science and Engineering, 4, No. 2, (May 1990) 155-168.

Peters, E.J. and Khataniar, S. : “The Effect of Instability on Relative Permeability Curves Obtained by the Dynamic-Displacement Method,” SPEFE (Dec. 1987) 469-474.

Pirson, S.J. : Oil Reservoir Engineering, McGraw-Hill Book Company, Inc., New York, 1958.

Poston, S.W., Ysreal, S.C., Hossain, A.K.M.S., Montgomery, E.F., III, and Ramey, H.J., Jr. : “The Effect of Temperature on Irreducible Water Saturation and Relative Permeability of Unconsolidated Sands,” Soc. Pet. Eng. J. (June 1970) 171-180; Trans., AIME (1970) 249.

Purcell, W.R. : “Capillary Pressures - Their Measurement Using Mercury and the Calculation of Permeability There From,” Trans., AIME (1949) 186, 39-48.

Rapoport, L.A. and Leas, W.J. : “Properties of Linear Waterfloods,” Trans., AIME (1953) 198, 139-148.

Richardson, J.G. : “Flow Through Porous Media,” Section 16, Handbook of Fluid Dynamics, Edited by V.I. Streeter, McGraw-Hill Book Company, Inc., New York, 1961.

Richardson, J.G., Kerver, J.K., Hafford, J.A. and Osoba, J.S. : “Laboratory Determination of Relative Permeability,” Trans., AIME (1952) 195, 187-196.

Rose, H.E. : “ An Investigation into the the Laws of Flow of Fluids Through Granular Material,” Proc. Inst. Mech. Eng. (1945) 153, 141-148.

Scheidegger, A.E. : The Physics of Flow Through Porous Media, University of Toronto Press, Toronto, 1960.

Stiles, W.E : “Use of Permeability Distributions in Water Flood Calculations,” Trans., AIME (1949) 186, 9-13.

van Genuchten, M.T. : "A Closed-Form Equation for Predicting the Hydraulic Conductivity of Unsaturated Soils," Soil Sci. Soc. Am. J., Vol. 44 (1980) 892-898.

Vauclin, M. : "Flow of Water and Air: Theoretical and Experimental Aspects," in Unsaturated Flow in Hydrologic Modeling: Theory and Practice, H.J. Morel-Seytoux (Ed.), NATO ASI Series C: Mathematical and Physical Sciences, Vol. 275, 1989.

Welge, H.J. : “A Simplified Method for Computing Oil Recovery by Gas or Water Drive,” Trans., AIME (1952) 195, 91-98.

Page 1016: +Peters Ekwere j. - Petrophysics

8-100

Willhite, G. P. : Waterflooding, SPE Textbook Series Vol. 3, Society of Petroleum Engineers, Richardson, Texas, 1986.

Wyllie, M.R.J. and Spangler, M.B. :”Application of Electrical Resistivity Measurements to Problems of Fluid Flow in Porous Media,” AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 359-403.

Wyllie, M.R.J. and Spangler, M.B. :”Application of Electrical Resistivity Measurements to Problems of Fluid Flow in Porous Media,” AAPG Bull., Vol. 36, No. 2 (Feb. 1952) 359-403.

Page 1017: +Peters Ekwere j. - Petrophysics

APPENDIX A

A Systematic Approach To Dimensional Analysis

Summary

This appendix presents a systematic method for obtaining the

complete set of independent dimensionless groups pertinent to a

problem. The algebraic theory underlying the method is presented. It is

shown that the dimensionless groups occupy the null space of the

dimensional matrix. The eigenvectors of this null space, which constitute

the bases vectors for the null space, contain a complete set of

independent dimensionless π groups. Other complete and independent

dimensionless π groups can be obtained systematically by a careful

navigation through this vector space. An example problem is presented

in detail to demonstrate the method.

Introduction

Dimensional analysis is a powerful tool for solving engineering

problems. It can be used to design cost effective experimental programs,

organize presentations of experimental as well as numerical simulation

results, and to scale the results of model experiments to predict the

performance of a large scale system, the prototype system.

Dimensional analysis is typically covered in a cursory manner in

fluid mechanics courses in which ad hoc techniques are presented for

A-1

Page 1018: +Peters Ekwere j. - Petrophysics

deriving the dimensionless groups. Perhaps, because of this cursory

treatment, most engineers have failed to appreciate the power of the

technique and have therefore not taken advantage of it in their work.

The objective of this appendix is to present a systematic method for

obtaining the complete set of independent dimensionless groups

pertinent to a problem. The algebraic theory underlying the method is

presented. An example problem is presented in detail to demonstrate the

method.

Algebraic Theory of Dimensional Analysis

Let a physical process be described by n variables .

Suppose each of these variables can be expressed dimensionally in terms

of the primary variables of mass (M), length (L) and time (T) such that

dimensionally,

1 2, ,..., nu u u

[ ] i i ia b ciu M L T= (A.1)

The dimensional analysis problem may now be stated as follows. We wish

to determine the exponents of real numbers 1 2, ,..., nx x x such that the

power product 1 21 2 ... nxx x

nu u u is dimensionless; that is, has the dimensions of

M0L0T0.

The dimension of the power product is given by

( ) ( ) ( )1 1 2 2 1 1 2 2 1 1 2 21 2 ... ... ...1 2 ... n n n n n nn a x a x a x b x b x b x c x c x c xxx x

nu u u M L T+ + + + + + + + +⎡ ⎤ =⎣ ⎦ (A.2)

The power product will be dimensionless if and only if the following

conditions are met:

A-2

Page 1019: +Peters Ekwere j. - Petrophysics

1 1 2 2 ... 0n na x a x a x+ + + = (A.3)

1 1 2 2 ... 0n nb x b x b x+ + + = (A.4)

1 1 2 2 ... 0n nc x c x c x+ + + = (A.5)

Eqs. (A.3) to (A.5) are derived from Eq.(A.2) by setting the dimensions of

mass (M), length (L) and time (T) to zero, respectively. Eqs.(A.3) to (A.5)

constitute a homogeneous system of linear algebraic equations for the n

unknowns 1 2, ,..., nx x x

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

. The system of equations is homogeneous because

the right-hand side is zero. Dimensional analysis always gives rise to a

homogeneous system of linear algebraic equations similar to Eqs.(A.3) to

(A.5).

