You are on page 1of 532

WATERFLOODING

By
William M. Cobb
James T. Smith

11/10

COPYRIGHT
By
William M. Cobb & Associates, Inc.
12770 Coit Road, Suite 907
Dallas, TX 75251
Telephone: (972) 385-0354
Fax: (972) 788-5165
E-Mail: office@wmcobb.com

ALL RIGHTS RESERVED


This book, or any part thereof, may not be reproduced
in any form without permission of William M. Cobb & Associates, Inc.

05/11

TABLE OF CONTENTS
PAGE
I. INTRODUCTION
The End of Primary Depletion ................................................................. 1-2
Factors Controlling Waterflood Recovery .............................................. 1-3
Waterflooding versus Pressure Maintenance ......................................... 1-5
Other References ....................................................................................... 1-6
II. REVIEW OF ROCK PROPERTIES AND FLUID FLOW
Wettability .................................................................................................. 2-1
Definition............................................................................................... 2-1
Importance............................................................................................ 2-3
Determination ....................................................................................... 2-4
Factors Affecting Reservoir Wettability ............................................ 2-5
Sandstone and Carbonates .................................................................. 2-6
Native-State, Cleaned, and Restored-State Cores............................. 2-6
Capillary Pressure ..................................................................................... 2-7
Definition............................................................................................... 2-7
Importance............................................................................................ 2-7
Sources of Data ..................................................................................... 2-7
Effect of Reservoir Variables .............................................................. 2-9
Fluid Saturation............................................................................... 2-9
Saturation History ........................................................................... 2-10
Pore Geometry ................................................................................. 2-11
Averaging of Data ................................................................................ 2-11
J-function ......................................................................................... 2-12
Correlate with Permeability ........................................................... 2-14
Relative Permeability ................................................................................ 2-17
Definition............................................................................................... 2-17
Air Permeability .............................................................................. 2-18
Absolute Permeability ..................................................................... 2-18
Effective Permeability ..................................................................... 2-18
Relative Permeability ...................................................................... 2-18
Importance............................................................................................ 2-19
Sources of Data ..................................................................................... 2-19
Effect of Reservoir Variables .............................................................. 2-20
Saturation History ........................................................................... 2-20
Wettability........................................................................................ 2-21
End-Point Values ................................................................................. 2-23
Averaging of Data ................................................................................ 2-24
PAGE

iii

Date Averaging Methods ................................................................ 2-24


Adjust Average Data to Account for Different Irreducible
Water Saturations ......................................................................... 2-25
Default Relative Permeability Relationships ..................................... 2-29
Problem ...................................................................................................... 2-38
III.

WATERFLOODABLE OIL IN PLACE


Oil Saturation............................................................................................. 3-2
Porosity ....................................................................................................... 3-6
Net Pay ........................................................................................................ 3-8
Net Pay Determination Using Air Permeability versus Oil
Permeability....................................................................................... 3-12
Net Pay Determination after Accounting for Data Scatter .............. 3-18
George and Stiles Fieldwide Net Pay Method .............................. 3-19
George and Stiles Individual Well Net Pay Method
(Net to Gross Method) ................................................................... 3-24
Waterflood Permeability Cutoff Determination Using a
Water Cut Method ............................................................................ 3-29
Comparison of Original Oil-In-Place Material Balance Versus
Volumetric Estimates........................................................................ 3-40
Primary Production Net Pay versus Secondary Floodable
Net Pay ............................................................................................... 3-41
Summary .................................................................................................... 3-45
Problems ..................................................................................................... 3-47

IV. MECHANISM OF IMMISCIBLE FLUID DISPLACEMENT


Introduction ............................................................................................... 4-1
Reservoir Response Incompressible vs. Slightly Compressible
Liquids ..................................................................................................... 4-5
Fractional Flow Equation ......................................................................... 4-8
Effect of Wettability ............................................................................. 4-17
Effect of Formation Dip and Direction of Displacement.................. 4-19
Effect of Capillary Pressure ................................................................ 4-20
Effect of Oil and Water Mobilities ..................................................... 4-21
Effect of Rate ........................................................................................ 4-22
Variations of Fractional Flow Equation ............................................ 4-23
Frontal Advance Equation ....................................................................... 4-24
Welge Analysis of the Buckley-Leverett Theory in Linear Systems .... 4-26
Welge Method Saturation at Flood Front ...................................... 4-27
PAGE

iv

Welge Method Average Water Saturation ..................................... 4-30


Performance at Water Breakthrough ........................................... 4-32
Performance after Breakthrough .................................................. 4-40
Application to Radial Flow ................................................................. 4-47
Effect of Free Gas Saturation ............................................................. 4-48
Production Performance................................................................. 4-55
Displacement Efficiency.................................................................. 4-55
Conditions for Development of an Oil Bank ................................ 4-56
Properties Fluid PVT ............................................................................. 4-59
Reservoir Pressure Distribution............................................................... 4-64
Maintain Optimum Reservoir Pressure to Minimize Sor ................... 4-74
Gravity Under-Running............................................................................ 4-75
Summary .................................................................................................... 4-76
Appendix A Development of Frontal Advance Equation ................... 4-79
Appendix B Buckley-Leverett Theory .................................................. 4-83
Buckley-Leverett Theory..................................................................... 4-83
Stabilized Zone Concept ...................................................................... 4-85
Welge Solution to Buckley-Leverett ................................................... 4-89
Water Saturation at the Front ....................................................... 4-89
Average Water Saturation ............................................................. 4-92
Problems ..................................................................................................... 4-98
V. FLOOD PATTERNS AND AREAL SWEEP EFFICIENCY
Introduction ............................................................................................... 5-1
Mobility Ratio ............................................................................................ 5-1
Water Displacing Oil ........................................................................... 5-2
Water-Oil Mobility Ratio after Breakthrough ................................. 5-6
Oil Displacing Gas................................................................................ 5-7
Basic Flood Patterns .................................................................................. 5-9
Direct Line Drive.................................................................................. 5-10
Staggered Line Drive ........................................................................... 5-11
Five-Spot ............................................................................................... 5-12
Nine-Spot............................................................................................... 5-13
Seven-Spot............................................................................................. 5-14
Areal Sweep Efficiency ............................................................................. 5-15
Causes and Effects ............................................................................... 5-16
Areal Sweep Efficiency at Breakthrough ................................................ 5-23
Isolated Pattern ............................................................................... 5-24
PAGE
Developed Pattern ........................................................................... 5-24

Normal Pattern ................................................................................ 5-24


Inverted Pattern ................................................................................... 5-24
Areal Sweep Efficiency after Breakthrough ..................................... 5-29
Effect of Free Gas Saturation on Areal Sweep.................................. 5-37
Water Zone ...................................................................................... 5-37
Oil Zone (Oil Bank) ......................................................................... 5-38
Re-Saturation Effects ...................................................................... 5-42
Other Factors Affecting Areal Sweep Efficiency .............................. 5-43
Fractures .......................................................................................... 5-43
Directional Permeability ................................................................. 5-43
Areal Permeability Variations ....................................................... 5-44
Formation Dip ................................................................................. 5-44
Off-Pattern Wells ............................................................................ 5-44
Sweep Beyond Edge Wells .............................................................. 5-45
Isolated Patterns .............................................................................. 5-45
Irregularly Spaced Wells ................................................................ 5-46
Peripheral and Line Floods ...................................................................... 5-47
Selection of Waterflood Pattern ............................................................... 5-48
Summary .................................................................................................... 5-49
Problems ..................................................................................................... 5-51
VI. INJECTION RATES AND PRESSURES
Factors Affecting Water Injection Rate .................................................. 6-1
Radial System, Unequal Mobilities .......................................................... 6-2
Regular Patterns ........................................................................................ 6-6
Unit Mobility Ratio .............................................................................. 6-6
Non-Unit Mobility Ratio...................................................................... 6-10
Regular Patterns, Unequal Mobilities ................................................ 6-16
Injectivity in Five-Spot Patterns .............................................................. 6-16
Prats, et al Method ............................................................................... 6-16
Craig Method........................................................................................ 6-17
Problem ...................................................................................................... 6-20
VII.

RESERVOIR HETEROGENEITY
Areal Permeability Variations ................................................................. 7-1
Detection of Areal Permeability Variations ...................................... 7-2
Effect of Areal Permeability Variations ............................................ 7-3
Vertical Permeability Variations ............................................................. 7-3
PAGE
Detection of Stratification ................................................................... 7-4
Quantitative Evaluation of Permeability Stratification ................... 7-4

vi

Single-Value Representation .......................................................... 7-5


Permeability Variation ................................................................... 7-6
Stiles Permeability Distribution ..................................................... 7-14
Lorentz Coefficient .......................................................................... 7-20
Miller-Lents Permeability Distribution ........................................ 7-21
Selection of Layers ............................................................................... 7-24
Geological Zonation ........................................................................ 7-25
Natural Barriers .............................................................................. 7-25
Equal Thickness............................................................................... 7-25
Equal Flow Capacity ....................................................................... 7-25
Statistical Zonation ......................................................................... 7-25
Effect of Crossflow between Layers ................................................... 7-26
Vertical Sweep Efficiency ......................................................................... 7-26
Mobility Ratio....................................................................................... 7-27
Crossflow............................................................................................... 7-27
Gravity Forces ...................................................................................... 7-27
Capillary Forces ................................................................................... 7-27
Problems ..................................................................................................... 7-29
VIII. PREDICTION OF WATERFLOOD PERFORMANCE
Simple Methods ......................................................................................... 8-1
Analogy ................................................................................................. 8-2
Rules of Thumb .................................................................................... 8-3
Empirical Relationships ...................................................................... 8-3
Reservoir Stratification ............................................................................. 8-3
Dykstra-Parsons Method..................................................................... 8-5
Stiles Method ........................................................................................ 8-8
Flow Capacity (C) and Permeability Distribution (k) ................. 8-9
Vertical Sweep Efficiency (Ev) ........................................................ 8-12
Water Cut and Water-Oil Ratio .................................................... 8-14
Oil and Water Producing Rates ..................................................... 8-16
Cumulative Oil Recovery ............................................................... 8-17
Procedure for Predicting Performance ......................................... 8-18
Confined Patterns with Stratification, Areal Sweep, and
Displacement Methods ............................................................................ 8-18
Numerical Simulation ............................................................................... 8-20
PAGE
CGM CRAIG-GEFFEN-MORSE METHOD
Introduction .......................................................................................... CGM-1

vii

Initial Calculations - Single Layer ...................................................... CGM-3


Stage 1: Performance Prior To Interference .................................... CGM-7
Stage 2: Performance from Interference To Fillup ......................... CGM-12
Stage 3: Performance from Fillup To Breakthrough ...................... CGM-15
Stage 4: Performance after Water Breakthrough ........................... CGM-18
Multi-Layer Performance ................................................................... CGM-35
Problems ............................................................................................... CGM-40
IX.

WATERFLOOD SURVEILLANCE
Introduction ............................................................................................... 9-1
Production and Injection Test Analyses.................................................. 9-2
Maps ...................................................................................................... 9-2
Production Well Test Procedures ....................................................... 9-3
Production and Injection Trend Analysis ......................................... 9-3
Production Wells ............................................................................. 9-4
Coordinate Graph ..................................................................... 9-5
Exponential Decline Curves (and Hyperbolic and
Harmonic) ............................................................................... 9-6
Injection Wells ................................................................................. 9-9
Patterns ............................................................................................ 9-10
Voidage Replacement Ratio (VRR) (Monthly and
Cumulative)................................................................................... 9-12
Spaghetti Graph .............................................................................. 9-14
Water/Oil Ratio Plot ....................................................................... 9-18
Oil Cut .............................................................................................. 9-20
X Plot ................................................................................................ 9-21
Oil Cut versus Cumulative Production (Coordinate Graph)...... 9-24
Recovery Factor versus Hydrocarbon Pore Volumes Injected .. 9-25
Multiple Trend Forecasting With Field Production
Constraints .................................................................................... 9-27
Summary of Production Graphs.................................................... 9-27
Pressure Transient Testing ....................................................................... 9-28
Pressure Buildup and Pressure Falloff Testing ................................ 9-29
Step Rate Test ....................................................................................... 9-30
Hall Method of Analyzing Injection Well Behavior ......................... 9-37
PAGE
Pattern Balancing ...................................................................................... 9-45
Volumetric Sweep Efficiency .................................................................... 9-58
Injection Profile Testing ........................................................................... 9-75
Interval Selection for Waterflood Monitoring........................................ 9-78

viii

Injection Profiles........................................................................................ 9-80


Alteration of Injection Profiles ................................................................. 9-84
Flood Front (Bubble) Maps ...................................................................... 9-85
Injection Analysis ...................................................................................... 9-91
Analysis Without Free Gas ( S g 0 ) ................................................... 9-93
Analysis With Free Gas ( S g 0 ) ........................................................ 9-101
Numerical Simulation .......................................................................... 9-114
Water Testing Program ............................................................................ 9-115
Dissolved Gases .................................................................................... 9-115
Microbiological Growth ...................................................................... 9-116
Minerals ................................................................................................ 9-116
Total Solids ........................................................................................... 9-116
Produced Water ................................................................................... 9-117
Pie Charts ................................................................................................... 9-117
Integrated Waterflood Monitoring .......................................................... 9-119
Project Review ........................................................................................... 9-124
Problem ...................................................................................................... 9-129

ix

INTRODUCTION

Waterflooding is the most widely used fluid injection process in the world today. It has
been recognized1 since 1880 that injecting water into an oil-bearing formation has the
potential to improve oil recovery. However, waterflooding did not experience fieldwide
application until the 1930s when several injection projects were initiated,2,3 and it was not
until the early 1950s that the current boom in waterflooding began. Waterflooding is
responsible for a significant fraction of the oil currently produced in the world. In fact, in
the 21st century, most operators begin to investigate the feasibility of water injection
within a short time following the initial field discovery.
Many complex and sophisticated enhanced recovery processes have been developed
through the years in an effort to recover the enormous oil reserves left behind by
inefficient primary recovery mechanisms. Many of these processes have the potential to
recover more oil than waterflooding in a particular reservoir. However, no process has
been discovered which enjoys the widespread applicability of waterflooding.

The

primary reasons why waterflooding is the most successful and most widely used oil
recovery process are4,5,6:

general availability of water

low cost relative to other injection fluids

ease of injecting water into a formation

high efficiency with which water displaces oil

The purpose of these notes is to discuss the reservoir engineering aspects of


waterflooding. It is intended that the reader will gain a better understanding of the
processes by which water displaces oil from a reservoir and, in particular, will gain the
ability to calculate the expected recovery performance and to manage the project to
maximize oil recovery with a minimum number of wellbores and injection volumes.
1-1

While written materials will be limited to the displacement of oil by water, the
displacement processes and computational techniques presented have application to other
oil recovery processes.
I.

The End of Primary Depletion.


If the cumulative water injection exceeds the cumulative production since the start
of injection, the reservoir pressure is no longer declining and in most instances
reservoir pressure begins to increase. For each time period (usually on a monthly
basis) in which the injection equals or exceeds production, measured at reservoir
conditions, the average reservoir pressure is maintained or increased. In those
instances where average pressure is maintained or increased, the primary depletion
stops. This is due to the fact that the predominate primary drive mechanisms
including liquid or rock expansion, gas evolving from solution, gas cap expansion,
or natural water influx are the result of declining reservoir pressure. When pressure
is being maintained or increased, these primary drive mechanisms no longer
function.
During the time of constant or increasing reservoir pressure resulting from water
injection, oil recovery is the result of a displacement process. It should be clear that
it is possible within localized areas of the field to have situations where injection
may be greater than production (pressure increasing) and in other areas injection
may be less than production (pressure declining). In those areas where injection is
less than production and where average reservoir pressure is only being partially
maintained, the reservoir is experiencing a combination of pressure depletion and
fluid displacement. When both recovery processes occur simultaneously, reservoir
analysis is very complicated and usually must be analyzed using a finely gridded
numerical simulation model which has been properly history matched with field
conditions.
The injection to production ratio on a pattern or fieldwide basis is frequently
referred to as the voidage replacement ratio (VRR). Reservoir voidage is measured
1-2

at reservoir conditions and includes oil, water, and free gas production.

As a

reminder, the free gas production should not be assumed negligible. Failure to
account for free gas in the voidage computation can be a major flaw in computing
total reservoir voidage.
II. Factors Controlling Waterflood Recovery
Oil recovery due to waterflooding can be determined at any time in the life of a
waterflood project if the following four factors are known.
A. Oil-in-Place at the Start of Waterflooding -- The oil-in-place at the time of initial
water injection is a function of the floodable pore volume and the oil saturation.
Floodable pore volume is highly dependent on the selection and application of net
pay discriminators such as permeability (and porosity) cutoffs. A successful
flood requires that sufficient oil be present to form an oil bank as water moves
through the formation. An accurate prediction of waterflood performance or the
interpretation of historical waterflood behavior can only be made if a reliable
estimate of oil-in-place at the start of waterflooding is available. Oil-in-place
considerations are discussed in Chapter 3.
B. Areal Sweep Efficiency -- This is the fraction of reservoir area that the water will
contact. It depends primarily upon the relative flow properties of oil and water,
the injection-production well pattern used to flood the reservoir, pressure
distribution between the injection and production wells, and directional
permeability.

The prediction of areal sweep efficiency will be discussed in

Chapter 5.
C. Vertical Sweep Efficiency -- Vertical sweep refers to the fraction of a formation
in the vertical plane which water will contact. This will depend primarily upon
the degree of vertical stratification existing in the reservoir and will be discussed
in Chapter 6.

1-3

D. Displacement Sweep Efficiency -- This represents the fraction of oil which water
will displace in that portion of the reservoir invaded by water. Chapter 4 will
discuss methods of determining the displacement sweep efficiency.
Methods for predicting oil recovery by waterflooding will be presented in Chapter 8.
The cumulative displaced waterflood oil can be computed at any time in the life of a
waterflood project from the following general equation:

N D N E A EV ED

(Eq. 1.1)

where

= the oil in place in the floodable pore volume at the start of water
injection, STB
E A = the fraction of the floodable pore volume area swept by the injected
water
EV = the fraction of the floodable pore volume in the vertical plane
swept by the injected water
ED = the fraction of the oil saturation at the start of water injection
which is displaced by water in that portion of the reservoir
invaded by water
If at the start of water injection, a free gas saturation has not formed within the oil
column, it can be assumed that the displaced waterflood oil is approximately equal
to waterflood oil production.

However, if at the start of injection, reservoir

pressure has declined below the initial bubble point pressure and a free gas
saturation has been developed, then the displaced oil described in Eq. 1.1 is less
than the produced waterflood oil. This subject is described in more detail in
Chapter 4.
Waterflood recovery is dependent on a number of variables. The variables which
usually have the greatest impact on waterflood behavior are listed below:
|

Oil saturation at the start of waterflooding,


1-4

So

Sor ( Sorw )

Residual oil saturation to waterflooding,

Connate water saturation,

Free gas saturation at the start of water injection,

Water floodable pore volume,

S wc

Vp ,

Sg

BBLS (This takes into account the

permeability or porosity net pay discriminator)

o and w

Oil and water viscosity,

Effective permeability to oil measured at the immobile connate water saturation,

( ko ) S

wir

Relative permeability to water and oil, krw and kro

Reservoir stratification, (Dykstra-Parsons coefficient, V )

Waterflood pattern (symmetrical or irregular)

Pressure distribution between injector and producer

Injection rate, BWPD

Oil formation volume factor,

Economics

Bo

III. Waterflooding versus Pressure Maintenance


Maximum combined primary and secondary oil recovery occurs when waterflooding
is initiated at or near the initial bubble point pressure.

When water injection

commences at a time in the life of a reservoir when the reservoir pressure is at a high
level, the injection is frequently referred to as a pressure maintenance project. On
1-5

the other hand, if water injection commences at a time when reservoir pressure has
declined to a low level due to primary depletion, the injection process is usually
referred to as a waterflood. In both instances, the injected water displaces oil and is
a dynamic displacement process. Nevertheless, there are important differences in
the displacement process when water displaces oil at high reservoir pressures
compared to the displacement process which occurs in depleted low pressure
reservoirs. The differences in the displacement mechanisms will be discussed in
Chapters 4 and 5.
IV. Other References
In June 2002, a search of the SPE e-library was conducted to obtain a listing of the
technical papers on the subject of waterflooding which have been presented at SPE
technical conferences or published in SPE journals. The listing is found at the end
of this chapter.

1-6

CHAPTER 1 REFERENCES
1. Carll, J.F.: The Geology of the Oil Regions of Warren, Venango, Clarion, and Butler
Counties, Pennsylvania, Second Geological Survey of Pennsylvania (1880) III, pp.
1875-1879.
2. History of Petroleum Engineering, API, Dallas, Texas (1961).
3. Fettke, C.R.: "Bradford Oil Field, Pennsylvania and New York," Pennsylvania
Geological Survey, 4th Series (1938) M-21.
4. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Dallas, Texas (1971) 3.
5. Willhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
6. Waterflooding, Reprint Series, SPE, Richardson, TX (2003) 56.

1-7

REVIEW OF ROCK PROPERTIES AND FLUID FLOW


An understanding of the basic rock and fluid properties which control flow in a porous
medium is a prerequisite to understanding how a waterflood performs and how a
waterflood should be designed, implemented, and managed. The purpose of this section
is not to teach the fundamentals of rock and fluid properties -- a basic knowledge of these
is assumed. However, certain multiphase flow properties will be discussed as they apply
to waterflood systems.
I. Wettability
A. Definition
In a rock/oil/brine system, wettability can be defined as the tendency of a fluid to
preferentially adhere to, or wet, the surface of a rock in the presence of other
immiscible fluids. In the case of a waterflood, the wetting phases can be oil or
water; gas will often be present, but will not wet the rock. When the rock is
water-wet, water occupies the small pores and contacts the rock surface in the
large pores. The oil is located in the middle of the large pores. In an oil-wet
system, the location of the two fluids is partly reversed from the water-wet case.
Water usually continues to fill the very small pores but oil contacts the majority of
the rock surface in the large pores. The water present in the large pores in the oil
wet rock is located in the middle of the pore, does not contact the large pore throat
surface, and is usually present in small amounts. Water fills the smallest pores
even in the oil-wet system because oil never enters the small pore system due to
capillary forces and consequently, the wettability of the small pores is not
expected to change.
Wettability concepts and the location of oil and connate water in the larger pores
can be illustrated with a simple diagram. Consider the "large" pore in Figure 2-1
which contains both oil and water.

2-1

FIGURE 2-1
PLANE VIEW, CROSS-SECTION VIEW, AND FLUID DISTRIBUTION IN A
HYPOTHETICAL WATER-WET, OIL-WET, AND FRACTIONAL-WET PORE
A
TORTUOUS PORE

A
PORE CROSS-SECTION AT POSITION A-A

CONNATE WATER

OIL

WATER-WET

OIL-WET

FRACTIONAL-WET

It is important to note, however, that the term wettability is used for the wetting
preference of the rock and does not necessarily refer to the fluid that is in contact
with the rock at any given time. For example, consider a clean sandstone core
that is saturated with a refined oil. Even though the rock surface is coated with
oil, the sandstone core is still preferentially water-wet.

Wettability is not a

parameter that is used directly in the computation of waterflood performance.


However, wettability can have a significant impact on such parameters as relative
permeability, connate water saturation, residual oil saturation, and capillary
pressure which directly affect waterflood performance. Anderson1-6 published a
series of excellent papers which discuss wettability and its impact on rock,
saturation, and fluid flow behavior.

2-2

B. Importance
The performance of a waterflood is controlled to a large extent by wettability.
Reasons for this are:
1. The wettability of the rock/fluid system is important because it is a major factor
controlling the location, flow, and distribution of fluids in a reservoir. In
general, one of the fluids in a porous medium of uniform wettability that
contains at least two immiscible fluids will be the wetting fluid. When the
system is in equilibrium, the wetting fluid will completely occupy the smallest
pores and be in contact with a majority of the rock surface (assuming, of
course, that the saturation of the wetting fluid is sufficiently high).

The

nonwetting fluid will occupy the centers of the larger pores and form globules
that extend over several pores. Since wettability controls the relative position
of fluids within the rock matrix, it controls their relative ability to flow. The
wetting fluid, because of its attraction to the rock surface, is in an unfavorable
position to flow. Furthermore, the saturation of the wetting fluid cannot be
reduced below some irreducible value when flooded with another immiscible
fluid. With all other things equal, a waterflood in a water-wet reservoir will
yield a higher oil recovery at a lower water-oil ratio (WOR) than an oil-wet
reservoir. Chapter 4 presents information that allows an engineer to quantify
the effects of wettability on flood performance.
2. Wettability affects the capillary pressure and relative permeability data used to
describe a particular waterflood system. It is found, in measuring multiphase
flow properties, that the direction of saturation change (saturation history)
affects the measured properties. If measurements are made on a core while
increasing the saturation of the wetting phase, this is referred to as the
imbibition direction.

Conversely, when the wetting phase saturation is

decreased during a test, it is referred to as the drainage direction. Different

2-3

capillary pressure and relative permeability curves are obtained depending upon
the direction of saturation change used in the laboratory to make measurements.
The direction of saturation change used to determine multiphase flow properties
should correspond to the saturation history of the waterflood.

Thus, it is

necessary to know the wettability of the reservoir. For example, a waterflood


in a water-wet reservoir is an imbibition process; whereas in an oil-wet
reservoir, it would be a drainage process. Different data would apply to these
two situations.
C. Determination
Historically, all petroleum reservoirs were believed to be strongly water-wet. This
was based on two major facts. First, most clean sedimentary rocks are strongly
water-wet. Second, most reservoirs were deposited in aqueous environments into
which oil later migrated. It was assumed that the connate water would prevent the
oil from touching the rock surfaces.
Reservoir rock can change from its original, strongly water-wet condition by
adsorption of polar compounds and/or the deposition of organic matter originally
in the crude oil. Some crude oils make a rock oil-wet by depositing a thick
organic film on the mineral surfaces. Other crude oils contain polar compounds
that can be adsorbed to make the rock more oil-wet. Some of these compounds
are sufficiently water soluble to pass through the aqueous phase to the rock.
The realization that rock wettability can be altered by absorbable crude oil
components led to the idea that heterogeneous forms of wettability exist in
reservoir rock. Generally, the internal surface of reservoir rock is composed of
many minerals with different surface chemistry and adsorption properties, which
may lead to variations in wettability.

Fractional wettability is also called

heterogeneous, spotted, or Dalmation wettability. In fractional wettability, crude


oil components are strongly adsorbed in certain areas of the rock, so a portion of

2-4

the rock is strongly oil-wet, while the rest is strongly water-wet. Note that this is
conceptually different from intermediate wettability which assumes all portions of
the rock surface have a slight but equal preference to being wetted by water or oil.
Several methods are available to determine the wettability of a reservoir rock.
These methods have been detailed in the literature2,7,8 and will not be discussed
here. They are:

Contact Angle

Imbibition -- Displacement Core Tests

Capillary Pressure Tests

Relative Permeability Tests

Others

D. Factors Affecting Reservoir Wettability


The original strong water-wetness of most reservoir minerals can be altered by the
adsorption of polar compounds and/or the deposition of organic matter that was
originally in the crude oil. The surface-active agents in the oil are generally
believed to be polar compounds that contain oxygen, nitrogen, and/or sulfur.
These compounds contain both a polar and a hydrocarbon end. The polar end
adsorbs on the rock surface, exposing the hydrocarbon end and making the surface
more oil-wet. Experiments have shown that some of these natural surfactants are
sufficiently soluble in water to adsorb onto the rock surface after passing through
a thin layer of water. In addition to the oil composition, the degree to which the
wettability is altered by these surfactants is also determined by the pressure,
temperature, mineral surface and brine chemistry, including ionic composition and
pH.

2-5

E. Sandstone and Carbonates


The types of mineral surfaces in a reservoir are also important in determining
wettability. Studies1 show that carbonate reservoirs are typically more oil-wet
than sandstone reservoirs. Laboratory experiments show that the mineral surface
interacts with the crude oil composition to determine wettability.
F. Native-State, Cleaned, and Restored-State Cores
Cores in three different states of preservation are used in core analysis: native
state, cleaned, and restored state.

Anderson1 indicates the best results for

multiphase-type flow analyses are obtained with native-state cores, where


alterations to the wettability of the undisturbed reservoir rock are minimized.
Anderson's1-6 work defines the term native-state as being any core that was
obtained and stored by methods that preserve the wettability of the reservoir. No
distinction is made between cores taken with oil- or water-based fluids, as long as
the native wettability is maintained.

Be aware, however, that some papers

distinguish on the basis of drilling fluid. Anderson further defined native-state to


be cores taken with a suitable oil-filtrate-type drilling mud, which maintains the
original connate water saturation. Fresh-state refers to a core with unaltered
wettability that was taken with a water-base drilling mud that contains no
compounds that can alter core wettability.
The second type of core is the cleaned core, where an attempt is made to remove
all the fluids and adsorbed organic material by flowing solvents through the cores.
Cleaned cores are usually strongly water-wet and should be used only for such
measurements as porosity and air permeability where the wettability will not
affect the results.
The third type of core is the restored-state core in which the native wettability is
restored by a three-step process. The core is cleaned and then saturated with brine
followed by reservoir crude oil. Finally, the core is aged in reservoir crude at

2-6

reservoir temperature for about 1,000 hours. The methods used to obtain the three
different types of cores are discussed in more detail in References 1 through 6.
II.

Capillary Pressure
A. Definition
Capillary pressure can be qualitatively expressed as the difference in pressure
existing across the interface separating two immiscible fluids. Conceptually, it is
perhaps easier to think of it as the suction capacity of a rock for a fluid that wets
the rock, or the capacity of a rock to repel a non-wetting fluid. Quantitatively,
capillary pressure will be defined in this text as the difference between pressure in
the oil phase and pressure in the water phase. For example:

Pc Po Pw

(Eq. 2.1)

B. Importance
1. Capillary forces, along with gravity forces, control the vertical distribution of
fluids in a reservoir. Capillary pressure data can be used to predict the vertical
connate water distribution in a water-wet system.
2. Capillary pressure data are needed to describe waterflood behavior in more
complex prediction models and in naturally fractured reservoirs.
3. Capillary forces influence the movement of a waterflood front and,
consequently, the ultimate displacement efficiency.
4. Capillary pressure data are used to determine irreducible (immobile) water
saturation.
5. Capillary pressure data provide an indication of the pore size distribution in a
reservoir.
C. Sources of Data
Unfortunately, capillary pressure data are not available for most reservoirs,
especially older reservoirs developed with no thought of subsequent enhanced
recovery projects. The only reliable sources of data are laboratory measurements
made on reservoir core samples. These measurements are seldom made due to the
2-7

time and expense of obtaining unaltered core samples and conducting necessary
tests. The laboratory tests4 most commonly used are:

Restored State (porous diaphragm) Method

Centrifuge Method

Mercury Injection Methods

Most laboratory measurements are made using either air-brine or air-mercury


systems. Consequently, the resulting data must be converted to actual reservoir
conditions, taking into account the difference between interfacial tensions of
laboratory and reservoir fluids and the difference in wettability effects of the fluids.
This conversion can be made using the relationship:

PcR PcL

cos R
cos L

(Eq. 2.2)

where:

PcR

= capillary pressure at reservoir conditions, psi

PcL

= capillary pressure measured in the laboratory, psi

= interfacial tension

= contact angle

Capillary pressure data from another reservoir having similar rock-fluid


characteristics can also be used but is not generally recommended. When this is
necessary, a correlating function such as the "J-function" (to be discussed later) is
generally used.

2-8

D. Effect of Reservoir Variables


1. Fluid Saturation
Capillary pressure varies with the fluid saturation of a rock, increasing as the
wetting phase saturation decreases. Accordingly, capillary pressure data are
generally presented as a function of wetting phase saturation.4

A typical

capillary pressure curve for a water-wet system is illustrated in Figure 2-2.


FIGURE 2-2
EFFECT OF SATURATION HISTORY ON OIL-WATER
CAPILLARY PRESSURE CURVES FOR A WATER-WET ROCK

20

Capillary Pressure, psia

15

10

Drainage

Imbibition

0
0

20

40

60

80

Water Saturation, percent

2-9

100

2. Saturation History
As noted previously, the direction in which the fluid saturation of a rock is
changed during measurement of multiphase flow properties has a significant
affect on measured properties. This hysteresis effect is obvious in Figure 2-2.
The direction of saturation change used in the laboratory, or in other models,
must match the direction of saturation change in the reservoir to which the data
will be applied.
3. Pore Geometry
Other factors being equal, capillary pressure is inversely proportional to the
radius of the pores containing the fluids.9 If all pores were the same size in a
rock, the capillary pressure curve would ideally be described by Curve 1 in
Figure 2-3. However, all rocks exhibit a range of pore sizes which causes a
variation in capillary pressure with fluid saturation. In general, the slope of the
capillary pressure curve will increase with increasing pore size heterogeneity.
This is illustrated by Curves 2, 3, and 4 on Figure 2-3 which represent a
homogeneous, moderately heterogeneous, and very heterogeneous reservoir,
respectively.

2 - 10

FIGURE 2-3
EFFECT OF RESERVOIR HETEROGENEITY ON
CAPILLARY PRESSURE CURVES

20

Capillary Pressure, psia

15

10

Curve 4
Curve 3
Curve 2

5
Curve 1

0
0

20

40

60

80

100

Water Saturation, percent

E. Averaging of Data
Even when good capillary pressure data are available, it is generally found that
each core sample tested from a reservoir gives a different capillary pressure curve
than every other core sample. Thus, an obvious question arises. How do we
determine which curve represents the average behavior of the reservoir to be
waterflooded? Two methods are commonly used to resolve this problem: (1) the
J-function and (2) correlation with permeability.
2 - 11

1. J-function
This function was developed by M. C. Leverett10 in an attempt to develop a
universal capillary pressure curve.

The dimensionless J-function relates

capillary pressure to reservoir rock and fluid properties according to the


relationship.

J S w

1
2

Pc k

f

(Eq. 2.3)

where:

J Sw

= J-function at a particular water saturation, dimensionless

Pc

= capillary pressure, dynes/cm2

= interfacial tension, dynes/cm

= permeability, cm2 (1.0 cm2 = 1.013 x 108 D)

= porosity, fraction

= wettability function, dimensionless

This equation was developed with the idea that, at a given saturation, the value
of

J Sw

would be the same for all rocks regardless of their individual

characteristics. For example, suppose the capillary pressure is measured for a


rock with permeability
tension

1 ,

and the

k1 , porosity 1 , using fluids with interfacial


wettability function is f cos 1.0 . The

capillary pressure for the rock will be some value

Pc1 at S w* .

Now suppose

we measure the capillary pressure in a second rock with properties

2 and f 1.0 .

At saturation

2 - 12

S w*

k2 , 2 ,

(same as for Core 1), a value of

capillary pressure

Pc2

will be obtained. If the J-function correlation works,

the J-function for Cores 1 and 2, at saturation

S w* , will be equal even though

the values of capillary pressure are different. For example:

J1 S w* J 2 S w*

1
k2 2

1
k1 2

Pc1
Pc 2

1 1.0 1
2 1.0 2

Further, this relationship would be true at all saturations so a plot of

S w should be the same for all rocks, as depicted by Figure 2-4.


FIGURE 2-4
J-FUNCTION VS WATER SATURATION

0
0

20

40

60

80

Water Saturation, percent

2 - 13

100

(Eq. 2.4)

versus

Ideally then, it would only be necessary to know the interfacial tension, average
porosity, and average permeability of the reservoir to be flooded to obtain the
proper capillary pressure curve for any reservoir.
Unfortunately, the method does not work universally, i.e., capillary pressure for
all cores, or reservoirs, will not plot on a common curve. This is due primarily
to the difference in pore size distributions and rock wettability between cores.
Rock samples of different permeability and porosity characteristics generally
would not be expected to have equivalent pore size distributions. Further,
because of handling, cleaning, and in situ variation in wettability, it is simply
not adequate to assume in Eq. 2.4 that

f 1.0 .

However, for a given

reservoir, or for a group of reservoirs with similar lithology, this plotting


technique is often satisfactory for smoothing capillary pressure data and
determining the capillary pressure curve that applies at average reservoir
conditions. Consequently, this method is probably used more commonly than
other techniques for averaging data.
2. Correlate with Permeability
This method is based on the following empirical observation. If capillary
pressure is determined for several cores from the same reservoir (so that

and

remain relatively constant) and the logarithm of permeability is plotted

as a function of water saturation for fixed values of capillary pressure, then


straight lines or smooth curves result. This is illustrated by Figure 2-5. If the
average effective permeability of the reservoir is known, the correct average
capillary pressure curve can be obtained by simply entering the subject graph
with the average permeability to read values of capillary pressure as a function
of saturation.

2 - 14

FIGURE 2-5
CORRELATION OF CAPILLARY PRESSURE WITH
PERMEABILITY

Permeability, md

1,000

100

10

Pc5 Pc4 Pc3

Pc2

Pc1

20

40

60

80

100

Water Saturation, percent

_________________________________________________________
EXAMPLE 2:1
Capillary pressure data measured on five cores from a sandstone reservoir are
presented below.

2 - 15

Water Saturations for Constant Capillary Pressure, percent


k, md
75 psi
50 psi
25 psi
10 psi
5 psi
470.0
18.5
22.0
29.0
39.0
49.5
300.0
22.5
25.5
34.0
45.5
56.0
115.0
30.0
34.0
41.0
53.5
65.0
50.0
36.0
40.5
51.0
64.0
77.0
27.0
41.0
44.0
55.0
69.0
81.5

The geometric mean permeability of the reservoir, based on 43 core samples, is


155 md. The interfacial tension,

of the air-brine system used to measure

capillary pressure, is 71 dynes/cm. The reservoir oil-water system has an


interfacial tension, R , equal to 33 dynes/cm. Find a capillary pressure curve
that will apply to average reservoir conditions, i.e., the geometric mean
permeability.
SOLUTION
Figure 2-6 shows that capillary pressure data can be correlated with
permeability. The laboratory values of capillary pressure versus saturation,
corresponding to k = 155 md, are shown in the following table. The values of
capillary pressure, converted to reservoir conditions, are also tabulated.

S w , percent
27.2
31.5
39.2
51.0
62.8

PcL, psi
75
50
25
10
5

2 - 16

PcR

R
P , psi
L cL
34.9
23.2
11.6
4.6
2.3

FIGURE 2-6
CORRELATION OF CAPILLARY PRESSURE, SATURATION,
AND PERMEABILITY FOR EXAMPLE 2.1

Permeability, md

1,000

k = 155 md
100

75 psi

50 psi

25 psi

10 psi

5 psi

10
0

20

40

60

80

100

Water Saturation, percent

III.

Relative Permeability
A. Definition
Before engaging in a discussion of relative permeability, a brief review of the
different permeability terms which frequently appear in technical reports or as part
of technical conversations is in order. The different permeability terms are:
2 - 17

air permeability, md

absolute permeability, md

effective permeability, md

relative permeability, dimensionless

1. Air Permeability - the routine permeability measured on a core sample. This


measurement is conducted using a gas, such as nitrogen or natural gas, and does
not usually take into account the Klinkenberg effect.9 Air permeabilities are
frequently used as estimates of absolute permeability, However, unless the
Klinkenberg correction is performed, air permeability can overstate the absolute
permeability by a factor of 1.5 or more.
2. Absolute Permeability - the permeability of a core sample when filled with a
single liquid such as water or oil. Absolute permeability is independent of the
fluid but is dependent on the pore throat sizes. Absolute permeability is most
applicable in aquifer studies because the aquifer usually contains a single fluid,
water.
3. Effective Permeability - the permeability to water, oil, or gas ( k w , ko , k g )
when more than one phase is present. Effective permeability of a phase is
dependent on fluid saturation. Application of Darcy's Law for determination of
production ( qo or

q w ) or injection ( iw ) rates utilize effective permeability.

Effective permeability to oil and water are most commonly used in waterflood
analysis.
4. Relative Permeability - the ratio of effective permeability to some base
permeability, usually the effective permeability to oil measured at the immobile
(irreducible) connate water saturation,

2 - 18

(ko ) S wir , kro ko /(ko ) S

wir

krw kw /(ko )S . Since the effective permeability of a rock depends


wir

on the fluid saturation, it follows that relative permeability is also a function of


fluid saturation. When the base permeability is

( k o ) S wir , then the relative

permeability to oil at the immobile connate water saturation,


1.0.

( k ro ) S wir , is

In relative permeability measurements prepared prior to about 1975,

laboratories frequently used the uncorrected air permeability as the base


permeability. The net effect is to cause the

( k r o ) S wir

value to be less

than 1.0, usually in the range of 0.6 to 0.8.