Eqs.(A.3) to (A.5) may be written in matrix notation as

(A.6)

1

21 2 3 4

31 2 3 4

41 2 3 4

... 0

... 0

... 0...

n

n

n

n

xx

a a a a ax

b b b b bx

c c c c c

x

⎡ ⎤⎢ ⎥⎢ ⎥⎡ ⎤⎢ ⎥⎢ ⎥ =⎢ ⎥⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦ ⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

or

0Ax = (A.7)

where the dimensional matrix A is given by

1 2 3 4

1 2 3 4

1 2 3 4

...

...

...

n

n

n

a a a a aA b b b b b

c c c c c

⎡ ⎤⎢ ⎥= ⎢ ⎥⎢ ⎥⎣ ⎦

(A.8)

A-3

Page 1020: +Peters Ekwere j. - Petrophysics

and x is the vector of the unknowns. In general, because there are more

unknowns than equations, the dimensional matrix is usually an

matrix in which . A homogeneous system of linear algebraic

equations with more unknowns than equations usually has an infinite

number of nontrivial solutions. It is these nontrivial solutions that give

rise to the dimensionless π groups we seek.

mxn m n<

If the rank of the dimensional matrix A is r, then we can solve for r

of the variables in terms of the remaing ( )n r− variables. Such a solution

will result in independent dimensionless groups, usually called

dimensionless π groups. Thus, to predict the number of independent

dimensionless groups in advance, we need to form the dimensional

matrix and then determine its rank. The rank r of a matrix A is the

largest submatrix of A with a nonzero determinant. Thus, for the mx

dimensional matrix A, the rank is m or less. We can easily determine the

rank of A by calculating the determinants of all possible

submatrices. If we fail to find an submatrix with a nonzero

determinant, then we calculate the determinants of all possible

submatrices until we determine the rank of A.

(n r−

)1

)

rxr

(1 m

n

mxm

mxm

( )m x− −

Equation (A.6) or (A.7) can be solved by simple row and column

operations on the dimensional matrix A. We know from linear algebra

that an system of homogeneous linear algebraic equations does not

have a unique nontrivial solution. Put another way, we know that an mx

system of homogeneous linear equations has an infinitely large number

of solutions. In the case in which the dimensional matrix has a rank of 3,

it is possible for the system of equations to be reduced to the following

row echelon form by row and column operations:

mxn

n

A-4

Page 1021: +Peters Ekwere j. - Petrophysics

(A.9)

1

2

14 15 1 3

24 25 2 4

34 35 3 5

1 0 0 ... 00 1 0 ... 00 0 1 ... 0

...

n

n

n

n

xx

a a a xa a a xa a a x

x

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎡ ⎤⎢ ⎥⎢ =⎢ ⎥⎢⎢ ⎥⎢ ⎥⎣ ⎦ ⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

⎡ ⎤⎥ ⎢ ⎥⎥ ⎢ ⎥

⎢ ⎥⎣ ⎦

Thus, the initial objective of the solution procedure is to reduce the

dimensional matrix using row and column operations such that an rxr

unit submatrix appears at the beginning of the transformed matrix as

shown in Eq.(A.9).

We can now solve Eq.(A.9) to obtain the following result:

1 14 4 15 5 1... n nx a x a x a x= − − − − (A.10)

2 24 4 25 5 2... n nx a x a x a x= − − − − (A.11)

3 34 4 35 5 3... n nx a x a x a x= − − − − (A.12)

We supplement Eqs.(A.10) to (A.12) with the following additional solutions:

4 4x x= (A.13)

5 5x x= (A.14)

...

n nx x= (A.15)

A-5

Page 1022: +Peters Ekwere j. - Petrophysics

Eqs.(A.10) to (A.15), which constitute the required solution to the

dimensional analysis problem, can be rearranged as a linear combination

of ( vectors as follows: )n r−

1 14 15 1

2 24 25 2

3 34 35 3

4 4 5

5

...1 0 00 1 0

... ... ... ...0 0 1

n

n

n

n

n

x a a ax a a ax a a ax x xx

x

− − −⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− − −⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− − −⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥= + + +⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦⎣ ⎦

x

)

(A.16)

Eq.(A.16) shows that the solution to the system of linear homogeneous

algebraic equations arising in dimensional analysis consists of a linear

combination of vectors. We can identify these vectors (n r−1 2, ,..., n re e e −

00

1

aaa

⎤⎥⎥⎥⎥⎥⎥⎥⎥⎥⎦

)

in

Eq.(A.16) as

(A.17)

14 15 1

24 25 2

34 35 3

1 2, ,...,1 00 1... ... ...0 0

n

n

n

n r

a aa aa a

e e e −

− − −⎡ ⎤ ⎡ ⎤ ⎡⎢ ⎥ ⎢ ⎥ ⎢− − −⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢− − −⎢ ⎥ ⎢ ⎥ ⎢= = =⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥ ⎢ ⎥ ⎢⎣ ⎦ ⎣ ⎦ ⎣

These vectors are the eigenvectors of the null space of the

dimensional matrix. These eigenvectors play a critical role in the

dimensional analysis. They contain a complete set of

(n r−

( )n r−

dimensionless π groups (Buckingham’s π theorem). Because, these

eigenvectors are linearly independent, the dimensionless π groups

derived from them are also linearly independent. Because the

A-6

Page 1023: +Peters Ekwere j. - Petrophysics

eigenvectors are the bases vectors for the null space of the dimensional

matrix, their linear combinations can be used to access every part of the

vector space thereby allowing us to derive other ( )n r− complete and

independent π groups. The π groups derived from the eigenvectors are

complete in the sense that any other π group that we derive will be

found to be a combination of this initial set of π groups. The π groups

are independent in the sense that each π group cannot be obtained by

algebraic manipulation of the other π groups.