B. Importance
As the name implies, relative permeability data indicate the relative ability of oil
and water to flow simultaneously in a porous medium. These data express the
effects of wettability, fluid saturation, saturation history, pore geometry, and fluid
distribution on the behavior of a reservoir system.5,6,7

Accordingly, this is

probably the single, most important flow property which affects the behavior of a
waterflood. When using

( k o ) S wir

as the base permeability, the relative

permeability to oil and water ranges between 0.0 and 1.0 when plotted versus
water saturation. This scale allows for easy comparison of one set of relative
permeability versus another set from a different core sample. The comparison is
made by a simple overlay.
C. Sources of Data
1. Laboratory measurement on representative core samples possessing appropriate
reservoir wettability
a. Steady-state method

2 - 19

b. Unsteady-state method
2. Use data from similar reservoir
3. Mathematical models
4. History matching
5. Calculate from capillary pressure data
D. Effect of Reservoir Variables
1. Saturation History
Figure 2-7 shows the effect of saturation history on a set of relative
permeability data. It is noted that the direction of flow has no effect on the
flow behavior of the wetting phase. However, a significant difference exists
between the drainage and imbibition curves for the non-wetting phase. This
again points out the need for knowing wettability. For a water-wet system, we
would choose the imbibition data; whereas, drainage data would be needed to
correctly predict the performance of an oil-wet reservoir.

2 - 20

FIGURE 2-7
EFFECT OF SATURATION HISTORY ON RELATIVE
PERMEABILITY DATA

100

Relative Permeability, percent

Wetting Phase

80
60
40
20
0
0

20

40

60

80

100

Wetting Phase Saturation, percent

2. Wettability
Wettability affects the fluid distribution within a rock and, consequently, has a
very important effect on relative permeability data. This is indicated on Figure
2-8 which compares data for water-wet and oil-wet systems.

2 - 21

FIGURE 2-8
EFFECT OF WETTABILITY
ON RELATIVE PERMEABILITY DATA

Relative Permeabilty, percent

100

Oil Wet

80

Water
Wet
60

40

20

0
0

20

40

60

80

100

Water Saturation, percent

Several important differences between oil-wet curves and water-wet curves are
generally noted.
a. The water saturation at which oil and water permeabilities are equal
(intersection point of curves) will generally be greater than 50 percent for
water-wet systems and less than 50 percent for oil-wet systems.
b. The connate water saturation for a water-wet system will generally be
greater than 20 percent; whereas, for oil-wet systems, it will normally be
less than 15 percent
2 - 22

c. The relative permeability to water at maximum water saturation (residual oil


saturation) will be less than about 0.3 for water-wet systems but will be
greater than 0.5 for oil-wet systems.
These observations may not hold true for intermediate wettability rocks.
Further, for high permeability values

o S wir

100 md

these findings

may not be true7. For example, water-wet rocks with large pore throats (high
permeability) sometimes exhibit immobile connate water saturation of less than
10 to 15 percent. Nevertheless, Figure 2-8 indicates the shape and magnitude
of relative permeability curves can give an indication of the wettability
preference of a reservoir for moderate to low levels of permeability; i.e.,

(ko ) S wir 100 md .


E. End-Point Values
Summary water-oil relative permeability tests are frequently conducted on core
samples. These summary tests are often referred to as "end-point" tests because
they reflect

Swir , Sor , (ko ) S

wir

, and ( k w )

Sor

. Results of these tests are

less expensive than normal relative permeability tests, but they can provide useful
information on reservoir characteristics. Listed below are end-point test data for
three sandstone cores.
Water-Oil End-Point Relative Permeability Tests*
Initial Conditions
Terminal Conditions

k A ,md
9.4
3.7
18.0

,% S wir ,% ko ,md Sor ,% k w ,md


14.5
15.8
13.8

27.5
37.6
24.7

6.4
2.4
13.0

*Tests conducted at confining overburden pressure


2 - 23

35.4
34.2
38.3

1.8
0.8
4.6

k ro

k rw

1.0
1.0
1.0

0.28
0.33
0.35

F. Averaging of Data
1. Data Averaging Methods
Again, we often face the problem of having several permeability curves for a
particular formation, all of which are different. It is desirable to select one set
of curves which will apply at average reservoir conditions, i.e., at the average
formation permeability. Methods to accomplish this are:
a. Determine the saturation at different values of

k ro or k rw / k ro for each

of the different sets of data (use same values of permeability or permeability


ratio in obtaining saturations from the different permeability curves). This
is probably done most often using

k rw / k ro .

The saturations obtained at

equal values of permeability are arithmetically averaged to define the


average set of permeability data.
b. In some cases, a plot of

k rw / k ro

versus water saturation for each core

will yield a correlation with permeability as shown in Figure 2-9. However,


smooth curves rather than straight lines will often result. If the effective
average permeability is known, an average permeability curve can be
determined from the correlation.

2 - 24

FIGURE 2-9
CORRELATION OF RELATIVE PERMEABILITY
DATA WITH ABSOLUTE PERMEABILITY
100

k1

k2

k3

10

0.1
0

20

40

60

80

100

Water Satuaration, percent

2. Adjust Average Data to Account for Different Irreducible Water


Saturations
This is not necessary for oil-wet systems, but in the case of water-wet systems,
the situation often occurs where the accepted value of irreducible water
saturation does not agree with the average relative permeability data chosen to
represent the reservoir. The procedure for converting the data to a different
irreducible water saturation is:
2 - 25

a. From the average relative permeability curves, read values of

k ro

and

k rw at different values of oil saturation.


b. Multiply each of the saturations from Step (a) by

So
1.0 S wir

c. Using the normalized curve obtained from Step (b), the permeability data
can be placed back on a total pore volume basis, using any desired value of
initial water saturation, by multiplying the normalized saturations by

1.0 S wir .
It is also possible to normalize the relative permeability data before the data are
averaged.
_______________________________________________________________
EXAMPLE 2:2
Relative permeability curves measured on three cores from the Levelland Field,
San Andres formation, in West Texas are shown in Figure 2-10. The average
initial water saturation of this reservoir is believed to be 15 percent. Find the
average oil and water relative permeability curves for this reservoir and adjust
the curves to the average connate water saturation.

2 - 26

FIGURE 2-10
RELATIVE PERMEABILITY DATA FOR EXAMPLE 2.2

100

Relative Perm eab ility

80

60

1 2

40

20

0
0

20

40

60

80

100

Water Saturation, percent

SOLUTION
The calculations necessary to average, normalize, and adjust the curves to a
new saturation basis are presented in the following tables for the oil and water
data. The average permeability curves, adjusted to 15 percent irreducible water
saturation, are presented in Figure 2-11.
2 - 27

(1)

(2)

Conversion of Oil Permeability Data


(All Values in Percent)
(3)
(4)
(5)
(6)
(7)

kro

Sw1

Sw2

Sw3

SwAVG

So
1.0 S wi

(6) * (1.0-0.15)

(Sw)NEW

1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.05
0.01
0.00

8.0
11.0
13.5
16.5
20.0
23.0
26.5
30.5
35.0
41.1
46.0
52.5
56.0

25.0
27.5
30.0
32.5
35.0
37.5
40.5
44.0
47.2
51.0
54.0
58.0
60.5

37.0
39.0
41.0
44.0
46.0
48.5
51.0
54.5
58.0
63.2
67.0
72.5
76.0

23.3
25.8
28.2
31.0
33.7
36.3
39.3
43.0
46.7
51.8
55.7
61.0
64.2

100.0
96.7
93.6
90.0
86.4
83.1
79.1
74.3
69.5
62.8
57.8
50.8
46.7

85.0
82.2
79.6
76.5
73.4
70.6
67.2
63.2
59.0
53.4
49.1
43.2
39.7

15.0
17.8
20.4
23.5
26.6
29.4
32.8
36.8
41.0
46.6
50.9
56.8
60.3

(1)

(2)

Conversion of Water Permeability Data


(All Values in Percent)
(3)
(4)
(5)
(6)
(7)

krw

Sw1

Sw2

Sw3

SwAVG

0.50
0.40
0.30
0.20
0.10
0.05
0.01
0.00

62.0
59.0
56.0
52.0
46.5
42.5
36.0
8.0

73.0
70.0
67.0
63.5
58.5
55.0
48.0
25.0

86.5
83.5
80.5
76.5
71.0
67.0
62.0
37.0

73.8
70.8
67.8
64.0
58.7
54.8
48.7
23.3

So
1.0 S wi
34.2
38.1
42.0
46.9
53.8
58.9
66.9
100.0

2 - 28

(8)

(8)

(6) * (1.0 - 0.15)

(Sw)NEW

29.1
32.4
35.7
39.9
45.7
50.1
56.9
85.0

70.9
67.6
64.3
60.1
54.3
49.9
43.1
15.0

FIGURE 2-11
NORMALIZED AND ADJUSTED
RELATIVE PERMEABILITY CURVES FOR EXAMPLE 2.2
100

Relative Perm eability

80

60

40

20

0
0

20

40

60

80

100

Water Saturation, percent

_______________________________________________________________
G.

Default Relative Permeability Relationships


The most reliable source of relative permeability data is from laboratory
measurements performed on cores obtained from the reservoir of interest. For the
measurements to be meaningful, considerable care and effort must be expended to
ensure that the in situ reservoir wettability is preserved during coring, surfacing,
2 - 29

storage, and measurement operations. Failure to preserve native wettability will


cause the measured relative permeability values to be of little use for reservoir
analysis.
Unfortunately, many reservoirs considered for waterflooding are characterized by
the absence of relative permeability or, at best, by unreliable data. In these
situations, it may be necessary to use certain "default" relative permeability
models for data.
Several authors have presented mathematical models which can be used to
describe relative permeability relationships for the simultaneous flow of oil and
water. The relationships are restricted to reservoirs in which flow is through the
matrix. Consequently, those results are not applicable for flow through reservoirs
possessing significant vugs or natural fractures.
Corey11 has suggested that for a drainage process (waterflood of an oil-wet rock):

4
k rw S we

(Eq. 2.5)

where:

S we

S w S wir
1.0 S wir

(Eq. 2.6)

with:

Sw

= water saturation, fraction

S wir

= irreducible water saturation, fraction

and:

2
k ro (1.0 S we ) 2 1.0 S we
2 - 30

(Eq. 2.7)

Where there is simultaneous flow of oil and water in a water-wet system during an
imbibition process, Smith12 suggests that:

1
4 S w S wir 2
krw S w
1.0 S

wir

(Eq. 2.8)

and:

S w S wir
kro 1.0
1.0 S wir Sor

2
(Eq. 2.9)

where:

Sor

= residual oil saturation, fraction

More recently, Hirasaki13 summarized certain relative data compiled by the 1984
National Petroleum Council14 (NPC). As part of a national enhanced oil recovery
study, it was necessary to forecast remaining waterflood recovery in many
reservoirs throughout the United States. In many instances, reservoir data such as
rock wettability and relative permeability were not available. Consequently, an
NPC technical committee recommended default relative permeability relationships similar to those presented by Molina15. These relationships are listed below.

S w S wir
k rw k rw S
or 1.0 S

or S wir

EXW

(Eq. 2.10)

and:

kro kro S

wir

1.0 S w Sor

1.0 Sor S wir

where:

2 - 31

EXO
(Eq. 2.11)

EXW

= water relative permeability exponent

EXO

= oil relative permeability exponent

(kro ) S

wir

= relative permeability to oil at the irreducible water saturation


(usually 1.0)

(krw ) S

or

= relative permeability to water at the waterflood residual oil


saturation (usually about 0.25 to 0.4 depending on
wettability)

Sor

= residual waterflood oil saturation, fraction

Sw

= water saturation, fraction

S wir

= irreducible water saturation, fraction

In addition to Eq. 2.10 and Eq. 2.11, the NPC also provided certain other default
data which are listed below.

Parameter
Oil relative permeability end-point
Water relative permeability end-point
Oil relative permeability exponent
Water relative permeability exponent
Residual oil saturation, percent

Sandstone
1.0
0.25
2
2
25

Carbonate
1.0
0.40
2
2
37

A comparison of these default end-point values with the statements listed on page
20 of Craig8 suggests a possible conclusion that carbonate reservoirs behave as if
they are oil-wet. This observation should not be interpreted as an indication of
rock wettability but the result of attempting to "average" a large amount of data.
Finally, Honapour16 provides a thorough review of the empirical equations used to
compute two phase (oil/water or gas/oil) and three phase (gas/oil/water) relative
permeability.

2 - 32

__________________________________________________________________

EXAMPLE 2:3
A carbonate oil reservoir is being considered for waterflooding. At the present
time, the immobile (irreducible) water saturation is estimated to be 25 percent.
Compute a pair of oil and water relative permeability curves that could be used in
the evaluation of the waterflood.

SOLUTION
In the absence of specific data, the default relative permeability relationships
described by Eq. 2.10 and Eq. 2.11 will be utilized. The following data are
estimated from analog fields or from the NPC default values.

Sorw

= 35 percent (analog field)

(kro ) S

wir

(krw ) S

or

= 1.0 (based on kbase

( ko ) S

wir

= 0.35 (assumes intermediate wettability)

EXO

= 2.0 (1984 NPC)

EXW

= 2.0 (1984 NPC)

EXW

k rw

S w S wir
k rw S

or 1.0 S S

or
wir

kro

1.0 S w Sor
kro S

wir 1.0 S

or
wir

2 - 33

EXO

Substituting:

k rw

S w 0.25
(0.35)
1.0 0.35 0.25

2.0

and:

1.0 S w 0.35
k ro 1.0
1.0 0.35 0.25
Finally, k rw and

2 .0

k ro can be computed and plotted as a function of water

saturation.

S w ,%

k rw

kro

25
30
35
40
45
50
55
60
65

0.000
0.001
0.022
0.049
0.088
0.137
0.197
0.268
0.350

1.000
0.766
0.562
0.391
0.250
0.141
0.062
0.016
0.000

2 - 34

FIGURE 2-12
OIL/WATER RELATIVE PERMEABILITY
1

R elative Perm eability

0.8

kro

0.6

0.4

krw

0.2

0
0

20

40

60

Water Saturation, percent

2 - 35

80

100

CHAPTER 2 REFERENCES

1. Anderson, W.G.: "Wettability Literature Survey - Part l: Rock/Oil/Brime Interactions and the Effects of Core Handling on Wettability," JPT (Oct. 1986) pp. 112544.
2. Anderson, W.G.: "Wettability Literature Survey - Part 2: Wettability Measurement,"
JPT (Nov. 1986) pp. 1246-62.
3. Anderson, W.G.: "Wettability Literature Survey - Part 3: The Effects of Wettability
on the Electrical Properties of Porous Media," JPT (Dec. 1986) pp. 1371-78.
4. Anderson, W.G.: "Wettability Literature Survey - Part 4: The Effects of Wettability
on Capillary Pressure," JPT (Oct. 1987) pp. 1283-1300.
5. Anderson, W.G.: "Wettability Literature Survey - Part 5: The Effects of Wettability
on Relative Permeability on Relative Permeability," JPT (Nov. 1987) pp. 1453-68.
6. Anderson, W.G.: "Wettability Literature Survey - Part 6: The Effects of Wettability
on Waterflooding," JPT (Dec. 1987) pp. 1605-20.
7. Willhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
8. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Dallas, Texas (1971) 3.
9. Amyx, J.W., Bass, D.M. Jr., and Whiting, R.L.: Petroleum Reservoir Engineering,
McGraw-Hill Book Company (1960).
10. Leverett, M.C.: "Capillary Behavior in Porous Solids," Trans., AIME (1941).
11. Corey, A.T.: "The Interrelation Between Gas and Oil Relative Permeabilities,"
Producers Monthly, (November 1954).
12. Smith, C.R.: Mechanics of Secondary Oil Recovery, Reinhold Publishing
Corporation, New York (1966).
13. Hirasaki, G.J., Morrow, F., Willhite, G.P.: "Estimation of Reservoir Heterogeneity
From Waterflood Performance," SPE Paper 13415, Unsolicited technical paper
submitted for publication during Fall 1984.
14. National Petroleum Council: Enhanced Oil Recovery, (June 21, 1984).

2 - 36

15. Molina, N.N.: "A Systematic Approach to the Relative Permeability in Reservoir
Simulation," SPE Paper 9234 presented at the 1980 SPE Annual Technical
Conference and Exhibition, Dallas.
16. Honarpour, M., Koederitz, L., and Harvey, A.H.: Relative Permeability of Petroleum
Reservoirs, CRC Press, Boca Raton , FL (1986).

2 - 37

PROBLEM 2:1
REVIEW OF ROCK AND FLUID PROPERTIES

A series of laboratory studies resulted in the following average relative permeability data
for an oil reservoir. (Note that the base permeability is the air permeability -- it is old
data.)

Sw , %

k rw

k ro

25
30
35
40
45
50
55
60
65
70

0.000
0.002
0.015
0.025
0.040
0.060
0.082
0.118
0.153
0.200

0.565
0.418
0.300
0.218
0.144
0.092
0.052
0.027
0.009
0.000

These data indicate the irreducible water saturation in the reservoir is 25 percent. Well
logs and core analysis suggest, however, that the true irreducible saturation is
approximately 15 percent. Adjust the permeability data so they represent an irreducible
water saturation of 15 percent and present the data in normalized form on a scale of 0.0
to 1.0.

2 - 38

2 - 39

WATERFLOODABLE OIL-IN-PLACE

To accurately predict and effectively manage waterflood operations, it is necessary to


estimate the reservoir oil-in- place within the floodable portion of the reservoir at the start
of water injection. As indicated earlier, the basic oil recovery evaluation equation used to
compute displaced oil by waterflooding can be summarized as:

N D N E A EV ED

(Eq. 3.1)

where:

ND

= oil displaced by water injection, STB


(It will be shown in later chapters that, in many instances, significant
amounts of displaced oil may not be produced due to gas resaturation
effects.)

= oil-in-place at start of waterflooding within the floodable zones, STB

EA

= areal sweep efficiency, fraction

EV

= vertical sweep efficiency, fraction

ED

= unit displacement efficiency, fraction

The oil-in-place at the start of waterflooding is given by:

7758 Ah So

(Eq. 3.2)

Bo

where:

= floodable area, acres

= floodable pay, feet

= porosity, fraction
3- 1

So

= oil saturation at start of the flood, fraction

Bo

= oil formation volume factor at start of the flood, RB/STB

Three major difficulties encountered in using Eq. 3.2 are the determination of
waterfloodable net pay, porosity, and oil saturation. Since

represents the oil in place at

the start of water injection, it would appear that this value could be obtained by simply
taking the differences in the original oil-in-place and the primary production up to the start
of injection. This simple approach can be misleading due to the fact that the net pay for
primary production is assumed to be the net pay for water injection. As will be shown
later in this chapter, the net pay cutoffs and subsequent net pay for water injection is much
different than for primary depletion. Another deficiency in computing

by taking the

difference in the original oil-in-place and primary production is that such a simple
calculation does not indicate how the gas saturation has increased and the oil saturation has
decreased if reservoir pressure is below the initial bubble point pressure.
I. Oil Saturation
Most waterfloods are implemented late in the life of the reservoir after significant
primary production has occurred and at a time when the reservoir pressure is below the
bubble-point pressure.

As primary production occurs, reservoir pressure declines

below the bubble-point, solution gas evolves from the oil in the reservoir, and a free
gas saturation forms within the oil zone. The development of a free gas saturation is
characterized by the production of a portion of the gas and an increase in the gas-oil
ratio. Despite production of the free gas, a large portion of it remains in the reservoir.
Consequently, the oil saturation at the start of waterflooding can be substantially less
than the oil saturation at the discovery of the field.
The oil saturation in the reservoir is constant at (1 S wc ) during those times when
average reservoir pressure is at or above the initial bubble point pressure. However,
the oil saturation begins to decrease and the free gas saturation begins to increase as
the average reservoir pressure declines below the initial bubble point pressure. To
compute the average oil saturation, let the initial bubble point pressure be the
3- 2

beginning reference point. Consider the following equation at any time when the
reservoir pressure is below the bubble point pressure.

So

Reservoir Oil Volume


Reservoir Pore Volume

(Eq. 3.3)

The reservoir oil volume consists of the number of barrels of oil in the reservoir at the
time of interest and can be estimated as:
Stock Tank Oil Volume = OOIP at bubble-point pressure - Primary Oil
Produced below bubble-point pressure
(Eq. 3.4)
or:
Reservoir Oil Volume =

( N ob N pp )Bo

(Eq. 3.5)

where:

Nob

= original oil-in-place at the bubble-point pressure, STB

N pp

= primary oil production between the bubble-point and current

Bo

reservoir pressure, STB


= oil formation volume factor at prevailing pressure, RB/STB

The reservoir pore volume can be estimated using a volumetric material balance
where:

N ob

V p ( 1.0 S wc )
Bob

(Eq. 3.6)

Solving for pore volume gives:

Vp

N ob Bob

( 1.0 S wc )
3- 3

(Eq. 3.7)

where:

Bob

= oil formation volume factor at the bubble-point pressure, RB/STB

S wc

= connate water saturation at the time of discovery, fraction

Substituting Eq. 3.5 and Eq. 3.7 into Eq. 3.3 leads to:

So

ob

N pp Bo

(Eq. 3.8)

N B
ob ob
1.0 S

wc

Rearranging results in the average oil saturation equation.

N pp B

So 1.0
o 1.0 S wc

N ob Bob

(Eq. 3.9)

This equation plays a very important role in estimating waterflood potential.


____________________________________________________________________

EXAMPLE 3:1
A reservoir is a candidate for waterflooding. The primary oil recovery factor below
the bubble-point pressure is 12 percent. The connate water saturation is 36 percent,
and the oil formation volume factors

Bo at the bubble-point and current pressure

are estimated from PVT charts to be 1.35 and 1.05 RB/STB, respectively. Estimate
the oil saturation at the bubble-point and current pressure.
At the bubble-point, no free gas is present within the oil zone. Consequently,

So 1.0 S wc 1.0 0.36 0.64 or 64%


The current oil saturation can be estimated using Eq. 3.9.
3- 4

N pp B

So 1.0
o 1.0 S wc

N ob Bob

1.05
So 1.0 0.12
1.0 0.36
1.35

So 0.438 or 43.8%
The gas saturation is:

S g 1.0 S wc So
S g 1.0 0.36 0.438

S g 0.202 or 20.2%
This example clearly indicates that the change in reservoir oil saturation is much
greater than the primary oil recovery factor of 12 percent.
____________________________________________________________________
Eq. 3.9 provides a means of computing the average oil saturation within the pore
volume. It is significant to recognize that the actual oil saturation may vary between
geological zones as a result of differential primary depletion, gas cap expansion, or
water influx. Further, if during primary depletion below the initial bubble point
pressure, there is water production, this produced water does not affect the average oil
saturation calculations. The produced water could cause the connate water saturation
to change which would have a direct effect on the gas saturation. Also, this procedure
leads to the determination of an average gas saturation but does not account for
possible gas segregation or the formation of a secondary gas cap. As will be shown in
Chapter 4, the gas saturation is important because it has a direct effect on the volume
of injected water necessary to achieve gas fillup, and thus waterflood oil response.
3- 5

II. Porosity
The most accurate determination of porosity is from cores when core porosity is
measured under overburden conditions. However, only a small percentage of the wells
in most fields will have cores. Consequently, porosity is usually determined from
logs. To provide the most reliable porosity values from logs, it is desirable to calibrate
the porosity logs using appropriate core data. The usual calibration technique is to plot
core porosity versus porosity log measurement such as sonic travel time,
density,

, and then develop a relationship between the parameters.

t or bulk

For example,

Figure 3-1 is a plot of core porosity versus sonic travel time. While there is scatter in
the data, it is clear a relationship exists.

In most instances, the relationship is

approximated by a straight line similar to that shown in Figure 3-1.

FIGURE 3-1
CORE POROSITY VERSUS
INTERVAL TRAVEL TIME FROM SONIC LOG

Core Porosity, fraction

0.20

A Bt

0.15

0.10

0.05

0.00

50

55
Sonic Travel Time, t

3- 6

60

The relationship is:

A Bt
where the constants
and

A and B

(Eq. 3.10)

are estimates from the data plot. The parameters

can be considered calibration constants from the reservoir under investigation.

Similar graphs could be made using density or neutron logs. Once the relationship
between porosity and log property is known, it can be used in the non-cored wells to
determine porosity as a function of the log measured parameters. When core data are
unavailable, the default relationship between

and

is the conventional Wylie-

time equation which is discussed in most logging textbooks.

III. Net Pay


Net pay,

h , is defined as those intervals of rock that contain moveable oil and from

which meaningful (economic) volumes can be recovered under a specific drive


mechanism.

Table 3-1 summarizes several uses of net pay.

It is this writers

experience of many years sin the petroleum industry that net pay determination is one
of the difficult parameters to estimate in a reservoir analysis. As a practical matter, net
pay is calculated after applying various cutoffs such as permeability (porosity), fluid
saturations, and shale volume. The difficulty associated with net pay lies in the
determination of the appropriate cutoffs. Net pay cutoffs must take into account water,
oil, and gas saturations, PVT properties (oil viscosity), reservoir drive mechanism
primary depletion, water injection, gas injection, etc.), oil/water permeability, and
variation in reservoir permeability above and below the median value.
The fact that reservoir drive mechanism impacts net pay determination leads to the
conclusion that the net pay cutoff criteria used to define net pay for primary depletion
is different from the criteria used to define net pay in a water injection project. Net
pay for original oil-in-place which contributes to primary depletion is always greater
3- 7

than net pay and oil-in-place that contributes to waterflood activities. Unfortunately,
this notion creates problems. That is, the original oil-in-place which contributes to
primary production is greater (an in some instances is substantially greater) than the
oil-in-place that contributes to waterflooding.
The value assigned to

in Eq. 3.2 has no meaning unless it contains oil-in-place which

can be recovered during primary, secondary, or enhanced recovery operations. It follows


that the value of

assigned to a well must represent that portion of the formation with

sufficient mobile oil saturation, lateral continuity from injection to producing wells, and
permeability to permit meaningful oil recovery. Figure 3-2 presents a flow chart that leads
to net pay determination. Those intervals in which the water saturation, gas saturation, and
shale volume are less than the cutoff values and

3- 8

3- 9

WellTestAnalysis

ReservoirRecovery
FactorAnalysis

DarcysLaw

Volumetrics

TABLE 3-1
USES OF NET PAY

FIGURE 3-2
POTENTIAL NET PAY METHODOLOGY

ko S wir ko Swir Cutoff


Cutoff

NO

YES
S g S g MAX
(Chapter 4)

NO

YES
S wc SwMAX
(Chapter 4)

NO

YES
VS VSMAX

YES
NET PAY

3- 10

NO

in which the permeability (porosity) is greater than the cutoff value and which are
laterally continuous between the injector and producer are defined as being net pay for
waterflooding purposes.
Determination of water and gas saturation cutoffs are discussed in more detail in
Chapter 4. A permeability cutoff procedure for waterflooding is presented later in this
chapter. Net pay cannot be directly determined from a permeability cutoff, in most
instances, due to the limited availability of cored wells.

Most geoscientists and

reservoir engineers consider themselves fortunate if a small percentage of the wells


within a given field have been cored. Nonetheless, if sufficient core data are available,
it is frequently possible to develop a correlation between porosity and permeability.
Typically, a semi-log plot is prepared with permeability plotted on the log scale, and
the permeability cutoff can then be used to define a corresponding porosity cutoff.
Figure 3-3 is a typical permeability-porosity plot.

FIGURE 3-3
TYPICAL PERMEABILITY-POROSITY RELATIONSHIP

Core Permeability, md

1000

100

10

0.1
0%

4%

8%
Core Porosity, percent

3- 11

12%

16%

Application of Figure 3-3 presents three major problems. First, air permeability
values from core data are usually plotted versus core porosity. It is well known that air
permeability overstates reservoir permeability.

A more technically correct

permeability is the effective permeability to oil measured at the immobile or


irreducible connate water saturation,

ko S

. Using oil/water, gas/oil, or end


wir

point measurements from special core analysis (SCAL), it is possible to develop a


relationship between oil permeability and air permeability when plotted on log-log
graph paper. Figure 3-4 is an example of a relationship between oil permeability and
air permeability for a tightly cemented sandstone reservoir. This transform can be
used to convert air permeability to oil permeability for other core samples obtained in
the field. Second, considerable scatter in the data points may create uncertainty about
the relationship between permeability and porosity. Figure 3-5 is an extreme case of
air permeability versus porosity showing considerable variation between the two
parameters. Third, net pay is highly dependent on the permeability cutoff which is
related to several other reservoir variables. Each of these three points is addressed
below.
A.

Net Pay Determination Using Air Permeability versus Oil Permeability


Figures 3-3 and 3-5 are typical permeability-porosity plots prepared by most
engineers and geologists. Permeability values used to make this type of graph
normally come from "routine" core analysis.

Routine core permeability,

sometimes called air permeability, is measured using a gaseous material such as


air, nitrogen, or natural gas and is intended to measure absolute permeability.
Moreover, the "air" permeabilities have usually not been corrected for Klinkenberg
effects.1 As a result, the routine air permeability,

ka , tends to overstate absolute

permeability. (Absolute permeability is the permeability of a rock when it is filled


with a single fluid. Absolute permeability has application in aquifer analysis
because water is the only fluid present but, absolute permeability has no practical
application within the oil column where multiple fluids coexist.)
3- 12

3- 13

FIGURE 34
Oil Permeability at Connate Water vs Air Permeability
Consolidated Sandstone

3- 14

Permeability, md

Porosity

FIGURE 35
Air Permeability vs Porosity
Consolidated Sandstone

Darcy's law for computing injection or production rates makes use of effective
permeability. Effective permeability is the permeability to water or oil when other
phases are present. As discussed in an earlier chapter, effective permeability to oil
or water is equal to the product of effective permeability to oil measured at the
immobile irreducible water saturation,

is due to the fact that


permeability

when

(kro and krw ) .

ko S

ko S

and relative permeability. This


wir

usually serves as the base or reference


wir

computing

relative

permeability

Table 3-2 compares

ko S

relationship of

ka

with

ko S

and

water

ko S

wir

wir

wir

and

ko S

for several core samples. It should be

is always less than

ka

oil

The appropriate and technically correct value of permeability

used in the construction of Figures 3-3 and 3-5 is

noted that

to

ka .

Figure 3-6 is a graph showing the

versus porosity for the data presented in Table


wir

3-2. It is noted that, for a given permeability cutoff, the porosity cutoff is increased
when using the

ko S

relationship.
wir

Since

ka

overstates effective

permeability, it leads to a porosity cutoff that is too low and thus results in an
optimistic estimate of net pay.

3- 15

TABLE 3-2
Comparison of

ka and ko

for a Sandstone

wir
Reservoir under Consideration for Waterflooding
Sample

k ,md

ko S

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

10.7
11.9
11.2
12.6
12.2
14.8
10.3
14.2
9.0
10.3
14.0
9.8
13.4
14.3
12.9
16.6
15.5
11.7
10.5

0.346
0.767
0.704
5.300
1.220
11.500
0.190
4.380
0.335
0.595
4.430
0.299
4.210
10.600
1.430
25.000
12.200
1.100
0.520

,md

wir

0.045
0.190
0.197
3.310
0.617
4.770
0.036
1.350
0.112
0.094
1.430
0.066
1.360
3.270
0.489
12.500
5.400
0.270
0.110

For example, consider Figure 3-6. If a permeability cutoff of 1.0 md is selected,


the porosity cutoff using the

ka

trend yields a value of 11.6 percent.

corresponding porosity cutoff using the

ko S

The

trend is 13.2 percent. For this


wir

particular reservoir, the total field pore volume using a porosity cutoff of 11.6
percent was estimated to be 30 percent greater than the corresponding pore volume
computed using a porosity cutoff of 13.2 percent.

3- 16

FIGURE 3-6

COMPARISON OF k a and k o S
VERSUS POROSITY FOR A
wir
CONSOLIDATED SANDSTONE RESERVOIR

Permeability, md

100
10

Air Permeability, ka

1
Oil Permeability, (ko ) S wir

0.1
0.01
8

10

12

14

16

18

Porosity, percent

Values of

ko S

are obtained from relative permeability tests or relative


wir

permeability end-point measurements conducted on core samples. It is noted that


these

ko S

values should be measured on core samples possessing


wir

appropriate reservoir wettability. Figure 3-4 is an example relating air permeability


to oil permeability on a log-log scale for a consolidated sandstone reservoir.
Once the porosity cutoff is estimated, it can be used with the available porosity
logs (after log porosity has been calibrated to core porosity at reservoir conditions)
to determine net pay.

All laterally continuous intervals containing adequate

moveable oil saturation and possessing porosity values greater than the porosity

3- 17

cutoff are considered net pay. All intervals possessing porosity less than the
porosity cutoff are considered non-pay and are neglected in all future calculations.
The net pay for each well can be combined with the thickness weighted average
porosity (above the porosity cutoff) to yield a net porosity-thickness for each well.
A map of

for each well can then be plotted and contoured to produce a

porosity-thickness map.

When planimetered, this map gives the desired pore

volume, A h .
B. Net Pay Determination After Accounting For Data Scatter
Permeability-porosity correlations, as described in Figures 3-3, 3-5, and 3-6 using

ka

either

or

ko S

to determine porosity cutoffs, are frequently


wir

characterized by significant data scatter. This scatter can introduce considerable


error in the selection of the porosity cutoff value and, subsequently, in the
floodable oil-in-place calculations.
George and Stiles2 noted in one West Texas carbonate reservoir that the porositypermeability relationship was so poor that the conventional permeability- porosity
technique previously described could not be used.

For example, when a

permeability cutoff of 0.1 md was used, it was found that some core samples with
porosity less than two percent had permeabilities greater than 0.1 md, while other
samples with porosities as high as eight percent had permeabilities less than 0.1
md.

To improve the oil-in-place calculations, George and Stiles offered new

procedures for estimating net pay. One procedure is applicable when only the total
field oil-in-place is needed or when most wells in the field have similar porosity.
This procedure is referred to as the fieldwide net pay determination method. A
second procedure is recommended when an accurate net pay determination is
required for each well. This second technique is referred to as the well net pay
determination solution and makes use of a net to gross
core data.
3- 18

N /G

factor derived from

1. George and Stiles Fieldwide Net Pay Method


In the normal process of analyzing a permeability versus porosity plot, a line
must be fit through the data points. Several different straight line (or curve
fitting) techniques can be used. Each technique will yield a different porosity
cutoff for a given permeability cutoff. The fieldwide net pay technique of
George and Stiles eliminates this problem. The fieldwide technique based on
core samples yields a single porosity cutoff that gives a pore volume that is
equal to the pore volume after applying a permeability cutoff.
The fieldwide net pay method requires core data be available and analyzed
according to the following procedures.
a.

Compute a permeability cutoff (for accurate results, the cutoff should be

ko S
b.

.
wir

From the core data, define net pay as being all cored footage possessing a
permeability greater than the permeability cutoff. Compute the net pay

for all pay above the permeability cutoff.


c.

From the core data, define gross pay as being all core footage possessing a
porosity greater than a porosity cutoff.

This step does not require the

selection of a single value of porosity cutoff. Instead, gross pay is computed


as a function of porosity.
d.

Select several values of porosity cutoff ranging from zero to the maximum
value of porosity. Usually, these values are selected in increments of one
porosity unit such as five to six, six to seven, seven to eight, etc. percent. For
each value of porosity cutoff, compute the pore volume,
pay. On coordinate paper, plot gross
define the gross

h .

h , for the gross

versus the porosity cutoff used to

The curve in Figure 3-7 is a hypothetical example of

this type of relationship.

3- 19

FIGURE 3-7
GROSS POROSITY-THICKNESS VS POROSITY CUTOFF
Gross Porosity-Thickness, feet

10

15

20

Porosity Cutoff, percent

e.

Enter the net

determined from Step b on Figure 3-8 and read the

corresponding porosity. This porosity represents the value of which the net
pay pore volume is equal to the pore volume computed after using a porosity
cutoff. This porosity value, representing the porosity cutoff when utilized in
all wells, will lead to the best estimate of
analysis..

3- 20

within the field based on core

FIGURE 3-8
GROSS POROSITY-THICKNESS VS POROSITY CUTOFF
Gross Porosity-Thickness, feet

6
Net h based on
Permeability Cutoff

2
Porosity Cutoff

10

15

20

Porosity Cutoff, percent

_________________________________________________________________

EXAMPLE 3:2
Twenty-nine wells in a consolidated sandstone reservoir have been cored.
Conventional

ka values have been measured on 2,551 core samples.

a conventional semi-log graph relating

ka to porosity.

Figure 3-9 is

Using the fieldwide net pay

method of George and Stiles described in the above paragraphs, estimate the
porosity cutoff which causes the pore volume based on a porosity cutoff to be equal
to the pore volume based on a permeability cutoff.

3- 21

FIGURE 3-9
CONVENTIONAL SEMI-LOG PLOT OF AIR PERMEABILITY VS
POROSITY FOR A CONSOLIDATED SANDSTONE (2551 SAMPLES)

Air Permeability, md

1000
100
10
1
0.1

Least Squares
Straight Line Fit

0.01
0

10

15

20

25

Porosity, percent

The permeability cutoff for waterflooding in this reservoir has been calculated to is
3.0 md to oil measured at irreducible (immobile) water saturation,
which corresponds to an air permeability,

ko S

wir

ka , of 6.0 md.

The pore volume, h , for all cores possessing

ka

values of 6.0 md or greater is

267 porosity-feet.
The pore volume is computed for different porosity cutoff values listed below.

3- 22

Porosity Cutoff
percent
2
4
6
8
10
12
14
16
18
20
22

Porosity-Thickness
feet
370
369
368
358
330
296
254
192
88
23
1

Pore volume versus porosity is shown in Figure 3-10.

FIGURE 3-10
GROSS POROSITY-THICKNESS VERSUS POROSITY CUTOFF FOR
A CONSOLIDATED SANDSTONE RESERVOIR

Gross Porosity - Thickness, feet

400

300

200

100

0
0

10

15

20

25

Porosity Cutoff, percent

Enter porosity-thickness of 267 feet, based on the permeability cutoff, on the vertical
scale and read the porosity cutoff value of 13.2 percent as shown in Figure 3-11.
3- 23

FIGURE 3-11
DETERMINATION OF APPROPRIATE POROSITY CUTOFF FOR FIELD PORE
VOLUME CALCULATIONS AFTER ACCOUNTING FOR DATA SCATTER

Gross Porosity - Thickness, feet

400

300
267

True Net Pay Pore Volume


Based on Permeability Cutoff

200

100

0
0

10

13.2 15

20

25

Porosity Cutoff, percent

This value of 13.2 percent takes into account the scatter in data. It is the porosity
cutoff that yields a pore volume which is equal to the pore volume in which
permeability is greater than the permeability cutoff. While this technique accounts
for data scatter and provides a more technically correct porosity cutoff, in this
example the 13.2 percent is almost identical to the 13.0 percent obtained by simply
using a 6.0 air permeability directly to Figure 3-9.
_______________________________________________________________
2. George and Stiles Individual Well Net Pay Method (Net to Gross Method)
George and Stiles noted that, while the procedure outlined above gives reliable pore
volume, there are some fields in which there are wells that have produced
significant amounts of oil which are given no pay because all porosity is below the
cutoff.

This failure to allocate pay created problems for certain wells during

waterflood unitization proceedings when unit participation formulas included net


pay. The problem was that wells which had produced primary oil were given little
or no credit for secondary operations because they contained no net pay when using
a straight porosity cutoff.
3- 24

To achieve a better distribution of porosity feet on a well-to-well basis, a second


method for net pay determination method was developed and is outlined below.
a.

Compute a permeability cutoff.

b.

Select a low porosity range, such as

2.0% 3.0% , and determine the

number of feet of core having a porosity within this range. Define this value as
gross pay. Compute the weighted average porosity of all core footage within
this porosity range. For example,

c.

1h1 2 h2 ..... hn
h1 h2 ..... hn

Determine how many feet of the gross pay from Step b have a permeability
greater than the permeability cutoff. Define this value as net pay.

d.

Compute the ratio of actual pay to apparent pay (the

N /G

N /G

factor) and plot

versus the weighted average porosity cutoff from Step b on Cartesian

coordinate paper.
e.

Repeat Steps b through d for increasingly larger values of the porosity cutoff,
i.e.,

3.0% 4.0% ; 4.0% 5.0% ;

etc.

Plot the

N /G

factor for the porosity range versus the average porosity within the range.
Figure 3-12 is an example. While some scatter will exist, George and Stiles
suggest that a straight line fit of the data will normally provide an adequate
description of the

N /G

factor versus porosity relationship. Note that there

may be a minimum value of porosity below which the

N /G

net pay) and a maximum value of porosity above which the

factor is zero (no

N /G

factor is

unity, that is the net pay will equal the gross pay. For example, in Figure 3-12,
all intervals possessing a value of porosity less than three percent have no pay,
whereas for all intervals possessing porosity greater than 20 percent, net pay will
equal gross pay.
3- 25

FIGURE 3-12
RATIO OF NET PAY TO GROSS PAY (N/G)
VERSUS POROSITY
1.00

N/G

0.75
0.50
0.25
0.00
0

f.

10
15
Porosity, percent

20

25

For each porosity interval identified on the porosity log, compute the net pay
using the following equation.