Eq.(A.16) shows that solutions to the dimensional analysis problem

can be constructed by arbitrarily choosing the values for 4 5, ,..., nx x x . Such

choices can be used to obtain an initial set of complete and independent

dimensionless π groups. The choices can also be used to transform this

initial complete set of independent dimensionless π groups into new

complete and independent sets, which may be more convenient for

applications than the initial set.

An initial complete set of π groups can be obtained systematically

as follows. Let us choose 4 5 61, ... 0.nx x x x= = = = The solution given by

Eq.(A.16) becomes

1 14

2 24

3 34

4 1

5

10

... ...0n

x ax ax ax ex

x

−⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥= =⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

(A.18)

The corresponding dimensionless π group is given by

A-7

Page 1024: +Peters Ekwere j. - Petrophysics

3414 24 1 0 01 1 2 3 4 5 ...aa a

nu u u u u uπ −− −=

Next, we choose The solution given by Eq.(A.16)

becomes

5 4 61, ... 0.nx x x x= = = =

1 15

2 25

3 35

4 2

5

01

... ...0n

x ax ax ax ex

x

−⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥= =⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

(A.19)

The corresponding dimensionless π group is given by

15 25 35 0 1 02 1 2 3 4 5...

a a anu u u u u uπ − − −=

We continue in this fashion until we finally choose 1nx = ,

The solution given by Eq.(A.16) becomes 4 5 1... 0.nx x x −= = = =

1 1

2 2

3 3

4

5

00

... ...1

n

n

n

n r

n

x ax ax ax ex

x

−⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥ ⎢ ⎥= =⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦

(A.20)

The last dimensionless π group is then given by

1 2 3 0 0 11 2 3 4 5 ...n n na a a

n r nu u u u u uπ − − −− =

A-8

Page 1025: +Peters Ekwere j. - Petrophysics

Transformations of the Dimensionless π Groups

The π groups π1, π2, ..., πn–r are complete and independent. However,

they are not unique. By appropriate choice of the values for x4, x5, ..., xn,

other complete and independent dimensionless π groups can be derived.

However, these new complete sets will be found to be transformations of

1 2 n–r

( )

independent, we must use a different eigenvector in forming each of the

1 2 n–r 1 1

π , π , ..., π . To ensure that the new set of π groups is complete, we

must derive of them. To ensure that the new π groups are

dimensionless groups. For example, if the new set of π groups is

designated as , ..., Π , then Π must contain the eigenvector

n r−

π

Π , Π e

either alone or as a linear combination with one or more of the other

eigenvectors; must contain Π2 2e either alone or in combination with the

other eigenvectors and Πn–r must contain n re − either alone or in

e or more of the other eigenvectors.

am

coefficient is a function of the

fluid properties, the porous medium and the displacement conditions.

et us write the functional relationship as

combination with on

Ex ple Problem

We wish to investigate experimentally the factors that affect

dispersion in porous media. A preliminary analysis shows that dispersion

in porous media in the direction of flow can be characterized by a

longitudinal dispersion coefficient. The preliminary analysis further

indicates that the longitudinal dispersion

L

( )1 , , , , , , ,L s p o o s oD f D u D gμ μ ρ ρ= (A.21)

A-9

Page 1026: +Peters Ekwere j. - Petrophysics

The objective is to derive a complete and independent set of

dimensionless π groups that affect the longitudinal dispersion coefficient

by dimensional analysis.

Procedure

Step 1.

The first step is to determine if the number of variables can be

reduced by combining some of them. For example, gravity segregation of

two fluids always involve the combination of variables (ρo–ρs)g or Δρg. So,

we can replace the three variables ρo , ρs, and g by the one variable Δρg.

Step 2.

Next, we form the power product of all the variables involved in the

problem to obtain

(A.22) ( ) 63 5 71 2 4 a dimensionless constantxx x xx x xs p o o LD u D g Dμ μ ρΔ =

Step 3.

Third, we express each of the variables in Eq.(A.22) in terms of the

primary dimensional variables of mass (M), length (L) and time (T). The

dimensions of each variable in Eq.(A.22) are shown in Table 1.

Step 4.

Fourth, we substitute the dimensions of each variable from Table 1

into Eq.(A.22) and insist that the power product become dimensionless

(i.e., has dimensions of M0L0T0). This step gives rise to the following

A-10

Page 1027: +Peters Ekwere j. - Petrophysics

system of linear homogeneous algebraic equations for the powers x1, x2,

x3, x4, x5, x6 and x7:

1 5 6 0x x x+ + = (A.23)

(A.24) 1 2 3 4 5 6 72 2 2x x x x x x x− + + + − − + = 0

0 (A.25) 1 3 4 5 6 72x x x x x x− − − − − − =

Table 1 Dimensions of the Variables in Example Problem

Variable Symbol Dimensions

Mean grain diameter Dp M0L1T0

Interstitial velocity u M0L1T–1

Diffusion coefficient Do M0L2T–1

Solvent viscosity μs M1L–1T–1

Oil viscosity μo M1L–1T–1

Buoyancy group Δρg M1L–2 T–2

Dispersion coefficient DL M0L2T–1

Eq.(A.23) comes from insisting that the power product have a mass

dimension of M0. Similarly, Eq.(A.24) comes from L0 and Eq.(A.25) comes

from T0. Let us organize Eqs.(A.23), (A.24) and (A.25) into a matrix

equation as follows:

A-11

Page 1028: +Peters Ekwere j. - Petrophysics

(A.26)

1

2

3

4

5

6

1 0 0 0 1 1 0 01 1 1 2 1 2 2 01 0 1 1 1 2 1 0

n

xxxxxxx

⎡ ⎤⎢ ⎥⎢ ⎥⎢ ⎥⎡ ⎤⎢ ⎥⎢− − − ⎢ ⎥⎢⎢ ⎥⎢ ⎥− − − − − −⎣ ⎦ ⎢ ⎥⎢ ⎥⎢ ⎥⎣ ⎦

⎡ ⎤⎥ ⎢ ⎥=⎥ ⎢ ⎥

⎢ ⎥⎣ ⎦

⎥⎥

Eq.(A.26) is a matrix equation similar to Eq.(A.7) where the dimensional

matrix is given in this case by

(A.27) 1 0 0 0 1 1 01 1 1 2 1 2 21 0 1 1 1 2 1

A⎡ ⎤⎢= − − −⎢⎢ ⎥− − − − − −⎣ ⎦

We could have derived the dimensional matrix directly from Step 2.

However, Steps 2 and 3 were presented to show the origin of the

dimensional matrix. Let us combine Steps 3 and 4 into one step and call

it Step 3(revised).

Step 3 (revised).

Next, we derive the dimensional matrix using the information in

Table 1. The result is shown in Table 2.

Table 2. Dimensional Matrix for Example Problem

x1 x2 x3 x4 x5 x6 x7

μs Dp u Do μo Δρg DL

M 1 0 0 0 1 1 0

L –1 1 1 2 –1 –2 2

T –1 0 –1 –1 –1 –2 –1

A-12

Page 1029: +Peters Ekwere j. - Petrophysics

It should be noted that the first row of the dimensional matrix is the

power of M in each variable, the second row is the power of L and the

third row is the power of T. This is what the first column of Table 2 is

designed to show. It should also be noted that the variables and their

powers in Eq.(A.2) have been placed on top of the dimensional matrix to

remind us of the relationship between the variables and the entries in

the dimensional matrix.

Step 4 (revised).

Next, we solve Eq.(A.26) to obtain the powers of the variables that

will make the power product dimensionless. An initial reaction to solving

Eq.(A.26) is to propose the solution x1 = x2 = x3 = x4 = x5 = x6 = x7 = 0.

This is the trivial solution in which we are not interested. Our interest is

in the nontrivial solutions.

There is extensive theory in linear algebra about the solutions of a

system of homogeneous linear algebraic equations such as Eq.(A.26).

Although we cannot review all of this theory at this point, we will use

some of it to enable us to solve Eq.(A.26) to obtain a complete set of

independent dimensionless π groups. Eq.(A.26) has more unknowns (7

unknowns) than equations (3 equations). A system of linear

homogeneous algebraic equations with more unknowns than equations

has many (infinite number of) solutions. These are the solutions we

want.

The dimensional matrix A has a rank of r where r is the largest rxr

submatrix of A with a nonzero determinant. Buckingham’s π–theorem

states that the number of complete and independent dimensionless π

groups is equal to the number of variables minus the rank of the

dimensional matrix, (n–r). Therefore, we can predict the number of

A-13

Page 1030: +Peters Ekwere j. - Petrophysics

complete and independent dimensionless π groups in advance by

determining the rank of the dimensional matrix without actually solving

the system of equations.

To determine the rank of the dimensional matrix in this example, we

compute the determinant of the first 3x3 submatrix of A. This gives a

determinant of –1 as shown in Eq.(A.27).

1 0 0

1 1determinant 1 1 1 1 determinant 1

0 11 0 1

x− =−

− −= − (A.28)

Because the determinant of a 3x3 submatrix of A is nonzero, the rank, r,

of A is 3. Therefore, there will be 7–3 or 4 complete and independent

dimensionless π groups in this example problem. In this example, the

rank of the dimensional matrix , r, is equal to the number of equations,

m. However, we should be aware that it is possible for the rank of the

dimensional matrix to be less than the number of equations if the rows of

the dimensional matrix are linearly dependent. Whatever the rank of A is,

we can find it by examining all possible 3x3 submatrices and if need be,

all possible 2x2 submatrices. If we examine all possible 3x3 submatrices

and fail to find a nonzero determinant, then the rank is obviously less

than 3. We will then examine all possible 2x2 submatrices to see if the

rank is 2.

Let us proceed to solve Eq.(A.26) by Gaussian elimination, which is

a standard method for solving a system of linear algebraic equations.

Gaussian elimination involves only simple row and column operations. In

this example, the variables in the dimensional matrix have been carefully

arranged initially to avoid column exchanges during the solution steps.

Therefore, only simple row operations will be required to solve the system

A-14

Page 1031: +Peters Ekwere j. - Petrophysics

of equations. Our initial objective in the solution process is to perform

the elimination to the point that the first 3x3 submatrix of A is

transformed into a unit matrix of the form

1 0 00 1 00 0 1

(28)

as shown in Eq.(A.9). Because the right hand side of Eq.(A.26) is zero

and will always remain zero no matter how we manipulate the equations,

we can ignore it in the solution procedure until the end and concentrate

on performing the row and column operations on the dimensional matrix

(A) alone.

Let us solve Eq.(A.26) systematically as follows. Add rows 1 and 2

and place the outcome in row 2. The result is shown in Table 3.

Table 3

x1 x2 x3 x4 x5 x6 x7

μs k u Do μo Δρg DL

M 1 0 0 0 1 1 0

L 0 1 1 2 0 –1 2

T –1 0 –1 –1 –1 –2 –1

Add rows 1 and 3 of Table 3 and place the outcome in row 3. The result

is shown in Table 4.

Table 4

x1 x2 x3 x4 x5 x6 x7

μs k u Do μo Δρg DL

A-15

Page 1032: +Peters Ekwere j. - Petrophysics

M 1 0 0 0 1 1 0

L 0 1 1 2 0 –1 2

T 0 0 –1 –1 0 –1 –1

Add rows 2 and 3 of Table 4 and place the outcome in row 2. The result

is shown in Table 5.

Table 5

x1 x2 x3 x4 x5 x6 x7

μs k u Do μo Δρg DL

M 1 0 0 0 1 1 0

L 0 1 0 1 0 –2 1

T 0 0 –1 –1 0 –1 –1

Multiply row 3 of Table 5 by –1 and place the outcome in row 3. The

result is shown in Table 6.