Pay thickness determined N


Net Pay = from log for particular
value of porosity
G
The ratio of net pay to gross pay or

N /G

(Eq. 3.10)

factor is obtained from Figure 3-12

for the particular value of porosity being considered. Suppose for example the
porosity of a particular interval is ten percent; hence from Figure 3-12, the

N /G

factor is approximately 0.4. This means that statistically throughout the

field, all intervals with a porosity of ten percent have a 40 percent probability of
possessing a permeability greater than the permeability cutoff. Consequently,
each one foot interval possessing ten percent porosity will be assigned 0.4 feet
of net pay.
3- 26

When this technique is used, wells with low porosities will not be excluded but
will be given a limited amount of pay. Both total pore volume and pore volume
distribution within the field will be realistic. This method for estimating net pay
is preferable to those methods previously described. However, more work is
required because each porosity interval in each well must be adjusted by a

N /G

factor based on the porosity of the interval. Since most porosity logs

have been digitized, it is a simple process to develop a mathematical relationship


between

N /G

and porosity which can be used in a computer analysis of the

logs on a foot by foot basis to determine net pay.


_______________________________________________________________

EXAMPLE 3:3
Consider the conventional permeability-porosity plot presented in Figure 3-13.

FIGURE 3-13
CONVENTIONAL SEMI-LOG PLOT OF PERMEABILITY VS POROSITY
1000

Ko @ Swir, md

100
10
2 md

0.1
15.3

0.01
8

10

12

14

16

18

Porosity, percent

3- 27

20

22

24

The permeability data represent

ko Swir

values. A permeability cutoff of

2.0 md is appropriate for this reservoir. It is desirable to develop a relationship


between

N /G

and porosity for this set of core data.

For this particular data set, the following table is established.

Porosity Range, %

N/G , fraction

0 - 12
12 - 14
14 - 16
16 - 18
18 or greater

0
3/11 = 0.273
11/20 = 0.550
15/18 = 0.833
1

Figure 3-14 presents a graph of

N /G

ratio versus average porosity within the

porosity range.

FIGURE 3-14
N/G VS POROSITY
1.00

N/G

0.75

0.50

0.25

0.00
8

10

12

14

16

18

20

22

Porosity, percent

The graph indicates for any foot of rock possessing a porosity value less than
about 11 percent, receives a

N /G
3- 28

factor of zero and is counted as non-pay.

Any foot of rock possessing a porosity less than 18 percent is adjusted with the
appropriate

N /G

factor from Figure 3-14 and is credited with a partial foot of

pay. For example, a foot of interval possessing a porosity of 14 percent is


characterized by a

N /G

factor of 0.38 resulting in a net thickness of 0.38 feet.

Similarly, a foot of rock possessing a porosity of 17 percent is credited with 0.83


feet of pay.
The gross interval in each well in this field can be adjusted with the
factor on a foot by foot basis. After adjusting by the

N /G

N /G

factor, h values

can be computed on a well basis for accurate reservoir pore volume


determination.

Statistically, this net to gross approach for computing pore

volume takes into account data scatter and when applied on a well by well basis
should provide more reliable estimates of pore volume than pore volumes
computed using a single porosity cutoff such as 15.3 percent obtained from
Figure 3-13.
_______________________________________________________________
C.

Waterflood Permeability Cutoff Determination Using a Water Cut Method


Determination of a permeability cutoff for waterflooding is not a simple task but it
is very important because the cutoff is used to define net pay and pore volume that
contributes to meaningful waterflood oil recovery. A reliable permeability cutoff is
essential in the efficient design and management of a waterflood project.
For a given reservoir, permeability cutoff changes depending upon the reservoir
drive mechanism. If net pay5,6,7 is defined as being those intervals that contain
mobile oil which can be recovered under a given drive mechanism, it is logical to
expect the permeability cutoff to change depending upon whether production is by
primary, secondary, or tertiary means. A major technical hurdle that reservoir
engineers, geologists, and managers must cross is an understanding that
permeability cutoffs for waterflooding are usually much greater than the cutoff
values used in primary analysis. For waterflooding, the permeability cutoff is
3- 29

influenced by other factors such as oil and water viscosity, mobility ratio, oil/water
relative permeability, oil, water, and gas saturations, and variation between high
and low permeability zones. With current high speed computers, it is a relatively
simple process to create a pattern waterflood model to calculate permeability
cutoffs. Such calculations take into account those factors which drive the outcome
of a waterflood project and which are based on waterflood principles.
A technique for computing permeability cutoff is described in this section and is
referred to as the water cut method.

This method utilizes a mathematical

waterflood model representative of the waterflood such as a pattern or a typical


group of injection and producing wells when a pattern does not exist. As an initial
start, it is usually desirable to select a simple model such as one with equal
thickness layers and no inter-layer cross flow. The model should incorporate
variation in oil permeability and porosity between the layers, oil/water relative
permeability curves, fluid properties, and saturations that exist at the start of
injection.
The permeability cutoff using the water cut method can be illustrated with the aid
of Figure 3-15. This figure presents a three-dimensional view of a typical injector
and offset producer of a five-spot pattern. The reservoir is subdivided into a
number of layers. There is no procedure that precisely defines the number of
layers which should be used but 20 to 25 layers should be sufficient. Appropriate
permeability, porosity, oil/water relative permeabilities, oil and water PVT
properties, and fluid saturations to the various layers are assigned. Appropriate
injection and production well pressures, and injection and production well skin
factors to the model are applied. The model is allowed to run until the producing
well reaches a water cut economic limit such as 96 percent. Permeability cutoff
can be defined as being that value of permeability where most (for example 98
percent) of the waterflood oil is produced. Those zones producing the remaining
two percent of the oil are considered negligible and can be ignored.

3- 30

3- 31

INJECTOR

OIL BANK

ZONE

(GAS)

UNAFFECTE

WATER

PRODUCER AT 96 PERCENT
WATER CUT

FOR PERMEABILITY CUTOFF


(PERMEABILITY VALUES ASSIGNED TO EACH LAYER
BASED ON DYKSTRA-PARSONS PERMEABILITY DISTRIBUTION)

FIGURE 3-15
SCHEMATIC DIAGRAM OF WATERFLOOD AT 96 PERCENT WATER CUT

________________________________________________________________

EXAMPLE 3:4
A waterflood is to be conducted in a field where the injection patterns approximate
a five-spot. A permeability cutoff is to be calculated to determine the net pay
which contributes to the waterflood and to determine floodable pore volume for
waterflood management purposes. The following information is given:

Pwi

= 6,000 psi

S wc

= 37%

Pwf

= 400 psi

Sorw

= 24%

Si

= -3

= 2.6 cp

Sp

= -4

= 0.6 cp

WOR Economic Limit = 30:1


The oil/water relative permeability data, air permeability cumulative distribution
plot (also known as Dykstra-Parson plot), oil permeability versus air permeability
transform, oil permeability cumulative distribution plot, and a graphical relationship
between porosity and permeability are provided in Figures 3-16 through 3-20. A
twenty layered, equal thickness model was developed with oil permeability values
(as determined from Figure 3-19) being assigned to each of the twenty layers. The
reference permeability for the oil/water relative permeability curves is permeability
to oil connate water

(kro at S wc 1.0) .

Therefore it is necessary that the

permeability of the layers be permeability to oil at connate water rather than air
permeability.

The layers are also assigned porosity values based on the

permeability porosity transform shown in Figure 3-20. Figure 3-21 summarizes


permeability and porosity for each layer. The model was waterflooded to a
water/oil ratio of 30:1. Figure 3-22 summarizes oil production

3- 32

3 - 33

Sw (%)
37.00
38.41
40.44
42.47
44.50
46.53
48.56
50.60
52.63
54.66
56.68
58.72
60.75
62.78
64.82
66.85
68.88
70.91
72.94
74.97
75.76

Krw
0.000
0.009
0.017
0.026
0.034
0.046
0.057
0.069
0.081
0.093
0.105
0.120
0.136
0.153
0.170
0.191
0.213
0.242
0.270
0.310
0.337

Kro
1.000
0.832
0.663
0.534
0.405
0.333
0.260
0.209
0.157
0.132
0.106
0.084
0.061
0.050
0.040
0.029
0.019
0.015
0.010
0.005
0.000

1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

Kro

Krw

Sw, %

10 20 30 40 50 60 70 80 90 100

FIGURE 3-16
Red Wine Field
Oil/Water Relative Permeability

Kr

3 - 34

100

1000

0.1

10

Permeability (mD)

K50 = 17.2 md

8 10

20

25 30 35 40 45 50 55 60 65 70 75

80

85

90 92 94

Cumulative Probability (Percent Equal To or Greater Than)

15

Max. Perm = 415 md; 20 Perm Data Points

V-Factor = 0.86

96

Log Normal Distribution

Kair

RedBig
Wine
Field -- Dykstra-Parsons Plot -- Air Permeability
Field -- Dykstra-Parsons Plot -- Air Permeability

FIGURE 3-17

98

3 - 35

Ko @ Swir [mD]

0.10

1.00

10.00

100.00

1,000.00

0.10

1.00

Kair [mD]

10.00

100.00

FIGURE 3-18
Red Wine Field
Ko @ Swir vs Kair from SCAL

1,000.00

3 - 36

100

1000

0.1

10

Permeability (mD)

K50 = 8.6 md

8 10

20 25 30 35 40 45 50 55 60 65 70 75 80

85

90 92 94

Cumulative Probability (Percent Equal To or Greater Than)

15

Max. Perm = 290 md; 20 Perm Data Points

V-Factor = 0.88

96

Log Normal Distribution

Koil

Red Wine Field -- Dykstra-Parsons Plot -- Oil Permeability

FIGURE 3-19

98

3 - 37

Porosity, %

0
0.01

10

15

20

25

30

0.1

10
Permeability to Oil, md

100

FIGURE 3-20
Porosity-Oil Permeability Relationship
Based on a Statistical Best Fit

1000

3 - 38

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

Layer
290
117
78.4
54.3
38.0
30.1
20.8
14.3
11.1
9.07
6.33
4.71
3.68
2.49
1.97
1.44
0.96
0.59
0.26
0.10

25.25
23.34
22.49
21.72
00.97
20.48
19.7
18.91
18.39
17.95
17.19
16.57
16.05
15.23
14.73
14.07
13.23
12.19
10.46
8.42

Oil
Permeability Porosity, %

FIGURE 3-21
Red Wine Field
Five-Spot Model

3 - 39

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

Layer
290
117
78.4
54.3
38.0
30.1
20.8
14.3
11.1
9.07
6.33
4.71
3.68
2.49
1.97
1.44
0.96
0.59
0.26
0.10

25.25
23.34
22.49
21.72
20.97
20.48
19.70
18.91
18.39
17.95
17.19
16.57
16.05
15.23
14.73
14.07
13.23
12.19
10.46
8.42

139
263
382
495
602
705
801
891
976
1057
1128
1185
1230
1261
1286
1305
1318
1326
1330
1332

10.4%
19.7%
28.7%
37.2%
45.5%
52.9%
60.1%
66.9%
73.3%
79.4%
84.7%
89.0%
92.3%
94.7%
96.5%
98.0%
98.9%
99.5%
99.8%
100.0%

Oil
Cumulative % of Total
Permeability Porosity, %
MBO
MBO

FIGURE 3-22
Red Wine Field
Results of the Five-Spot Model @ WOR = 30:1

from each layer. The permeability cutoff was chosen as being the permeability
above which 98 percent of the waterflood oil was produced. Examination of Figure
3-22 indicates that 98 percent of the oil production was produced in layers
possessing a permeability to oil of 1.44 md and greater.

Consequently, the

permeability cutoff is 1.44 md.


D.

Comparison of Original Oil-In-Place - Material Balance Versus Volumetric


Estimates
Comparisons of volumetric original oil-in-place (OOIP) versus estimates of
material balance OOIP are difficult to make. Volumetric estimates of OOIP are
complicated by ancient or insufficient data, old logs, limited core data, uncertain
fluid contacts and difficulty in establishing net pay cutoff criteria. Material balance
calculations are difficult to perform due to incomplete and inaccurate pressure
measurements, uncertainty associated with fluid analysis, concerns related to
aquifer and/or gas cap size, and uncertainty in produced volumes (especially water
as gas production). In those reservoirs where sufficient production and pressure
data are available to make both material balance and volumetric calculations, there
are times when the volumetric oil in place is greater than the material balance value
and there are also situations where the material balance estimate of OOIP is greater
than volumetric estimates. When there is a major difference between the two
estimates of OOIP, both methods should be re-evaluated..
Volumetric calculations assume that all formation thickness with permeability
greater than the permeability cutoff (or porosity cutoff) can be used to compute
OOIP. Unfortunately, many porosity stringers are not continuous between wells,
and in fact, some zones are not penetrated by wells. Because of the lack of rock
continuity, only those intervals connected to a wellbore will affect material balance
calculations. Furthermore, those porosity zones which are continuous between
wells but have no effective well completions will have no affect on the material
balance. Thus, OOIP calculated by material balance depends on well spacing,
porosity continuity, and effective well completions.
3- 40

George and Stiles suggested that the ratio of material balance to volumetric OOIP
can be considered as a measure of reservoir continuity resulting from a
combination of well spacing and effective completion intervals.

Stiles4 has

indicated that in one West Texas field, the material balance OOIP was estimated to
be 738 MMSTB. A volumetric OOIP of 1,029 MMSTB was calculated using a six
percent porosity cutoff. If both values are assumed reasonably correct, the ratio of
material balance to volumetric OOIP of 0.72 is a measure of rock continuity and
effective well completions. Stiles indicated that continuity calculations indicated
75 percent of the total pay was continuous for primary spacing of 40 acres. Hence,
most of the difference between material balance and volumetric original oil-inplace can be reconciled by the lack of continuity. However, for this conclusion to
be correct it must be assumed that the net pay cutoff factors used in the volumetrics
for primary depletion are correct.
Determination of lateral continuity is difficult.

Appropriate estimates of this

parameter can only be made after considering all relevant data including material
balance studies, geological cross-sections, stratigraphy, and pressure test data.
E.

Primary Production Net Pay Versus Secondary Floodable Net Pay


During primary production, all effectively perforated intervals possessing
permeability greater than the permeability cutoff contribute to production.
Furthermore, some zones possessing permeability less than the permeability cutoff
may contribute to production. Production from the low permeability zones is best
described with the aid of Figure 3-23.

3- 41

FIGURE 3-23
CROSS SECTION VIEW ILLUSTRATING WATERFLOOD PAY AND NON-PAY
Layer 1

( k 0 ) S wir WF Perm Cutoff

Layer 2

(k0 ) S

Layer 3

( k 0 ) S wir WF Perm Cutoff

wir

WF Perm Cutoff

Layer 4

( k 0 ) S wir WF Perm Cutoff

Layer 5

( k 0 ) S wir WF Perm Cutoff

Layer 6

( k 0 ) S wir WF Perm Cutoff

Layer 7

(k0 ) S

wir

WF Perm Cutoff

The figure represents a cross-section cartoon between two producing wells during
primary production in a reservoir characterized by several porosity intervals.
Layers 1 and 3 are continuous between the production wells and possess
permeability values greater than the permeability cutoff. Layer 2 is continuous but
possesses a permeability which is less than the permeability cutoff. With respect to
conventional horizontal flow, Layer 2 is treated as being non-productive for the
waterflood
Following some primary production from Layers 1 and 3, they become partially
pressure depleted. If small values of vertical permeability are present, oil under
primary depletion may travel a short distance in the vertical direction, as in Layer
2, until it enters a zone of high permeability zone such as Layers 1 and 2 and then
may move laterally to a producing well. This vertical cross flow can account for
production that is normally not anticipated and results in primary production being
more favorable than would otherwise be predicted. From Figure 3-23, Layers 4, 5,
6, and 7 contribute to primary production. However, only a portion of Layer 6
contributes to waterflooding.
3- 42

Under water injection, water will flow through Layers 1 and 3 with negligible
injection into Layer 2. As the water front moves laterally through Layer 1 or 3, it
is possible that injected water may migrate into Layer 2 through capillary
imbibition and oil will then flow into Layers 1 or 3. This counter-current flow of
water and oil occurs only in water wet rocks.

Significantly, oil production

resulting from counter-current imbibition is slow, occurs at high watercuts, and is


assumed to be negligible in these notes.
In many reservoirs, it is uncommon to find porosity zones that are continuous over
large distances. In fact, some zones may be continuous over several thousand feet
while others extend only a few feet. To be flooded, a pay interval must:
1. possess a permeability greater than the waterflood permeability cutoff and
a water saturation, gas saturation, and shale volume below their respective
cutoffs.
2. be continuous between an injection well and producing well,
3. be injection supported,
4. be effectively completed in a producing well, and
5. have the fluid levels pumped off.
It is usually difficult to map individual stringers accurately. This difficulty is
illustrated with the aid of Figure 3-24.

It helps to illustrate the continuous-

discontinuous nature of thin porosity zones. Recently, Stiles4 reviewed a statistical


technique used to estimate reservoir continuity.

In his approach, continuity

between wells was defined as the fraction of total pay in a well connected to
another well. Each stringer was considered continuous if it correlated between
pairs of wells and discontinuous if it could not be correlated.

3- 43

FIGURE 3-24

The upper curve in Figure 3-25 is an example of a continuity curve in one West
Texas field. As can be seen, rock continuity decreases as the distance between
wells increases.

FIGURE 3-25
CONTINUOUS AND FLOODABLE PAY FOR MEANS FIELD
(WEST TEXAS)

Percent Continuity

1.00

0.75

Continuous Pay
Floodable Pay

0.50

0.25

0.00
0

1000

2000

3000

4000

Horizontal Distance between Wells, feet

3- 44

5000

6000

Because of irregularities in layer geometry, all continuous zones are not floodable.
Consider Layer 6 in Figure 3-23. It is apparent that the zone, while continuous
between wells, is not completely floodable. Since the shape of the porosity zone
between wells is not known, it is difficult to predict performance in this layer.
Stiles used a Monte Carlo technique to determine the fraction of the irregular layer
thickness which could be expected to flood. The overall result is shown in the
lower curve of Figure 3-24.
After accounting for lateral continuity for a specific well spacing, waterflood net
pay will usually be less than waterflood pay based only on permeability, water
saturation and shale volume cutoffs. Practical application of the floodable pay
concept shows that as average distance between injectors and producers decrease,
floodable pay increases. This concept becomes important when evaluating infill
drilling or pattern changes.
IV. Summary
This chapter provides insight into the oil and gas saturations after primary depletion.
Importantly, this chapter describes problems associated with net pay calculation and
gives procedures which, if followed, allow reservoir prediction calculations and
production management techniques to be implemented with a higher level of
confidence.

3- 45

CHAPTER 3 REFERENCES

1. Amyx, J.W., Bass, D.M., and Whiting, R.L.: Petroleum Reservoir Engineering,
McGraw-Hill Book Company, New York, (1960) - Chapter 2.
2. George, C.J. and Stiles, L.H.: "Improved Techniques for Evaluating Carbonate
Waterfloods in West Texas," Journal of Petroleum Technology (November 1978), p.
1547.
3. Willhite, F.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
4. Stiles, L.H.: "Optimizing Waterflood Recovery in a Mature Waterflood, The Fullerton
Clearfork Unit," paper SPE 6198 presented at the 1976 SPE Annual Technical
Conference and Exhibition, New Orleans.
5. AAPG Methods in Exploration Series, No. 10, Development Geology Reference
Manual, D. Morton-Thompson and A.M. Woods (ed.) AAPG, Tulsa, Oklahoma,
(1993).
6. Cobb, W.M. and Marek, F.J.: Net Pay Determination for Primary and Waterflood
Depletion Mechanisms, paper SPE 48952 presented at the 1998 SPE Annual
Technical Conference and Exhibition, New Orleans.
7. Worthington, P.F. and Cosentino, L.: The Role of Cutoffs in Integrated Reservoir
Studies, paper SPE 84387 presented at the 2003 SPE Annual Technical Conference
and Exhibition, Denver.

3- 46

PROBLEM 3:1
OIL IN PLACE
The original discovery pressure of an oil reservoir was above the bubble point pressure.
The primary producing mechanism was fluid expansion and solution gas drive.
Cumulative primary production is 3,200,000 STBO (3,200 MSTBO) of which 700
MSTBO was produced as the reservoir pressure declined from the original discovery
pressure to the bubble point pressure. Given the following rock and fluid property data,
estimate the current average oil and gas saturation in the reservoir.

S wc

= 26%

Boi

= 1.35 RB/STB

Bob

= 1.41 RB/STB

Bo

= 1.10 RB/STB

= 880 acres

= 24 feet

= 16%

3 - 47

PROBLEM 3:2
NET PAY WEIGHTING FACTOR - MIDDLE EAST RESERVOIR
A Middle East oil reservoir is being evaluated for waterflood potential. Figure 3:2-1
presents a semi-log graph of

ka and (ko ) Swir

versus porosity.

1. Compute and compare the porosity cutoff for a 10 md permeability cutoff using the

ka and (ko ) Swir

correlation.

2. Figure 3:2-2 is a plot of

(ko ) Swir

versus porosity. Compute and plot

versus porosity using a 10 md cutoff.

3 - 49

N /G

3 - 50

10

100

1,000

10,000

Permeability, millidarcies

10

12

14

K air
( K o ) S wir

16

18
20
22
Porosity, percent

24

PERMEABILITY VERSUS POROSITY


FOR A MIDDLE EAST RESERVOIR

FIGURE 3:2-1

26

Kair

28

(Ko)Swir

30

3 - 51

Permeability, millidarcies

10

100

1,000

10,000

10

12

14

16

20

22
Porosity, percent

18

24

PERMEABILITY VERSUS POROSITY


FOR A MIDDLE EAST RESERVOIR

FIGURE 3:2-2

26

28

(Ko)Swir

30

MECHANISM OF IMMISCIBLE FLUID DISPLACEMENT


I. Introduction
The purpose of this chapter is to discuss the mechanism by which a fluid is displaced
from a reservoir by an immiscible injection fluid. The primary emphasis of this text is
on the process of waterflooding, and accordingly, equations and solution techniques are
presented specifically for the process of oil displacement by water. The reader should
be aware, however, that the methods presented are also applicable to other displacement
processes involving immiscible fluids.

Other applications, for example, are the

immiscible displacement of oil by gas, primary recovery by gravity drainage, and


primary recovery by natural bottom-water drive.
This discussion will be concerned with the determination of how much oil can be
displaced from a portion of reservoir rock which has been contacted by water. As
indicated in Chapter 1, if water injection exceeds reservoir voidage in a part of the field
(pattern, well cluster, tank battery area, etc.), the reservoir pressure in that area is
maintained or increased. Also, when reservoir pressure stops declining, conventional
primary reservoir depletion stops. If injection maintains or increases reservoir pressure,
then oil production is the result of a displacement process. It is recognized that without
water injection, a portion of the displaced oil would have been produced by a
continuation of primary depletion. If water injection is less than reservoir voidage,
reservoir pressure continues to decline but at a slower rate.

In this situation, oil

recovery is the result of two drive mechanisms primary depletion and displacement.
Conventional waterflood theory assumes that reservoir pressure is maintained or
increased so that only a single drive mechanism displacement exists. For the partial
pressure maintenance in which reservoir pressure is above the bubble point pressure,
the reservoir can usually be treated as if reservoir pressure is maintained because
primary depletion, without natural water influx, above the bubble point pressure is
usually negligible when compared to oil recovery resulting from displacement. On the
other hand, if only partial pressure maintenance occurs and the reservoir is below the
4-1

bubble point and gas is evolving out of solution in the reservoir, the oil recovery
behavior is much more complex and may require numerical simulation analysis to
evaluate reservoir performance. In this section, it is assumed that injection matches or
exceeds reservoir voidage.
Oil which is displaced can be predicted at any time in the life of a waterflood if the
following information is known:
1. Oil-in-place at start of the waterflood,
2. Areal sweep efficiency,

EA

3. Vertical sweep efficiency,

EV

4. Displacement sweep efficiency,

ED

If this information is know at a particular time in the life of a project, the oil displaced

N D due to waterflooding can be computed according to the following equation.

N D N E A EV ED

(Eq. 4.1)

If the gas saturation at the beginning of waterflood operations can be neglected, then
the displaced oil,

N D , is approximately equal to the produced oil. If a free gas

saturation exists, the displaced oil will not be produced until gas fillup is achieved.
This behavior is described near the end of this chapter.
Determination of oil-in-place at the start of water injection is generally based upon
geological information (pore volume), fluid saturation estimates
PVT information Bo . Procedures for computing

So , S g , S wc and

N are discussed in Chapter 3.

Areal and vertical sweep efficiencies refer, respectively, to the fraction of reservoir area
and the fraction of vertical reservoir section which is contacted by water. These sweep
4-2

efficiencies are influenced by many factors including well pattern, well spacing,
pressure distribution, fluid and rock properties, and reservoir heterogeneity. Methods
used to estimate these efficiencies will be discussed in subsequent chapters.
Collectively, areal and vertical sweep efficiencies determine volumetric sweep
efficiency (the fraction of reservoir volume which will be contacted by injected water).
Finally, the fraction of the oil saturation which will be displaced from that portion of
the reservoir contacted or swept by water is the displacement sweep efficiency,

ED .

To study the mechanism of immiscible fluid displacement, assume the following flow
system:
1. Linear Flow (one-dimensional flow)
2. At water breakthrough at the outlet face, the areal sweep and vertical sweep
will be 100 percent that is, 100 percent volumetric sweep at initial water
breakthrough.
3. The gas saturation is zero. This assumption will be modified later in the
chapter.
4. The connate water saturation is immobile. This assumption will be modified
later in the chapter.
5. The oil and water density are constant which means oil and water (and
reservoir rock) are assumed to be incompressible.
Consider a typical injection and offset production well arrangement in a waterflood as
shown in Figure 4-1.

4-3

FIGURE 4-1
PLANE AND CROSS-SECTION VIEWS
OF A LINEAR STRIP OF THE RESERVOIR

h, , (hko ) S wir

It is recognized that radial flow occurs near the injector and producer. This chapter
focuses only on fluid flow and displacement behavior directly between the injector and
producer wells within a single layer of constant thickness,
permeability,

ko Swir .

porosity,

and

Further, it is assumed (for simplicity) that flow is linear (1-

D) within the strip whose width is


developed for computing

h,

w.

It will be shown later that the methods

E D in a linear flow can easily be adapted to other

geometries having radial, elliptical flow, or irregular flow..


Water injection into an oil reservoir containing only oil and immobile connate water is
initially considered. An enlarged side view of the single layer under study is presented
in Figure 4-2.

4-4

FIGURE 4-2
WATER INJECTION INTO A LINEAR STRIP
CONTAINING OIL AND IMMOBILE CONNATE WATER

Injected
"Water Zone"

Unaltered
"Oil Zone"

qw
iw

q prod
qo

Two zones are depicted. The water zone is a two-phase flow zone in which oil and
injected water are flowing simultaneously. In the oil zone, only oil is flowing because
the connate water is immobile in this example. The presence of moveable connate
water and/or a free gas will be discussed later.
II. Reservoir Response Incompressible vs. Slightly Compressible Liquids
It is convenient and reasonable in many liquid filled waterflood projects to assume the
rock, connate water, injected water and oil to be incompressible.

Under these

conditions, steady-state flow exists and it can be assumed that one barrel of effective
injection equals one barrel of liquid production when measured at reservoir conditions.
The liquid filled condition occurs when the free gas, if any, within the oil column has
been displaced from the reservoir or has been re-dissolved in the oil. The displacement
and dissolving of the gas is referred to as gas fillup and is discussed in more detail
later in this chapter.
4-5

If the reservoir liquids are assumed to be incompressible, it is implied that reservoir


production will be equal to effective injection into the net pay. For example, if the
injection rate into the reservoir is 1,000 BWPD, the reservoir liquid production will be
1,000 BLPD. Further, if the injection is suddenly increased to 2,000 BWPD, the liquid
production will immediately increase to 2,000 BLPD. Anyone with experience in
waterflooding will know this sudden increase in production does not occur. However,
given sufficient time, the production will increase to 2,000 BLPD. The reason for the
delay in response is due to the fact that oil, water, and rock are slightly compressible.
Pressure transient theory introduces the concept of a radius of investigation (also
referred to as the radius of drainage). It can be shown from pressure transient theory1
that:

1
2

kt
ri 0.029

(Eq. 4.2)

where

ri

= radius of investigation, feet

= effective permeability to flow, md

= time, hours

= porosity, fraction

= viscosity, cp

ct

= the reservoir isothermal coefficient of compressibility, 1/psia

Rearranging Eq. 4.2 and solving for time leads to:

4-6

1190 ri ct
t
k

(Eq. 4.3)

Equation 4.3 can be used to estimate the time for response at a production well caused
by the change in injection rate. It is observed from Eq. 4.3 that the time for response is
independent of the magnitude of rate change at the injection well.
______________________________________________________________________
EXAMPLE 4:1
A producing well is located 1,000 feet from an injector. The reservoir consists of three
flow units whose permeabilities are 100, 10, and 1 md, respectively. Assume each
flow unit possesses the same porosity, viscosity, and compressibility as listed below:
Estimate the delay in production response if the injection rate is suddenly increased by
50 percent.

20% ; 1.5cp ; ct 15 106 / psia


SOLUTION
The time for production response is governed by Eq. 4.3.

1190 ri ct
t
k

1190 1000 0.20 1.5 15 106


2

5,355
k
4-7

k , md t , hours t , days
1.0

5,355.00

223.00

10.0

535.50

22.30

100.0

53.55

2.23

______________________________________________________________________
III. Fractional Flow Equation
The fractional flow equation relates the fraction of displacing fluid (water) in the total
fluid stream in the two-phase flow water zone or the fraction of mobile connate water
(if any) at any point in the reservoir including the oil zone to the properties of the
reservoir. According to Darcys linear flow equation, the flow rate of water at any
location in the reservoir and, in particular, the water zone is:

k A p

qw 0.001127 w w 0.00694 w sin


w s

(Eq. 4.4)

qw w
pw

0.00694 w sin
s
0.001127k w A

(Eq. 4.5)

or:

Similarly, the pressure gradient in the oil phase is:

po
qo o

0.00694 o sin
s
0.001127ko A
where:

qo

= oil flow rate at reservoir conditions, bbl/day

qw

= water flow rate at reservoir conditions, bbl/day


4-8

(Eq. 4.6)

po

= pressure in oil phase, psia

pw

= pressure in water phase, psia

= oil viscosity, cp

= water viscosity, cp

= distance to point of interest in the reservoir, measured from some


reference point along the direction of flow, feet

ko , k w

= effective water and oil permeabilities at the water saturation which


exists at a distance, S, from some reference point in the reservoir,
md

= cross-sectional area of the linear reservoir through which fluid is


flowing

w h , ft

w , o

= density of reservoir water and oil at reservoir conditions, lbm/ft

= angle measured between horizontal (positive x-axis) and the


direction of flow in the counterclockwise direction, degrees

The sign convention of Eqs. 4-4, 4-5, and 4-6 is illustrated in Figure 4-3.

4-9

FIGURE 4-3
SIGN CONVENTION FOR INCLINED FLOW

Updip Flow
s

Downdip Flow

Recall that capillary pressure was defined by Eq. 2-1 as:

Pc Po Pw

(Eq. 2.1)

Pc Po Pw

s
s
s

(Eq. 4.7)

Thus,

or:

Pc
s

qw w

qo o

0.001127k w A 0.001127ko A

0.00694( w o ) sin
4 - 10

(Eq. 4.8)

The total reservoir throughput rate,

qt , is the sum of the oil and water flow rates, and

it is equal to the water injection rate, iw . For example,

qt qo qw iw

(Eq. 4.9)

Therefore, the fraction of water flowing in the total stream,

f w , at a specific cross-

section is:

fw
Also,

qw
q
w
qo qw iw

(Eq. 4.10)

f w is frequently referred to as water cut. Likewise, the fraction of oil flowing

or oil cut is:

q
f o o 1.0 f w
iw

(Eq. 4.11)

Introducing the definitions of Eqs. 4.9 and 4.10 into Eq. 4.8 results in the following
relationship for the fraction of water flowing at any point,

1.0
fw

s , in a linear flow system.

0.001127ko A Pc

0.00694
sin

w
o
oiw
s

w ko
1.0
o k w

(Eq. 4.12)
Equation 4.12 is commonly referred to as the fractional flow equation or water cut
equation.
The fractional flow equation is a very important relationship because it makes possible
the determination of the relative flow rates of oil and water at any point in a porous
4 - 11

flow system. Furthermore, it incorporates all factors which affect the displacement

o , w , o , w , Pc ,
rock properties ko , k w , S o , S w , total throughput rate iw , pressure gradient
p / s , and structural properties of the reservoir , direction of flow . If the
efficiency of a waterflood project; such as fluid properties

total flow rate is constant, and if fluid properties can be assumed constant (i.e., not
functions of pressure), it is important to note that fraction flow is a function only of
saturation.
If sufficient reservoir data are available, it is possible to use Eq. 4.12 to compute the
fraction of water flowing in a reservoir as a function of water saturation. This data,
when plotted as

f w versus S w on Cartesian paper, forms what is widely referred to

as a fractional flow curve. A typical fractional flow curve is depicted by Figure 4-4.

4 - 12

FIGURE 4-4
TYPICAL FRACTIONAL FLOW CURVE

1.0

1.0 Sor

fw

S wir

0.0
0

100
Water Saturation, percent

It will be shown in subsequent sections that this plot forms the basis of evaluating

ED

and is very useful in the prediction and analysis of reservoir behavior during a waterflood.
4 - 13

______________________________________________________________________________
EXAMPLE 4:2

Data for an oil reservoir which is proposed for waterflooding is presented. Construct the
fractional flow curve for this reservoir. Capillary pressure gradients can be assumed
negligible.

= 18 percent

= 2.48 cp

S wc

= 30 percent

Bo

= 1.37 RB/STB

= 0.62 cp

Bw

= 1.04 RB/STB

iw

= 1000 bbl/day

kbase = 45 md

= 0.8

= 1.03

= 50,000 ft2

= 30 degrees

S w , percent k ro
30
40
50
60
70
80

0.940
0.800
0.440
0.160
0.045
0.000

k rw
0.000
0.040
0.110
0.200
0.300
0.440

SOLUTION
The general fractional flow equation was presented previously as Eq. 4.12. If the
capillary pressure gradient is neglected, this equation reduces to:

1.0
fw

0.001127ko A

oiw

( 0.00694( w o )sin )

w ko
1.0
o k w
4 - 14

where:

ko

kbase kro 45kro md

62.4 w ( 62.4 )( 1.03 ) 64.3 lbm/ft 3

62.4 o ( 62.4 )( 0.8 ) 49.9 lbm/ft 3

w o 14.4 lbm/ft 3
sin sin 30 0.5
thus:

1.0

( 0.001127 )( 45 )( kro )( 50,000 )

fw

( 2.48 )( 1000 )

[ 0.00694( 14.4 )( 0.5 )]

0.62 kro
1.0
2.48 krw

and:

1.0 0.05kro
fw
k
1.0 0.25 ro
krw
Calculations of

f w versus S w are summarized in the following table and are

presented graphically in Figure 4-5.

4 - 15

S w , percent

kro

k rw

1.0 0.05kro
fw
kro
1.0 0.25
krw

30
40
50
60
70
80

0.940
0.800
0.440
0.160
0.045
0.000

0.000
0.040
0.110
0.200
0.330
0.440

0.000
0.160
0.489
0.827
0.962
1.000

FIGURE 4-5
FRACTIONAL FLOW CURVE FOR
EXAMPLE 4:2

1.0

fw

0.0
0

100
Water Saturation, percent

___________________________________________________________________________________________

4 - 16

To obtain a high displacement efficiency,

ED ,

and, correspondingly, an efficient

waterflood, it is required that the fraction of water flowing at any reservoir location be
minimized. We want

f w to be as small as possible at a particular value of water

saturation. Recognizing this fact, it is possible by analysis of Eq. 4.12 to determine the
effect different reservoir variables will have on displacement efficiency.
A. Effect of Wettability
At a particular water saturation, the effective permeability to water,

k w will be

smaller in a water-wet rock than in an oil-wet rock. Accordingly, the denominator


of Eq. 4.12 will be larger for a water-wet rock, and the corresponding value of

fw

will be smaller. This relationship is depicted graphically by Figure 4-6 which


shows a comparison of fractional flow curves for a reservoir under both oil-wet and
water-wet conditions.

4 - 17

FIGURE 4-6
COMPARISON OF FRACTIONAL FLOW CURVES
FOR OIL-WET AND WATER-WET RESERVOIRS

1.0

0.8
Oil Wet

0.6

fw
Water Wet

0.4

0.2

0.0
0

20

40

60

80

100

Water Saturation, percent

In this example, it is assumed immobile connate water and residual oil saturation
are identical. The difference between the water-wet/oil-wet

f w curves is due to

wettability effects on the relative permeability curve shape and end-point.


Since it is desirable to minimize

f w at a particular saturation condition, it can be

seen from Figure 4-6 that water-wet reservoirs will yield a higher displacement
efficiency and higher oil recovery than comparable oil-wet reservoirs.

4 - 18

B. Effect of Formation Dip and Direction of Displacement


When a waterflood is conducted in a reservoir with significant dip, the magnitude
of dip and the direction of water injection relative to the dip angle can have
considerable influence upon oil recovery. The effect of formation dip is dictated by
the gravity term,

w o sin , in Eq. 4.12. When the sign of this term is

positive, the effect of gravity will be to minimize


water displaces oil updip so that

f w . This can only occur when

0 180 .

Conversely, when

180 360 , for example when water displaces oil downdip, the effect of
gravity is to decrease the displacement efficiency. Figure 4-7 shows the effect of
formation dip on the fractional flow curve. The conclusion from these observations
is that water should be injected downdip so as to displace oil updip to obtain
maximum oil recovery. If the density of the oil is greater than the density of water,
water should be injected into updip crestal locations. Further, if the angle of dip is
small (less than about ten degrees), the gravity term is usually negligible.
Moderate to high dip angles must be analyzed using numerical simulation methods.

4 - 19

FIGURE 4-7
EFFECT OF FORM ATION DIP ON FRACTIONAL FLOW

1.0

0.8

0.6

fw
0.4
Downdip
Zero Dip
Updip

0.2

0.0
0

20

40

60

Water Saturation, percent

C. Effect of Capillary Pressure


Capillary pressure was defined previously by Eq. 2.1 as:

Pc Po Pw
The capillary pressure gradient in the s-direction is:

Pc Po Pw

s
s
s
4 - 20

80

In a water-wet rock, this gradient will be a positive number; accordingly, its effect
will be to increase the value of

f w and decrease the efficiency of the waterflood.

It would be desirable in a waterflood to decrease, or eliminate, the capillary


pressure gradient. This can be accomplished by altering the wettability of the rock
or by decreasing, or eliminating, the interfacial tension between oil and water.
Several enhanced recovery processes have the capability to accomplish this; these
processes are beyond the scope of this text, however, and will not be discussed
further. It has been shown in numerical simulation studies that waterflood recovery
in most cases is not strongly dependent on capillary pressure. Consequently, it can
frequently be neglected without introducing appreciable error. On the other hand,
capillary pressure can be a major factor affecting waterflood recovery in naturally
fractured reservoirs and must be considered.
D. Effect of Oil and Water Mobilities
Improved oil recovery results from decreasing the water mobility,
increasing the oil mobility, ko

o .

k w w , or by

The effective permeabilities to oil and water

are affected primarily by the fluid saturations existing in the reservoir. These can
be controlled to some extent by the time in the life of a reservoir when a waterflood
is conducted. For example, if a solution gas drive reservoir is permitted to undergo
significant pressure depletion before initiating a waterflood, a large free gas
saturation will exist in the oil zone at the time of flooding. The effect of this gas
will be to reduce the effective permeability to oil. This in turn has the effect of
increasing

f w . This problem can be eliminated by initiating the flood earlier in

the life of the reservoir before the gas saturation develops.


A displacement process can be improved by increasing the water viscosity or by
decreasing the oil viscosity. Water viscosity, for example, can be increased by the
addition of polymers. Oil viscosity can be decreased by using various thermal
recovery processes such as steam flooding. The effect of oil viscosity on the
4 - 21

fractional flow curve is depicted by Figure 4-8 for a particular set of reservoir
conditions.

FIGURE 4-8
EFFECT OF OIL VISCOSITY
ON FRACTIONAL FLOW OF WATER

1.0

o = 10.0 cp
6.0 cp
2.0 cp
0.5 cp

0.8

0.6

fw
0.4

0.2

0.0
0

10

20

30

40

50

60

70

80

Water Saturation, percent

E. Effect of Rate
The effect of rate varies depending upon whether water is moving updip or
downdip. Keeping in mind that the objective is to minimize

f w , it is clear from

Eq. 4.12 that a low value of iw is desirable if water is moving updip. Conversely,
4 - 22

a large rate should be used for downdip displacement. From a practical standpoint,
the rate will generally be controlled by economics and the physical limitations of
the injection equipment and reservoir.
It is concluded that the fractional flow equation gives valuable insight into the
factors which affect the efficiency of a waterflood or other displacement processes.
To summarize, observations made from this equation are:
1. Updip displacement of oil by water leads to a lower

f w and better

displacement. The displacement improves as the angle of dip increases.


2. Downdip displacement results in a larger

f w and poorer displacement. The

displacement becomes less efficient as the downdip angle increases.


3. The capillary pressure gradient increases
4. A large density difference

f w and results in lower displacement.

w o improves updip recovery but decreases

downdip recovery.
5. Improved oil recovery results from a small water mobility,
oil mobility, ko

k w w , or a large

o .

6. Increasing the rate improves the efficiency of a downdip flood but causes lower
efficiency in an updip flood.
F. Variations of Fractional Flow Equation
Many situations exist where insufficient information is available to evaluate the
capillary pressure gradient.

In other cases, the effect of capillary pressure is

negligible. The fractional flow equation in both of these situations reduces to the
following form.

4 - 23

1.0

7.83 10 6 ko A w o sin

fw

oiw
w ko
1.0
o k w

(Eq. 4.13)

If it can be further assumed gravity effects are negligible, Eq. 4.13 reduces to:

fw

1.0

1. 0 w
o

ko
kw

(Eq. 4.14)

or to the equivalent form:

fw

1.0

w kro
1.0
o k r w

(Eq. 4.15)

Equation 4.15 is the most widely used form of the fractional flow equation.
IV. Frontal Advance Equation
The fractional flow equation relates the fraction of oil and water flowing at any point
in the reservoir to the fluid saturation at that point. However, a complete waterflood
analysis requires that we know the saturation distribution of the various phases at any
given time as well as the manner in which this distribution changes with time. The
frontal advance equation will provide this information. Consider Figure 4-9. It is
desirable to develop a procedure which will allow the determination of water
saturation and oil saturation with distance in a linear flow system. The formula for
computing water saturation in the water invaded portion of the linear system is
referred to as the frontal advance equation. This equation is developed in Appendix
A of Chapter 4 as Eq. 4-A.11 and it is also developed in References 2 and 3.
4 - 24

5.615 iwt df w 5.615 Wi df w

A dS w
A dS w

= distance traveled by a fixed saturation,

iw

= injection rate, bbl/day

Wi

= cumulation water injected at time t , reservoir barrels

(Eq. 4.16)

where

S w , during time t , feet

time interval of interest, days

4 - 25

FIGURE 4-9
WATERFLOOD SATURATION DISTRIBUTION
IN A LINEAR RESERVOIR PRIOR TO BREAKTHROUGH
WITH NO FREE GAS OR MOBILE CONNATE WATER
Injector

Producer
Water Zone

Oil Zone

LINEAR STRIP OF RESERVOIR


Water Zone

Oil Zone

} Sor

Saturation, percent

100

Oil Saturation

Sw

Swf

Water Saturation
0

Swc

Length

V. Welge Analysis of the Buckley-Leverett Theory in Linear Systems


It was shown by Buckley-Leverett2 that the frontal advance equation, Eq. 4.16, can be
used to compute saturation distribution in a linear waterflood system as a function of
time. The work of Buckley-Leverett followed by a publication by Terwilliger3 and
later another publication by Welge4 led to the following conclusion: As the injection
water front moves through the reservoir, the water saturation at the leading edge of
the water front,

S wf , and the average water saturation behind the front, S w , are

constant up to and including the moment of water breakthrough. Further, Welge


4 - 26

developed mathematical expressions for computing

S wf

and

S w.

Significantly,

Welge also shows a graphical procedure can be used to solve the mathematical
equation for computing the saturation values.
A. Welge Method Saturation at Flood Front
Welge demonstrated that a line drawn tangent to the fractional flow curve from the
point

fw/ S

wc

, Swc

with a point of tangency equal to

w/ Swf

, Swf

That is, the point of tangency is the water saturation at the front as illustrated by
Figure 4.10.

4 - 27

FIGURE 4-10
DETERMINATION OF WATER SATURATION AT THE
FRONT FROM THE FRACTIONAL FLOW CURVE

1.0

Tangent
Point

f w / Swf

fw

Swf
0.0
0

100
Water Saturation, percent

In regard to Figure 4-10, two important points are noted.

4 - 28

1. The tangent line to the fractional flow curve should always be drawn from
the initial water saturation. In some cases, the initial water saturation will be
greater than the irreducible water saturation and the tangent line will not
originate from the end of the fractional flow curve. Construction of the
tangent line in this situation is illustrated by Figure 4-11.

FIGURE 4-11
CONSTRUCTION OF TANGENT LINE
WHEN S wc IS GREATER THAN S wir
.

1.0

Tangent
Point

f wf
fw

S wir

S wc

S wf

0.0
0

100
Water Saturation, percent

4 - 29

2. The saturation,

S wf , is constant from the time the flood begins until

breakthrough. At the instant of breakthrough, the water saturation of the


producing well will suddenly increase from the connate value,

S wc , to

S wf . As additional injection takes place, the water saturation at the


producing well will increase until it reaches a maximum value,

S wm , which

is equivalent to 1.0 S or .
B. Welge Method Average Water Saturation
Usually, the most desirable water saturation is the average saturation in the water
swept portion of the reservoir,
compute

ED .

S w.

It is this average value that will be used to

Welge (in Appendix B), showed that:

S w S wf

1 f wf
df w
dS
w f

(Eq. 4.17)

A graphical solution to Eq. 4.17 is obtained by simply extending the tangent line
to the point where

f w 1.0 as illustrated in Figure 4-12.

4 - 30

FIGURE 4-12
GRAPHICAL DETERMINATION OF

Sw

1.0

fw

Swf

Swir

Sw

0.0
0

100
Water Saturation, percent

Significantly,

S wf and S w

remain constant until initial water breakthrough.

Consider Figure 4-13. This shows the water saturation at three different periods,

t1 , t2 ,

and tbt . It is observed that in each instance,

4 - 31

S wf

and

Sw

are

constant. Moreover,

Sw

including breakthrough,

is constant until breakthrough. Therefore, up to and

S w is usually denoted as S wbt .

FIGURE 4-13
SATURATION DISTRIBUTION BETWEEN
INJECTOR AND PRODUCER AT THREE DIFFERENT TIMES
INCLUDING WATER BREAKTHROUGH
100
Saturation, percent

} Sor
S wbt

Oil

t1

tbt

t2

S wf
Water
0

Connate Water

Distance

1. Performance at Water Breakthrough


The average water saturation in the water swept portion of the reservoir at
any time prior to breakthrough and at the instant of breakthrough is a
constant value,

S wbt .

This means the water saturation in the water swept

portion of the reservoir increases by an amount


Displacement efficiency,

ED , is defined by:

4 - 32

S wbt Swc .

ED

Decrease in Oil Saturation in Water Swept Zone (Eq. 4.18)


Oil Saturation at the Start of Waterflooding

or:

ED

So
So

(Eq. 4.19)

The change in oil saturation can be expressed in terms of change in water


saturation. Up to the time of breakthrough, the average water saturation is

S wbt .

Thus:

EDbt

1.0 S wc 1.0 S wbt


1.0 S wc

(Eq. 4.20)

or

EDbt

S wbt S wc
1.0 S wc

(Eq. 4.21)

Equation 4.21 is applicable until initial water breakthrough when there is no


free gas present. A later section discusses the free gas case.
At breakthrough,

x L , and Eq. 4.16 can be rewritten as:


1

5.615iwtbt df w


AL
dS
wf

Considering the left-hand side of this equation, it is observed that:

4 - 33

(Eq. 4.22)

5.615iwtbt
AL

Pore volumes of

bbls water injected

water injected Qibt


bbls/pore volume

at breakthrou gh
(Eq.4.23)

Therefore,
1

Qibt

df
w
dS w f

(Eq. 4.24)

Equation 4.24 shows that the number of pore volumes of water injected at
breakthrough is simply equal to the inverse of the slope of the tangent to the
fractional flow curve. With a constant flow rate, the time to breakthrough
can be computed as the ratio of cumulative water injected to water injection
rate.
For example:

tbt

Wibt ALQ ibt

iw
5 .615 iw

(Eq. 4.25)

_____________________________________________________________________

EXAMPLE 4:3
A waterflood is to be conducted in a homogeneous under saturated oil reservoir which
has dimensions that will result in linear flow. The average cross-sectional area is
approximately 78,000 square feet. Additional reservoir data are:

iw

= 7000 bbl/day

Bw

= 1.02 RB/STB

S wc

= 25 percent

= 1.39 cp

4 - 34

= 22 percent

= 0.50 cp

kbase

= 50 md

=0

Bo

= 1.25 RB/STB

S w , percent

kro / krw

25.0
30.0
35.0
40.0
45.0
50.0
55.0
60.0
65.0
70.0
72.0

36.950
11.120
4.840
2.597
1.340
0.612
0.292
0.098
0.017
0.000

If the first row of producers is located 1,320 feet from the injection wells:
(a) determine the oil recovery (STB) at the time of breakthrough,
(b) determine the time until breakthrough in days,
(c) determine displacement sweep efficiency at the time of breakthrough,
(d) how many barrels of water must be injected to obtain breakthrough?
For a single layer reservoir with linear flow, areal and vertical sweep efficiencies can be
assumed unity. Further, the capillary pressure gradient can be neglected.

SOLUTION
Neglecting gravity and capillary forces, the fractional flow equation reduces to the form
of Eq. 4.15.
4 - 35

fw

1 .0
k
1 .0 ro w
k rw o

The fractional flow data for this reservoir are summarized in the following table.

S w , percent

fw

25.0

0.000

30.0

0.070

35.0

0.200

40.0

0.365

45.0

0.517

50.0

0.674

55.0

0.820

60.0

0.905

65.0

0.966

70.0

0.994

72.0

1.000

These data are plotted in Figure 4-14.

4 - 36

F IG U R E 4 -1 4
F R A C T IO N A L F L O W C U R V E F O R E X A M P L E 4 .2

1.0

0.8

0.6

fw
0.4

S w bt 6 1.4%

0.2

0.0
0

20

40

60

80

100

W ater Saturation, percent

a) For a liquid filled reservoir (no free gas):

N pbt NE A EV E Dbt
1.0

1.0

N pbt NE A EV EDbt
From Eq. 4.26:

N pbt

AL 1.0 S wc S wbt S wc

5.615Bo
1.0 S wc

or:
4 - 37

(Eq. 4.26)

N pbt

AL
5.615Bo

S wbt S wc

The average water saturation in the reservoir is determined by drawing a line


tangent to the fractional flow curve.

f w 1.0

defines

S wbt .

The intersection of this line with

As depicted by Figures 4-14,

S wbt 0.614

for this reservoir. Consequently,

N pbt

0.22 78,000 ft 2 1320 ft

ft 3
bbl
5.615
1.25

bbl
STB

0.614 0.25

Thus:

N pbt 1.175 106 STB


(b) Based upon Eq. 4.23:

tbt

ALQibt
5.615iw

where:

Q ibt

df
w
dS w

S wbt S wc
f

Thus:

Qibt 0 .614 0 .25 0 .364


Therefore:
4 - 38

tbt

(0.22 )( 78,000 )(1320 )( 0.364 )


(5.615 )( 7000 )

tbt 209.8 days


(c) The displacement sweep efficiency at breakthrough is defined by Eq. 4.21
as:

EDbt

S wbt S wc 0.614 0.25

1.0 S wc
1.0 0.25

EDbt 0.485
(d) Eq. 4.24 defines the pore volumes of water injected as:

Qibt

df
w 0.364
dS w f

The cumulative water injection, Wibt , is:

Wibt QibtV p
AL
Wibt Qibt

5.615
(0.22)(78,000)(1320)
Wibt 0.364

5
.
615

Wibt 1.468 106 bbls


______________________________________________________________________
4 - 39

2. Performance after Breakthrough


After breakthrough, the saturation at the outlet will increase continuously
from

S wf

to

S wm ( S wm 1.0 Sor ) as noted in Figure 4-15.

FIGURE 4-15
SATURATION DISTRIBUTION
AT WATER BREAKTHROUGH, AFTER BREAKTHROUGH,
AND AT WATERFLOOD RESIDUAL OIL
100

Saturation, percent

S or

Sw
S w2

S wbt

S wf

Water
0

At

the

Distance

time

the

saturation

at

the

outlet

is

S w2 ,

where

S wf S w2 S wm , Welge4 showed that:


i. The average water saturation in the reservoir at the time the saturation at
the outlet is

S w2 and is given by the equation:

S w S w2

fo2
1.0 f w2
S w2
df w
df w

dS w 2
dS w 2
4 - 40

(Eq. 4.27)

Graphically, this means

Sw

can be determined by drawing a tangent to

the fractional flow curve at the saturation


tangent to

f w 1.0

gives the value of

S w2 .

Sw .

Extrapolation of the

Knowing this saturation,

the oil recovery at this time can be computed.


computations at a number of saturations between

S wf

By making these
and 1.0 Sor ,

a composite of recovery versus outlet saturation can be obtained. This is


illustrated by Figure 4-16.

4 - 41

FIGURE 4-16

S w AFTER BREAKTHROUGH

DETERMINATION OF

S wbt S w

1.0

f w2

df w

dS w S w 2

fw

S wf

S w2

0.0
0

100
Water Saturation, percent

ii. After breakthrough, water is produced at a surface producing water-oil


ratio (WOR) equal to:

WOR

qw Bo iw f w2 Bo
f w2
Bo

qo Bw iw f o 2 Bw 1.0 f w2 Bw
4 - 42

(Eq. 4.28)

where

f w2 is determined at S w2 .

If a mobile water saturation exists in

the reservoir when the flood is initiated, water will be produced before
breakthrough. A modification for this saturation was shown in a previous
section.
iii. The number of pore volumes of water injected at the time the water
saturation at the producing end is

S w2 and is given by the relationship:

df
Qi w
dS w S w 2

(Eq. 4.29)

Knowing this quantity and the water injection rate, the time required to
reach this stage of the flood can be computed.
iv. Oil and water flow rates at the time the saturation at the outlet end of the
linear system is

qo
qw

S w2 are given by the following equations.


( 1.0 f w2 )iw
Bo
f w2iw
Bw

, STB/D

(Eq. 4.30)

, STB/D

(Eq. 4.31)

Finally, it is noted that after water breakthrough, the average water


saturation in the swept part of the reservoir,
throughput.

Sw ,

increases with

Accordingly, the displacement efficiency,

increases. For any value of

Sw :

4 - 43

ED ,

also

ED

S w S wc
1.0 S wc

(Eq. 4.32)

In summary, the Welge method can be used to predict oil recovery, wateroil ratio, displacement efficiency, and cumulative water injected as a
function of time for a linear waterflood. These calculations are illustrated
by Example 4:4.
_________________________________________________________________
EXAMPLE 4:4
Example 4:3 presented data for a reservoir which was subjected to a waterflood.
Predictions of oil recovery at the time of water breakthrough were also presented
in that example.

Extend these calculations to include after breakthrough

performance and compute:


(a) recovery as a function of producing WOR (STB),
(b) recovery as a function of cumulative water injection (STB),
(c) recovery as a function of time (STB)
SOLUTION
The fractional flow data for this reservoir were computed in Example 4:3 and
were presented graphically in Figure 4-14. Calculations in Example 4:3 resulted
in the following information at the time of breakthrough.

N pbt 1.175 106 STB

tbt

209.8 days

S wbt 0.614
4 - 44

Wibt 1.468 106 bbls


For computations beyond breakthrough, that portion of the fractional flow curve
representing the non-stabilized zone

Swf

S w2 S wm

is shown

enlarged in Figure 4-17.


FIGURE 4-17
TANGENT CONSTRUCTION
TO THE FRACTIONAL FLOW CURVE FOR
CALCULATIONS BEYOND BREAKTHROUGH (EXAMPLE 4:4)
1

1.00

3 4

5 6 7 8

0.95

0.90

fw
0.85

0.80

S wf 53%

f wf 0.775

0.75
50

55

60

65

70

75

Water Saturation, percent

By selecting a number of saturations at the producing outlet between

S wf

and

S wm 1.0 S or , a history of oil and water production can be computed


4 - 45

using the slope and average water saturation corresponding to each value of

S w2 chosen.

These computations are summarized in the following tables.

N p V p S w S wbt
Bo
Key (see
Fig. 4.17)

S w2 ,%

f w2 df w / dS S w ,%
2.747

61.4=

Incremental Recovery beyond


Breakthrough, STB

53.0 =
Swf

0.775

55.0

0.820

2.093

63.6

71,000

57.5

0.865

1.753

65.2

122,600

60.0

0.905

1.462

66.5

164,600

62.5

0.940

1.132

67.8

206,500

65.0

0.965

0.875

69.0

245,000

67.5

0.983

0.548

70.6

297,000

70.0

0.994

0.400

71.5

306,000

S wbt

4 - 46

Wi V pQi

N p ,STB 10 6

S w2 ,%

N pbt N p

df
Qi w
dS
w

bbls 10 6

53.0 = Swf

1.175

0.364

1.468

55.0
57.5
60.0
62.5
65.0
67.5
70.0

1.246
1.298
1.340
1.382
1.420
1.472
1.482

0.498
0.570
0.684
0.883
1.143
1.825
2.500

2.008
2.299
2.759
3.562
4.611
7.362
10.085

S w2 ,%

Wi

53.0 = Swf

, days
7000
210 = tbt

55.0

WOR

qo ,STB

qw ,STB

4.2

1,260

5,319

287

5.6

1,008

5,627

57.5

328

7.9

756

5,936

60.0

394

11.7

532

6,211

62.5

509

19.2

336

6,451

65.0

659

33.8

196

6,623

67.5

1,052

70.9

95

6,746

70.0

1,441

203.0

34

6,822

__________________________________________________________________
C. Application to Radial Flow
Felsenthal and Yuster5 extended the frontal advance method to radial systems
and found the average water saturation behind the front and the saturation at the
front could be determined in the same manner as for linear flow. This same
observation should apply to any waterflood regardless of the flow geometry.
4 - 47

D. Effect of Free Gas Saturation


If the reservoir pressure declines below the initial bubble-point pressure prior to
the start of waterflooding, a free gas saturation will develop within the oil
column. In this situation, the average oil saturation at the start of waterflooding
can be computed by using Eq. 3.9 as discussed in Chapter 3. That is:

N pp Bo

1.0 S wc
So 1.0
N
B

ob ob

(Eq. 3.9)

and:

S g 1.0 So S wc
Eq. 3.9 assumes primary depletion is the result of solution gas drive and fluid
expansion. It neglects water influx and gravity segregation. For most reservoirs
under consideration for waterflooding, negligible water influx and gravity
segregation are acceptable assumptions. In those situations where water influx
and gravity segregation are important, it may be necessary to account for their
impact on recovery with a numerical simulation model. In this section, it is
assumed that Eq. 3.9 sufficiently describes the average oil (and gas) saturation in
the area of the field under consideration and that water influx is negligible.
Figure 4-18 is a fluid saturation distribution between an injector and a producer
in a reservoir containing a free gas phase resulting from primary depletion at the
start of injection.

4 - 48

FIGURE 4-18
SATURATION DISTRIBUTION
BETWEEN INJECTOR AND PRODUCER WITH FREE GAS
100

Saturation, percent

Free Gas

Oil

(S g )

( So )

Connate Water

( S wc )

Distance

Early in the life of the water injection project before gas fillup, Figure 4-19
depicts a possible saturation profile.

4 - 49

FIGURE 4-19
SATURATION DISTRIBUTION
EARLY IN THE LIFE OF A WATERFLOOD
HAVING AN INITIAL GAS SATURATION

Saturation, percent

100

Water Zone

} S ohc

Oil Zone or Oil Bank

Unaffected
Gas Zone

Free Gas

Trapped Gas

Oil

S wbt

S wf
Water
0

Connate Water
Distance

It is observed that three different fluid regions develop and are defined as the
injected water zone, oil bank zone, and unaffected gas zone. Figure 4-19 is
similar to Figure 3-21 presented by Willhite6. It is significant to note that while
a trapped gas saturation may exist within the water zone or oil bank, in practice,
it is not likely to be important. During most waterfloods, reservoir pressure
within the water zone and oil bank increases. Craig7 presented an equation (Ref.
7, Eq. 3.20) which can be used to calculate the pressure level at which the
trapped gas dissolves into the oil bank.

Usually, an increase in reservoir

pressure of 200 to 300 psi is sufficient. In most waterfloods, an increase in


reservoir pressure of more than 200 to 300 psi occurs resulting in the trapped gas
being re-dissolved in the oil.
During the waterflood process, a portion of the initial free gas will usually be
displaced by the leading edge of the oil bank. This will occur if the initial gas
4 - 50

saturation exceeds the critical saturation (usually in the range of three to five
percent) due to a favorable mobility ratio (discussed in the next chapter) between
the displacing oil and the displaced gas. Recognizing the free gas will be
displaced to the production well and/or dissolved within the oil, the saturation
profile in Figure 4-19 can be simplified to Figure 4-20.

FIGURE 4-20
SATURATION DISTRIBUTION
EARLY IN THE LIFE OF A WATERFLOOD
WITHOUT TRAPPED GAS

Saturation, percent

100

Water Zone

} Sor

Oil Zone or Oil Bank

Unaffected
Gas Zone

Free Gas

Oil

S wbt

S wf
Water
0

Connate Water

Distance

In Figure 4-20, the water saturation distribution in the water zone is identical
to the distribution when there is no free gas as illustrated in Figure 4-13. The
oil bank which lies immediately ahead of the water zone is denoted by an oil
saturation which is

1.0 Swc . This is equivalent to the oil saturation at the

bubble-point pressure.
4 - 51

The increase in oil saturation in the oil bank is exactly equal to the decrease in
the initial free gas saturation,

Sg .

Also, the increase in oil saturation in the oil

bank is the result of water displacing oil from the water zone. The buildup or
increase in the oil saturation in the oil zone is sometimes called an oil resaturation effect. During this re-saturation process, oil is displaced from the
water zone and re-saturates pore space in the oil bank previously filled with free
gas. During the re-saturation process, the oil displaced from the water zone is
not produced. It is simply displaced from the water zone to a different part of
the reservoir (the oil bank). The re-saturation process is also referred to as the
gas fillup process8.
With continued water injection, the leading oil bank front reaches the producing
well. This is referred to as the gas fillup or more commonly is simply referred to
as fillup. When fillup is achieved (the oil bank arrives at the production well),
steady-state injection and production concepts can be used to describe injection
and production behavior. Figure 4-21 is an example of the fluid saturation
distribution at gas fillup and shows only the water and oil zones.

4 - 52

FIGURE 4-21
SATURATION DISTRIBUTION AT FREE GAS FILLUP

} Sor

Saturation, percent

100

Water Zone

Oil Zone

Oil

S wbt

S wf
Water
Connate Water
0

Distance

The saturation distribution in Figure 4-21 is identical to Figure 4-9.


cumulative barrels or injected water necessary to reach gas fillup,

The

Wif , can be

approximated as:

Wif V p S g
where:

Vp

= pore volume, barrels

Sg

= free gas saturation at the start of water injection, fraction

4 - 53

(Eq. 4.33)

_______________________________________________________________

EXAMPLE 4:5
An 80-acre direct line drive pattern (to be discussed in Chapter 5) is under
consideration for waterflooding. Given the following information, compute
the barrels of water which must be injected to reach gas fillup and the time to
reach fillup.

= 80 acres

So

= 55 percent

= 4 feet

Sg

= 15 percent

= 20 percent

iw

= 200 BWPD

S wc

= 30 percent

From Eq. 4.33, the barrels of water to reach fillup are:

Wif V p S g
Wif 7758AhSg
Wif 7758(80)(4)(0.20)(0.15)
Wif 74,477 barrels
The time to reach fillup is given by:

tf

Wif
iw

4 - 54

tf

74,477 bbls
200 bbls/day

t f 372 days
_______________________________________________________________
1.

Production Performance
From the start of water injection until gas fillup, the oil bank will not have
arrived at the producing well. Moreover, during fillup, the saturations (and
pressure) at the producing well will not be materially changed. Therefore,
during fillup, primary production will continue almost as if a waterflood
had not commenced. Upon fillup, the oil bank arrives at the producing well
and free gas has disappeared (it has been produced or re-dissolved). At this
time, the reservoir can be approximately treated as being filled with
incompressible liquids and the steady-state concept of one barrel in-one
barrel out can be applied.

Thus, after fillup, total production when

measured at reservoir conditions is equal to total effective injection.


2. Displacement Efficiency
Waterflood displacement efficiency in the water swept reservoir in the
presence of an initial free gas saturation can be computed as:

ED

Change in oil saturation in water swept zone So

Initial oil saturation at start of waterflood


So

The initial oil saturation is:

So 1.0 S wc S g
4 - 55

and the average oil saturation in the water swept zone is

1.0 S w .

Therefore,

ED

1.0 S

wc

S g 1.0 S w

1.0 S wc S g

(Eq. 4.34)

or:

ED

S w S wc S g
1.0 S wc S g

Prior to the injection water breakthrough,


break-through,

S w S wbt

S w S wbt .

(Eq. 4.35)

After water

and can be computed from fractional flow

theory as discussed in Example 4:4.


3. Conditions for Development of An Oil Bank
In a depleted reservoir possessing a large free gas saturation, it is possible
that an oil bank may not form7. Consider a radial flow system in which a
water zone and oil bank are created around an injection well.

4 - 56

FIGURE 4-22
WATER ZONE AND OIL BANK
SURROUNDING AN INJECTION WELL
EARLY IN THE LIFE OF A WATERFLOOD WITH FREE GAS

OIL
BANK
WATER
ZONE

re

rw

The radii of the water and oil banks depend on the cumulative volume of
water injected,

Wi .

Within the radius, re , the injected water effectively

fills the gas space from rw to re . Thus, for a single layer of thickness, h ,
and porosity, , it is possible to write:

re2h S g 5.615Wi

(Eq. 4.36)

or:

1
2

5.615Wi

re
hS g

where:

4 - 57

(Eq. 4.37)

Wi = cumulative water injected, bbls


All of the injected water is within the water bank of radius,
average water saturation in the water bank is

r.

Since the

S wbt , it is possible to write:

r 2 h S wbt S wc 5.615Wi

(Eq. 4.38)

Equating Eq. 4.36 and 4.38 leads to:

r 2 h S wbt S wc re2 hS g

(Eq. 4.39)

or:

re2

S wbt S wc
r

Sg

(Eq. 4.40)

or:
1
2

S wbt S wc

re r

Sg

(Eq. 4.41)

or:
1
2

re S wbt S wc

r
Sg

If re

r , no oil bank is formed.

(Eq. 4.42)

From Eq. 4.42, it follows that no oil

bank will develop when:

S g S wbt S wc
4 - 58

(Eq. 4.43)

As a practical matter, if the ratio of re

/r

is less than about 1.25, the oil

bank will be small and water breakthrough will occur soon after fillup.
Assuming an re

/r

ratio of 1.25 to be a minimum for a practical oil bank

to form, then a small or negligibly small oil bank develops when:

1
2

S wbt S
wc
1.25

r
Sg

re

(Eq. 4.44)

S wbt S wc
1.5625

S
g

(Eq. 4.45)

S g 0.64 S wbt S wc

(Eq. 4.46)

or:

The concept of oil bank development is discussed in more detail in a later


section.
VI. Properties Fluid PVT
The key oil fluid properties9 in a black oil reservoir under consideration for a
waterflood are the solution gas-oil ratio,
volume factor,

Bo .

Rs , oil viscosity, o , and the oil formation

An important question that frequently needs to be answered is

At what pressure should these factors be evaluated?.


Consider Figures 4-23, 4-24, and 4-25. These figures relate solution gas, oil viscosity
and formation volume factor to pressure. For these figures, the initial reservoir
pressure is 3,000 psi and the initial bubble point pressure is 2,200 psi. The primary
depletion is predominantly a solution gas drive mechanism. At the start of water
4 - 59

injection, average reservoir pressure declined to 1,200 psi. Water injection increased
the reservoir pressure to its current value of 2,500 psi. At this time, what is the
appropriate value to use for

Rs , o , and Bo ?

Unfortunately, the exact answer may

not be known or easily calculated. However, if the values are evaluated at the average
reservoir pressure existing at the start of the waterflood, they can serve as reasonable
estimates for the entire life of a waterflood even if the reservoir pressure is ultimately
increased to near the original pressure.

FIGURE 4-23
SOLUTION GAS, Rs, vs PRESSURE
600
Solution Gas, SCF/STBO

Primary

500
400
Waterflood

300
200

Start WF

100
0
0

600

1200

4 - 60

P, psi

1800

2400

3000

FIGURE 4-24
FORMATION VOLUME FACTOR vs PRESSURE
1.3

FVF, Bo, RB/STBO

Primary

1.2
Waterflood

1.1

Start WF

1.0
0

600

1200

P, psi

1800

2400

3000

FIGURE 4-25
OIL VISCOSITY vs PRESSURE
1.6

Oil Viscosity, CP

1.4
1.2
1.0

Waterflood

0.8
0.6
Start WF

0.4

Primary

0.2
0.0
0

600

1200

4 - 61

P, psi

1800

2400

3000

___________________________________________________________________

EXAMPLE 4:6
A partly depleted single layer of a 160 acre five-spot pattern is to be waterflooded.
The reservoir is characterized by the following data:

= 160 Acres

So

= 61%

= 4 feet

= 400 psia

= 18%

= 180o F

S wc

= 24%

= 0.95

Sg

= 15%

Bo

= 1.15 RB/STBO @ 400 psia

1. Compute the SCF of free gas and STBO in the layer at the start of
waterflooding.
2. If the free gas in re-dissolved during the fillup period, what is the increase in
the solution gas-oil ratio.

SOLUTION
1. Free gas =

43,560 Ah S g Bg , SCF

where

Bg 35.3 p / z T 460 (assumes standard conditions to be


14.7 psia and 60oF)

Bg 35.3 400 psia / 0.95 180 460


4 - 62

Bg 23.2 SCF/ft 3
and
Free gas =

43,560 160 4 0.18 0.15 23.2

Free gas = 17,500,000

SCF

OIP at start of WF 7758 Ah So / Bo

OIP 7758 160 4 0.18 0.61 /1.15

OIP 474,000 STBO at start of waterflood


2. If all free gas is re-dissolved in the oil, the average increase in

Rs

Rs

would be:

17,500,000 SCF
474,000 STBO

Rs 37 SCF/STBO which in most waterfloods would be


considered negligibly small.
_____________________________________________________________________
The oil bank will usually displace most of the free gas saturation from the reservoir
leaving only a small portion of the gas to re-dissolve in the oil. However, as Example
4-6 indicates, even if all of the gas is re-dissolved into the oil, the increase in

Rs

will

usually be negligibly small. As a result, the PVT properties of the crude oil at the
start of injection usually represent reasonable estimates of the oil PVT properties
throughout the project life even if reservoir pressure is increased to the original
pressure. Consequently, in the oil bank the oil saturation will be similar to the oil
saturation that existed at the time of field discovery but the PVT properties will
4 - 63

correspond to those properties measured at the reservoir pressure at the start of the
waterflood.
Following gas fillup, it would be desirable to obtain a new PVT sample. The PVT
characteristics should be representative of the reservoir oil and results will be more
reliable than estimating PVT properties measured at reservoir pressure at the start of
the waterflood. The key for a reliable fluid sample is the requirement that gas fillup
has occurred.
VII. Reservoir Pressure Distribution
The pressure distribution existing throughout a reservoir during primary depletion is
dependent upon many variables such as permeability, porosity, thickness, rock and
fluid compressibilities, wellbore skin factors, and well production rates.

In

waterflooding, the pressure distribution is impacted by these same factors plus


additional factors such as injection pressure, mobility ratio, and the degree of gas
fillup.
As a general statement, for either primary depletion or waterflooding, most of the
decrease in reservoir pressure occurs near the producing or injection wells. Reservoir
engineering texts10,11 describe in considerable detail the nature of pressure distribution
from the wellbore to the external drainage radius. For example, it can be shown for
pseudo-steady state radial flow under primary production of a constant
compressibility fluid that reservoir pressure at any point within the external drainage
radius, re , can be calculated by equation 4.47 as:

p pwf

141.2qB r
ln s
kh
rw

where

= pressure at radius
4 - 64

r , psi

(Eq. 4.47)

pwf

= bottomhole producing pressure, psi

= production rate, STB/D

= net thickness, feet

= effective permeability to the produced fluid, md

= skin factor at producing well, dimensionless

rw

= producing wellbore radius, feet

= fluid formation value factor, RB/STB

= fluid viscosity, cp

Equation 4.47 can be used to compute the pressure distribution between the producing
well radius and the external drainage radius. Under pseudo-steady state conditions
during primary depletion, each well drains a reservoir pore volume in proportion to its
production rate. For example, if two well are producing at 100 and 200 STB/D
respectively, the well producing at the higher rate drains a pore volume that is twice
the pore volume of the well producing at the lower rate.
_____________________________________________________________________

EXAMPLE 4:7
Consider two wells within the same reservoir producing at equal rates from a single
layer of constant thickness, porosity, and permeability.

The two wells will be

draining equivalent pore volumes. For a line connecting the two wells, the drainage
distance for each producing well can be assumed to be at the mid-point. For the
conditions listed below, compute the pressure distribution along a line connecting the
wells for the following conditions:

4 - 65

= 45 STBO/D

pwf

= 100 psi (both wells)

ko

= 20 md

s1 s2

= 0.0

= 5 feet

Bo

= 1.2 RB/STBO

re

= 660 feet (40 acre density)

= 1.5 cp

rw

= 0.25 feet

SOLUTION
Pressure versus radius can be computed from Eq. 4.47.

p 100

141.2 45 1.2 1.5 r


ln
0

20
5
0.25


or

p 100 114.4 ln r 1.39


r , ft

p, psi

r , ft

p, psi

0.25

100

100

785

1.0

259

200

865

5.0

443

300

911

10.0

522

400

944

25.0

627

500

970

50.0

706

660

1000

Figure 4-26 presents the pressure distribution along a line connecting the two wells.
It is significant to note that of the total pressure drop for a given well and external
4 - 66

drainage radii, in this case 0.25 feet and 660 feet respectively, about 50 percent of the
total pressure drop occurs within 12.7 feet of the producing well radius representing
less than two percent of the total drainage radius. For radii of 0.25 feet and 660 feet,
this 50 percent pressure drop is valid for different flow rates, permeabilities, and fluid
viscosities although the total pressure drop would be different. Also, if the skin factor
at the producing wells are different from zero, the pressure drop near the wellbore
could be more or less than the 50 percent factor depending upon if the skin factor is
negative or positive.

Figure 4-26
Pseudo Steady State Pressure Distribution Between Two Wells Producing at Equal
Rates From a Flow Unit of Constant Permeability, Thickness, and Porosity With Zero
Skin Factors at Each Well

1200

Pressure, psi

1000
800
600
400
200
0
0

10

20

30

40

50

60

70

80

90

100

Percent Distance

_____________________________________________________________________
At the start of water injection, the pressure distribution begins to look very different
and changes rapidly until gas fillup is achieved. Assume a gas saturation is present at
the start of injection, an oil bank forms, and the water and oil within the water and oil
banks are incompressible. The free gas existing in that portion of the reservoir
beyond the outer radius of the oil bank usually exists at moderate to low pressures and
4 - 67

is very compressible.

(Large gas saturations usually result from low reservoir

pressure.) As a result of the compressible gas, pressure is not effectively transmitted


through the gas zone to the producing well.

Therefore, the reservoir pressure

distribution in those parts of the reservoir containing the gas saturation is not
materially altered from the pressure distribution created by primary depletion at the
start of injection. Consequently, near the producing well, reservoir pressure (and fluid
saturations) and production rates are not significantly altered until gas fillup occurs
within a flow unit.
During the gas fillup period, the only part of the reservoir flow unit in which there is
an impact on pressure behavior is that portion of the reservoir within the water bank
and oil bank.

In fact, the pressure at the outer radius of the oil zone can be

approximated by the pressure that existed at that position at the start of the
waterflood. As long as the oil and water regions are circular around the injection well
as depicted in Figure 4-22, and if the mobility ratio (discussed in a later chapter) is
unity, then Eq. 4.47 can be modified as:

p pwi

141.2iw w r

ln
s
i
kwh
rw

where

pwi

= bottomhole injection pressure, psi

iw

= water injection rate, BWPD ( Bw 1.0 )

kw

= effective permeability to water at

si

= injection well skin factor, dimensionless

= water viscosity, cp
4 - 68

S wbt , md

(Eq. 4.48)

Because Eq. 4.48 assumes a unit mobility ratio,


to

ko / o

kw / w in the water zone is equal

in the oil zone.

____________________________________________________________________

EXAMPLE 4:8
Consider the two wells described in Example 4:7. Compute the pressure distribution
between the injector and producer when the oil bank radius is midway between the
two wells.
Mobility Ratio = 1.0

Si

= 0.00

pwi

= 4200 psi

kw / w ko / o 20 /1.5 13.33 md/cp


h

= 5 feet

p at midpoint

= 1000 psi

iw

= 192 BWPD at the initial oil radius of 660 feet

SOLUTION
For a unit mobility ratio, the pressure distribution within the water zone and oil zone
is computed by Eq. 4.48.

p pwi

141.2iw w r

s
ln
i

k wh rw

4 - 69

p 4200

141.2 192
ln r 1.39 0.00
13.33 5

p 4200 406.8 ln r 1.39

r , ft

p, psi

r , ft

p, psi

0.25

4200

100

1763

1.0

3635

200

1481

5.0

2981

300

1315

10.0

2699

400

1198

25.0

2326

500

1105

50.0

2044

660

1000

Further, the pressure distribution from the mid-point through the gas zone to the
producing well has not materially changed from the pressure distribution existing at
the start of injection. Figure 4-27 shows the pressure distribution between the injector
and producer.

4 - 70

Figure 4-27
Pressure Distribution Between an Injection and Producing Well Before Gas Fillup for
a Unit Mobility Ratio When the Radius to the Oil Front is 50% of the Total Distance
from the Injector to Producer with Zero Skin Factors at Both Wells
Radius to Oil Front

4500

Unaffected Gas Zone

Pwi = 4200 psi

4000
Pressure, psi

3500
3000
2500
2000
1500
1000
500
Pwf = 100 psi

0
0

10

20

30

40

50

60

70

80

90

100

Percent Distance

_____________________________________________________________________
At gas fillup, the oil bank arrives at the producing well and the gas is no longer
present. If the skin factors at each well are identical and the mobility ratio is unity,
the pressure distribution near the injection well is the reciprocal of the pressure
distribution at the producing well.
_____________________________________________________________________

EXAMPLE 4:9
For Example 4:8, plot the pressure distribution between the injection and production
wells. For convenience, after gas fillup the pressure at the producing well is set at
200 psi so the total pressure drop between the injector and producer is 4000 psi. The
pressure at the mid-point is 2200 psi.

4 - 71

SOLUTION
Figure 4-28 presents the steady state pressure distribution after fillup. It is observed
that of the total pressure drop between the two wells, about 40 percent occurs over the
first 10 percent of the distance and an additional 40 percent of the pressure drop
occurs over the final 10 percent of the distance. The pressure drop covering the
distance from the 10 percent to 90 percent marker is only 20 percent of the total.
Figure 4-28
Pressure Distribution Between an Injection and Producing Well After Gas Fillup for
an Incompressible Reservoir System with a Mobility Ratio of Unity and Zero Skin
Factors at Both Wells
Liquid Filled (Incompressible) Reservoir

4500
4000

40% of Total Pressure


Drop = 1600 psi

Pressure, psi

3500
3000

40% of Total Pressure


Drop = 1600 psi

20% of Total Pressure Drop = 800 psi

2500
2000
1500

Average Pressure = 2200 psi

1000
500

Pwf=200 psi

0
0

10

20

30

40

50

60

70

80

90

100

Percent Distance

Figure 4-29 presents a composite pressure distribution at three different times, prior to
the start of injection, mid-way to gas fillup and after gas fillup.

4 - 72

Figure 4-29
Composite Pressure Distribution at 3 Different Times:
1) Immediately Prior to Start of Injection
2) Midway to Gas Fillup
3) After Gas Fillup

4500
4000
Pressure, psi

3500
3000
2500
2000
1500
1000
500
0
0

10

20

30

40

50

60

70

80

90

100

Percent Distance

_____________________________________________________________________
Example 4:9 illustrates the pressure distribution assuming skin factors at each well are
zero. If the skin factors are net zero in each well, then pressure distribution is altered.
The pressure drop at each well due to the skin factor can be shown from well testing1
to be:

pskin

141.2qB
s
kh

The pressure drop due to skin at the injection and production wells can be computed
as:

psi

141.2iw w
si
k wh
4 - 73

and

psp

141.2qo Bo o
sp
ko h

If the total pressure drop between the injector and producer is denoted by

ptotal ,

then the pressure drop through the reservoir, after excluding the pressure loss due to
the skin is

pres ,

pres ptotal psi psp


For a layer of uniform thickness and permeability and possessing a unit mobility ratio,
the

pres

is similar to Figure 4-28 with 40 percent of the pressure drop occurring

near the injection and producing wells.