Table 6

x1 x2 x3 x4 x5 x6 x7

μs k u Do μo Δρg DL

M 1 0 0 0 1 1 0

L 0 1 0 1 0 –2 1

T 0 0 1 1 0 1 1

We have achieved our initial objective of obtaining a 3x3 unit

submatrix at the beginning of the modified dimensional matrix as shown

in Table 6. Let us now bring the right hand side of Eq.(A.26) into the

picture to complete our solution. The right hand side of Eq.(A.26) has

remained zero and has been unaffected by the row operations on the

A-16

Page 1033: +Peters Ekwere j. - Petrophysics

dimensional matrix A. With the right hand side of Eq.(A.26) brought into

the picture, we can now solve for the 7 unknowns as follows:

1 5 6x x x= − − (A.29)

2 4 62 7x x x x= − + − (A.30)

3 4 6 7x x x x= − − − (A.31)

4 4x x= (A.32)

5 5x x= (A.33)

6 6x x= (A.34)

7 7x x= (A.35)

The solutions presented in Eqs.(A.29) to (A.35) appear somewhat

confusing and unclear. Let us reorganize them into a linear combination

of the eigenvectors of the null space of the dimensional matrix as follows:

1

2

3

4 4 5 6

5

6

7

0 1 1 01 0 2 11 0 1 1

1 0 0 00 1 0 00 0 1 00 0 0 1

s

p

o

o

L

xxDxux x x xDxxgxD

μ

μρ

− −⎡ ⎤⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎢ ⎥⎢ ⎥

−⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥= + + +⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢

⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦⎣ ⎦ ⎣ ⎦

7x⎥⎥⎥⎥⎥

(A.36)

The variables associated with the solution have been placed in column 1

of Eq.(A.36) for orientation purposes. We can readily identify the (n–r)

eigenvectors of the null space of A in Eq.(A.36) as

A-17

Page 1034: +Peters Ekwere j. - Petrophysics

(A.37) 1 2 3 4

0 1 11 0 21 0 1

, , ,1 0 00 1 0 00 0 10 0 0 1

e e e e

− −⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥− −⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥= = = =⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦

011

0

0

We are now ready to obtain the predicted 4 (n–r) complete and

independent dimensionless π groups from the solutions given in

Eq.(A.36).

Step 5.

Next, we obtain the predicted (n–r) complete and independent

dimensionless π groups by constructing nontrivial solutions of Eq.(A.26)

as follows. Eq.(A.36) shows that nontrivial solutions can be constructed

by arbitrarily choosing numerical values for x4, x5, x6 and x7. Let us

choose x4 = 1, x5 = 0, x6 = 0 and x7 = 0. The solution becomes

1

2

3

4 1

5

6

7

011

1000

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥= =⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.38)

The corresponding dimensionless π group is obtained by substituting this

solution into Eq.(A.25) to obtain

( )00 1 1 1 0 0 01 0s p o L

p

DD u D g DuD

π μ μ ρ− −= Δ = (A.39)

A-18

Page 1035: +Peters Ekwere j. - Petrophysics

The dimensionless π group could have been obtained directly by

inspection of Eq.(A.38). The first column of Eq.(A.38) contains the

variables and the last column contains the powers of the variables in the

dimensionless π group. Therefore, 01

p

DuD

π = could have been written down

directly by inspection. Next, let us choose x5 = 1, x4 = 0, x6 = 0 and x7 =

0. The solution then becomes

1

2

3

4 2

5

6

7

1000100

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

−⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥= =⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.40)

The corresponding dimensionless π group is

( )01 0 0 0 1 0 02 0s p o L

s

D u D g D μπ μ μ ρμ

−= Δ = (A.41)

Next, let us choose x6 = 1, x4 = 0, x5 = 0 and x7 = 0. The solution then

becomes

1

2

3

4 3

5

6

7

121

0010

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

−⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥= =⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.42)

The corresponding dimensionless π group is

A-19

Page 1036: +Peters Ekwere j. - Petrophysics

( )2

11 2 1 0 0 03 0

ps p o L

s

D gD u D g D

π μ μ ρμ

− − Δ= Δ = (A.43)

Finally, let us choose x7 = 1, x4 = 0, x5 = 0 and x6 = 0. The solution given

by Eq.(A.36) then becomes

1

2

3

4 4

5

6

7

011

0001

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥−⎢ ⎥⎢ ⎥ ⎢ ⎥= =⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.44)

The corresponding dimensionless π group is

( )00 1 1 0 0 14 0

Ls p o L

p

DD u D g DuD

π μ μ ρ− −= Δ = (A.45)

The dimensionless groups π1, π2, π3 and π4 constitute a complete set

of independent dimensionless groups. The π groups are complete

because any other dimensionless groups derived from the solutions will

be some combination of π1, π2, π3 and π4. The groups are independent

because each of the π groups cannot be formed by combining the

remaining π groups. For example, π1 cannot be formed by combining π2,

π3 and π4. Similarly, π2 cannot be formed by combining π1, π3 and π4.

Further, π3 cannot be formed by combining π1, π2 and π4. Finally, π4

cannot be formed by combining π1, π2 and π3. A physical argument to

convince oneself that π1, π2, π3 and π4 are independent is to observe that

Do appears only in π1; μo appears only in π2; Δρg appears only in π3; and

DL appears only in π4. Since Do appears only in π1, it is impossible to

A-20

Page 1037: +Peters Ekwere j. - Petrophysics

manipulate π2, π3 and π4 to get π1. A similar reasoning establishes the

independence of π2, π3 and π4.

Transformations of the Dimensionless π Groups for Example

Problem

The dimensionless π groups π1, π2, π3 and π4 are complete and

independent but are not unique. Other legitimate complete sets of π

groups can be formed. However, these other groups will be found to be

transformations of the original basic group π1, π2, π3 and π4.