Finally, if the permeability or thickness of a flow unit changes from the injector to
producer or if the mobility ratio is greater than or less than one, these factors will
influence pressure distribution. However, most of the total pressure drop will usually
occur near the injector or producer wells with only a small portion of the pressure
drop taking place in the main portion of the reservoir where most of the pore volume
is located. Pressure distribution through the reservoir is significant because it impacts
injection rate, production rate, flow paths (streamlines), and ultimately areal sweep
efficiency.
VIII. Maintain Optimum Reservoir Pressure to Minimize

Sor

Willhite6 and Craig7 indicate, in some instances, oil recovery can be increased if the
reservoir pressure is carefully controlled so as to leave a trapped gas saturation within
the oil bank. The idea is to reduce the

Sor

value by an amount equal to the trapped

gas saturation. For example, if the residual hydrocarbon saturation is 30 percent and
if a trapped gas saturation can be maintained at 5 percent, the residual liquid
4 - 74

hydrocarbon saturation would be 25 percent. In this case, the residual liquid would be
reduced by 16.67 percent. Chik, et al12 described a case study involving this concept.
In theory, the concept seems to have application. However, selecting and maintaining
the optimum reservoir pressure to maintain this critical gas saturation is difficult to
achieve in practice.
IX. Gravity Under-Running
When flooding zones of relatively large thicknesses, the question of gravity underrunning occurs. A method has been presented by Dietz13 which can be used to
determine the angle at which water can be expected to enter an oil bearing sand. This
method applies to floods with favorable mobility ratios in reservoirs which are
horizontal or have relatively low dips. According to Dietz, the angle,

, between

the formation and the flood front will be:


2045iw w o
k rw k ro
Tan
Ak ( w o )

4 - 75

(Eq. 4.45)

FIGURE 4-30
WATER UNDER-RUNNING DUE TO GRAVITY

WATER

OIL

Practical units, as previously defined, are used in the equation.

The relative

permeability to oil is defined at the initial oil saturation, and the relative permeability
to water is expressed at the average water saturation behind the front. It is noted that
as iw increases, the front becomes more vertical (i.e.,

approaches 90 degrees).

An application of this method is given by Cronquist14.


X. Summary
The basic frontal advance theory for immiscible fluid displacement has been developed
and illustrated. From this material, it is seen that, in most cases, relative permeability
and viscosity ratio control the displacement behavior. However, factors such as gravity
and capillary pressure must be considered in special cases. Procedures were shown
which could be used to approximate saturation distributions, production history, and
WOR behavior of a waterflood process. The presence of a free gas saturation must be

4 - 76

considered. Free gas resulting from primary pressure depletion leads to the concept of
gas fillup.
The limitations of the linear flow calculations for a homogeneous reservoir having
constant rock and fluid properties must be considered. It was assumed that the total
throughput rate is constant and equal to the water injection rate. Areal and vertical
coverage were taken to be unity, and it is assumed all injected water contributes to the
displacement process (i.e., no water lost at the boundaries or to other formations).
Thus, the main contribution of the calculations in this chapter is to give us an
understanding of the displacement behavior in that portion of the reservoir contacted
by the injected water. When flow is not linear, or when areal or vertical heterogeneities
exist which reduce coverage of the injected water, modifications to the model must be
made or a completely different model must be used. These adjustments are discussed
in the next several chapters.

4 - 77

CHAPTER 4 REFERENCES
1. Lee, J.: Well Testing, Textbook Series, SPE, Dallas (1982) 1.
2. Buckley, S.E. and Leverett, M.C.: Mechanism of Fluid Displacements in Sands,
Trans., AIME (1942) 146, pp. 107-116.
3. Terwilliger, P.L., Wilsey, L.E., Hall, H.N., Bridges, P.M., and Morse, R.A.: An
experimental and Theoretical Investigation of Gravity Drainage Performance,
Trans., AIME (1951) 192, pp. 285-296.
4. Welge, H.J.: A Simplified Method for Computing Oil Recovery by Gas or Water
Drive, Trans. AIME(1952) 195, pp. 91-98.
5. Felsenthal, M. and Yuster, S.T.: A Study of the Effect of Viscosity in Oil Recovery
by Waterflooding, paper SPE 163-G presented at the 1951 SPE West Coast
Meeting, Los Angeles, Oct. 25-26.
6. Wilhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
7. Craig, F.F.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Richardson, TX (1971) 3, p. 39.
8. Callaway, F.H.: Evaluation of Waterflood Prospects, JPT (Oct. 1959) pp. 11-16.
9. McCain, W.D.: The Properties Of Petroleum Fluids, PennWell Publishing
Company, Tulsa, OK (1990).
10. Craft, B.C. and Hawkins, M.F.: Applied Petroleum Reservoir Engineering,
Prentice-Hall, Inc., Englewood Cliffs, N.J. (1959).
11. Dake, L.P.: Fundamentals of Reservoir Engineering, Elsevier Science B. V.,
Amsterdam (1978) p. 372.
12. Chik, A.N., Selamat, S., Elias, M.R., White, J.P. and Wakatake, M.T.: Guntong
Field: Development and Management of a Multiple-Reservoir Offshore
Waterflood, JPT (Dec. 1996) p. 1139.
13. Dietz, D.N.: A Theoretical Approach to the Problem of Encroaching and
Bypassing Edge Water, Proc., Koninkl. New. Akad. Wetenschap (1953) B56, 38.
14. Cronquist, C.: Waterflooding by Linear Displacement in Little Creek Field,
Mississippi, Trans., AIME (1968) 243, pp. 525-533.

4 - 78

APPENDIX A CHAPTER 4
DEVELOPMENT OF FRONTAL ADVANCE EQUATION

Consider the simultaneous linear flow of oil and water in the water zone in a
porous system of cross-sectional area,

A , and length, x , as shown in Figure 4A-

1.
FIGURE 4A-1
LINEAR MODEL FOR
DERIVATION OF FRONTAL ADVANCE EQUATION

ko S wir ,

fw x

f w x x

X
X + X

A material balance for this segment of the reservoir rock can be written as:

of Water Flow Rate of Flow Rate of


Rate
Accumulation Water In Water Out
These terms can be expressed symbolically as:
4 - 79

(Eq. 4A.1)

Flow rate of water in = iw f w x , bbls


Flow rate of water out =

iw f w

Rate of water accumulation =

xx

, bbls

Ax S w

, bbls

5.615 t x x

With substitution of these terms, the material balance becomes:

Sw x x
2

5.615iw f w / x x f w / x

Take the limit of this equation as

(Eq. 4A.2)

x approaches zero to obtain:

5.615iw f w 5.615iw f w S w
S w

A
x

A
S
x

w t
(Eq. 4A.3)
Equation 4A.3 gives the water saturation as a function of time at a particular
location,

x , within the linear system.

A more useful expression, however, would

be one that gives the saturation as a function of location at a particular time.


This is possible if it is observed that the water saturation is a function of both
position and time. For example:

S w S w x, t
Therefore, the total derivative of

S w is:

4 - 80

(Eq. 4A.4)

S w
S w
dS w
dx
dt
x t
t x

Eq. 4A.5)

Since we are interested in determining the saturation distribution in the reservoir,


the procedure taken here will be to trace the movement of a particular water
saturation. If a fixed water saturation,

S w , is considered, then dS w 0 , so that:

S
S
0 w dx w dt
t t
t x

(Eq. 4A.6)

and:

S w
S w dx

t x
x t dt S w

(Eq. 4A.7)

This mathematical relationship can be substituted into Eq. 4A.3 to yield:

5 . 615 i w
dx

A
dt S w
If the total injection rate is constant,

f w

S w

(Eq. 4A.8)

f w is independent of time. Accordingly:

f w
df

w
S w t dS w

(Eq. 4A.9)

and:

5.615iw df w
dx

A dS w
dt S w

(Eq. 4A.10)

Integration of this expression yields the following equation which is widely


referred to as the frontal advance equation.
4 - 81

5.615iwt df w 5.615Wi df w

A dS w
A dS w

(Eq. 4A.11)

where:

= distance traveled by a fixed saturation,

iw

= injection rate, bbl/day

= time interval of interest, days

Wi

= cumulative water injected at time t , reservoir bbls

S w , during time, t , feet

df w
= slope of the fractional flow curve at the water saturation of interest
dS w
Eq. 4A.11 is the equation used to compute the water saturation distribution in the
water zone up to water breakthrough. While this is an important relationship, its
greatest value in waterflood analysis is that it is the basis for computing

Sw .

4 - 82

S wf and

APPENDIX B CHAPTER 4
BUCKLEY-LEVERETT THEORY
I. Buckley-Leverett Theory
It was shown by Buckley and Leverett2 that the frontal advance equation (Eq. 4.16)
can be used to compute the saturation distribution in a linear waterflood system as a
function of time. According to Eq. 4.16, the distance,

x , moved by a given saturation

in the time interval, t , is proportional to the slope of the fractional flow curve at the
particular saturation of interest. Therefore, if the slope of the fractional flow curve is
graphically determined at a number of saturations, it is possible to calculate the
saturation distribution in the reservoir as a function of time.

Furthermore, the

saturation distribution can be used to predict oil recovery and required water injection
on a time basis. However, this procedure must be used with some caution as the
Buckley and Leverett analysis may give a multi-saturation distribution that is
physically impossible.
The problem arises because of the shape of the fractional flow curve. It is noted on
Figure 4-4 that equal values of slope,

df w dS w , can occur at two different water

saturations. According to Eq. 4.16, this means two different saturations can occur at
the same location in the reservoir at the same time---this is not possible. Moreover,
under some conditions it can be shown that theory predicts a triple-valued
distribution. An example of the multi-valued saturation distribution resulting from
this situation is shown in Figure 4B-1.

4 - 83

Figure 4B-1
MULTI-VALUED SATURATION PROFILES

Water Saturation, percent

100

S wm

50

Reservoir Oil

Flood Water

t1

t2

t3

Initial Water

Distance

To rectify this mathematical difficulty, it was suggested by Buckley and Leverett that
a portion of the calculated saturation distribution curve is imaginary and the real
curve contains a discontinuity at the front. The method for finding the real curve is
illustrated by Figure 4B-2. The imaginary portion of the curve is shown as a dashed
line. The real curve is shown as a solid line which becomes discontinuous at a
distance

x f . This distance is based on a material balance of the injected water and

can be determined graphically by locating the front in such a position that areas A and
B are equivalent.

4 - 84

Figure 4B-2

LOCATION OF FLOOD FRONT BY BUCKLEY-LEVERETT PROCEDURE

Water Saturation, percen

100

S wm

t t1

50

0
Distance

xf

The Buckley and Leverett procedure illustrated in Figure 4B-2 neglects capillary
pressure. Consequently, in a practical situation, the flood front will not exist as a
discontinuity but will exist as a stabilized zone of finite length with a large saturation
gradient. This was recognized and presented in a paper of fundamental importance by
Terwilliger, et al3.
II. Stabilized Zone Concept
The first of many papers which confirm the frontal advance theory was presented by
Terwilliger, et al. While applying this theory to a gravity drainage system, they found
that, at the leading edge of the front, a zone exists where displacing fluid saturations
all moved at the same velocity. Accordingly, the shape of the front was observed to
be constant with respect to time. This zone was termed the stabilized zone. Further,
it was found that by using the complete fractional flow equation (including capillary

4 - 85

effects) along with the frontal advance equation, the saturation distribution computed
using the Buckley-Leverett theory matched the saturation distribution observed
experimentally. The stabilized zone is illustrated in Figure 4B-3.

FIGURE 4B-3
SATURATION DISTRIBUTION SHOWING EXISTENCE
OF STABILIZED AND NONSTABILIZED ZONES

Water Saturation

swm

Nonstabilized zone where

sw
0
x

swf
Stabilized Zone

x

t Sw

Front at Time t1

Front at Time t2

= constant

swc
Distance

In theory, the flood zone has a measurable width with a large saturation gradient. In
practice, the transition zone at the leading edge of the water front can be closely
approximated as a step function as illustrated in Figure 4B-4.

4 - 86

FIGURE 4B-4
SATURATION DISTRIBUTION
WITH A STEP FUNCTION AT WATER FRONT
Water Zone

Saturation, percent

100

Oil Zone

Stabilized
Step Front

S wf
S wc
0

Distance

It was also observed that the saturation at the leading edge of the stabilized zone,

S wf , could be defined as the tangent point on the fractional flow curve obtained by
drawing a tangent line originating at the point
proven by Welge3.

S w S wc , f w 0.

This was later

Accordingly, the velocity of this particular saturation is

proportional to the slope of the fractional flow curve at this point, i.e., the slope of the
tangent line. Since all saturations in the stabilized zone move at the same velocity, it
follows that

df w dS w must be the same for all saturations in the stabilized zone.

This slope is defined by a line drawn tangent to the fractional flow curve from the
initial water saturation. The fractional flow curve with the described tangent line is
illustrated in Figure 4B-5.

4 - 87

FIGURE 4B-5
FRACTIONAL FLOW CURVE
SHOWING STABILIZED ZONE
EFFECT

1.0

Curve Demanded
By Stabilized Zone

f w 0.5

S wf

S wir

0.0
0

50

100

Water Saturation, percent


Thus, it is concluded that the saturation distribution in the stabilized zone

S wc S w S wf should be computed based on the slope of the tangent to the

fractional flow curve. Many mathematical and experimental studies conducted more
recently have verified the presence of the stabilized zone. Also, several studies have
considered the effect that the stabilized zone has on waterflood performance. It is
4 - 88

generally accepted that the length of the stabilized zone is negligible at practical flood
rates and the method of Welge (to be covered later) can be used to predict linear flood
results.
Behind the flood front is a zone where the saturation distribution does change with
time. Appropriately, the zone is referred to as being non-stabilized. In contrast to the
stabilized zone, saturations change very little with distance in this zone and we can
write

S w
x

0 . Since the capillary pressure term in the fractional flow equation

can be written, according to the chain rule, as:

Pc
x

Pc S w

S w x

(Eq. 4B.1)

it follows that the capillary pressure gradient can be neglected in this zone. The nonstabilized zone was illustrated in Figure 4B-3.
III. Welge Solution to Buckley-Leverett
A. Water Saturation at the Front
Welge4 greatly simplified the graphical procedure of Buckley and Leverett but
requires the initial water saturation be uniform.

At some time after the

displacement process begins, the saturation distribution will appear as depicted


by Figure 4B-6.

4 - 89

FIGURE 4B-6
SATURATION PROFILE DURING FLOOD

Water Saturation

S wm
Front

S wf

Swc

Xf

The area of the shaded rectangle between

S wc and S wf is:

x f S wf S wc S wf x dS w
wc

(Eq. 4B.2)

where:

S wf = water saturation at the front


Substituting x from Eq. 4.14 into Eq. 4B.2:

S wf

x f S wf S wc S

wc

4 - 90

5.615iwt df w

dS
A dS w w

or:

x f S wf S wc

5.615iwt

f
f
w / S wc
A w / Swf

Thus,

5.615iwt f w / Swf f w / S wc

xf
A
S wf S wc

If Eq. 4.16 is written for the special case where

5.165iwt df w
xf

A dS w
S

(Eq. 4B.3)

x xf :
(Eq. 4B.4)

S wf

Equating Eq. 4B.3 and 4B.4:

f w / S wf f w / S wc
df w

dS
S wf S wc
w S wf

(Eq. 4B.5)

The graphical interpretation of Eq. 4B.5 is that a line drawn tangent to the
fractional flow curve from the point
tangency equal to

fw / S

wf

f w / S wc , S wc will have a point of

, S wf . That is, the point of tangency is the water

saturation at the front. This is illustrated by Figure 4B-7.

4 - 91

FIGURE 4B-7
DETERMINATION OF WATER SATURATION AT THE
FRONT FROM THE FRACTIONAL FLOW CURVE

1.0

Tangent
Point

f w / S wf

fw

S wf
0.0
0

100
Water Saturation, percent

B. Average Water Saturation


The average water saturation behind the flood front can also be determined
using the fractional flow curve. Again consider the saturation distribution at
some time during the flood as illustrated by Figure 4B-6. The total water
(injected and connate) in the reservoir behind the front is:
4 - 92

Total H 2O= Ao wf S wdx Ao wm x dS w

(Eq. 4B.6)

where:

S wm maximum water saturation 1.0 Sor


and
S
S

Total H 2O A owf x dS w S wm x dS w
wf

(Eq. 4B.7)

S
Total H O AX S A wm x dS
S
2
f wf
w
wf

(Eq. 4B.8)

Substitute Eqs. 4.16 and 4B.3, respectively, into Eq. 4B.8 to obtain:

df
w 5.615i t S wm df
w S wf
w
dS
w

Total H 2O 5.615iwtS wf

(Eq. 4B.9)
By definition, the average water saturation behind the front is:

Sw

total H 2O
total water behind front

total flooded pore volume


Ax f

(Eq. 4B.10)

or:

Sw

5.615iwtS wf df

w
dS
AX f
w S

wf

4 - 93

5.615iwt

AX f

S wm
S d f w
wf

(Eq. 4B.11)

By substituting Eq. 4B.4 into this expression, the following expression is


obtained for

S w.

S w S wf

1.0 f wf

(Eq. 4B.12)

df w

dS
wf

All of the information required to compute

Sw

using Eq. 4B.12 is available

from the tangent point of the fractional flow curve.

However, an easier

graphical procedure can be developed by considering the fractional flow curve


depicted by Figure 4B-8.

4 - 94

FIGURE 4B-8
DETERMINATION OF SLOPE
RELATIONSHIPS FOR THE FRACTIONAL
FLOW CURVE
1.0

SwA ,1.0

SwA , f wf

f wf

fw

S wf
0.0
0

100
Water Saturation, percent

It is observed on Figure 4B-8 that the tangent line intersects the line
corresponding to

S wA .

f w 1.0

at a saturation which is arbitrarily defined as

The slope of the tangent line can be defined in terms of this saturation

according to the equation:


4 - 95

1.0 f wf
df w


dS w f S wA S wf
This can be rearranged to solve for

S wA S wf

S wA .

1.0 f wf
df w

dS w f

(Eq. 4B.13)

Comparing Eqs. 4B.12 and 4B.13, it is evident that S w SwA . It is concluded,


therefore, that S w can be obtained by simply extending the tangent line to the
fractional flow curve to the point where f w 1.0 . This is illustrated by Figure
4B-9.

4 - 96

FIGURE 4B-9
GRAPHICAL DETERMINATION OF

Sw

1.0

fw

Swf

Swir

Sw

0.0
0

100
Water Saturation, percent

4 - 97

PROBLEM 4:1
Oil is being displaced by water in a horizontal, linear flow system where capillary
pressure effects are negligible. The relative permeability data for oil and water are
presented in the following table.

Sw , %
20
25
30
35
40
45
50
55
60
65
70
75
80

k rw

k ro

0.000
0.002
0.009
0.020
0.033
0.051
0.075
0.100
0.132
0.170
0.208
0.251
0.300

0.800
0.610
0.470
0.370
0.285
0.220
0.163
0.120
0.081
0.050
0.027
0.010
0.000

Pressure is being maintained at a constant value for which

Bw 1.0 RB/STB .

Bo 1.3 RB/STB

The connate water saturation is immobile. Compute and

compare the values of water saturation at the producing well,


saturation,

and

S wf , the average water

S wbt , and displacement sweep efficiency at the time of water breakthrough

for the following fluid combinations.


Case
1
2
3

o , cp

w , cp

50.0
5.0
0.5

0.50
0.50
0.50

w o
0.01
0.10
1.00

The fractional flow curves for each of these cases are presented in Figure 4P1.1.

4 - 98

FIGURE 4P1.1
FRACTIONAL FLOW CURVES FOR PROBLEM 4:1

1.0
Case 1
w
0.01
o

0.8

Case 2
w
0.10
o

0.6

fw
0.4
Case 3
w
1 . 00
o

0.2
0.0
0

20

40

60

80

Water Saturation, percent

4 - 99

100

PROBLEM 4:2
Consider the following data for a linear reservoir which is to be waterflooded.

iw

1000 bbl/day

2.48 cp

18 percent

400 feet

S wc

20 percent

Bo

1.15 RB/STB

50,000 ft2

Bw

1.0 RB/STB

0.62 cp

Sw

kro

krw

0.20
0.30
0.40
0.50
0.55
0.60
0.70
0.80
0.85

0.930
0.600
0.360
0.228
0.172
0.128
0.049
0.018
0.000

0.000
0.024
0.045
0.124
0.168
0.222
0.350
0.512
0.600

If areal and vertical sweep efficiency at initial water breakthrough are unity and capillary
pressure is negligible, determine the following information at water breakthrough and at
those times when the saturation at the producing well,

S w2 , is 55, 60, 65, 70 and 75

percent.
a.

Cumulative oil recovery, STB versus time

b. Producing water-oil ratio versus time


c.

Cumulative water injection, STB versus time, bbls

d. Oil producing rate, STB/D versus time


e.

Water producing rate, STB/D versus time


4 - 102

4 - 103

4 - 104

FLOOD PATTERNS AND AREAL SWEEP EFFICIENCY


I. Introduction
The frontal advance theory developed in Chapter 4 assumes flow between injection
and producing wells is linear (all flow paths are straight lines) and 100 percent of the
reservoir pore volume is contacted by injected water. Although this behavior may be
approximated in some elongated reservoirs, ideal linear flow would be possible only
if fluids could be injected into, and produced from, the entire reservoir cross section,
such as in a core flood, rather than through the limited area of a wellbore. This
problem is complicated further by the fact that most fields are developed, and
waterflooded, using some well pattern, either regular or irregular. Looking at fields
areally, both injection and production take place at points.

If the patterns are

symmetrical, the shortest travel path and the largest pressure gradient will occur along
a straight line between producers and injectors. Accordingly, injected water which
travels along the most direct flow path, defined as a streamline, will reach the
producing wells first. Water traveling along longer flow paths (streamlines) will not
have reached the producing well at the time of breakthrough on the short direct
streamline and, consequently, part of the reservoir will not have been contacted by
water at that time. That fraction of a waterflood pattern which has been areally
contacted by water at a given time during a flood is referred to as the areal sweep
efficiency,

EA.

In general, areal sweep efficiency will depend upon (1) mobility

ratio, (2) geometric configuration of the flood pattern, (3) reservoir pressure
distribution, (4) reservoir heterogeneities, and (5) cumulative volume of water
injected into the pattern area.
II. Mobility Ratio
Mobility is the ease with which a fluid moves in the reservoir. One of the most
important characteristics of a waterflood is the mobility ratio. It is defined1,2 in terms

5-1

of the effective permeability and viscosity of the displacing and displaced fluids
involved in the flood according to the following relationship.

M Mobility Ratio

Displacing
Displaced

Mobility of Displacing Phase


Mobility of Displaced Phase
k

Displacing

Displaced

(Eq. 5.1)

(Eq. 5.2)

A. Water Displacing Oil


For a conventional waterflood where injected water is displacing oil, the mobility
ratio is:

k w / w k w o krw o

ko / o ko w kro w

(Eq. 5.3)

It is important to note that the relative permeabilities to water and oil in Eq. 5.3
are defined at two separate points in the reservoir, i.e.,

krw is

the relative

permeability to water in the water-contacted (flooded) portion of the reservoir and

kro

is the relative permeability to oil in the oil bank ahead of the water zone.

It should be pointed out that the definition of mobility ratio expressed by Eq. 5.3
has been standardized by the Society of Petroleum Engineers (SPE) since 1957.
Prior to this time, however, when many studies of waterflooding were conducted,
mobility ratio was defined at the user's discretion.

Therefore, when using

technical literature related to mobility ratio, an engineer is advised to be fully


aware of the definition employed, particularly if the literature was published prior
to 1957.
5-2

Equation 5.3 shows mobility ratio is a function of effective or relative


permeability which, in turn, is a function of fluid saturation. This presents a
dilemma because, as shown in Chapter 4 by the frontal advance theory of oil
displacement, a saturation gradient exists within the water zone. Since

krw

is

the relative permeability to water behind the water front, the following question
arises. Which value of water saturation behind the front should be used to
determine

krw ?

conclusion that

A significant study by Craig, et al3 led to the widely accepted

krw

should be evaluated at the average water saturation,

It was further determined in this same study that


connate water saturation,

kro

S wbt .

should be evaluated at the

S wc , in the oil bank ahead of the water front.

Thus,

based on these conclusions:

o (krw ) S wbt
w (kro ) Swc

(Eq. 5.4)

Average water saturation behind the water front remains constant until the time of
water breakthrough. Based upon Eq. 5.4, it follows that mobility ratio will also
remain constant until breakthrough. When engineers use the term mobility ratio,
they are usually referring to the pre-breakthrough value. The mobility ratio after
water breakthrough is not constant; instead, it increases continuously in response
to the increasing average water saturation in the reservoir which, in turn, causes

krw to increase.
Mobility ratio is generally termed favorable or unfavorable depending on
whether its value is less than or greater than unity.

When

M 1.0 ,

the

mobilities of oil and water are identical, and they encounter the same resistance to
flow within the reservoir. When

M 1.0 , oil flows better than water and it is

easy for water to displace oil. This condition generally results in high sweep
5-3

efficiencies and good oil recovery. Conversely, when

M 1.0 , water flows

better than oil and is less effective in displacing oil. In general, sweep efficiency
and oil recovery tend to decrease as mobility ratio increases. In our experience,
the most common values of mobility ratios encountered during waterflooding
range from about 0.2 to 5.0.
Consider the definition of mobility ratio in Eq. 5.4. It is known that the viscosity
of water is dependent on water salinity and temperature as shown in Figure 5-1.

5-4

FIGURE 5-1
EFFECT OF TEMPERATURE ON
VISCOSITY OF SALT WATER

2.0

Viscosity, centipois

1.5

250,000 ppm
200,000 ppm
150,000 ppm
100,000 ppm
50,000 ppm
0 ppm

1.0

0.5

0.0
30

40 50

60

70

80 90 100 110 120 130 140 150 160 170

Temperature, degrees Fahrenheit


However, water viscosity is usually in the range of 0.4 to 0.8 cp. A value of 0.6
can be used as a reasonable approximation in many instances.
connate water saturation is immobile, or nearly immobile, then

5-5

Also, if the

kro Swc

can

be approximated as 1.0. Finally,

krw Swbt is dependent on items such as the

fractional flow graph, oil and water viscosity, and rock wettability; but it
frequently lies in the range of 0.1 to 0.3. Therefore, for many waterfloods,

krw Swbt can be approximated as being 0.2.

This leads to the following

calculation.

0.20
0.60 1.0

(Eq. 5.5)

or:

M 0.333 o

(Eq. 5.6)

In those instances where a rapid estimate of waterflood mobility ratio is needed,


Eq. 5.6 can be helpful. However, in performing a comprehensive waterflood
analysis, mobility ratio should be computed using Eq. 5.4.
B. Water-Oil Mobility Ratio after Breakthrough
At the time of water breakthrough on a flow path (streamline), the oil zone on that
streamline is no longer present.

Hence, on that flow path, the concept of

displacing fluid and displaced fluid tends to lose its meaning.

Following

breakthrough, the injected water tends to "drag" the oil. This behavior leads to
the concept of a drag waterflood.
After water breakthrough on a streamline, average water saturation,

S w S wbt ), begins to increase.


increases,

Sw

(where

Some publications indicate that since

k rw will increase and mobility ratio increases.

Sw

The increasing value

of mobility ratio is not consistent with the pre-water breakthrough value. After
breakthrough, the definition of mobility ratio changes. The oil bank has been
5-6

produced. Accordingly, the post-water breakthrough value of

should be used

with caution.
C. Oil Displacing Gas
When a free gas saturation exists prior to the start of waterflooding, an oil bank
usually develops as described in Eq. 4.61. Figure 5-2 presents the saturation
distribution between an injector and producer prior to gas fillup.

FIGURE 5-2
SATURATION DISTRIBUTION
EARLY IN THE LIFE OF A WATERFLOOD
WITHOUT TRAPPED GAS

Saturation, percent

100

Water Zone

} Sor

Oil Zone or Oil Bank

Unaffected
Gas Zone

Free Gas

Oil

S wbt
S wf
Water
0

Connate Water
Distance

In the previous paragraph and in the petroleum literature, mobility ratio


discussions usually refer to the mobility of the displacing water to the mobility of
the displaced oil. There is, however, another mobility ratio consideration. The
oil bank displaces all (or a significant portion) of the free gas. Therefore, an
alternate mobility ratio between the oil and gas is:

5-7

MO /G

( kro ) S wc
(krg ) S g

(Eq. 5.7)

where:

MO /G

= mobility ratio between displacing oil and displaced gas

( k rg ) S g

= relative permeability to gas at the gas saturation,

= gas viscosity, cp

Sg

For most waterfloods with oil displacing gas, the following approximations can
be made.

( k ro ) S wc 1.0
(krg ) S g 0.1

g 0.02 cp
Thus,

MO /G

1.0 0.02

0.1 o

(Eq. 5.8)

or:

MO /G

0.2

(Eq. 5.9)

Therefore, the mobility ratio between the displacing oil bank and the displaced
gas is strongly dependent on the oil viscosity as shown in the following table.
5-8

o , cp M O / G
0.50
1.00
10.00

0.40
0.20
0.02

These results indicate that in most waterfloods of moderate oil viscosity, the

MO /G

will be less than about 0.2.

In general, mobility ratio between the displacing phase (water or oil) and the
displaced phase (oil or gas) is used to estimate areal sweep efficiency for certain
patterns. Also, mobility ratio affects water injection rate. These two definitions
of mobility ratio play an important role in calculating areal sweep efficiency and
oil recovery as shown later in this chapter. Use of the term mobility ratio in
subsequent text will refer to the normal mobility ratio between water displacing
oil except where noted.
III. Basic Flood Patterns
Many fields have been developed for primary and secondary recovery using an
irregular well spacing. Generally, it is believed that relatively uniform well spacing
and drilling patterns lead to better recovery than with irregular patterns. The ideal
drainage or secondary pattern must consider reservoir continuity, directional
permeability, and reservoir pressure distribution. At the time a waterflood begins, a
field may be completely developed. Since infill wells are expensive to drill and
equip, waterflood development generally utilizes well patterns (or lack thereof)
existing at the start of injection. When irregular patterns exist, it may be desirable to
infill drill and regularize the flood. This process usually results in incremental oil
recovery, but this additional recovery can be difficult to quantify without a complete
geological and numerical simulation study. Few (if any) ideal regular patterns exist
in field operations. Wellbores may exist such that an apparent pattern is present, but
a careful analysis of such items as directional permeability, imbalances between
5-9

injection and withdrawal, and lateral discontinuity of reservoir rock cause the "actual
flow" within the reservoir to be quite different than indicated by consideration of
only wellbore locations. Even so, to understand waterflood behavior, it is essential
that an understanding of commonly used flood patterns be learned.
A. Direct Line Drive
As noted previously, the only way to achieve 100 percent areal sweep at the time
of breakthrough would be to inject fluid over an entire vertical plane so as to
achieve linear flow.

This is not physically possible but can be approached

somewhat with a pattern where the producing and injection wells directly offset
each other. Figure 5-3 is a direct line drive pattern. As depicted by Figure 5-4,
the sweep efficiency1 of this direct line drive pattern improves as the
increases, where

is the distance between adjacent wells in a row.

FIGURE 5-3
DIRECT LINE DRIVE PATTERN

a
d

Pattern Boundary
Injection W ell
Producing W ell

5 - 10

d /a

ratio

The relationship between

d /a

and

E Abt

is presented in Figure 5-4 for a unity

mobility ratio. It should also be noted that the ratio of producers to injectors is
unity for this pattern.

Areal Sweep Efficiency at Breakthrough

FIGURE 5-4
AREAL SWEEP EFFICIENCY OF DIRECT LINE
AND STAGGERED LINE DRIVE
1
0.8
Mobility Ratio = 1.0
(Reference 1)

0.6
0.4
0.2

Direct Line
Staggered
Line

0
0.0

0.4

0.8

1.2

1.6

2.0

d/a

2.4

2.8

3.2

3.6

B. Staggered Line Drive


As shown by Figure 5-5, the staggered line drive1 is simply a modification of the
direct line drive where rows of producing and injection wells are displaced onehalf the inter-well distance. The effect of this staggering, as shown by Figure 5-4,
is to significantly increase the breakthrough areal sweep efficiency as compared
to the direct line drive, especially for low

d /a

ratios. Accordingly, if field

development permits, this flood pattern is preferable to the direct line drive.

5 - 11

4.0

FIGURE 5-5
STAGGERED LINE DRIVE PATTERN

C. Five-Spot
The five-spot pattern, depicted by Figure 5-6, is a special case of the staggered
line drive where the

d /a

ratio is 0.5. This is the most commonly used

flooding pattern resulting primarily from the regular well spacing required, or at
least used, in most areas. Note that the drilling pattern required to have a fivespot is square, and the ratio of producers to injectors is unity. The five-spot is a
highly conductive pattern since the shortest flow path is a straight line between
the injector and producer. Also, the pattern gives good sweep behavior. The
square drilling pattern which yields the five-spot is also flexible enough that other
flood patterns can be generated simply by rearranging the position of the injection
and production wells. Examples of other patterns are the skewed four-spot, ninespot, and inverted nine-spot.

5 - 12

FIGURE 5-6
DEVELOPED FIVE-SPOT PATTERN

D. Nine-Spot
The normal nine-spot pattern, illustrated by Figure 5-7, can be developed from a
square drilling pattern. The injection well placement for the nine-spot pattern
leads to an injection-production well ratio of three. This type of system is very
useful if a high injection capacity is needed due to low permeability or similar
problems. The inverted nine-spot is used more than the normal nine-spot. In this
case, producing wells out number injection wells by a factor of three. The
inverted pattern is useful where fluid injectivity is high.

5 - 13

FIGURE 5-7
NINE-SPOT PATTERN

One of the major advantages of the nine-spot is flexibility.

Directional

movement of water and premature breakthrough in certain wells can necessitate


major conversions in flooding patterns. Some patterns are very difficult and
expensive to convert and may require extensive infill drilling. However, with
minimum effort, the inverted nine-spot can be revised to result in a 1:1 injectorproducer ratio pattern, either five-spot or line drive.
E. Seven-Spot
The normal seven-spot pattern, depicted by Figure 5-8, has two injection wells
per producer and has merit where injectivity is low. The pattern required is an
equilateral triangle or can be considered a staggered line pattern with a

d /a

ratio of 0.866. If a field is not developed on this pattern, a large number of infill
wells are generally required to make the pattern feasible.

5 - 14

FIGURE 5-8
SEVEN-SPOT PATTERN

An inverted seven-spot is also used occasionally. This pattern, also termed a


four-spot, has two producers per injector.
IV. Areal Sweep Efficiency
The amount of oil which can be displaced by waterflooding is directly proportional
to areal sweep efficiency. This was indicated previously by Eq. 1.1 which shows
that:

N D N E A EV ED

(Eq. 1.1)

It was shown in Chapter 4 how frontal advance theory can be used to predict the
displacement sweep efficiency,

ED .

The purpose of this section is to analyze those

factors which affect areal sweep efficiency and show how it can be determined using

5 - 15

commonly available information.

The vertical sweep efficiency,

EV

will be

discussed in a subsequent chapter.


A. Causes and Effects
Four major factors which affect areal sweep efficiency are mobility ratio,
directional permeability, well pattern, and pressure distribution between injectors
and producers. It was indicated previously that a decreasing mobility ratio causes
the sweep efficiency to increase.

Correlations of sweep efficiency versus

mobility ratio will be presented in a subsequent section for several commonly


occurring well patterns. Unfortunately, we do not have much control over the
mobility ratio or directional permeability. It will be shown later that if directional
permeability trends can be identified, injection and production wells can be
arranged to take advantage of the trends to enhance areal sweep efficiency. On
the other hand, it may be possible to maximize areal sweep through a careful
management of pressure distribution and proper injection-production pattern
selection.
Consider a developed five-spot pattern located in a homogeneous reservoir with
uniform permeability. Assume the bottomhole injection pressure to be 1000 psi,
the bottomhole producing pressure to be zero psi (any pressure drop,

p ,

is

acceptable), and the skin factors at the injector and producer to be identical.
Finally, assume the gas saturation at the start of water injection is zero and the
mobility ratio is unity. For these conditions, Figure 5-9 shows the pressure
distribution for one quarter of the five- spot.4 It is noted that about 40 percent of
the total pressure drop occurs near the injection well and another 40 percent
occurs nears the production well. This pressure behavior is consistent with the
concept of radial flow and occurs in both regular and irregular patterns. The lines
represent lines of constant pressure and are called isopotential or isobaric lines.
The lines in Figure 5-10 represent flow paths or streamlines.

5 - 16

FIGURE 5-9
ISOPOTENTIAL LINES FOR A DEVELOPED FIVE-SPOT
PATTERN WITH A MOBILITY RATIO EQUAL TO 1.0
Isopotential
Lines

Injector

Pwi 1000 psi


I1

P1

Producer

Pwf 0 psi
5 - 17

FIGURE 5-10
ISOPOTENTIAL LINES AND STREAMLINES FOR A
DEVELOPED FIVE-SPOT PATTERN WITH A MOBILITY
RATIO EQUAL TO 1.0
Injector
Streamline

I1

F
E
D
C
B
A
B
C
D
F

P1

Producer
5 - 18

The injected water and displaced oil travel along the streamlines. A basic law of
fluid flow is that flow lines (streamlines) are perpendicular to isopotential lines.
There are an infinite number of both isopotential lines and streamlines. However,
only eleven streamlines are shown in Figure 5-10. Significantly, flow along a
particular streamline is independent of fluid movement along another streamline.
Streamlines can be parallel or tangent (adjacent) to other streamlines but streamlines never intersect. Fluids moving along a streamline are not allowed to mix or
contact fluids in another streamline. Each streamline is an independent flow
path.
A simplified form of Darcy's Law helps to understand fluid movement along any
streamline.

For example, the fluid velocity along a particular streamline is

computed by:

q k P

A L

(Eq. 5.10)

Therefore, fluid velocity along a particular streamline is proportional to the fluid


mobility and pressure gradient.

For the case of uniform permeability (no

directional permeability trends) and unit mobility ratio, as described for Figure 59, all streamlines are subject to the same pressure drop (they all start from the
injector and end at the producer). However, the streamlines are of different
lengths.

The shortest distance between injector and producer is along the

diagonal (Streamline A) connecting the wells. It follows that the greatest pressure
gradient and highest fluid velocity, V A will occur along the shortest streamline,

A.

Moreover, it is clear that V A

VB VC VD VE VF .

Figure 5-11 shows the position of the injected flood front on each streamline at
several different times ( t1, t2 , t3

... tn ) up to breakthrough, tbt .

A smooth

curve connecting the flood front points on the different streamlines represents the
position of the injected water front within the pattern.
5 - 19

FIGURE 5-11
ISOPOTENTIAL LINES AND STREAMLINES FOR A
DEVELOPED FIVE-SPOT PATTERN WITH A MOBILITY
RATIO EQUAL TO 1.0
Injector

t1

t2

t3 t4 t5 t6 .....tbt

I1

P1

Producer
5 - 20

It is noted that the water saturation at the flood front on each streamline is
Moreover,

S wf

is the same value for each streamline because

dependent only on the fractional flow

( fw )

S wf .

S wf

is

graph discussed in Chapter 4 and

not the position of the flood front. Also, the average water saturation between
the injection well and flood front is

S wbt

for each streamline. Therefore, prior

to water breakthrough on the shortest streamline, the average water saturation in


the area swept by water (which is the area behind the front) is

S wbt .

Thus, up

to and including initial water breakthrough, a material balance on the injected


water for any pore volume can be expressed as:

Wi V p E A ( S wbt S wc )

(Eq. 5.11)

where:

EA

= areal sweep efficiency of the pore volume, fraction

S wc

= connate water saturation at the start of injection, fraction

S wbt

= average water saturation in water swept area as computed


from fractional flow theory in Chapter 4, fraction

Vp

= pore volume, barrels

Wi

= cumulative water injected, barrels

Rearranging Eq. 5.11 leads to:

EA

Wi

V p ( S wbt S wc )
5 - 21

(Eq. 5.12)

From Eq. 5.12, it is possible to compute

E A as a function of Wi

up to the time

of initial water breakthrough.


As additional water is injected, the water front on each streamline advances
toward the production well. Water flowing along the diagonal flow path (Streamline A) will be the first to breakthrough at the producing well. It is noted on
Figure 5-11 that at the time of water breakthrough along Streamline A, the water
front moving along Streamlines B, C, D, E, and F is still a significant distance
from the producing well.
Because of the slower fluid movement along these longer streamlines, a part of
the reservoir remains unswept at the time of initial water breakthrough on
Streamline A. A confined five-spot waterflood system with a unit mobility ratio
will have an areal sweep efficiency at water breakthrough of approximately 70
percent.
If water injection continues following initial water breakthrough, the water front
on Streamline B will breakthrough, and then breakthrough will occur on Streamline C. Finally, when water breakthrough occurs on all streamlines, the areal
sweep efficiency for the five-spot will be 100 percent.
Figure 5-12 depicts the waterflood front in a quadrant of a five-spot at several
times during the life of a flood. The areal sweep efficiency at any time during the
flood is defined as the ratio of the swept area to the total area.

5 - 22

FIGURE 5-12
FLOOD FRONT LOCATION AT SUCCESSIVE TIMES
IN A FIVE-SPOT PATTERN
1

E =
A
Area
Area
+ Area

Finally, it should be noted that any factor which alters the reservoir pressure
distribution such as an irregular pattern, non-unit mobility ratio, directional
permeability, or uneven pressure gradients between injector and surrounding
producers will thereby alter the position of the streamlines and affect areal sweep
efficiency.

For ideal patterns developed in homogeneous reservoirs with

symmetrical pressure gradients,

E Abt

has been computed by several

investigators and plotted versus mobility. In those instances where directional


permeability trends, irregular patterns, or unbalanced patterns exist,

E A can best

be estimated using numerical simulation procedures or with front-tracking


models.
B. Areal Sweep Efficiency at Breakthrough
The measurement of areal sweep efficiency has received considerable attention in
the literature. Excellent summaries of these studies for ideal homogeneous patterns are presented by Craig1 and Willhite2. Typically, areal sweep efficiency
5 - 23

data at initial breakthrough,

E Abt

are presented on a plot of

on a semi-log graph for a particular well pattern.

E Abt

versus

When using these data,

however, you must be aware of what type well pattern the data represent. Four
types of well patterns occur.
1. Isolated Pattern
This is a pattern which exists in a liquid filled reservoir which has no
boundaries and no other wells. It is possible with isolated patterns to have an
areal sweep efficiency at breakthrough greater than 100 percent.

This is

because fluids from the injection well can sweep oil from outside the pattern.
2. Developed Pattern
This is a pattern in a field where the total field is developed on the same
pattern.

Sweep efficiency data for developed patterns have the widest

application for waterflood predictions.


3. Normal Pattern
This is a pattern which contains one producing well.
4. Inverted Pattern
This is a pattern with one injection well.
Figure 5-131,2 presents an areal sweep efficiency correlation for a developed fivespot pattern.

5 - 24

FIGURE 5-13
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
FOR A DEVELOPED FIVE-SPOT PATTERN (REFERENCE 1)
Areal Efficiency At Breakthrough, percent

100
90
80
70
60
50
40
0.1

1
Mobility Ratio

10

E ABT 0.5460 0.0317 / M 0.3022 / e M 0.0051M (Ref. 2 - Page 115)

This graph represents a summary of data from five-spot patterns prepared by a


number of investigators. While there is some disagreement in the results of the
many studies, it is generally agreed that the line on Figure 5-13 is probably the
most representative for reservoir flooding operations. Figure 5-13 shows that for
an

M of

however,

less than about 0.15, the

E Abt

is 100 percent. If

increases,

E Abt decreases.

Data for an isolated five-spot pattern, both inverted and normal, are presented in
Figure 5-14. This figure shows that, as mentioned previously, isolated patterns
can have sweep efficiencies greater than 100 percent.

5 - 25

FIGURE 5-14
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
FOR AN ISOLATED FIVE-SPOT PATTERN (REFERENCE 1)
Areal Efficiency At Breakthrough, percent

240
Normal Pattern

200
160

Inverted Pattern

120
80

Inverted
Normal

40

92% at M

0.1

10

100

Mobility Ratio

Sweep efficiency data are presented1 in Figures 5-15 and 5-16 for the developed
normal and developed inverted seven-spot pattern.

5 - 26

Areal Efficiency At Breakthrough, percent

FIGURE 5-15
AREAL SW EEP EFFICIENCY AT BREAKTHROUGH
FOR A DEVELOPED NORMAL SEVEN-SPOT PATTERN (REFERENCE 1)

100
Pattern Area

90

80

70

60
0.1

Mobility Ratio

10

100

FIGURE 5-16
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
FOR A DEVELOPED INVERTED SEVEN-SPOT PATTERN (REFERENCE 1)

Areal Efficiency At Breakthrough, percent

100
Pattern Area

90

80

70

60
0.1

Mobility Ratio

5 - 27

10

100

The sweep efficiency of direct line drives and staggered line drives depends upon

d is the distance between adjacent rows of wells and a is


the distance between like wells. The relationship between E Abt and d / a was
the

d /a

ratio, where

shown previously in Figure 5-4 for a mobility ratio of unity. Figures 5-17 and 518 depict the relationship between

E Abt

and mobility ratio for the developed

direct line drive and staggered line drive, respectively, for

d / a = 1.0.

FIGURE 5-17
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
FOR A DEVELOPED DIRECT LINE DRIVE, d/a = 1.0 (REFERENCE 1)

Areal Efficiency At Breakthrough, percent

100
Pattern Area

90

a
d

80
70
60
50
40
0.1

Mobility Ratio

5 - 28

10

100

FIGURE 5-18
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
FOR A DEVELOPED STAGGERED LINE DRIVE, d/a = 1.0 (REFERENCE 1)
Areal Efficiency At Breakthrough, percent

100

Pattern Area

90
80

70
60
50
40
0.1

Mobility Ratio

10

100

C. Areal Sweep Efficiency After Breakthrough


With continued injection beyond breakthrough, the areal sweep efficiency of
a developed pattern will continue to increase until it reaches 100 percent.
The producing water-oil ratio also increases rapidly after breakthrough;
however, the water-oil ratio economic limit is likely to be reached before
complete areal coverage is attained. The increase in areal sweep beyond
breakthrough will be a function of how much water is injected into the
system. It is desirable in planning a waterflood to know the relationship
between these two variables. A good method for computing areal sweep
beyond breakthrough in a developed five-spot pattern is presented in Figure
5-19.

5 - 29

FIGURE 5-19
EFFECT OF INJECTED FLUID VOLUME ON AREAL SWEEP EFFICIENCY
AFTER BREAKTHROUGH FOR A DEVELOPED FIVE-SPOT PATTERN
(REFERENCE 1)

100
90

EA

80
70
60
50

Wi / Wibt

9 10

This correlation, developed by Craig et al3, requires knowledge of

E Abt ( E A at Wi / Wibt 1.0) and

the ratio of cumulative water

injected to cumulative water injected at breakthrough. Application of this


correlation will be presented in an example problem. The correlation for

EA

presented by Figure 5-19 can also be expressed by the following

equation.

W
E A 0.2749 ln i E Abt
Wibt

(Eq. 5.13)

_________________________________________________________________

EXAMPLE 5:1
A thin oil reservoir is to be waterflooded with a well arrangement that
approximates a developed five-spot. Assume vertical sweep efficiency at initial
5 - 30

water breakthrough to be 100 percent. Given the following relative permeability


values, reservoir data, and fractional flow graph in Figure 5-20, compute the
following.
1. The waterflood mobility ratio prior to water breakthrough.
2. The areal sweep efficiency of the injected water at initial water breakthrough.
3. The barrels of water required to reach initial water breakthrough.
4. The barrels of water required to achieve gas fillup.
5. The areal sweep efficiency of the injected water at gas fillup.
6. The areal sweep efficiency when the cumulative water injected is 1.5 times
the volume of water injected at breakthrough.
7. The barrels of water to achieve 100 percent areal sweep.
8. The areal sweep of the leading edge of the oil bank at gas fillup.

Sw , %

k rw

k ro

fw

22
27
32
37
42
47
52
57
62
67
72

0.000
0.004
0.014
0.032
0.058
0.090
0.130
0.176
0.230
0.291
0.360

1.00
0.81
0.64
0.49
0.36
0.25
0.16
0.09
0.04
0.01
0.00

0.00
0.03
0.13
0.31
0.52
0.71
0.84
0.93
0.97
0.99
1.00

5 - 31

FIGURE 5-20
FRACTIONAL FLOW GRAPH VERSUS WATER SATURATION
1.0

S wbt 54.2%

0.8

fW

0.6
0.4

f w 0.133

0.2

S wc 32%

0.0
10

20

30

40

50

60

70

80

Water Saturation, percent

Pattern area = 60 acres

= 4.0 feet

= 22 percent

S wir

= 22 percent

S wc

= 32 percent

So

= 56 percent (calculation from primary depletion data and Eq. 3.9)

Sg

= 12 percent

= 0.60 cp

= 4.0 cp
5 - 32

SOLUTION
1. The mobility ratio is defined by Eq. 5.4 as:

To compute

(krw ) S wbt
(kro ) Swc

o
w

it is necessary to calculate

in this reservoir,
percent.

S wc S wir .

k rw

at

S wbt .

Thus, the tangent line is initiated at 32

From the fractional flow graph in Figure 5-20,

percent. Also,

It is noted that

S wbt

= 54.2

krw S wbt = 0.149 and (k ro ) Swc = 0.640, or:

0.149 4.0

1.55
0.640 0.6

2. From Figure 5-13,

E Abt = 62 percent.

3. The volume of injected water to reach initial water breakthrough can be


computed from Eq. 5.11 as:

Wibt V p E Abt ( S wbt S wc )


or:

Wibt 7758(60)( 4)(0.22)(0.62)(0.54 0.32)


Wibt 55,872 barrels of water
4. The volume of water to reach fillup as computed from Eq. 4.52 is:

Wif V p S g
5 - 33

or:

Wif 7758(60)( 4)(0.22)(0.12)


Wif 49,155 barrels
5. Prior to water breakthrough, the areal sweep efficiency is calculated from
Eq. 5.12. Thus, at gas fillup,

E Af

Wif
V p ( S wbt S wc )

or:

E Af

49,155
7758(60) 4 (0.22)(0.54 0.32)

E Af 55 percent
6. After water breakthrough, areal sweep is governed by Eq. 5.13, that is:

W
E A 0.2749 ln i E Abt
Wibt
For

Wi
1.5 , Wi 83,808 barrels water
Wibt

E A 0.2749 ln(1.5) 0.62


or:

E A 0.73 or 73 percent
5 - 34

7. When

E A 1.0 , from Eq. 5.13:


W
1.0 0.2749 ln i 0.62
Wibt

or:

Wi
3.98
Wibt
and:

Wi ( E A 1.0) (3.98)(55,872) barrels


or:

Wi ( E A 1.0) 222,370 barrels of water


8.

MO /G

( kro ) S wc
(krg ) S g

(krg ) S g 0.1(estimated)

g 0.02(estimated)
thus:

MO /G

0.640 0.02

0.1
4.0
5 - 35

M O / G 0.032
From Figure 5-13, the areal sweep efficiency of the leading front of the oil
bank is 100 percent at gas fillup.
_________________________________________________________________
_
Figure 5-21 presents an alternate correlation of areal sweep efficiency versus
mobility ratio for the developed five-spot pattern which is applicable after breakthrough. Similar correlations are presented in Appendix D of SPE Monograph 31
for other well patterns.
FIGURE 5-21
EFFECT OF MOBILITY RATIO ON DISPLACEABLE VOLUMES
INJECTED FOR THE DEVELOPED FIVE-SPOT PATTERN (REFERENCE 1)

Areal Sweep Efficiency, percent

100

VD

90

2.5
2.0
1.5

80

1.2
1.0

70
0.75

60
50

kt
l B re a
I n itia

0.1

0.2

h ro u g

0.3

0.5

1.0

2.0

3.0

Recriprocal of Mobility Ratio (1.0/M)

5.0

10

A new factor used in the correlation presented in Figure 5-21 is the displaceable
pore volume, VD , where:

5 - 36

VD
VD

= displaceable pore volume

Wi

V p pattern So max V p pattern


Wi

Wi
1.0 S wir Sor

= cumulative water injected, bbls

(V p ) pattern =
( So ) max

pore volume in pattern, bbls

= maximum displaceable oil saturation

D. Effect of Free Gas Saturation on Areal Sweep


1. Water Zone
Most studies of areal sweep behavior have been conducted in reservoir
systems which were initially filled with liquid (oil and connate water).
However, most reservoirs which are waterflooded contain an initial gas
saturation.

What effect does a gas saturation have on the sweep

efficiency correlations previously considered?


When water is injected into a reservoir with an initial gas saturation,
studies2 show that the injected water will move out radially until either (1)
the leading edge of the oil bank contacts an oil bank formed about an
adjacent injector or (2) the oil bank encounters a producing well. When
either of these events occur, the water front will begin to cusp toward the
nearest producer. If at this time the water zone flood front would also
have been essentially radial in an initially liquid-saturated reservoir, the
areal sweep of the water zone at water breakthrough with initial gas
present would be the same as in a system with no gas. Accordingly, the
5 - 37

performance at and after breakthrough would be the same for these two
systems. However, total oil produced from the system with initial gas
would be less than in the liquid filled reservoir by an amount equal to the
initial volume occupied by the gas.
If gas fillup occurs at a higher sweep than that at which radial flow would
occur in an equivalent liquid-filled system, the areal sweep performance
of the system with initial gas present would be better than that predicted
by the sweep efficiency correlations.
Most waterfloods are conducted in reservoirs where the gas saturation is
such that gas fillup occurs before the flood front would cusp in an
equivalent liquid-filled system.

Accordingly, initial gas does not

materially affect areal sweep of the water zone in most waterfloods.


2. Oil Zone (Oil Bank)
An oil bank is usually created in most waterfloods when an initial free gas
saturation is present at the start of water injection. If

Sg

is too high, an

oil bank will not form, as discussed in Chapter 4. The concept of a water
zone and oil zone (oil bank) was discussed extensively in Chapter 4. Of
course, if there is no free gas, the entire reservoir behaves as an oil bank
at the start of injection because the reservoir contains only oil and
connate water. In conventional waterflood theory, the concept of the
creation of an oil bank only makes practical sense when there is a free gas
saturation. This is the focus of this section.
As water injection volume increases, the leading edge of the oil bank
precedes the water zone and sweeps through the pattern or reservoir area.
Consider two five-spot patterns within a developed five-spot flood as
shown in Figure 5-22.

5 - 38

FIGURE 5-22
FLOOD FRONT EARLY IN THE LIFE
OF A DEVELOPED FIVE-SPOT WATERFLOOD
WITH AN INITIAL FREE GAS SATURATION
UNAFFECTED AREA

UNAFFECTED AREA

OIL
BANK

OIL

WATER
ZONE
re

WATER
r

Figure 5-22 shows the flood fronts while they are radial about the
injection well. As injection volume increases, the radii of the water and
oil zones increase. It was shown in Chapter 4 that:
1
2

re S wbt S wc

r
Sg

In time, the oil bank from the two wells meet. This is defined as oil bank
interference and is depicted in Figure 5-23.

5 - 39

FIGURE 5-23
FLOOD FRONT AT OIL BANK INTERFERENCE
UNAFFECTED PATTERN AREA

OIL ZONE

WATER
ZONE
rei

rei

At interference, the radius of the oil bank, rei , is one-half the distance
between two adjacent injectors. Continued injection leads to a rapid
change in the shape of the oil bank. At a later time, the shape of the
flood fronts are shown in Figure 5-24.

5 - 40

FIGURE 5-24
FLOOD FRONT AFTER INTERFERENCEBUT BEFORE GAS FILLUP
UNAFFECTED PATTERN AREA
OIL BANK

OIL BANK

WATER
ZONE

WATER
ZONE

In Figure 5-24, it is noted that the leading edge of the oil bank is now
concentric or radial about the production wells.

At gas fillup, the

leading edge of the oil bank reaches the producing well as depicted in
Figure 5-25.

FIGURE 5-25
FLOOD FRONT AT GAS FILLUP

OIL BANK

WATER
ZONE

WATER
ZONE

5 - 41

It is significant to note that at fillup, the areal sweep of the leading edge
of the oil bank is 100 percent. Recall from Chapter 4 the mobility ratio
between the oil bank and unaffected gas zone is usually very low (less
than about 0.2 to 0.4). Moreover, from Figure 5-13, sweep efficiency
theory suggests that the areal sweep of the oil bank at fillup should be
approximately 100 percent. At gas fillup, the areal sweep of the injected
water can be computed as follows.

Wif V p S g

(Eq. 4.52)

and from Eq. 5.12,

E Af

Wif

(Eq. 5.12)

V p ( S wbt S wc )

Combining Eq. 4.52 and Eq. 5.12 results in:

E Af

Sg
S wbt S wc

where:

E Af

= areal sweep efficiency of the injected water at fillup

3. Re-Saturation Effects
After gas fillup, the reservoir can be separated into two regions. These
two areas are the water swept portion and the portion not yet swept by
water. The oil saturation in the water swept area is

1.0 S wc

in the unswept area.

5 - 42

1.0 S w

In the swept area,

and

S w S wbt

before breakthrough. After water breakthrough,

S w S wbt

in the

water swept area.


During the fillup period, oil will be displaced into the oil bank, but
production of the displaced oil will not begin until after the fillup
period. During the fillup period, the injected water displaces oil and resaturates the gas pore space with liquid.
E. Other Factors Affecting Areal Sweep Efficiency
As indicated previously, most areal sweep correlations were developed for an
ideal reservoir with no free gas.

When using this information or when

predicting the sweep efficiency by other available methods, one must be


aware of the many factors related to a waterflood system that can cause
significant variation in predicted results.
1. Fractures
If a flood pattern is established so that the direction between injectors and
producers corresponds to the fracture orientation, the results will probably
be disastrous. This arrangement results in early water breakthrough and a
low sweep efficiency. The problem can be rectified by arranging the
injectors and producers so the direction of a line connecting them is
perpendicular to the fracture orientation. It is possible that recovery with
this arrangement will exceed that of a homogeneous system due to the
fracture acting as a plane-source of water.
2. Directional Permeability
When the permeability is much greater in one direction than in other
directions, fluid will obviously attempt to flow in the direction of
maximum permeability. The effect of this directional permeability is the
same as the effect of a fracture, although probably not as drastic.
5 - 43

Accordingly, the injectors and producers should be arranged along a line


perpendicular to the direction of greatest permeability. Craig1 compares
the sweep efficiency of a five-spot system with directional permeability
when operating under the most favorable arrangement and under the least
favorable arrangement.
3. Areal Permeability Variations
Many different types and variations of areal permeability changes can
occur across a reservoir. These may occur due to changes in compaction,
cementation, depositional environment, etc.

This type of areal

heterogeneity must be handled on an individual basis with the effect on


sweep being determined from numerical simulation methods.

The

detection of areal heterogeneities will be discussed in a subsequent


section.
4. Formation Dip
It was shown in Chapter 4 that water should be injected so as to displace
oil in an updip direction in a non-horizontal reservoir to maximize oil
recovery. However, if formation dip is generally less than about ten
degrees, dip angle will probably have only minor or negligible influence.
If dip is more severe or is important, its effect on areal sweep can be
considered through the use of numerical simulation.
5. Off-Pattern Wells
When a well is irregularly spaced, the effect is to cause early water breakthrough in the wells nearest the injector and late breakthrough in other
wells.

The overall effect is to reduce areal sweep at initial water

breakthrough. Moreover, the water-oil ratio economic limit is likely to be


reached at an earlier time and, thus, effectively reduce areal (or vertical)
sweep. Nevertheless, reasonable estimates of waterflood behavior can be
5 - 44

approximated by treating the irregular patterns as regular patterns. If the


wells create highly irregular patterns, it may be necessary to employ
numerical simulation methods to estimate areal sweep.
6. Sweep Beyond Edge Wells
A significant portion of a reservoir generally lies between the edge wells
and the reservoir boundary. If the boundary is within the well spacing of
the edge wells, it has been found by Caudle5 that essentially all the oil in
this part of the reservoir will eventually be contacted by water. However,
in field application, this will depend upon the geology of the reservoir and
the amount of water throughput beyond breakthrough which is
economically feasible. Again, numerical simulation methods must be
used.
7. Isolated Patterns
When a field is developed using a balanced repetitive flood pattern, each
pattern tends to behave independently of the others since the oil and water
in that portion of the reservoir are confined by the influence of adjacent
patterns.

If a flood pattern is established in an otherwise infinite

reservoir, fluids are not confined to the pattern area, and it is possible to
compute a sweep efficiency greater than 100 percent based on produced
fluids. This situation is depicted by Figure 5-26. An understanding of the
sweep behavior of isolated patterns is especially important when
interpreting the results of pilot waterfloods.

5 - 45

FIGURE 5-26
SWEEP BEHAVIOR
IN AN INVERTED, ISOLATED FIVE-SPOT PATTERN

8. Irregularly Spaced Wells


Unfortunately, many fields are drilled using random well locations. In
these instances, the areal sweep studies may be of little help and each case
must be handled individually. If the project is large enough to justify
mathematical model studies of the field, these are recommended.
Otherwise, the concept of using stream tubes, as introduced by Higgins, et
al6, might be used.

This technique has been utilized in several

applications7,8,9,10 to predict the behavior of enhanced recovery projects.


Beyond this, one must rely on experience, common sense, and luck.

5 - 46

V. Peripheral and Line Floods


In contrast to the use of repetitive patterns, a peripheral flood utilizes the edge wells
along all or a part of the reservoir boundary as injection wells. If a single line of
wells along one side or down the middle of the field is used, we often refer to it as a
line flood. This type of flood generally requires fewer injection wells per producer
than most pattern floods, thereby requiring a smaller initial investment. Also, this
type of flood generally results in less produced water than a pattern flood. This is
particularly true when operators shut in the production wells which experience water
breakthrough and continue to produce only those wells ahead of the water front. It
was shown by Ferrell, et al11 in a study of end-to-end floods that less injected water is
required to recover the oil and good areal sweep is still obtained if producing wells
are shut in soon after water breakthrough. However, if this procedure is used, it
should be obvious that the reservoir permeability must be high enough for water to
move at the desired rate over long distances from the injection well under the
imposed injection pressure. If this is not possible, the production wells can be
converted to injectors after breakthrough. This can involve long injection lines and
considerable expense, however, and is generally not desirable.
A further advantage of a peripheral type flood is its flexibility. Maximum advantage
of dipping reservoirs and reservoirs with permeability variations can be utilized.
Also, line or peripheral patterns are generally well suited to conversion to a more
dense injection pattern if performance dictates such a change.
A major disadvantage of peripheral floods occurs when a reservoir has a high gas
saturation. No significant recovery response will occur in a reservoir until the gas
space is effectively filled with water. Consequently, there may be a long time delay
and considerable water injection expense before this type reservoir responds to water
injection. This can be critical to an operator who needs a quick return on his
investment.

5 - 47

VI. Selection of Waterflood Pattern


The choice of waterflood pattern is one of the most important decisions an engineer
must make when planning a flood. This decision must be consistent with the existing
well pattern, the geology of the reservoir, and the injection and production objectives
of the flood. The economics of most floods will dictate that the flooding pattern be
consistent with existing wells or a minimum of infill drilling be required. This will
automatically eliminate some patterns from consideration.

However, most

development patterns offer several possibilities of injection-production well


arrangement. A square development pattern, for example, permits the use of fivespot, skewed four-spot, normal or inverted nine-spot, or line drive or peripheral
drive.

The decision in this situation would be dictated primarily by reservoir

characteristics.
The relative injection-production capacity of a reservoir will often dictate the pattern.
Suppose for example that we have a square development pattern and are considering
a five-spot, skewed four-spot, or normal or inverted nine-spot. All of these patterns
offer different ratios of producing to injection wells. In particular, the ratio is 1:1 for
a five-spot, 2:1 for a skewed four-spot, 3:1 for an inverted nine-spot, and 1:3 for a
normal nine-spot. If, for example, high injection capacity was needed to increase
reservoir pressure, the normal nine-spot would be a likely choice. This decision,
however, would have to be compatible with reservoir geology.
If a field contains significant heterogeneities such as fractures or permeability trends,
this will generally be the overriding factor in pattern selection. It is essential in such
situations to prevent adjacent injectors and producers from lying along a line parallel
to the direction of maximum permeability or fracture orientation. This will cause
early water breakthrough and result in very low areal sweep. The optimum pattern
in this situation will be one where the line connecting adjacent injectors is parallel
to the direction of the permeability or fracture trend.

5 - 48

In summary, a good waterflood pattern should meet the following criteria1.


1. Provide sufficient water injection capacity to yield desired oil production rate.
2. Maximize oil recovery with minimum water production.
3. Take advantage of reservoir non-uniformities such as fractures, permeability
trends, dip, etc.
4. Be compatible with existing well pattern and require a minimum of new wells.
5. Be compatible with flooding operations of other operators on adjacent leases.
6. Allow gas fillup time to occur within a reasonable time.
VII. Summary
A successful waterflood operation requires that areal sweep efficiency be reasonable
high. This is best achieved if the mobility ratio is sufficiently low and if a flood
pattern is chosen that takes advantage of reservoir heterogeneities such as fractures
and directional permeability and allows for sufficient injection and production
capacity. Many sweep studies have been made that aid in the prediction of sweep
efficiency for basic flood patterns in horizontal, homogeneous, liquid-filled
reservoirs undergoing steady-state flow.

Sweep predictions for reservoirs with

irregularly spaced wells, dipping reservoirs, or reservoirs with heterogeneities shall


be studied using numerical simulation or stream tube models.

5 - 49

CHAPTER 5 REFERENCES
1. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Richardson, TX (1971) 3.
2. Willhite, F.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
3. Craig, F.F., Jr., Geffen, T.M., and Morse, R.A.: "Oil Recovery Performance of
Pattern Gas or Water Injection Operations from Model Tests," Trans., AIME
(1955) 204, pp. 7-15.
4. Craft, B.C. and Hawkins, M.F.: Applied Petroleum Reservoir Engineering,
Prentice-Hall, Inc., Englewood Cliffs, N.J. (1959).
5. Caudle, B.H., Erickson, R.A., and Slobod, R.L.: "The Encroachment of Injected
Fluids Beyond the Normal Well Pattern," Trans., AIME (1955) 204, pp. 79-85.
6. Higgins, R.V. and Leighton, A.J.: "A Computer Method to Calculate Two-Phase
Flow in Any Irregularly Bounded Porous Medium," Trans., AIME (1962) 225, pp.
679-683.
7. Baldwin, D.E., Jr.: "Prediction of Tracer Performance in a Five-Spot Pattern,"
Trans., AIME (1966) 237, pp. 513-517.
8. Davies, L.G., Silberg, I.H., and Caudle, B.H.: "A Method of Predicting Oil
Recovery in a Five-Spot Pattern," Trans., AIME (1968) 243, pp. 1050-1058.
9. LeBlanc, J.L. and Caudle, B.H.: "A Streamline Model for Secondary Recovery,"
SPEJ (March 1971) pp. 7-12.
10. Doyle, R.E. and Wurl, T.M.: "Stream Channel Concept Applied to Waterflood
Performance Calculations for Multiwell, Multizone, Three-Component Cases,"
JPT (March 1971) pp. 373-380.
11. Ferrell, H., Irby, T.L., Pruitt, G.T., and Crawford, P.B.: "Model Studies for
Injection-Production Well Conversion During a Line Drive Water Flood," Trans.,
AIME (1960) 219, pp. 94-98.

5 - 50

PROBLEM 5:1

Presented below are the data for an oil reservoir being considered for a waterflood.

S w ,%

krw

kro

fw

28.0
32.2
36.4
40.6
46.9
51.1
55.3
61.6
65.8
70.0

0.000
0.003
0.012
0.027
0.061
0.091
0.127
0.192
0.271
0.300

1.000
0.810
0.640
0.490
0.303
0.202
0.123
0.040
0.003
0.000

0.000
0.011
0.053
0.142
0.376
0.573
0.669
0.935
0.986
1.000

= 1.50 cp

Bw

= 1.0

= 0.50 cp

= 20 percent

S wc

= 28.0 percent

= 15 feet

Bo

= 1.35 RB/STB

Well spacing = 40 acres

The fractional flow curve for this reservoir is presented in Figure 5P.1.
1. Compute the mobility ratio prior to breakthrough for a waterflood in this reservoir.

5 - 51

FIGURE 5P.1
DETERMINATION OF

S wbt
.

FRACTIONAL FLOW CURVE


1
0.9
0.8
0.7
0.6

fw 0.5
0.4
0.3
0.2
0.1
0
0.1

0.2

0.3
0.4 0.5
0.6 0.7
Water Saturation, fraction

5 - 52

0.8

PROBLEM 5:2

Consider a partly depleted single layer of a 160 acre five-spot pattern that is to be
waterflooded. The layer is characterized by the following data.

160 acres

= 2.0

5 feet

= 400 psi

18%

= 180 F

S wc

24%

= 0.95

Sg

15%

API =

So

61%

Bo

S wf

50%

Sorw =

S wbt =

58%

28

= 1.1 RB/STBO @ 400 psi


30%

= 5 cp

1. Compute the volume of injected water to reach gas fillup, Wif .


2. Compute the areal sweep efficiency

E A of the injected water at fillup.

3. Compute the areal sweep efficiency

E Abt

at water breakthrough.

4. Compute the volume of injected water necessary to reach water breakthrough, Wibt .
5. What is areal sweep of the injected water when the cumulative water injection is
twice the volume required to reach breakthrough?
5 - 55

6. How many barrels of water are required to reach 100 percent areal sweep?
7. If the oil production during the fillup period is negligible, how many STBO will have
been displaced at fillup?
8.

How many STBO will have been produced at water breakthrough?

9.

What is the maximum theoretical recoverable oil?

5 - 56

INJECTION RATES AND PRESSURES


Advance information on the relationship between injection rate and injection pressure is
useful, and often critical, in the design and analysis of any enhanced oil recovery project.
In particular, an estimate of injection rates and pressures is needed during the planning
stage of a waterflood for the purpose of sizing injection equipment and pumps and for the
purpose of predicting oil recovery rates. Further, it is possible in low permeability
reservoirs that the injection rate required for a project to be economically feasible will
necessitate injection pressures which exceed the fracture pressure of the subject
formation and which, if imposed on the formation, could result in poor reservoir sweep
efficiency (areal and vertical) and substantially decreased oil recovery. In those cases
where a high degree of uncertainty exists, it might become necessary to conduct a pilot
flood to determine injection rates and pressures required for economic operations. Such
pilot injection tests should be carefully designed and analyzed due to the fact that short
term injection tests lasting only a few days may lead to substantial and misleading results.
The purpose of this chapter is to present methods which can be used to predict injection
rates and pressures in terms of information commonly available for a waterflood project.
When the mobility ratio of a flood is unity, this can be accomplished using simple
analytical relationships which require only a knowledge of the waterflood pattern and
properties of the reservoir. Calculations for mobility ratios different from unity are more
difficult and require the use of approximate analytical techniques or experimental
correlations. In general, prediction of rates and pressures are more difficult after water
breakthrough than before.
I. Factors Affecting Water Injection Rate
During a flooding operation, the injection rate at which water can be injected depends
upon the following factors1,2,3.

6-1

A. Physical properties of the reservoir rock and fluids, such as:

ko , k w , kro , krw , o , w , h, si , s p , S g
B. Area swept by the injected water and oil bank.
C. Fluid mobilities in the water zone and oil bank
D. Well geometry, pattern, spacing, and wellbore radii.
E. Bottom-hole injection pressure, producing well pressure, and average reservoir
pressure at the start of injection.
Some of these factors cannot be changed. Others, however, such as the flood pattern,
injection well pressure, producing well pressure, and skin factors can be selected or
altered to best achieve the desired injection-production performance. The effect of
these factors on water injection rate will be considered in the remainder of this
chapter.
II. Radial System, Unequal Mobilities
Since fluid mobilities are equal throughout the reservoir in unit mobility waterfloods,
the position of the flood front has no effect upon water injectivity after gas fillup.
When mobility ratio is different from unity, however, resistance to fluid injection
varies depending upon the relative amounts of oil and water in the reservoir. When
the mobility ratio,

M , is less than unity, oil flows better than water; when M

is

greater than unity, water flows better than oil. It follows that total fluid mobility in
the reservoir will change as increasing amount of water are injected, thereby causing
the injection rate to change.

These functional relationships between injectivity,

mobility ratio, and flood front position can be shown explicitly by analyzing a simple
geometric pattern.
Early in the life of an injection well and prior to gas fillup, both the water zone and oil
zone are radial about the injection well. The zones will continue to be circular about
the injection well until the radius of the oil bank reaches a distance of about 70

6-2

percent of the distance between the injector and producer. Consider the radial system
depicted by Figure 6-1 which has a central injection well of radius

rw .

FIGURE 6-1
IDEAL FLOW SYSTEM
WITH RADIAL OIL AND WATER BANKS
OIL

ko / o

re

WATER
kw / w

pw , rw

pe

Applying Darcys steady state radial flow equation for incompressible fluids, it can be
shown2 that the injection rate at any mobility ratio,

M , and any injection well skin

factor is equal to:

iw

0.00707 khp
w
r
r
ln o ln e
krw rw kro r

where:

iw

= water injection rate, bbls/day

= net pay, feet

6-3

(Eq. 6.1)

= base permeability used to define relative permeability, md


usually the effective permeability to oil at irreducible water,

ko Swir , md

S wc

k ro

= relative permeability to oil in oil bank at

krw

= relative permeability to water in water bank at

re

= radius of oil bank, feet (see Eq. 4.37)

= radius of water bank, feet (see Eq. 4.42)

rw

= effective wellbore radius, feet =

rw

= wellbore radius, feet

si

= Skin factor at injection well, dimensionless

= applied pressure differential, psi


(difference between pressure at formation face of injection well,
pw , and pressure in reservoir at the outer edge of oil bank, pe ,
usually assumed as the average reservoir pressure at start of
injection)

= oil viscosity, cp

= water viscosity, cp

S wbt

rwe si

_____________________________________________________________________
EXAMPLE 6:1
1. A new injection well is to be placed in service in an oil reservoir where the
reservoir pressure has declined below the bubble-point pressure. Current reservoir
pressure is 800 psi. Bottom-hole injection pressure is expected to be 2600 psi.
Compute the water injection rate early in the life of the well when the radius to the

6-4

water and oil banks are 200 and 388 feet, respectively. Assume the injection well
skin value is zero. Other data are given below.

( ko ) S wir

= 10 md

= 8 feet

(kro ) Swc

= 1.0

= 0.45

( krw ) S wbt

= 0.30

Sg

= 8%

= 0.9 cp

S wbt

= 56%

= 0.6 cp

S wc

= 26%

rw

= 0.33 feet

Early in the life of an injection well during which the flood fronts are circular
about the injector, water injection can be computed using Eq.6.1 where:

iw

where

0.00707 khp
w

r
r
ln
o ln e
krw r 'w kro r

rw = rw for s = 0.
iw

0.00707(10)(8)(2600 800)
76 BWPD
0.6
200 0.9 388
ln
ln

0.30 0.33 1.0 200

2. If the injection well is effectively stimulated such that a negative skin of 4 exists,
compute the water injection rate for the conditions described above. First, the
effective injection well radius is computed.

rw rwe si

6-5

rw 0.33e ( 4)

rw 18.0 feet
and:

iw

0.00707(10)(8)(2600 800)
188 BWPD
0.60 200 0.9 388
ln
ln

0.30 18 1.0 200

_____________________________________________________________________
III. Regular Patterns
A. Unit Mobility Ratio
When fluid mobilities in the water zone and oil zone portions of the reservoir
are equal, i.e.,

M=

1, fluid injectivity does not change as the flood front

advances after gas fillup. Further, injectivity of a particular well pattern is


independent of the size of the area swept by water but is directly proportional
to the fluid mobility involved. The determination of injectivity under these
conditions reduces to a geometric problem which results in simple analytical
relationships.
Deppe1 and Muskat2 have developed simple mathematical formulas which
relate injection rate and injection pressure for a number of regular well
patterns. In addition to assuming a unit mobility ratio, these equations assume
steady state flow and are limited to reservoirs where no gas is present or to
reservoirs following gas fillup. These equations are summarized in Table 6-1.

6-6

Table 6-1: Injectivity Equations for Regular Patterns with Unit Mobility Ratios

0.00354 ko S

Direct Line Drive2

iw

1
a

o ln

wir

hp

a
d
1.570 1.837 0.5 si s p
rw
a

d
a

0.003541 ko S

Staggered
2

Line Drive

1
a

iw

o ln

wir

hp

a
d
1.570 1.837 0.5 si s p
rw
a

6-7

d
a

0.003541 ko S

Five-Spot2

iw

o ln

hp

d
0.619 0.5 si s p
rw

6-8

wir

Seven-Spot2

iw

0.004723 ko S

wir

hp

d
0.569 0.5 si s p
rw

o ln

Nine1

Spot

iw

iw

0.003541 ko S

wir

h p i ,c

1 R d
s
s
ln
0.272
0.5

i
p o

2 R rw

0.00708 ko S

wir

h p i ,s

3 R d
0.693

ln
0.272
0.5
s
s

o
i
p

2 R
2 R rw

R=
p i,c

Ratio of producing rates of corner well(c) to side well (s)

p i,s

Difference in pressure between injection well and side well (s).

Difference in pressure between injection well and corner well (c).

6-9

These equations have also been summarized by Willhite3.


B. Non-Unit Mobility Ratio
The equations listed in Table 6-1 are valid after fillup when the mobility ratio
is unity. The permeability term is the effective permeability to oil measured at
the irreducible water saturation,

ko Swir .

If, as in most waterfloods, the

mobility ratio is different from unity, the calculated injection rate obtained
from these equations must be adjusted using a correction factor defined as the
conductance ratio. The actual injection rate is computed as:

iw ibase

(Eq. 6.2)

where:

= conductance ratio

ibase =

base water injection rate, bbls/day (steady-state water


injection rate in an oil-filled pattern with immobile connate
water for a unit mobility ratio)

and

ibase

is defined by the equations listed in Table 6-1 for the various

patterns.

6 - 10

The conductance ratio,

, is an experimentally determined factor based on the

work of Caudle and Witte4 which, when used in Eq. 6.2, gives the correct
injection rate. The conductance ratio is presented in Figure 6-2 as a function of
mobility ratio,

M,

and areal sweep efficiency of the injected water,

Note in Figure 6-2 that for


For

M >1.0,

M < 1.0,

and

and

iw

iw

M = 1.0, = 1.0,

and

iw

EA.

are constant.

increase with increasing sweep efficiency. When

decrease with increasing sweep efficiency.

6 - 11

6-12

0.1

10

1.0

0.9

EA

0.1

0.1

0.7

0.5

0.9

0.3

10

Moblity Ratio
CR 1/(1+E A ((1/ M ) 1))(Re f. Dr. Ben Caudle - Private Communication)

0.1

0.5
0.7

0.3

1.0

EA

FIGURE 6-2
CONDUCTANCE RATIO FOR LIQUID FILLED FIVE-SPOT PATTERNS
(REFERENCE 4)

Conductance Ratio

The areal sweep of the injected water required by Figure 6-2 can be computed
as:
Before water breakthrough:

EA

Wi

V p S wbt S wc

(Eq.6.3)

After water breakthrough3,5,6:

E A 0.2749ln

Wi
E Abt
Wibt

(Eq. 6.4)

The conductance ratio and Equation 6.4 have been established for a developed
five-spot pattern. Nevertheless, both can be combined with the equations in
Table 6-1 to compute injection rates for other patterns with a high degree of
accuracy.
__________________________________________________________________
Example 6:2
1. For the injection well described in Part 1 of Example 6:1, compute the
water injection at gas fillup at which time

E A = 0.27.

Assume the well

is part of an 80-acre five-spot pattern in which the diagonal distance,


between the injector and producer is 1,320 feet.

d,

The producing well

pressure is set at 500 psi. After gas fillup, water injection is computed
using Eq. 6.2 where:

iw ibase
For a five-spot pattern,

ibase is obtained from Table 6-1 to be:

6 - 13

ibase

0.003541( ko ) S wir hp

o ln

d
0.619 0.5( si s p )
rw

which is the injection rate after gas fillup for

ibase

M 1.

0.003541(10)(8)(2600 500)
1320

0.619 0.5(0 0)
0.9 ln
0.33

ibase = 86 BWPD
Next, compute the conductance ratio,

M 0.45 .

At fillup,

determined to be

0.80 .

E A 0.27

to correct for the actual

and from Figure 6-2,

is

The actual water injection rate at fillup is:

iw = (0.80)(86) = 69 BWPD
2. Compute the water injection rate at fillup for the well conditions described
above except that both the injection and production wells are effectively
stimulated and possess negative skin values of 4.

ibase

0.00354110 8 2600 500


180 BWPD
1320

0.9 ln
0.619 0.5 4 4
0.33

3. and:

iw 0.80 180 144 BWPD


__________________________________________________________________

6 - 14

To summarize, prior to fillup the injectivity

iw / p will rapidly decrease up

to fillup. After fillup, the injectivity will increase if

< 1.

M > 1 or decrease if M

This behavior is shown in Figure 6-3. Also, as indicated in Figure 6-3, the

most dramatic injectivity changes occur during the early part of the flood, whereas
changes become less pronounced during latter stages of the flood.

From a

practical viewpoint, it is noted that short term injectivity tests conducted in


depleted fields can result in overly optimistic injection rates which cannot be
sustained during the major portion of the life of the flood.