There are occasions in which the transformations of the original π

groups into new ones are necessary. Let us demonstrate the need for

transformations of the π groups in our example problem. The result of

the dimensional analysis obtained so far is given by

( )4 2 1 2 3, ,fπ π π π= (A.46)

or

2

2 , , po oL

p p s s

D gDD fuD uD u

ρμμ μ

⎛ ⎞Δ= ⎜⎜

⎝ ⎠⎟⎟ (A.47)

Suppose we wish to conduct an experimental study to investigate the

variation of π4 as a function of π1, π2 and π3. In other words, we wish to

establish the nature of the function f2 experimentally. A good

experimental strategy requires that we vary one π group at a time to

determine its influence. For example, we should vary π1 while keeping π2

and π3 constant, then vary π2 while keeping π1 and π3 constant and

finally, vary π3 while keeping π1 and π2 constant.

A-21

Page 1038: +Peters Ekwere j. - Petrophysics

For a given fluid pair and porous medium, μo, μs, Do, Δρ, and of

course, g are fixed. The only variable available for varying π1 is u.

However, we see in Eq.(A.47) that u occurs in π4, π1 and π3. The

occurrence of u in π4 poses no problem. However, its occurrence in π1

and π3 poses a serious problem in controlling the experiments if we were

to plan the experiments based on the current π groups. A change in u

will simultaneously change π1 and π3. Therefore, it would impossible to

tell if an observed change in π4 was caused by the change in π1 or by the

change in π3. What we need is a new set of complete and independent π

groups that allows us to exert more experimental control by being able to

vary one π group while keeping the other π groups constant. Such a new

set of dimensionless π groups can be obtained by transforming the

current set of π1, π2, π3 and π4.

pD

To perform the transformation systematically, we return to the

vector space spanned by the eigenvectors 1e , 2e , 3e and 4e . We recall that

the complete set of dimensionless π groups was obtained by our

judicious choice of the values of x4, x5, x6 and x7 to obtain the solutions

to Eq.(A.26). We can choose these values in some fashion to access

different parts of the vector space spanned by 1e , 2e , and . This

vector space contains an infinite set of complete and independent (n–r)

dimensionless π groups. The challenge in the transformations is to find

these complete sets of dimensionless π groups without including the

linearly dependent ones. Careful navigation through the vector space

allows us to meet this challenge.

3e 4e

Let us form a new set of complete and independent dimensionless π

groups that we shall call Π1, Π2, Π3 and Π4 by choosing the values x4, x5,

A-22

Page 1039: +Peters Ekwere j. - Petrophysics

x6 and x7 in the following manner. Let us choose x4 = –1, x5 = 0, x6 = 0

and x7 = 0. The solution given by Eq.(A.36) becomes

1

2

3

4 1

5

6

7

0111

000

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥= − = −⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.48)

The corresponding dimensionless π group is

( )00 1 1 1 0 01 0

0 1

1ps p o L

uDD u D g D

Dμ μ ρ

π−Π = Δ = = (A.49)

Let us choose x5 = 1, x4 = x6 = x7 = 0. The solution then becomes

1

2

3

4 2

5

6

7

1000100

s

p

o

o

L

xxDxux eDxxgxD

μ

μρ

−⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥= =⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.50)

The corresponding dimensionless π group is

( )01 0 0 0 1 0 02 0s p o L

s

D u D g D 2μμ μ ρ πμ

−Π = Δ = = (A.51)

Let us choose x6 = 1, x4 = –1, x5 = 0 and x7 = 0. The solution becomes

A-23

Page 1040: +Peters Ekwere j. - Petrophysics

1

2

3

4 1 3

5

6

7

1301

010

s

p

o

o

L

xxDxux e eDxxgxD

μ

μρ

−⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥= − + = −⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.52)

The corresponding dimensionless π group is

( )3

11 3 0 1 0 0 33 0

1

ps p o L

o s

D gD u D g D

Dρ πμ μ ρμ π

− − ΔΠ = Δ = = (A.53)

Finally, let us choose x7 = 1, x4 = –1, x5 = 0 and x6 = 0. The solution then

becomes

1

2

3

4 1 4

5

6

7

0001

001

s

p

o

o

L

xxDxux e eDxxgxD

μ

μρ

⎡ ⎤⎡ ⎤ ⎡ ⎤⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥= − + = −⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥

Δ ⎢ ⎥⎢ ⎥ ⎢ ⎥⎢ ⎥⎢ ⎥ ⎢ ⎥⎣ ⎦⎣ ⎦ ⎣ ⎦

(A.54)

The corresponding dimensionless π group is

( )00 0 0 1 0 1 44 0

1

Ls p o L

o

DD u D g DD

πμ μ ρπ

−Π = Δ = = (A.55)

The dimensionless groups Π1, Π2, Π3 and Π4 constitute a new set of

complete and independent dimensionless groups. However, note that

they are transformations of the original complete set of π1, π2, π3 and π4.

The new set was obtained by a careful manipulation of the vector space

A-24

Page 1041: +Peters Ekwere j. - Petrophysics

as a linear combination of the eigenvectors. To ensure that the π groups

are complete, we formed the theoretically required (n–r) new groups. To

ensure that the groups are linearly independent, we used a different

eigenvector to form the different π groups. We used 1e to form Π1, we

used to form Π2, we used 2e 3e (in combination with 1e ) to form Π3, and

we used (in combination with 4e 1e ) to form Π4. These choices of

eigenvectors and eigenvalues preserve the linear independence of the

dimensionless π groups.

The transformations lead to the result

( )4 3 1 2 3, ,fΠ = Π Π Π (A.56)

or

3

3 , ,p poL

o o s o

uD D gD fD D D

ρμμ μ

⎛ ⎞Δ= ⎜⎜

⎝ ⎠s⎟⎟ (A.57)

We see that the new set of dimensionless π groups allows us more

experimental control. Note that u now appears in Π1 only. It is now

possible to vary Π1 while keeping Π2 and Π3 constant. Similarly, we can

vary Π2 while keeping Π1 and Π3 constant. We can also vary Π3 while

keeping Π1 and Π2 constant. Therefore, we are now in a position to

experimentally investigate the separate effects of Π1, Π2 and Π3 on Π4.