Water Injectivity Index (I.I.), BW/D/psi

FIGURE 6-3
WATER INJECTION RATE VARIATION

M > 1.0
M = 1.0
M < 1.0

Fillup of Gas Space

Cumulatiave Injected Water Volume (or Time)

6 - 15

C. Regular Patterns, Unequal Mobilities


Studies by Muskat2 of steady-state pressure distributions in various well
patterns with unit mobility ratio show most of the pressure change between
injection and producing wells occurs in areas near the wells where flow is
essentially radial. Even for the complex nine-spot pattern, radial flow occurs
in the vicinity of injection and producing wells1. Even when mobility ratios
differ from unity, experimental studies7 indicate that near-well flow patterns
are radial.
Recognizing that radial flow occurs near injection and producing wells and, as
indicated in the previous section, the largest changes in injectivity occur in
these radial flow regions, it was concluded by Deppe1 that the injection rates in
any pattern can be approximated by dividing the pattern into regions where
radial and linear flow predominate. As a result, Deppe showed that simple
equations could be developed to compute injection rate for a variety of
geometrical configurations including both regular and irregular patterns.
IV.

Injectivity in Five Spot Patterns


The five-spot pattern is the most commonly used flooding pattern for reasons
discussed in previous chapters. It follows that this pattern has also been subject to
more extensive theoretical and experimental injectivity studies than other patterns.
A. Prats, et al Method
Prats, et al8 developed an analytical method whereby injection rates can be
calculated for an enclosed five-spot well pattern where oil, gas, and water
saturations are present. Most reservoirs which have undergone significant
pressure depletion during primary recovery will have a free gas phase at the
time secondary recovery is initiated. This is one of the few methods which has
attempted to quantify the effect of an initial gas saturation.

Figure 6-4

illustrates an idealized picture of fluid regions which will exist between the

6 - 16

producing and injection well3. The flood is divided into three displacement
periods.
1. Start of flood to oil bank interference.
2. Oil bank interference to oil breakthrough (gas fillup).
3. After oil breakthrough this also includes after water breakthrough.
4. The positions of oil and water banks at the beginning and end of each of
these periods are shown in Figure 6-4.
B. Craig Method
Craig3,6 developed an excellent method for predicting injection performance
which can be applied to stratified systems with or without free gas present.
This method, which uses the correlations of Caudle and Witte4 to predict
injection rate as a function of mobility ratio and areal sweep efficiency,
considers water injection in four states which are similar to the periods
presented in Figure 6-4.
They are:
1. Stage 1: Start of the flood to interference.
2. Stage 2: Interference to gas fillup.
3. Stage 3: Fillup to water breakthrough.
4. Stage 4: After breakthrough.
A detailed description of this method will be presented in a later chapter as part
of the Craig-Geffen-Morse method5 of waterflood prediction.

6 - 17

FIGURE 6-4
Stage 1

Interference
Between Oil Banks

Stage 2

Stage 3

Water Bank

6 - 18

Gas Region

Oil Bank

REFERENCES
1. Deppe, J.C.: Injection RatesThe Effect of Mobility Ratio, Area Swept, and
Pattern, Trans, AIME (1961) 222, pp. 81-91.
2. Muskat, M.: Physical Principles of Oil Production, McGraw-Hill Book
Company, Inc., N.Y. (1949) 650.
3. Willhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
4. Caudle, B.H. and Witte, M.D.: Production Potential Changes During SweepOut in a Five-Spot System, Trans., AIME (1959) 216, pp. 446-448.
5. Craig, F.F., Jr., Geffen, T.M. and Morse, R.A., Oil Recovery Performance of
Pattern Gas or Water Injection Operations from Model Tests, Trans, AIME
(1955) 204, pp. 7-15.
6. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding,
Monograph Series, SPE, Dallas (1971) 3.
7. Dyes, A.B., Caudle, B.H. and Erickson, R.A.: Oil Production After
Breakthrough as Influenced by Mobility Ratio, Trans, AIME (1954) 201, pp.
81-86.
8. Prats, M., Matthews, C.S., Jewett, R.L., and Baker, J.D.: Prediction of
Injection Rate and Production History for Multifluid Five-Spot Floods, Trans,
AIME (1959) 216, pp. 98-105.

6 - 19

PROBLEM 6:1
A new waterflood is planned for a 6,000 foot reservoir which has been partially depleted.
Original reservoir pressure was 2700 psi and current reservoir pressure is 1000 psi. The
flood is to be implemented on 160-acre five-spot patterns. The distance between an
injector and producer is 1,867 feet. It is estimated that the reservoir fracture gradient is
0.62 psi/ft and that bottom hole injection pressure will be 3720 psi. Other data are listed
below.
Other Data
Mobility Ratio

= 3.0

( ko ) S

= 30 md

wir

(kro ) S
(krw )

wc

S wbt

= 1.0
= 0.25

= 22 ft

rw

= 0.25 ft

= 6.0 cp

= 0.5 cp

Sg

= 14%

S wc S wir

= 24%

S wbt

= 56%

Pwi 6,000' x 0.62 psi / ft 3,720 psi

6 - 20

1. Determine the instantaneous water injection rate early in the life of the waterflood
when the radius of the water and oil banks are 20 feet and 30 feet respectively.
Next, compute the injection rate at a later time when the radius of the water and oil
banks are 400 feet and 600 feet respectively from the injection well. The injection
well skin is zero.
2. If a skin is allowed to develop at the injection well and reaches a value of +8, what
is the maximum injection rate that can be obtained when the water and oil banks
are 400 feet and 600 feet respectively from the injection well?
3. At the time of gas fillup, the areal sweep efficiency of the injected water is 0.44.
If the producing well pressure is maintained at 500 psi, compute the water
injection rate at this time for the case of a zero skin at both the injection and
production wells.
4. Compute the water injection rate at water breakthrough if the producing well
pressure is maintained at 500 psi and the skin factor at both the injector and
producer is maintained at zero. For

M 3.0 , the E Abt

is about 0.56 for a five-spot pattern.

6 - 21

of the injected water

RESERVOIR HETEROGENEITY
Throughout our previous discussions, the reservoir has basically been considered as a
single-layered homogeneous porous system. Using this ideal reservoir, we have been
able to predict the efficiency with which water displaces oil from the water-contacted
portion of the reservoir. We can also predict the fraction of the reservoir area that will be
contacted by the injected water as a function of reservoir geometry and reservoir fluid
properties. Further, procedures have been developed for computing water injection rate
from the start of water injection (with or without free gas) to the economic limit of the
layer. These observations must be tempered, however, by the fact that no reservoir can
be considered homogeneous or acting as a single layer on a macroscopic scale.
Reservoir heterogeneity probably has more influence than any other factor on the
performance of a fluid injection project. At the same time, it is the most difficult effect to
quantify. Our purpose in this chapter is to discuss how areal and vertical permeability
variations can be determined and how these variations can be quantified for inclusion into
waterflood prediction and performance calculations.
I. Areal Permeability Variations
Areal changes in permeability affect pressure distribution and fluid velocity along the
streamlines and thus areal sweep. In a similar manner, variation in permeability
within the different layers affects vertical sweep. Interestingly, areal variations in
permeability tend to affect the outcome of a flood less than vertical changes. This
perhaps is not surprising because we expect a formation, especially sandstones, to
exhibit relatively high lateral continuity; the material deposited during a given
geologic period should be of the same physical nature over a relatively large surface
area. This is fortunate because, due to the large spacing between wells, we have few
test points with which to define the areal characteristics of a reservoir.
This is not meant to imply areal permeability variations are not important. To the
contrary, changes in environment or the process of deposition, compaction, tectonic
7-1

processes (which can cause fractures), or cementation can cause large areal variations
in the permeability of a reservoir which should be accounted for in the selection of
flood patterns and in the prediction of performance. The most severe problems
involve fractures and directional permeability as previously discussed.
Some carbonate rocks are particularly difficult to describe because much of the
permeability development occurs after deposition due to solution, dolomitization, recrystallization, etc. Lateral continuity of net pay as well as physical properties cannot
be assumed in this environment.
A. Detection of Areal Permeability Variations
Methods which are commonly employed to detect and quantify areal variations in
permeability are:
1. Mapping of core data, log data, and well test data
2. Detailed lithological studies
3. Pressure transient tests (including pulse tests and interference tests to detect
and quantify directional permeability trends)
4. Environment of deposition - recognition of depositional environment (channel
sediment, delta sediment, beach sediment, etc.) allows us to infer probable
directional changes in grain size, grain orientation, permeability, etc.
5. Injection and production well behavior
6. Performance history matching using mathematical simulators
7. Fracture detection - areal photo interpretation, pressure transient analysis,
tectonics analysis, inflatable packers, step-rate tests, core studies, etc.

7-2

B. Effect of Areal Permeability Variations


The best way to account for the effect of areal permeability variations on
waterflood performance is to determine its effect on areal sweep efficiency. As
indicated in Chapter 5, numerous studies have been made to determine areal
sweep efficiency, many of these for systems with areal heterogeneities. Perhaps,
by some fortuitous set of circumstances, one of these studies will match the
conditions in the reservoir being considered. More probably, you will have to
conduct your own study. Several possibilities exist:
1. Numerical simulation model - probably the best approach but can be
expensive. Results will be dependent on quality and quantity of input data.
2. Streamtube model - this type of model, while similar to simulation, may be
less expensive but can be very useful.
3. Analogy - extrapolate performance based on behavior of reservoirs with
similar characteristics.
II. Vertical Permeability Variations
Whereas a given layer of rock may exhibit lateral similarity due to its deposition
from a common source at a common geologic time, we find a reservoir may exhibit
many different layers in the vertical section that have highly contrasting properties.
This stratification can result from many factors including change in depositional
environment, change in depositional source, and particle segregation.
When vertical permeability stratification occurs to a significant degree, as it does
in many reservoirs, it will probably have more influence on waterflood
performance than any other physical reservoir characteristic. Water injected into a
stratified system will preferentially enter the layers of highest permeability and will
move at a higher velocity in these layers.

Consequently, at the time of water

breakthrough in high permeability zones, a significant fraction of the less permeable


zones will remain unflooded. Although a flood will generally continue beyond
7-3

breakthrough, the economic limit is often reached at an earlier time. Unless an


engineer has initiated a program to combat the effects of stratification, a large
fraction of the reservoir oil may remain untouched by water and unrecovered at the
time the project is terminated.
Recognizing the effect that stratification can have on ultimate waterflood recovery, it
is important that we be able to detect stratification and to quantify the effect it may
have.
A. Detection of Stratification
In contrast to areal heterogeneities, vertical stratification is easier to detect since
the producing horizon is generally penetrated by several wells. Within the small
area sampled in each well, log and core data give a good picture of the vertical
variation in properties. Further information is obtained from pressure transient
tests, production logs, and the behavior of production and injection wells. If a
particular strata occurs in several wells, we may confidently believe that it is
continuous across the field, and we can estimate what it is like between the wells.
If the strata cannot be traced from well to well, we have no idea what it is like
between wells and performance predictions become very difficult. The following
methods of quantifying vertical permeability variation assume each strata exhibits
areal continuity over that portion of the reservoir being studied. This may involve
the entire reservoir or simply the wells within a single flooding pattern.
B. Quantitative Evaluation of Permeability Stratification
The question which concerns us here is how to express quantitatively the effect of
permeability stratification on the injection and production behavior of a
waterflood system.

Several different techniques are commonly used to

accomplish this.

7-4

1. Single-Value Representation
One approach to the problem is to ask the following question: what single
value of permeability should be assigned to a homogeneous reservoir, having
the same size as the stratified reservoir, for it to behave in the same manner as
the stratified reservoir? It has been common for engineers to determine this
single-value permeability by simply taking a weighted average of the
permeabilities of each layer. For example:

k1h1 k2 h2 kn hn
h1 h2 hn

(Eq. 7.1)

where:

k1, k2 , kn
h1, h2 , hn

= permeabilities of individual layers which


compose the formation of interest
= thickness of individual layers

It is unfortunate that this procedure has been used so often because it gives
optimistic results. This averaging method is not recommended.
Several model studies1 using simulated flow patterns in variable permeability
media have shown that the best single-value representation of permeability is
obtained by taking the geometric mean of the available data. For example:

k ( k1 * k2 * k3 * ----*kn )1/ n

(Eq. 7.2)

If this relationship is applied to a vertical section, the formation should be


divided into intervals of equal thickness so that each value of permeability is
weighted equally.
Equation 7.2 can also be applied to find the best areal average permeability.
For example, suppose several wells penetrate what has been determined by
7-5

well-to-well correlation to be a common strata and it is desired to assign a


single value of permeability to that strata.

If each permeability value

represents an equal area, then Eq. 7.2 will give a good representation of the
average areal permeability.
Although it is convenient for mathematical purposes to replace a variable
permeability reservoir with an equivalent homogeneous reservoir having a
single permeability, it must be realized that this simplified model has severe
limitations. For example, it can be used to study the potential productivity or
injectivity of a well. It cannot be used, however, to study important facets of a
waterflood such as the water-oil ratio (WOR) behavior after water
breakthrough, cumulative water requirements, etc. Calculations of this type
require a prediction technique which accounts in detail for the permeability
contrast in the reservoir. The following models attempt to accomplish this.
2. Permeability Variation
The first statistical approach to predicting the effects of variable permeability
was presented by Law2 who showed that a random sample of permeability
data will generally have a log-normal distribution. Dykstra and Parsons3, in a
paper of fundamental significance, utilized this ideal model to compute a
coefficient of permeability variation. This method assumes the reservoir is
composed of a number of strata, or layers, each having a different
permeability with no cross-flow between the layers. The basic procedure for
determining the permeability variation using this layer-cake model is:
a. Divide permeability samples so all samples represent layers of equal
thickness, i.e., one foot.
b. Arrange the permeability data in the order of decreasing value.

7-6

c.

Calculate for each sample the percent of samples which have a greater
permeability and express this number as percent greater than. This is
illustrated by the following table.

k, md

% greater than

10
9
8
7
6
6
6
5
4
3

0
10
20
30
40
40
40
70
80
90

d. Plot the data from Step (c) on log-probability paper. Plot

on the log

scale and percent greater than on the probability scale. This plot is
illustrated by Figure 7-1

7-7

FIGURE 7-1
TYPICAL LOG PROBABILITY PLOT OF PERMEABILITY DATA

Permeability, md

100

10

6 8 10

15

20

30

40

50

60

70

80 85

90 92 94 96

Cumulataive Probability (Percent Greater Than)

e.

From the best straight line fit of the data, determine

at 84.1 and 50.0

percent probability.
f.

Compute the permeability variation, V , as:

V
The value of

k50 k84.1
k50

(Eq. 7.3)

computed in Step (f) is a quantitative indicator of the degree

of reservoir heterogeneity. Dykstra and Parsons3, as well as other authors4,5,


have used

to predict the expected performance of a waterflood. These

methods will be discussed in Chapter 8.

7-8

98

________________________________________________________________________

EXAMPLE 7:1
Table 7.1 presents the core data for a hypothetical reservoir presented in Reference 4.
Use the Dykstra-Parsons method to determine the permeability variation of this reservoir.
TABLE 7.1: Core Analysis for Hypothetical Reservoir
Cores from 10 Wells, A Through J
Each Permeability Value (md) Represents One Foot of Pay
Depth, ft
A
B
C
D
E
F
G
H
I
6,791
2.9
7.4 30.4
3.8
8.6 14.5 39.9
2.3 12.0
6,792
11.3
1.7 17.6 24.6
5.5
5.3
4.8
3.0
0.6
6,793
2.1
21.2
4.4
2.4
5.0
1.0
3.9
8.4
8.9
6,794
167.0
1.2
2.6 22.0 11.7
6.7 74.0 25.5
1.5
6,795
3.6 920.0 37.0 10.4 16.5 11.0 120.0
4.1
3.5
6,796
19.5
26.6
7.8 32.0 10.7 10.0 19.0 12.4
3.3
6,797
6.9
3.2 13.1 41.8
9.4 12.9 55.2
2.0
5.2
6,798
50.4
35.2
0.8 18.4 20.1 27.8 22.7 47.4
4.3
6,799
16.0
71.5
1.8 14.0 84.0 15.0
6.0
6.3 44.5
6,800
23.5
13.5
1.5 17.0
9.8
8.1 15.4
4.6
9.1

J
29.0
99.0
7.6
5.9
33.5
6.5
2.7
66.0
5.7
60.0

SOLUTION
The Dykstra-Parsons coefficient of permeability variation requires that all permeability
values, irrespective of their position in the reservoir, be combined and arranged in the
order of decreasing permeability. When this ordering is completed, we calculate for each
permeability the percentage of permeability values which are greater in magnitude than
the subject value -- this percentage is reported as percent greater than. Table 7.2 shows
the percent greater than calculations for this reservoir.

7-9

Table 7.2: Dykstra-Parsons Calculations for Example. 7.1

k , md

% Greater Than

k , md

% Greater Than

k , md

% Greater Than

920.0
167.0
120.0
99.0
84.0
74.0
71.5
66.0
60.0
55.0
50.4
47.4
44.5
41.8
39.9
37.0
35.2
33.5
32.0
30.4
29.0
27.8
26.6
25.5
24.6
23.5
22.7
22.0
21.2
20.1
19.5
19.0
18.4
17.6

0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33

17.0
16.5
16.0
15.4
15.0
14.5
14.0
13.5
13.1
12.9
12.4
12.0
11.7
11.3
11.0
10.7
10.4
10.0
9.8
9.4
9.1
8.9
8.6
8.4
8.1
7.8
7.6
7.4
6.9
6.7
6.5
6.3
6.0

34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66

5.9
5.7
5.5
5.3
5.2
5.0
4.8
4.6
4.4
4.3
4.1
3.9
3.8
3.6
3.5
3.3
3.2
3.0
2.9
2.7
2.6
2.4
2.3
2.1
2.0
1.8
1.7
1.5
1.5
1.2
1.0
0.8
0.6

67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99

These data are plotted on log probability paper as shown by Figure 7-2.

7 - 10

7 - 11

Permeability, md

0.1 2

10

100

1000

8 10

15

k 50

30

40

50

60

70

80

85

Cumulative Probability (Percent Greater Than)

20

k 84.1

LOG PROBABILITY PLOT FOR EXAMPLE 7:1

FIGURE 7-2

90 92 94

96

98

From Figure 7-2 it is observed that:

k50 10.2 md
k84.1 3.0 md
The permeability variation is computed from Eq. 7.3. For example:

k50 k84.1
k50

10.2 3.0
10.2

V 0.706
There are at least four major concerns related to computation of the DykstraParsons coefficient. First, to compute a V value, it must be assumed that the
permeability measurements possess a log-normal distribution.

Second,

practical geological concepts would suggest that only those core samples from
zones or areas across the field possessing similar environments of deposition
and/or diagenesis should be analyzed together. Third, only those core samples
possessing a permeability greater than the permeability cutoff should be used.
Finally,

( ko ) S wir

values should be used. Unfortunately,

ka

values which

do not represent true permeability are frequently used in the analysis.


Theoretically, the

factor ranges from 0.0 (homogeneous reservoir) to

0.9999 (very, very heterogeneous). As a practical matter, experience tells us


the V factor seldom is less than about 0.6 to 0.7. In carbonate reservoirs, the

factor is frequently 0.85 or greater. As the

factor increases, waterflood

oil recovery factor decreases for a specific WOR value.


7 - 12

Figure 7.3 is a graph relating WOR versus waterflood recovery factor for

factors of 0.7, 0.8, and 0.9. For each of the three curves, all other reservoir
variables were held constant. The

k50 permeability is 25 md, oil viscosity is

1.0 cp, water viscosity is 0.5 cp, gas saturation is 0.0, and the rock is
characterized by water wet relative permeability curves. It is observed that for
a WOR value of 20:1. the oil recovery factor ranges from 28 percent to 47
percent.

FIGURE 7-3
WATER-OIL RATIO versus OIL RECOVERY FACTOR
as a Function of Vertical permeability Heterogeneity
100

WATER-OIL RATIO

90
80

V=0.8

V=0.9

70

V=0.7

60
50
40
30
20
10
0
0

10

20

30

40

50

60

RECOVERY FACTOR, Np/N,%

Clearly, as the V factor increases for a specific set of reservoir conditions, oil
recovery factor decreases for a given WOR.

7 - 13

3. Stiles Permeability Distribution


The Stiles method6 utilizes a layer-cake model as did the method of Dykstra
and Parsons.

The Stiles procedure for expressing vertical variation in

permeability is as follows.
a. Arrange all permeability data, regardless of which well it came from or its
vertical position within the formation, in the order of decreasing
permeability.
b. Determine the distribution of flow capacity,

kh , within the formation.

It

is convenient to express this distribution in dimensionless form as is


illustrated by the following table.

h1
h2
h3
ht h
c.

k1
k2
k3

k h

k h 1
k h 2
k h 3
k h

k h i
k h

C1
C2
C3

C C h hi

C1
C2
C3 1.0

h1
h2
h3 1.0

h1
h2
h3

h
ht

1.0

Plot the capacity distribution curve for the reservoir; i.e.,


is illustrated by the solid curve,

C vs h' .

This

ABC , in Figure 7-4.

d. Use the capacity distribution curve to determine the permeability


distribution curve. You will note that, since capacity is defined as the
permeability-thickness product, permeability is the derivative of capacity
with respect to thickness. For example:

dC
dh

(Eq. 7.4)
7 - 14

so that dimensionless permeability is:

k'

dC
dh'

(Eq. 7.5)

Therefore, the permeability distribution curve can be obtained by


differentiating the capacity distribution curve. This can be accomplished
graphically by dividing the

h'

axis into equal increments (ten should be

sufficient). The necessary calculations are illustrated in the following table.

h'

C1
C2

h'1
h'1

h '

C1
h1 '
C2 C1 h' 2 h' 2

k'

C
h'

Plot Point

k '1
k '2

h'1 / 2
h'1 ( h' 2 h'1 ) / 2

k '10

h' 9 1.0 h' 9 / 2

C10 1.0 h'10 1.0 1.0 C9 1.0 h' 9

The permeability distribution curve is obtained by plotting

k'

versus the

plot point indicated in the table. Note that the plot point is simply the
midpoint of the interval used to compute k ' . The permeability distribution
is indicated by the dashed curve on Figure 7-4.

7 - 15

FIGURE 7-4
STILES CAPACITY AND PERMEABILITY DISTRIBUTION CURVES
1.0

kmax

Y
X
C

k 0.5

0.0

0.0
0.0

0.5

h'

1.0

A major criticism of this method is that it does not account for the position
from which each permeability value was obtained, i.e., each sample is
treated as random data. However, this method has been successfully used
and is one method of expressing permeability variation. We will show in
Chapter 8 how this type of permeability distribution is used to predict
waterflood behavior.
_______________________________________________________________

EXAMPLE 7:2
Shown in the following table are permeability data7 for a reservoir to be
waterflooded.

The data have already been rearranged in the order of

decreasing permeability. Plot the capacity and permeability distribution curves


for this reservoir.

7 - 16

Sample
No.

Thickness

Permeability

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
1
9

776
454
349
308
295
282
273
262
228
187
178
161
159
148
127
109
88
87
77
49

h, feet

k , md

SOLUTION
Calculations of capacity and permeability distribution are shown in Table 7.3
and Table 7.4, respectively. A plot of these data are presented in Figure 7-5.

7 - 17

Table 7.3: Calculation of Capacity Distribution

h, ft k , md kh, md ft
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
1
9

h 29

776
454
349
308
295
282
273
262
228
187
178
161
159
148
127
109
88
87
77
49

776
454
349
308
295
282
273
262
228
187
178
161
159
148
127
109
88
174
77
441

kh 5076

7 - 18

k h
h
h'
h
kh

0.1529
0.2423
0.3111
0.3717
0.4299
0.4854
0.5392
0.5908
0.6357
0.6726
0.7076
0.7394
0.7707
0.7998
0.8249
0.8463
0.8637
0.8980
0.9131
1.0000

0.0345
0.0690
0.1034
0.1379
0.1724
0.2069
0.2414
0.2759
0.3103
0.3448
0.3793
0.4138
0.4483
0.4828
0.5172
0.5517
0.5862
0.6552
0.6897
1.0000

Table 7.4: Calculation of Permeability Distribution


Plot Point

h'

h '

0.1529
0.2423
0.3111
0.3717
0.4299
0.4584
0.5392
0.5908
0.6357
0.6726
0.7076
0.7394
0.7707
0.7998
0.8249
0.8463
0.8637
0.8980
0.9131
1.0000
----

0.0345
0.0690
0.1034
0.1379
0.1724
0.2069
0.2414
0.2759
0.3103
0.3498
0.3793
0.4138
0.4483
0.4828
0.5172
0.5517
0.5862
0.6552
0.6897
1.0000
----

0.1529
0.0894
0.0688
0.0606
0.0582
0.0555
0.0538
0.0516
0.0449
0.0369
0.0350
0.0318
0.0313
0.0291
0.0251
0.0214
0.0174
0.0343
0.0151
0.0869
----

0.0345
0.0345
0.0344
0.0345
0.0345
0.0345
0.0345
0.0345
0.0344
0.0345
0.0345
0.0345
0.0345
0.0345
0.0344
0.0345
0.0345
0.0690
0.0345
0.3103
----

7 - 19

k'

C
h'

4.4334
2.5938
1.9939
1.7597
1.6854
1.6111
1.5597
1.4968
1.3026
1.0684
1.0169
0.9198
0.9084
0.8455
0.7256
0.6227
0.5028
0.4970
0.4399
0.2799
0.0000

h'i 1

h'i
2

0.0170
0.0520
0.0860
0.1210
0.1550
0.1900
0.2240
0.2590
0.2930
0.3280
0.3620
0.3970
0.4310
0.4660
0.5000
0.5340
0.5670
0.6030
0.6720
0.8450
1.0000

FIGURE 7-5
STILES CAPACITY AND PERMEABILITY DISTRIBUTION CURVES
FOR EXAMPLE 7:2

0.8

0.6

0.4

0.2

0.2

0.4

h'

0.6

0.8

4. Lorentz Coefficient
Another method of expressing vertical permeability variation which utilizes
the Stiles permeability distribution was presented by Schmalz and Rahme8.
They observed that the area between the capacity distribution curve and the
diagonal (Figure 7-4) is a measure of reservoir heterogeneity.

For a

homogeneous system, the capacity distribution curve would overlay the


diagonal.

As a measure of heterogeneity, a number called the Lorentz

coefficient was defined as:

Lorentz Coefficient

area WXYW
area WZYW

(Eq. 7.6)

The method experiences limited application, however, because the coefficient


is not unique; i.e., several different permeability distributions can yield the
same Lorentz coefficient.
7 - 20

5. Miller-Lents Permeability Distribution


Miller and Lents9 suggested an approach that retains the positional identity of
the permeability source.

They believed a reservoir rock is deposited in

essentially horizontal layers, and permeability data derived from a particular


vertical position in a well should be averaged only with data from a similar
position in other wells to define the characteristics of that particular layer.
This is in contrast to the Dykstra-Parsons and Stiles techniques which lump all
the permeability data together regardless of its original position in the
reservoir.
In application, individual strata or layers which compose the reservoir must
first be identified by well-to-well correlation or by statistical methods. The
permeability data from all wells penetrating a particular layer are then
averaged (geometric mean) to define the permeability of that layer. This is
repeated for other existing layers. The end result is still a layer-cake model,
but the positional identity of the data is preserved. At this point, the Stiles or
Dykstra-Parsons methods can still be applied to the averaged data, but the
results will be different than when the data are grouped in the order of
decreasing permeability.
Published studies by Elkins, et al10,11 indicate that the Miller-Lents approach
to handling heterogeneities yields a better match with actual field performance
than does the Stiles methods when applied to cycling operations in gas
condensate reservoirs. Published comparisons are not available, however, for
waterflood systems. The Stiles approach seems to be more commonly used
throughout the industry.

7 - 21

_______________________________________________________________

EXAMPLE 7:3
Consider the permeability data in Table 7.1. These data were analyzed in
Example 7.1 to determine the Dykstra-Parsons permeability variation.
Determine the permeability variation of these data using the Miller-Lents
positional approach and compare the results with the Dykstra-Parsons
variation.

SOLUTION
Whereas the Dykstra-Parsons method disregards the positional identity of the
data, this methods requires that the layer identity be retained.

Table 7.1

presents permeability data for ten wells, each of which contain ten layers. The
approach used will be to determine the geometric mean permeability of each
layer. The average permeability of each layer will be plotted on log probability
paper to determine the permeability variation.
The geometric mean permeability is defined by Eq. 7.2.

Applying this

relationship to the permeabilities in the first layer at each well:

k 1 (( 2.9)(7.4)(30.4)(3.8)(8.6)(14.5)(39.9)( 2.3)(12.0)( 29.1))1 / 10

k1 10.0 md
Average permeabilities in the remaining layers are determined in the same
manner. These values are summarized in the following table.

7 - 22

Layer

k , md

1
2
3
4
5
6
7
8
9
10

10.0
6.8
4.7
10.4
20.5
12.1
8.6
18.4
14.3
10.9

The average permeabilities are now rearranged in the order of decreasing


permeability and the percent greater than is computed for each value. These
calculations are presented in the following table.

k , md

Percent Greater Than

20.5
18.4
14.3
12.1
10.9
10.4
10.0
8.6
6.8
4.7

0
10
20
30
40
50
60
70
80
90

These data are plotted in Figure 7-6.

7 - 23

FIGURE 7-6
LOG PROBABILITY PLOT OF PERMEABILITY DATA FOR
EXAMPLE 7:3

Permeability, md

100

k50

10

k84.1
1

6 8 10

15 20

30

40

50

60

70

80 85

90 92 94 96

98

Cumulative Probability (Percent Greater Than)

Using the data from Figure 7-6, the permeability variation is computed to be:

k50 k84.1 10.0 5.95

k50
10.0

V 0.405
This compares to

V 0.706

computed using the conventional Dykstra-

Parsons method.
_______________________________________________________________
C. Selection of Layers
A big question which is encountered early in the effort to predict stratification
effects is how to recognize and select the individual zones which comprise the
reservoir. The basic Dykstra-Parsons and Stiles methods result in layer selections
7 - 24

which have no physical meaning, although these methods can still be applied
when the layers are chosen based on position. It seems logical that a zonation
technique should be used which recognizes the actual location of strata within the
reservoir. Several possibilities exist.
1. Geological Zonation - Zones are selected based on similar lithological
characteristics. This approach requires much detailed information from cores,
well logs, lithological analyses, etc. but results in the most natural zonation
possible. A very good approach but is time consuming and expensive.
2. Natural Barriers - Zone selection is sometimes made easy by the occurrence
of shale barriers which break the reservoir into natural zones. However, life is
seldom this simple.
3. Equal Thickness - This is often used because it is simple, and it retains the
positional identity of the strata. The major limitation is that it does not
account for natural zonation within the reservoir.
4. Equal Flow Capacity

(kh )

- Probably better than (3) since it better reflects

the effect of high permeability zones which control the water-oil ratio
behavior of a flood.
5. Statistical Zonation - A statistical method which eliminates much of the bias
in zone selection was suggested by Testerman12. The permeability data from
each well are statistically divided into zones so as to provide maximum
permeability contrast between zones and yet have minimum permeability
variation within a given zone. The zones are then traced from well to well by
statistical correlation. This method has received considerable use, but it does
require the use of a computer.

7 - 25

D. Effect of Crossflow Between Layers


The methods previously discussed contain the common assumption that a
reservoir is made up of a series of layers, each having horizontal continuity,
where each layer is insulated from its neighbor except at the wellbore of
producing and injection wells. Accordingly, the crossflow of fluid between layers
is neglected. Although some reservoirs contain shale beds or shale streaks which
prevent crossflow, most reservoirs have sufficient vertical permeability and
vertical continuity to experience flow between layers.
Many experimental and mathematical studies have been conducted to evaluate
crossflow effects. These studies, summarized in References 4 and 5, indicate that
for favorable mobility ratios

( M 1) ,

cross flow effects tend to improve

recovery performance beyond that predicted by the layer-cake models. However,


unfavorable mobility ratios tend to reverse this trend.

Most studies which

compared field performance with studies predicted using layer-cake models show
the predicted results are generally pessimistic. Crossflow can only be considered
with numerical simulation methods.
III. Vertical Sweep Efficiency
As a result of permeability stratification, and other effects to be discussed, injected
water is seldom able to contact the entire vertical cross-section of a reservoir. As a
measure of the efficiency with which water covers a reservoir in the vertical plane,
we define the term vertical sweep efficiency,

Ev .

This term is also sometimes

referred to as invasion sweep efficiency and designated by the symbol,

EI

Vertical sweep efficiency is defined as the hydrocarbon pore space contacted by


injected water divided by the hydrocarbon pore space behind the water front (the
water front is defined by its most forward position).

7 - 26

As noted in previous discussions, the vertical sweep efficiency is significantly


affected by stratification due to the preferential movement of fluids in the more
permeable zones. This is complicated further by several additional factors.
A.

Mobility Ratio - Vertical sweep improves with decreasing mobility ratio.

B.

Crossflow - Discussed previously.

C.

Gravity Forces - Gravity effects can significantly reduce vertical sweep in


some reservoirs. However a general correlation is impossible due to effects of
rate, vertical permeability, stratification, etc. Simulation studies and lab tests
indicate that increasing the rate tends to minimize gravity effects. In practice,
however, rate increases of several-fold are required to cause significant
changes in performance.

Changes of this magnitude are generally not

practical.
D.

Capillary Forces - A summary of studies on vertical sweep efficiency is


provided in References 4 and 5.

7 - 27

CHAPTER 7 REFERENCES

1. Warren, J.E. and Price, H.S.: "Flow in Heterogeneous Porous Media," Trans., AIME
(1961) 222, pp. 153-169.
2. Law, J.: "Statistical Approach to the Interstitial Heterogeneity of Sand Reservoirs,"
Trans., AIME (1944) 155, pp. 202-222.
3. Dykstra, H. and Parsons, R.L.: "The Prediction of Oil Recovery by Waterflood,"
Secondary Recovery of Oil in the United States, 2nd Ed., API (1950) pp. 160-174.
4. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Richardson, TX (1971) 3.
5. Willhite, F.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
6. Stiles, W.E.: "Use of Permeability Distribution in Waterflood Calculations," Trans.,
AIME (1949) 186, pp. 9-13.
7. Garb, F.A.: "Waterflood Calculations for Hand-Held Computers," World Oil (June,
1980) pp. 205-210.
8. Schmalz, J.P. and Rahme, H.D.: "The Variation in Waterflood Performance with
Variation in Permeability Profile," Prod. Monthly (1950) 15, No. 9, pp. 9-12.
9. Miller, M.G. and Lents, M.R.: "Performance of Bodcaw Reservoir, Cotton Valley
Field Cycling Project: New Methods of Predicting Gas-Condensate Reservoir
Performance Under Cycling Operations Compared to Field Data," Drilling and
Production Practices, API (1947) pp. 128-149.
10. Elkins, L.F. and Skov, A.M.: "Some Field Observations of Heterogeneity of
Reservoir Rocks and Its Effects on Oil Displacement Efficiency," paper SPE 282
presented at the 1962 SPE Production Research Symposium, Tulsa, April 12-13.
11. Elkins, L.F., Brown, R.C. and Skov, A.M.: "Comparison of Performance During
Cycling and Blowdown with Various Prediction Methods - Washington Cockfield
"D" Gas Condensate Reservoir," paper SPE 5531 presented at the 1975 SPE Annual
Technical Conference and Exhibition, Dallas, Sept.28-Oct.1.
12. Testerman, J.D.: "A Statistical Reservoir Zonation Technique," JPT (August 1962)
pp. 889-893.

7 - 28

PROBLEM 7:1

A sandstone oil reservoir under study for waterflooding has an average thickness of 30
feet. For reservoir engineering study, it has been subdivided into ten 3-foot intervals.
Routine air permeability data,
Also,

( ko ) S

7P.1-1.

( ko ) S

wir

ka , obtained from several wells, are available for analysis.

measurements are available for 15 samples which are tabulated in Table

These data are cross-plotted in Figure 7P.1-1.


wir

is 0.3 md. This is equivalent to

ka

The permeability cutoff,

of 1.0 md. The air permeability values

have been analyzed in the order of decreasing permeability as listed in Table 7P.1-2 after
applying the 1.0 md cutoff. Figure 7P.1-2 is a Dykstra-Parsons plot of the

ka

data for

those samples with permeability above the cutoff.


1. Compute the Dykstra-Parsons coefficient for this reservoir.
2. Compute the permeability values,

( ko ) S

wir

, which should be assigned to

a 10 layer equal-thickness waterflood prediction model to analyze


waterflood performance.

7 - 29

TABLE 7P.1-1
SPECIAL CORE ANALYSIS DATA

ka , md

( ko ) S

0.346
0.767
0.704
5.300
1.220
11.500
0.190
4.380
0.335
0.595
4.430
0.299
4.210
10.600
1.430

wir

, md

0.045
0.190
0.197
3.310
0.617
4.770
3.036
1.350
0.112
0.094
1.430
0.066
1.360
3.270
0.489

7 - 30

7-31

0.01

0.1

10

0.1

ka , md

10

(ko.) Swir versus ka from Special Core Analysis

FIGURE 7P1.1

100

TABLE 7P.1-2
REARRANGED CORE PERMEABILITY AFTER
APPLYING THE 1.0 PERMEABILITY CUTOFF
Air
Permeability
md
186.0
38.0
34.0
24.0
22.0
20.0
19.0
18.0
17.0
16.0
15.0
15.0
14.0
13.0
12.0
11.0
10.0
8.9
8.6
8.5
7.7
7.5
7.0
6.8
6.4
6.0
5.8
5.7
5.5
5.3
5.1
4.7
4.5

Cumulative
Number of Number of
Samples
Samples
1
1
1
1
1
2
1
1
1
2
1
1
1
1
2
3
1
1
1
2
1
2
1
2
1
1
1
1
2
1
1
2
3

1
2
3
4
5
7
8
9
10
12
13
14
15
16
18
21
22
23
24
26
27
29
30
32
33
34
35
36
38
39
40
42
45

7 - 32

Cumulative Samples
for

ka 1.0 md

percent greater than


0.00
0.97
1.85
2.78
3.70
4.63
6.48
7.41
8.33
9.26
11.11
12.04
12.96
13.89
14.81
16.67
19.44
20.37
21.30
22.22
24.07
25.00
26.85
27.78
29.63
30.56
31.48
32.41
33.33
35.19
36.11
37.04
38.89

TABLE 7P.1-2 (CONTINUED)


REARRANGED CORE PERMEABILITY AFTER
APPLYING THE 1.0 PERMEABILITY CUTOFF
Air
Permeability
md
4.2
4.1
4.0
3.9
3.8
3.7
3.6
3.4
3.3
3.1
3.0
2.8
2.6
2.4
2.3
2.2
2.1
2.0
1.9
1.8
1.7
1.6
1.5
1.4
1.3
1.2
1.1
1.0

Permeability cutoff is

Cumulative
Number of Number of
Samples
Samples

( ko ) S

2
1
1
2
1
1
2
2
1
1
2
1
1
1
2
1
1
6
1
1
1
2
4
3
3
10
6
3

wir

47
48
49
51
52
53
55
57
58
59
61
62
63
64
66
67
68
74
75
76
77
79
83
86
89
99
105
108

Cumulative Samples
for

ka 1.0 md

percent greater than


41.67
43.52
44.44
45.37
47.22
48.15
49.07
50.93
52.78
53.70
54.63
56.48
57.41
58.33
59.26
61.11
62.04
62.96
68.52
69.44
70.37
71.30
73.15
76.85
79.63
82.41
91.67
97.22

0.3 md or ka 1.0 md

7 - 33

FIGURE 7P1.2
Dykstra-Parsons Graph
For an Air Permeability Greater than 1.0 md

Air Permeability, md

100

10

6 8 10 15 20

30

40

50

60

70

80 85 90 92 94 96

Cumulative Probability (percent greater than)

7 - 34

98

Log Probability Graph Paper

Air Permeability, md

100

10

6 8 10 15 20

30

40

50

60

70

80 85 90 92 94 96

Cumulative Probability (percent greater than)

7 - 35

98

PREDICTION OF WATERFLOOD PERFORMANCE

This chapter is concerned with the problem of predicting waterflood behavior. Given a
particular waterflood prospect, it is desirable to predict information such as the time
required for initial secondary oil response (gas fillup), water breakthrough, oil recovery at
breakthrough,

water-oil

ratio

performance

after

breakthrough,

production-time

performance, oil production-water injection performance, etc. Numerous methods have


been proposed to accomplish this, each differing in the manner of handling heterogeneity,
areal sweep calculations, water injection performance, displacement efficiency, and other
variables which can affect waterflood performance.
For purposes of description, waterflood prediction methods have been categorized into
four groups. These groups are defined as being:

Simple methods

Reservoir stratification methods

Confined patterns with stratification, areal sweep, and displacement mechanism


methods

Numerical simulation methods

I. Simple Methods
Waterflood estimates are sometimes computed using simple techniques such as (1)
analogy, (2) rules of thumb such as a secondary to primary (s/p) ratio method, and (3)
empirical relationships. These methods may work in localized areas; however, in
general it is believed these methods can lead to significant errors because they
frequently do not take into account data that is unique to the reservoir being studied.
The simple waterflood calculation techniques should not represent the principal
procedure for waterflood forecasting, except in circumstances where there is a lack of
technical data or when time constraints prevail.
8-1

A. Analogy
When computing waterflood performance behavior using the analogy technique,
it is important to remember that to be analogous, at the start of injection, projects
should be similar in many respects such as:

well density

flood pattern

fluid saturations,

oil and water viscosity,

relative permeability to water and oil,

reservoir stratification, Dykstra-Parsons coefficient, V

net pay cutoff criteria

injection rate per foot of net pay, BWPD/foot

oil formation volume factor,

residual oil saturation to waterflooding,

So , S g , S wc ,

and

S wir

o and w
krw and kro

Bo
Sor

It is significant to note that two reservoirs are analogous only when rock and fluid
properties and saturations are similar. In many instances, reservoirs have been
considered analogous due to the fact that their rock properties are similar. This is
usually incorrect! For example, if two reservoirs possess similar rock properties
but the free gas saturation in one reservoir at the start of injection is 20 percent
and the free gas saturation in the second reservoir is 10 percent, the waterflood
performance of the two reservoirs should be quite different. Therefore, similarity

8-2

in rock and fluid properties and fluid saturations are required for two reservoirs
to be analogous.
B. Rules of Thumb
Rules of thumb represent "generalized" statements which frequently are not
applicable to any particular reservoir. For example, a common phrase that is used
to describe waterflood potential is that the secondary recovery will be equal to the
primary recovery. In other words, the s/p value is 1.0. This is a false statement.
There is no technical basis to conclude the s/p ratio will be 1.0 except under very
limited conditions. In fact, the s/p ratio can range from less than 0.25 to greater
than 2.5 -- a ten-fold difference. The s/p ratio is influenced by many factors such
as the fluid saturations at the start of waterflooding, reservoir stratification
(Dykstra-Parsons coefficient), injection efficiency, directional permeability
trends, reservoir continuity, and waterflood pattern. Generalized statements or
experience guidelines should be used with caution.
C. Empirical Relationships
Several empirical relationships relating waterflood recovery factor to reservoir
variables such as connate water saturation, average permeability, and oil viscosity
have been developed. These empirical procedures are summarized in Reference
1. It is recommended these relationships be used with extreme caution because
they do not incorporate the important reservoir rock and fluid variables that
determine the outcome of a waterflood project.
II. Reservoir Stratification
It has been long recognized in the petroleum industry that the recovery factor in a
waterflood project is a combination of E A ,

EV , and ED .