In the laboratory experiment for determining the dispersion

coefficient, the viscosity of the injected solvent must be matched with

that of the displaced fluid. Therefore, o

s

μμ

⎛ ⎞⎜⎝ ⎠

⎟ is a constant equal to 1. The

effect of this dimensionless group is eliminated from the experiment.

A-25

Page 1042: +Peters Ekwere j. - Petrophysics

Also, the density of the injected solvent must be matched with that of the

displaced fluid. Therefore, the dimensionless group 3p

o s

D gD

ρμ

⎛ ⎞Δ⎜⎜⎝ ⎠

⎟⎟ is a

constant equal to zero. Its effect is eliminated from the experiment. With

these restrictions imposed on the experiments, Eq.(A.57) becomes

4pL

o o

uDD fD D

⎛ ⎞= ⎜

⎝ ⎠⎟ (A.58)

A similar equation can be written for the transverse dispersion coefficient

as

5pT

o o

uDD fD D

⎛ ⎞= ⎜

⎝ ⎠⎟ (A.59)

The experimental program will then consist of determining the functions

f4 and f5 by varying the Peclet Number, p

o

uDD

⎛ ⎞⎜⎝ ⎠

⎟ . The results of such

experimental programs are shown in Figures 5.27 and 5.28, which are

reproduced here for convenience.

Figure 5.27. Correlations for dimensionless longitudinal and transverse

A-26

Page 1043: +Peters Ekwere j. - Petrophysics

dispersion coefficients plotted on the same scale (Perkins and Johnson, 1963).

Figure 5.28. Correlation for dimensionless longitudinal dispersion coefficient by various authors (Pfannkuch, 1963; Saffman, 1960).

Some Practical Considerations

There are several interesting observations that can be made about

solving the dimensional analysis problem.

1. In general, there are usually more unknowns than equations. In

the example problem, there are 7 unknowns and 3 equations. This

is an over-determined system with an infinite number of nontrivial

solutions.

2. Reduction of the dimensional matrix to row echelon form without

any row becoming zero shows that the rank r of the dimensional

A-27

Page 1044: +Peters Ekwere j. - Petrophysics

matrix is equal to the number of equations. In the example

problem, the rank of the dimensional matrix was determined in

advance to be equal to 3. This is confirmed by the fact that upon

reduction to row echelon form (Table 6), none of the rows became

zero. Thus, the rank of the dimensional matrix in this case is equal

to the number of equations. If one of the rows had become zero

during the reduction to row echelon form, then the rank of the

dimensionless matrix would have been 3-1 or 2. A row will become

zero during the row operations if that row is a linear combination

of the other rows. In such a case, no new information is contained

in the extra row. These are important considerations because

according to Buckingham’s π-theorem, the number of independent

dimensionless groups is equal to the number of variables minus

the rank of the dimensional matrix.

3. Because there are n unknowns and r equations, r unknowns can

be solved for in terms of the remaining (n–r) unknowns. In the

example problem, we solved for x1, x2 and x3 in terms of x4, x5, x6

and x7.

4. The choice of which r variables to solve for in terms of the

remaining (n–r) variables can be used to good advantage. The

variables whose powers are given by the remaining (n–r) unknowns

will each appear in only one dimensionless group. Thus, this

choice can be used to isolate the most important variables of the

problem into one dimensionless group. In the example problem, in

forming the initial dimensionless π groups, we solved for x1, x2 and

x3 in terms of x4, x5, x6 and x7. Therefore, in the initial set of

dimensionless π groups, Do whose power is x4 appeared only in π1;

μo whose power is x5 appeared only in π2; Δρg whose power is x6

appeared only in π3; and DL whose power is x7 appears only in π4.

A-28

Page 1045: +Peters Ekwere j. - Petrophysics

On the other hand, the variables whose powers are given by the r

variables solved for can appear in more than one dimensionless

group.

5. The initial arrangement of the variables is important in

constructing the dimensional matrix. Since the initial objective of

the solution of the homogeneous system of the linear algebraic

equations arising in dimensional analysis is to reduce the first

mxm or rxr submatrix of the dimensional matrix to a unit matrix,

it is advantageous to arrange the variables in such a way that the

first mxm or rxr submatrix of A is a diagonal matrix if possible.

There are two significant advantages from such an arrangement.

The first advantage is that the rank of the matrix can be

determined easily by inspection because the determinant of a

diagonal matrix is equal to the product of the diagonal terms. The

second advantage is that solution of the resulting system of

homogeneous system of algebraic equations is considerably

simplified if the first rxr submatrix of the dimensional matrix is a

diagonal matrix. Reduction to row echelon is then easily achieved

by dividing each row of the dimensional matrix by the first nonzero

entry of that row. Of course, the dimensional analysis can still be

performed with the variables arranged in any order. In such a case,

both row and column operations may be necessary to solve the

resulting homogeneous system of equations. Finally, since the

dependent variable of the problem needs to be isolated into only

one dimensionless π group, this variable should be placed in the

rightmost column of the dimensional matrix. In the example

problem, the dependent variable DL was placed in the last column

of the dimensional matrix to ensure that it appeared in one

dimensionless π group only.

A-29

Page 1046: +Peters Ekwere j. - Petrophysics

Concluding Remarks

A systematic procedure has been presented to derive the complete

and independent set of dimensionless π groups pertinent to a given

problem. The theoretical basis of the method has been presented and

discussed. It is shown that a complete set of independent dimensionless

π groups is given by the eigenvectors of the null space of the dimensional

matrix. These eigenvectors contain a set of complete and independent

dimensionless π groups. By appropriate choice of eigenvalues, the initial

set of dimensionless π groups can be transformed systematically to

obtain other complete and independent dimensionless π groups that may

be more convenient to use in experiments than the initial set. It is hoped

that the systematic approach to dimensional analysis presented here will

inspire those who have hitherto neglected this powerful tool to employ it

in solving their engineering and scientific problems.

Nomenclature

DL = longitudinal dispersion coefficient

Dp = mean grain diameter of the porous medium

f1 = an unknown function we wish to determine experimentally

f2 = an unknown function we wish to determine experimentally

f3 = an unknown function we wish to determine experimentally

f4 = an unknown function we wish to determine experimentally

f5 = an unknown function we wish to determine experimentally

Do = binary diffusion coefficient between the solvent and oil

g = gravitational acceleration

u = interstitial displacement velocity

μo = viscosity of the displaced fluid

A-30

Page 1047: +Peters Ekwere j. - Petrophysics

μs = solvent viscosity

ρo = density of the displaced fluid

ρs = solvent density

π = initial set of dimensionless groups (has nothing to do with 3.141..)

Π = transformed set of dimensionless groups

References

Birkoff, G.: “Dimensional Analysis and Partial Differential Equations,” Elec. Eng., Vol. 67, p. 1185-1188, 1948.

Brand, L.: “The pi-Theorem of Dimensional Analysis,” Am. Math. Monthly, Vol. 59, p. 516, 1952.

Bridgman, P.W.: Dimensional Analysis, Yale University Press, New Haven, 1937.

Buckingham, E. : “On Physically Similar Systems: Illustrations of the use of Dimensional Equations,” Phys. Rev., Vol. 4, pp. 345-376, 1914.

Catchpole, J.P. and Fulford, G.: “Dimensionless Groups,” Industrial and Engineering Chemistry, Vol. 58, No. 3 (March 1966) 46–60.

Chen, W.: “Algebraic Theory of Dimensional Analysis,” J. Franklin Inst., Vol. 292, No. 6, pp. 403-422, Dec. 1948.

Corrsin, S.: “A Simple Geometrical Proof of Buckingham’s p-Theorem,” Am. J. Phys., Vol. 19, pp. 180-181, 1951.

Decius, J.C.: “Dimensional Analysis: An Approach from Transformation Theory and a Criterion for Scaling Model Experiments,” J. Franklin Inst., Vol. 245, pp. 379-387, 1948.

Drobot, S.: “On the Foundations of Dimensional Analysis,” Studia Math., Vol. 14, pp. 84-99, 1953.

Happ, W.W.: “Computer-oriented Procedures for Dimensional Analysis,” J. Appl. Phys., Vol. 38, pp. 3918-3926, 1967.

Hellums, J.D. and Churchill, S.W.: “Dimensional Analysis and Natural Circulation,” Chemical Engineering Symposium Series, pp. 75-80.

Huntley, H.E.: Dimensional Analysis, New York, Dover, 1967.

Langhaar, H.L.: “ A Summary of Dimensional Analysis,” J. Franklin Inst., Vol. 242, pp. 459-463, 1946.

Langhaar, H.L.: Dimensional Analysis and Theory of Models, John Wiley and Sons Inc., New York, 1953.

A-31

Page 1048: +Peters Ekwere j. - Petrophysics

Luchsinger, W.: “Dimensional Analysis,” Brown Boveri Rev., pp. 175-180, 1940.

Marle, C.M.: Multiphase Flow in Porous Media, Gulf Publishing Company, 1981.

Moon, P. and Spencer, D.E.: “A Modern Approach to Dimensions,” J. Franklin Inst., Vol. 248, pp. 495-521, 1949.

Pankhurst, R.C.: Dimensional Analysis and Scale Factors, Chapman and Hall Ltd., London, 1964.

Schmidt, R. and Housen, K.: “Problem Solving with Dimensional Analysis,” The Industrial Physicist, 1995, 21–24.

Rayleigh, J.W.S., Lord: “On the Stability of the Flow of Fluids,” Phil. Mag., 34, 59, 1892.

Rayleigh, J.W.S., Lord: “The Principle of Similitude,” Nature, 95, 66, 1915.

Ruark, A.E.: “Inspectional Analysis: A Method Which Supplements Dimensional Analysis,” J. Elisha Mitchell Sci. Soc., 51, 127 (1935).

Sedov, L.I.: Similarity and Dimensional Methods in Mechanics, New York, Academic Press, 1959.

Silberberg, I.H. and McKetta, J.J., Jr.: “Learning to Use Dimensional Analysis, Part I. Which Dimensional System Should You Use?” Petroleum Refiner, Vol. 32, No. 3, 179-183, April 1953.

Silberberg, I.H. and McKetta, J.J., Jr.: “Learning to Use Dimensional Analysis, Part II. The Correct Use of Dimensional Systems” Petroleum Refiner, Vol. 32, No. 3, 147-150, May 1953.

Silberberg, I.H. and McKetta, J.J., Jr.: “Learning to Use Dimensional Analysis, Part III. The Rayleigh Method?” Petroleum Refiner, Vol. 32, No. ??, 101-103, June 1953.

Silberberg, I.H. and McKetta, J.J., Jr.: “Learning to Use Dimensional Analysis, Part IV. The Buckingham Pi Theorem” Petroleum Refiner, Vol. 32, No. 3, 179-183???, July 1953.

Shook, M., Li, D. and Lake, L.W.: “Scaling Immiscible Flow Through Permeable Media by Inspectional Analysis,” In Situ (1992) 16, No. 4, 311-349.

Sloan, A.D. and Happ, W.W.: “Literature Search: Dimensional Analysis,” NASA Rept. ERC/CQD 68-631, Aug. 1968.

Sloan, A.D. and Happ, W.W.: “Matrix Methods of Dimensional Analysis,” NASA Rept. ERC/CQD 68-6231, Sept. 1968.

A-32

Page 1049: +Peters Ekwere j. - Petrophysics

A-33

Sloan, A.D. and Happ, W.W.: “Computer Program for Dimensional Analysis,” NASA Rept. ERC/CQD 68-621, Aug. 1968.

Van Driest, E.R.: “On Dimensional Analysis and the Presentation of Data in Fluid Flow Problems,” J. Appl. Mech., Vol. 68, A-34 (March 1946) pp. A-34-A40.

Zlokarnik, Marko: Dimensional Analysis and Scale–up in Chemical Engineering, Springer–Verlag, New York, 1991.