In 1949 and 1950, two

waterflood methods2,3 were developed which attempted to take into account EV .


The techniques are similar and are referred to as the Dykstra-Parsons3 method and
the Stiles2 method. The methods utilize Equation 8.1 to calculate oil production.
8-3

N p N E A EV ED

(Eq. 8.1)

It is the intent of the two prediction methods to account for the effects of reservoir
stratification (variation in permeability) and thus EV . Significantly, both methods
assume piston displacement (residual oil saturation in the water contacted portion of
the reservoir). Furthermore, both methods assume linear flow which implies the
areal sweep at initial water breakthrough is 100 percent. Dykstra-Parsons and Stiles
recognized that piston displacement and linear flow resulted in an optimistic estimate
of oil recovery. However, it should be recalled that their prediction methods were
developed at a time prior to the widespread availability of powerful desktop
computing capabilities.

Therefore, their methods were rather simplistic.

Nevertheless, if reasonable estimates of

EA

(say 60 to 80 percent) and

ED

could

be made for a particular reservoir, Equation 8.1 reduces to:

N p A EV

(Eq. 8.2)

where:

A N E A ED
Dykstra-Parsons and Stiles developed mathematical procedures for computing

EV

versus water to oil ratio (WOR ) for no cross flow multi-layered reservoirs of
different permeabilities. After computing

EV

versus

WOR ,

it is possible to

calculate oil production versus water-oil ratio using Equation 8.2. Results can be
tabulated as shown below.

8-4

WOR

EV

N p (Eq. 8.2), STBO

N p 1
N p 2

WOR 1 EV 1
WOR 2 EV 2
-----

-----

-----

N p N

WOR N EV N

The assumption of piston displacement and constant areal sweep at water


breakthrough for a particular layer means that production from that layer after
breakthrough is 100 percent water. Only oil is produced from the layers in which
breakthrough has not occurred.

The Dykstra-Parsons and Stiles methods are

summarized in the following paragraphs.


A. Dykstra-Parsons Method
Dykstra and Parsons developed a method of predicting waterflood behavior in
stratified systems which may be helpful in some cases if a rough approximation
of waterflood recovery is needed.

This method requires knowledge of the

variation in permeability between the various layers and the mobility ratio,

M.

The original method is subject to several assumptions and limitations which affect
the accuracy of waterflood predictions. The method assumes:

A layer-cake model with no crossflow between layers

Equal porosity layers

Piston-like displacement with no oil production after water breakthrough

Linear flow

8-5

Assumes 100 percent areal sweep at initial water breakthrough (This value can
be adjusted to be less than 100 percent at breakthrough, but there is no simple
method to allow for the increase in

E A after breakthrough.

Steady-state flow

Fluid properties and saturations are identical for each layer

Gas saturation at the start of injection is negligible

Permeability and thickness are the only variations between layers

The Dykstra-Parsons linear flow model consists of a series of layers arranged in


order of decreasing permeability. This is illustrated by Figure 8-1 which depicts
the reservoir at the time of water breakthrough in the most permeable layer.
FIGURE 8-1
LINEAR FLOW RESERVOIR
FOR DYKSTRA-PARSONS WATERFLOOD METHOD
Linear Reservoir

i w1

Water Zone

i w2

Water Zone

k1, h1

k2 , h 2

Oil Zone

q w1

q o2

Water Zone

i wn

Oil Zone

k n , hn

k1
1 > k2
2 > .... > kn
n

8-6

q on

In the Dykstra-Parsons method, when water breakthrough occurs in any layer


of a

layered system, only water is produced from layers 1 through

oil is produced from layers


in layer

x,

x 1 through n .

and only

Moreover, at water breakthrough

Dykstra-Parsons showed that water-oil ratio (WOR ) could be

computed for any value of

as:

WOR

qwi

i 1
n

(Eq. 8.3)

qoi

i x 1
or:

WOR

hi ko S

i 1
n

i x 1

wir ,i

hi ko S

(Eq. 8.4)

wir ,i

1
2

ko Swir ,i
M 2
1 M 2
ko Swir , x

Also, at the time of water breakthrough in layer x , Dykstra-Parsons showed that


the vertical sweep (or vertical coverage) could be computed as:

1
2

ko Swir ,i
n
x M n x
1
2
EV
1 M 2

n M 1 n M 1 n i x 1
ko Swir , x

(Eq. 8.5)
where:

8-7

ko Swir

= effective permeability to oil at the irreducible connate water


saturation for layer

or x .

n layered reservoir, Equations 8.4 and 8.5 can be used to compute values
of WOR and EV when water breakthrough occurs in each layer. Results can
For an

be combined with Eq. 8.2 to estimate waterflood production as a function of

WOR .
and

Garb4 presented an example problem illustrating how

N p can be computed.

WOR , EV ,

Johnson5 and Mobarak6 have published papers on

this recovery method.


B. Stiles Method
The Stiles method2,7 is an older technique used for predicting waterflood behavior
in stratified reservoirs.

The method was first published in 1949 but is not

commonly used at the present time because of the development of better


prediction methods such as the Craig-Geffen-Morse technique and the use of
numerical simulation procedures. The Stiles method is similar to the DykstraParsons method and is also limited by the following assumptions.*

A layer-cake model with no crossflow between layers

Equal porosity layers

Piston-like displacement with no oil production after water breakthrough in each


layer

Linear flow

Assumes 100 percent areal sweep at initial water breakthrough (This value can
be adjusted to be less than 100 percent at breakthrough but there is no simple
method to allow for the increase in E A after breakthrough.)

Steady-state flow

A more comprehensive development of the Dykstra-Parsons method is available by contacting


Bill Cobb
*

8-8

Fluid properties and saturations are identical for each layer

Gas saturation at the start of injection is negligible

Permeability and thickness are the only variations between layers

Distance of flood front penetration into each layer is proportional to the flow
capacity

k
h
o
S

wir

of the layer. This is equivalent to assuming the

mobility ratio is unity.


1.

Flow Capacity

C and Permeability Distribution k

The Stiles method utilizes a layer-cake model as did the method of Dykstra and
Parsons.

The first step is to prepare a dimensionless flow capacity and

permeability distribution curve for the reservoir. These curves are computed in
the following manner.
a. Arrange all permeability data, regardless of its vertical position within the
formation, in the order of decreasing permeability.
b. Determine the distribution of flow capacity,
simply as

kh

k
h
o
S

wir

(denoted

in this section for convenience) within the formation. It is

convenient to express this distribution in dimensionless form as is illustrated


by the following table.

k1

h1

k2
k3

h2
h3

k h

k h 1
k h 2
k h 3

ht k h k h

k h i

k h
C1

C 2
C

C3 C h hi

C1
C2

C3 1.0

1.0
8-9

h1
h2
h3

h
h1
h2

h
h

h3 1.0

c. Plot the flow capacity distribution curve for the reservoir; that is

C vs. h .

This is illustrated by the solid curve in Figure 8-2.

FIGURE 8-2
TYPICAL STILES PERMEABILITY AND
FLOW CAPACITY DISTRIBUTION CURVE

kmax

d. Use the flow capacity distribution curve to determine the permeability


distribution curve.

You will note, since capacity is defined as the

permeability-thickness product, permeability is the derivative of flow


capacity with respect to thickness.

dC
dh

(Eq. 8.6)

8 - 10

so that dimensionless permeability is:

dC
dh

(Eq. 8.7)

Therefore, the permeability distribution curve can be obtained by


differentiating the flow capacity distribution curve.
accomplished graphically by dividing the

This can be

h axis into equal increments.

For

example, consider 10 increments. The necessary calculations are illustrated


in the following table:

C
C1

h1

C2

h2

C1

h1

C2 C1 h2 h1

C Plot Point
h

k1

h1 / 2

k 2

h1 h2 h1 / 2

C10 1.0 h 1.0 1.0 C9 1.0 h9

k10

h9 1.0 h9 / 2

The permeability distribution curve is obtained by plotting

k versus

the

plot point indicated in the table. Note that the plot point is simply the
midpoint of the interval used to compute k . The permeability distribution
is indicated by the dashed curve on Figure 8-2. Craft and Hawkins8 provide
independent discussions on the development of the flow capacity and
permeability distribution plots.
8 - 11

A major criticism of the Stiles method is that it does not account for the
position from which each permeability value was obtained. In other words,
each sample is treated as random data. However, this method was used to
approximate waterflood behavior in the 1950s and 1960s but the method is not
commonly used at the present time.
2.

Vertical Sweep Efficiency

EV

It is assumed in this method that flow is linear and that the distance of
penetration of the flood front is proportional to permeability. This means that
the front of advancing water will have the same shape as the permeability
distribution curve.

Consider Figure 8-3 which depicts the permeability

distribution of a reservoir; if this distribution curve is thought of as the flood


front, then for clarity we can assume that the line
well and line

cd

ab

represents the injection

represents the producing well. The position of the flood

front after h1 layers have reached water breakthrough is cfb ; the fraction of

X Y . Since
the total reservoir volume is equivalent to the area X Y Z , therefore
the reservoir flooded at this time is proportional to the area

the vertical sweep efficiency is equal to:

Vertical Sweep EV

8 - 12

X Y
X Y Z

(Eq. 8.8)

FIGURE 8-3
USE OF STILES PERMEABILITY
DISTRIBUTION CURVE TO DEPICT SHAPE OF
FLOOD FRONT
g

k1

Z
X
a

h1

It can be shown7 that the area under the permeability distribution curve is
unity, that is:

W X Y 1.0

(Eq. 8.9)

Since the flow capacity distribution is the integral of the permeability curve,
the capacity corresponding to the dimensionless formation thickness, h1 is:

C W X

(Eq. 8.10)
8 - 13

Combining Equations 8.9 and 8.10,

Y 1.0 W X

Y 1.0 C

(Eq. 8.11)

It is further observed from Figure 8-3 that

X ae ac h1k1
In the general case where

fraction of the total formation thickness has

achieved water breakthrough,

X hk

(Eq. 8.12)

It follows that the vertical sweep,

EV ,

defined by Equation 8.8, can be

rewritten as:

EV

k h 1 C
k

(Eq. 8.13)

Equation 8.13 can be used to compute vertical sweep of the water front as a
function of the fraction of formation which has achieved water breakthrough.
The only information required for this calculation is the flow capacity and
permeability distribution curve.
3.

Water Cut and Water-oil Ratio


Referring again to Figure 8-3, that portion of the formation with permeabilities
greater than k1 will be flowing water. The formation flow capacity producing
water, therefore, is C , and the formation flow capacity producing oil is

1 C . According to Darcy's Law, the water reservoir production rate from


that portion of the formation with a flow capacity
8 - 14

is:

k
qw C rw
w

(Eq. 8.14)

Further, the reservoir oil producing rate can be expressed as:

k
qw 1 C rw
w

(Eq. 8.15)

Thus, the total reservoir production rate is:

qt iw qo qw
The reservoir water cut,

Ckrw

f wR ,

1 C kro
o

(Eq. 8.16)

defined as the fraction of total reservoir

production which is water, can be computed as

f wR

CA
CA 1 C

(Eq. 8.17)

where

krw o
w kro

The surface water cut,

(Eq. 8.18)

fw ,

defined as the fraction of the total surface

production which is water, can be computed as:

fw

CS
CS 1 C

(Eq. 8.19)

8 - 15

where

krw o Bo
w kro Bw

(Eq. 8.20)

The surface water-oil ratio can be computed as:

WOR
4.

qw Bo
CS

qo Bw 1 C

(Eq. 8.21)

Oil and Water Producing Rates


If steady state flow is assumed, the total reservoir flow rate will be equivalent
to the water injection rate, therefore:

qo qw iw

(Eq. 8.22)

It follows that the surface production rate of water can be computed as:

qw f w iw , STB/D

(Eq. 8.23)

Accordingly, the oil production rate, expressed at reservoir conditions is:

qo iw qw Bw , RB/D

(Eq. 8.24)

The stock tank oil production rate is:

qoS
5.

qo
, STB/D
Bo

(Eq. 8.24)

Cumulative Oil Recovery


Cumulative oil recovery can be computed at any time in the life of a flood in
terms of the vertical sweep at that time.
variables has been shown previously to be:
8 - 16

The relationship between these

N p N E A EV ED

(Eq. 8.1)

or

Np

7758 Ah So
E A EV
Bo

or

Np

7758 Ah So
So Sor E A EV
Bo

S Sor
o

So

where

6.

= floodable area, acres

= floodable net pay, ft

= porosity, fraction

So

= oil saturation at start of injection, fraction

Sor

= residual oil saturation to waterflooding, fraction

EA

= estimated areal sweep, fraction

EV

= Equation 8.13

Procedure for Predicting Performance


a. Arrange the permeability data in the order of decreasing permeability and
prepare a plot of dimensionless permeability,
capacity,

C,

k ,

and dimensionless

as a function of dimensionless formation thickness,

h as

illustrated in Figure 8-2.


b. Divide the dimensionless permeability and capacity curves into increments
of equal thickness (such as ten layers) and select from the curves values of
8 - 17

k and C to represent each


h 0.1, 0.2, ..., 1.0 .

layer; i.e., read values of

and

at

c. The cumulative oil production and water-oil ration can be calculated as


outlined in the following table for different mobility ratios,

h1
h2
...
...
...
h10

k1 C1
k 2 C2
... ...
... ...
... ...
k10 1.0

M.

EV , Eq.8.13 N p , Eq.8.1 WOR*, Eq.8.21


EV 1

N p1
N p2
...
...
...
N p10

EV 2
...
...
...
1.0

WOR 2

...
...
...
WOR 10

* indicates value before breakthrough in the indicated interval

III. Confined Patterns With Stratification, Areal Sweep, and Displacement Methods
Craig, Geffen, and Morse (CGM)1,9 and Higgins-Leighton10,11,12,13 developed procedures
for estimating waterflood performance in multi-layer reservoirs without crossflow. The
CGM methods is one of the most thorough and practical prediction methods available
for five-spot systems. It is believed the technique is also applicable to other patterns if a
few minor adjustments are made. The method utilizes a modified Welge14 equation to
consider the displacement mechanism in the swept area. Variations in injectivity for
constant pressure water injection are accounted for using the conductance ratio

by

Caudle and Witte15, and the effects of increases in areal sweep efficiency beyond
breakthrough are included on the basis of the correlations presented by Craig, Geffen,
and Morse16.

8 - 18

Although the original CGM paper did not consider multi-layered systems, subsequent
modifications by Hendrickson17, and by Wasson and Schrider18, permit application to
stratified systems. A detailed discussion of this method is presented in SPE Monograph
31 and in Wilhite's SPE Textbook9, along with an example application. The CGM
technique is presented in detail in Appendix CGM at the end of Chapter 8.
The Higgins-Leighton method basically applied the displacement theory of Buckley and
Leverett19 to any flooding pattern for which the isopotential and flow streamlines are
available. It is more complicated to use than previously discussed methods and requires
the use of a computer. To apply the method, the reservoir is divided into flow channels
based on flow streamlines as determined from numerical simulation studies. Each
stream channel is subdivided into equal volume cells. Assuming unidirectional flow, a
Buckley-Leverett type material balance on each cell yields the rate of water
accumulation and oil displacement from which saturation gradients can be determined.
From individually calculated flow resistance for each cell, and the total pressure drop
between wells, instantaneous oil and water flow rates can be computed.
Data required for the Higgins-Leighton method of analysis includes oil/water relative
permeability, viscosity, effective layer permeability to oil

ko Swir , layer thickness,

applied differential pressure, and the isopotential and streamline configuration for the
particular well pattern studied.
A major limitation of the method is its dependence on the resistance factors (shape
factor) which must be known for each cell to properly account for sweep variations
induced by the different cell geometry. These resistance factors have been presented in
the literature for many commonly used flooding patterns. A major assumption in setting
up the cell models is that stream channels determined using a unit mobility
can be applied to any system. This method is also reviewed by Wilhite.

8 - 19

M 1

IV. Numerical Simulation


A complete solution to the multiphase, multidimensional partial differential equations
which govern fluid flow in a porous and permeable media is probably the best
waterflood prediction model that can be used. Such a mathematical model (numerical
simulator) can account for such items as directional variation in fluid and rock
properties, layering effects, crossflow, gravity, capillary pressure, irregular boundaries,
and individual well behavior. The effects of varying injection patterns, horizontal wells,
well locations, injection and producing rates, plus many other factors, can be studied
which were not possible using previously discussed waterflood prediction methods.
In general, mathematical numerical models are relatively inexpensive to develop and
run. Moreover, they are available for lease on a short term basis (month to month) and
can be run on high speed personal computers or work stations. Significantly, extensive
amounts of data are generally required to take advantage of the flexibility and accuracy
afforded by these models. Numerical simulation guidelines have been summarized by
Mattax and Dalton20 in SPE Monograph 13. It is beyond the scope of this seminar to
discuss numerical simulation methods.

8 - 20

CHAPTER 8 REFERENCES

1. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph


Series, SPE, Dallas (1971) 3.
2. Stiles, W.E.: "Use of Permeability Distribution in Waterflood Calculations," Trans.,
AIME (1949) 186, pp. 9-13.
3. Dykstra, H. and Parsons, H.L.: "The Prediction of Oil Recovery by Waterflooding,"
Secondary Recovery of Oil in the United States, 2nd ed., API, New York (1950) pp. 160174.
4. Garb, F.A.: "Waterflood Calculations for Hand-Held Computers," World Oil, (July
1980), pp. 155-159.
5. Johnson, C.E., Jr.: "Prediction of Oil Recovery by Waterflood -- A Simplified Graphical
Treatment of the Dykstra-Parsons Method," Trans., AIME (1956) pp. 207, 345-346.
6. Mobarak, S.: "Waterflooding Performance Using Dykstra-Parsons As Compared with
Numerical Model Performance," Journal of Petroleum Technology, (January 1975) pp.
113-115.
7. Cole, F.W.: Reservoir Engineering Manual, Gulf Publishing Co., Houston (1969).
8. Craft, B.C. and Hawkins, M.F.: Applied Petroleum Reservoir Engineering, PrenticeHall Inc., Englewood Cliffs, N.J. (1959).
9. Wilhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
10. Higgins, R.V. and Leighton, A.J.: "A Computer Method to Calculate Two-Phase Flow
in Any Irregularly Bounded Porous Medium," Journal of Petroleum Technology, (June,
1962) pp. 679-683.
11. Higgins, R.V. and Leighton, A.J.: "Computer Prediction of Water Drive of Oil and Gas
Mixtures Through Irregularly Bounded Porous Media -- Three Phase Flow," Journal Of
Petroleum Technology, (September 1962) pp. 1048-1054.
12. Higgins, R.V. and Leighton, A.J.: "Waterflood Prediction of Partially Depleted
Reservoirs," paper SPE 757 presented at the 33rd Annual California Regional Fall
Meeting, Santa Barbara, October 24-25, 1963.
13. Higgins, R.V., Boley, D.W. and Leighton, A.J.: "Aids in Forecasting the Performance of
Water Floods," Journal of Petroleum Technology, (September 1964) pp. 1076-1082.
14. Welge, H.J.: "A Simplified Method for Computing Oil Recovery by Gas or Water
Drive," Trans., AIME (1952) 195, pp. 91-98.
8 - 21

15. Caudle, B.H. and Witte, M.D.: "Production Potential Changes During Sweepout in a
Five-Spot Pattern," Trans., AIME (1959) 216, pp. 446-448.
16. Craig, F.F., Jr., Geffen, T.M. and Morse, R.A.: "Oil Recovery Performance of Pattern
Gas or Water Injection Operations from Model Tests," Trans., AIME (1955) 204, pp. 715.
17. Hendrickson, G.E.: "History of the Welch Field San Andres Pilot Waterflood," Journal
of Petroleum Technology (August 1961) pp. 745-749.
18. Wasson, J.A. and Schrider, L.A.: "Combination Method for Predicting Waterflood
Performance for Five-Spot Patterns in Stratified Reservoirs," Journal of Petroleum
Technology (October 1968) pp. 1195-1202.
19. Buckley, S.E. and Leverett, M.C.: "Mechanism of Fluid Displacement in Sands,"
Trans., AIME (1942) 146, pp. 107-116.
20. Mattax, C.C. and Dalton, R.L.:
Richardson, TX (1990) 13.

Reservoir Simulation, Monograph Series, SPE,

8 - 22

APPENDIX
CRAIG-GEFFEN-MORSE METHOD
I. Introduction
The Craig-Geffen-Morse1,2,3 method of waterflood prediction is a steady state technique
which combines areal sweep effects, displacement mechanism, stratification, and
variable injectivity4 to predict waterflood performance in a five-spot pattern.

The

method is valid with or without free gas initially present, provided there is no trapped
gas behind the front. The calculations can be adopted for use in other pattern floods but
do not account for edge or bottom water influx. The method assumes 100 percent
vertical sweep efficiency within each layer of the stratified reservoir and cross-flow
between layers is negligible.. Experimentally derived correlations are used to determine
areal sweep efficiency at breakthrough and after breakthrough.
Calculations are made in four stages:
w Stage 1 - This stage begins with the start of water injection and ends when oil
banks formed around adjacent injectors meet. This meeting of oil banks is termed
interference. If there is no free gas present at the start of the flood, skip Stage 1
and Stage 2 and go directly to Stage 3. Oil production during this time period is
simply a continuation of previously existing primary production. No secondary
oil is recovered during this part of the flood.
w Stage 2 - This period extends from interference until all pre-existing gas space is
filled by injected water. Only primary oil production occurs during this stage.
w Stage 3 - This period extends from gas fillup to water breakthrough at producing
wells. Oil production caused by the waterflood begins at the start of Stage 3.
Furthermore, oil production during this stage is a combination of incremental
waterflood recovery and a continuation of primary recovery but the total oil
production rate is equal to the effective injection rate measured at reservoir
conditions. Injected water production begins at the end of Stage 3.
w Stage 4 - This stage extends from water breakthrough to the economic limit.
CGM-1

Stages 1, 2, and 3 are illustrated in Figure CGM-1.

FIGURE CGM-1
Stage 1

Interference
Between Oil Banks

Stage 2

Stage 3

Water Bank

CGM-2

Gas Region

Oil Bank

We will show first how waterflood predictions are made for a five-spot pattern reservoir
with only one layer. Extended calculations for multi-layered five-spot reservoirs will be
presented in a subsequent section.
II. Initial Calculations - Single Layer
Before considering the detailed procedures necessary to predict flood performance
during each of the four stages, it is convenient to present the following calculations.
A. Calculate pattern pore volume, V p

V p 7758 Ah

(Eq. CGM.1)

where:

Vp

= pore volume, bbls

= reservoir area, acres

= average net thickness, feet

= average porosity, fraction

B. Calculate stock tank oil-in-place at the beginning of the waterflood,

No

V p So

(Eq. CGM.2)

Bo

where:

No =

No

oil-in-place at start of flood, STB

So

= oil saturation at start of flood, fraction

Bo

= oil formation volume factor at start of flood, RB/STB

CGM-3

C. Calculate mobility ratio,

M , prior to water breakthrough using Eq. CGM.3 and

fractional flow data.

(krw ) S

wbt

( kro ) S

wc

o
w

(Eq. CGM.3)

where:

krw

the relative permeability to water evaluated at the average water


saturation in the water swept region,
fraction

kro

S wbt , at waterbreakthrough,

the relative permeability to oil at the connate water saturation, S wc


at the start of waterflooding, fraction

Water viscosity can be estimated from Figure CGM-2.

CGM-4

FIGURE CGM-2
EFFECT OF TEMPERATURE ON
VISCOSITY OF SALT WATER
2.0

Viscosity, centipois

1.5

250,000 ppm
200,000 ppm
150,000 ppm
100,000 ppm
50,000 ppm
0 ppm
1.0

0.5

0.0
30

40

50

60

70

80

90

100 110 120 130 140 150 160 170

Temperature, degrees Fahrenheit

D. Determine sweep efficiency at water breakthrough,

E Abt

from Step C and the correlation shown in Figure CGM-3.


CGM-5

using the mobility ratio

Areal Efficiency at Breakthrough, percent

FIGURE CGM-3
AREAL SWEEP EFFICIENCY AT BREAKTHROUGH
(DEVELOPED FIVE-SPOT PATTERN)
100
90
80
70
60
50
40
0.1

10

Mobility Ratio

E Abt 0.5460 .0.0317 / M 0.3022 / e M 0.0051M


E. Calculate cumulative water injected at the time of interference.

Wii

rei2 h S g

(Eq. CGM.4)

5.615

where:

Wii

cumulative water injected at interference, bbls

rei

half the distance between adjacent injectors, feet

F. Calculate cumulative water injected at gas fillup.

Wif V p S g

(Eq. CGM.5)

where:
CGM-6

Wif

cumulative water injected at gas fillup, bbls

Sg

gas saturation at start of flood, fraction

G. Calculate cumulative water injected at the time of water breakthrough.

Wibt V p E Abt S wbt S wc

(Eq. CGM.6)

where:

Wibt

cumulative water injected at breakthrough, bbls

S wbt =

average water saturation in swept region at breakthrough, fraction

S wc

connate water saturation at start of flood, fraction

III. STAGE 1: PERFORMANCE PRIOR TO INTERFERENCE


It is assumed during this period that the water and oil banks are radial in shape and that
Darcy's radial flow equation can be used to predict water injection into the reservoir.
Consider the injection wells depicted by Figure CGM-4.
FIGURE CGM-4
RADIAL WATER AND OIL BANKS
ASSOCIATED WITH INJECTION WELLS DURING STAGE 1

re

OIL

OIL

WATER

WATER

rei

CGM-7

For a constant pressure differential ( p ) , the water injection rate prior to interference
will be:

iw

0.00708khp
w r o re
ln
ln
krw rw kro r

(Eq. CGM.7)

where:

iw

water injection rate, bbls/day

net pay, feet

base permeability used to define relative permeability, md [usually


the effective permeability to oil at irreducible water, ( ko )
,
S wir

md]

kro

relative permeability to oil in oil bank at

krw

relative permeability to water in water bank at

radius of water bank, feet

re

radius of oil bank, feet

rw

effective wellbore radius, feet = rwe

rw

wellbore radius, feet

si

skin factor at injection well, dimensionless

applied pressure differential, psi (difference between pressure at


formation face of injection well and pressure in reservoir at the
outer edge of oil bank, usually assumed as the average reservoir
pressure at start of injection)

oil viscosity, cp

CGM-8

S wc
S wbt

si

water viscosity, cp

The radii of the water and oil banks required by Eq. CGM.7 depend upon the
cumulative water injection,

Wi .

Since all water injected during Stage 1 effectively

fills the gas space in the region from rw to re we can write:

re2 h Sg 5.615Wi
re2

(Eq. CGM.8)

5.615Wi
h Sg

(Eq. CGM.9)

1
5.615Wi 2

re

h Sg

(Eq. CGM.10)

where:

Wi

cumulative water injected, bbls

All of the water injected will be within the water bank of radius,
water saturation in the water bank is

S wbt

Since the average

we can write:

r 2h S wbt S wc re2h Sg

(Eq. CGM.11)

Sg
r 2 re2

S
wbt
wc

(Eq. CGM.12)

1
2

Sg
r re

S wbt S wc

(Eq. CGM.13)

CGM-9

Summary - Stage 1 Calculations


1.

Select values of Wi from zero toWii . There are no rigorous rules for making this
selection; generally ten intervals of equal

Wi

increments will be adequate.

2.

Compute re for each value of Wi using Eq. CGM.10.

3.

Compute

4.

Compute iw for each value of Wi using Eq. CGM.7.

5.

Compute the average water injection rate for each increment of water injection.

for each value of Wi using Eq. CGM.13.

[(iw ) avg ]

(iw ) n (iw ) n 1
2

6.

(Eq. CGM.14)

Compute the time required for each increment of water injection.

t n

Wi n Wi n1
iw
avg

(Eq. CGM.15)

7.

Compute cumulative time for each value of Wi

tn t n

(Eq. CGM.16)

A summary of calculations required in Stage 1 is summarized by Table CGM-1.

CGM-10

TABLE CGM-1
SUMMARY OF STAGE 1 CALCULATIONS
(1)

(2)

(3)

(4)

(5)

Wi , bbls

re2

re ,feet

r ,feet

( w / krw )ln( r / rw )

W1

re2

re

r1

rei 2

rei

W2

Wii
(6)

(7)

(8)

(9)

( o / kro )ln(re / r )

(5) + (6)

iw , B/D

(iw ) avg [(iw ) n (iw ) n 1 ]/ 2

iwi
(10)

(11)

t [Wi /(iw ) avg ],days

t t ,days

ti

CGM-11

IV. STAGE 2: PERFORMANCE FROM INTERFERENCE TO FILLUP


At the time of interference, the shape of the oil bank and water bank is radial; however,
from interference to fillup, the shape of the oil bank must continuously change as the
remaining gas space within the five-spot pattern is filled. Because of the changing
geometry of the banks during this period, it is not possible to write a simple equation to
predict water injection behavior. Fortunately, the length of Stage 2 is generally short
compared to other stages; accordingly, we are going to compute the water injection
rates at the end of Stage 1 and the beginning of Stage 3 and assume that iw changes
linearly between these two values. Therefore, the time differential between interference
and fillup is:

t
Values of

Wif

Wif Wii

(Eq. CGM.17)

0.5(iwi iwf )
and

Wii

are known from initial calculations. The water injection rate

at interference, iwi , corresponds to the rate at the end of Stage 1.


The injection rate at fillup, iwf , as well as injection rates from fillup to water
breakthrough, and to the end of the economic life of the project, can be calculated as3,4:

iw ibase

(Eq. CGM.18)

where:

ibase =

conductance ratio
base water injection rate, bbls

For a five-spot pattern, ibase is defined as:

CGM-12

ibase

0.003541 ko S

wir

hp

o ln 0.619 0.5s p 0.5si


rw

(Eq.CGM.19)

where:

ibase

= base water injection rate (steady-state water injection rate in an


oil-filled, five-spot pattern with a unit mobility ratio), bbls/day

d
( ko ) S

= diagonal distance between adjacent injection and production


wells, feet
wir

= effective permeability to oil at immobile connate water


saturation, md.

sp

= skin factor at producing well, dimensionless

si

= skin factor at injection well, dimensionless

= bottom-hole pressure difference between the injection and


production well after fillup, psi

The conductance ratio,

, is an experimentally determined factor based on the work of

Caudle and Witte4 which, when used in Eq. CGM.18, gives the correct injection rate.
The conductance ratio is presented in Figure CGM-5 as a function of mobility ratio,

M , and areal sweep efficiency, E A


1.0 and iw is a constant. For
efficiency. When

< 1.0,

Note in Figure CGM-5 that for


> 1.0,

and iw

= 1.0,

increase with increasing sweep

and iw decrease with increasing sweep efficiency.

CGM-13

CGM-14

0.1

10

0.1

0.7
0.9

0.5 0.3

1.0

0.1

EA

Mobility Ratio

EA
1.0
0.9
0.7
0.5 0.3
0.1
0

FIGURE CGM-5
CONDUCTANCE RATIO FOR LIQUID FILLED FIVE-SPOT
PATTERNS (REFERENCE 4)

10

CR 1/(1 E A ((1/ M ) 1))(Ref. Dr. Ben Caudle - Private Communication)

Conductance Ratio

EA

Wi

(Eq. CGM.20)

V p ( S wbt S wc )

Summary - Stage 2 Calculations


1. Obtain values of Wif and Wii from initial calculations.
2. Obtain value of iwi from Stage 1 calculations where Wi
3. Compute

Wii

E A at fillup using Eq. CGM.20.

4. Obtain the mobility ratio,


5. Determine

M , from Step C of the initial calculations.

at fillup from Figure CGM-5.

6. Compute ibase using Eq. CGM.19.


7. Compute water injection rate at fillup,

iwf using Eq. CGM.18.

8. Compute time interval required for Stage 2 using Eq. CGM.17.


V. STAGE 3: PERFORMANCE FROM FILLUP TO BREAKTHROUGH
The end of the gas fillup period marks the beginning of secondary oil production. It is
assumed that, on a reservoir volume basis, the total oil producing rate during this stage is
equal to the water injection rate. The water injection rate can be determined using Eq.
CGM.18. Thus, the oil producing rate in STB/D is:

qo
where

fo

iw
fo Swc
Bo

(Eq. CGM.21)

is the oil cut at the producing well before water breakthrough.

S wc S wir , then f o 1.0 .


CGM-15

If

Cumulative oil production,

N p,

since the beginning of Stage 3 (fillup) can be

computed in terms of cumulative water injected during Stage 3 as:

Np

Wi Wif
Bo

fo

(Eq. CGM.22)

Summary - Stage 3 Calculations


1. Select values of Wi from Wif to Wibt using a convenient interval.
2. Determine

E A for each value of Wi

3. Determine

using Eq. CGM.20.

for each value of Wi using Figure CGM-5.

4. Compute iw using Eq. CGM.18.


5. Compute average value of iw for each interval.
6. Compute incremental and cumulative times associated with each interval.
7. Compute

qo using Eq. CGM.21.

8. Determine

fo

at

S wc

from the fractional flow graph.

9. Compute cumulative oil recovery using Eq. CGM.22.


A summary of these calculations is tabulated in Table CGM-2.

CGM-16

TABLE CGM-2
SUMMARY OF STAGE 3 CALCULATIONS
(1)

(2)

(3)

Wi , bbls

(Eq. CGM.21)

(4)

EA

(Figure CGM-5)

iw

(Eq. CGM.19)

Wif

EAbt

Wibt

iwbt

(5)

(6)

(7)

(iw )avg [(iw ) n (iw ) n 1 ]/ 2

t Wi /(iw ) avg

t t

tbt
(8)

4
qo
f o ,STBD
B
o

(9)

Wi Wif (1) Wif

(10)

Np

9 fo ,STB
Bo

N pbt

CGM-17

VI. STAGE 4: PERFORMANCE AFTER WATER BREAKTHROUGH


This stage, which marks the beginning of injected water production, is characterized by
an increasing mobility ratio, increasing areal sweep efficiency, increasing water-oil
ratio, and decreasing oil producing rate. The producing water-oil ratio is governed by
the amount of oil and water flowing from the previously swept region of the reservoir
plus the amount of oil displaced as the swept area increases. Oil and water production
from the previously swept region is governed by fractional flow data and can be
predicted using previously developed methods. Oil displaced from the newly swept
portion of the reservoir is assumed to be that displaced by the water saturation
immediately behind the stabilized zone,

S wf .

Considering a given time interval, incremental oil produced from the previously
unswept portion of the reservoir,
areal sweep efficiency,

N pu will depend upon the incremental increase in

E A the change in water saturation in the newly swept area

( S wf S wc ) and the pore volume,V p .

N pu E A S wf S wc V p
The term

E A / Wi / Wibt

(Eq. CGM.23)

is a convenient term to include in these calculations.

Accordingly, multiplying both sides of Eq. CGM.23 by this term results in:

N pu

Wi
E A
S wf S wc V p

Wi / Wibt
Wibt

(Eq. CGM.24)

and

Wi
N pu S wf S wc V p

W
ibt

CGM-18

(Eq. CGM.25)

where:

E A
Wi / Wibt

(Eq. CGM.26)

These calculations can be put on the basis of one barrel of total injection (or production,
since injection and production rates are assumed equal at reservoir conditions) by
setting

Wi 1.0
N pu

Thus,

V p S wf S wc

(Eq. CGM.27)

Wibt

The water injected at breakthrough is:

Wibt V p E Abt S wbt S wc

(Eq. CGM.28)

Substitution of Eq. CGM.28 into Eq. CGM.27 yields the result:

N pu

S wf S wc

E Abt S wbt S wc

The oil produced from the previously unswept area,

Wi

N pu during

the time that

barrels of water are injected can be predicted using Eq. CGM.29 provided that

is known. However, notice that

efficiency,

depends upon the increase in areal sweep

E A which occurs as a result of injecting Wi

Geffen and Morse1 found experimentally that


of Wi

(Eq. CGM.29)

/ Wibt .

barrels of water. Craig,

E A increases linearly with the logarithm

This relationship was shown previously in Chapter 5 and the equation is

repeated as follows:

CGM-19

E A 0.2749 ln Wi / Wibt E Abt

(Eq. CGM.30)

It follows that:

E A dE A 0.2749

Wi
Wi dWi

(Eq. CGM.31)

and:

W
0.2749 i
Wibt

1
(Eq. CGM.32)

Thus,

N pu 0.2749

S wf S wc

E Abt S wbt

Wi

S wc Wibt

Incremental oil from the previously swept region,

N ps

1
(Eq. CGM.33)

based upon one barrel of

total production is:

N ps f o 2 1 N pu

where:

fo 2

= fraction of oil in the producing stream = 1.0 f w2

f w2 =

fraction of water in the producing stream

CGM-20

(Eq. CGM.34)

Since

N pu

is known (Eq. CGM.33), it is obvious from Eq. CGM.34 that

can be determined if

fo 2 can be defined.

N ps

The big question at this point is -- how can

fo 2 (or f w2 ) be determined at any time after breakthrough?


It will be recalled from frontal advance theory that
fractional flow curve (Figure CGM-6) if
known. Unfortunately we do not know

S w2
S w2

f w2

can be determined from the

the saturation at the producing well, is


however, we do know that

S w2

is the

tangent point on the fractional flow curve defined by a tangent line of slope

df w
1

dS w Sw2 Qi Sw2

(Eq. CGM.35)

where:

Qi
If

Qi

= pore volumes of water injected at the time in question.

were known, it would be possible to compute the slope of the tangent line using

Eq. CGM.35.

S w2 and f w2

could then be determined from the fractional flow curve

(See Figure CGM-6).

CGM-21

FIGURE CGM-6
FRACTIONAL FLOW CURVE DEMONSTRATING
USE OF Qi TO DETERMINE f w2.
1

1 Sor

f w2

fw

fw

df w
dS w

df w
dS w

0
0

S wc

df w
1

dS w Qi

Sw
CGM-22

S w2

The water injected at breakthrough, Wibt was defined by Eq. CGM.28 to be:

Wibt V p E Abt S wbt S wc

(Eq. CGM.28)

When expressed in water-contacted pore volumes, this becomes:

Qibt

Wibt
S wbt S wc
E AbtV p

(Eq. CGM.36)

The cumulative water injected at any time beyond breakthrough is equal to the water
injected at breakthrough plus the additional water injected beyond breakthrough.

Wi Wibt W beyond

(Eq. CGM.37)

breakthrough

Expressed in terms of water-contacted pore volumes:

Qi Qibt Qbeyond

(Eq. CGM.38)

breakthrough

The sweep efficiency after breakthrough increases as

Wi

increases. Accordingly, the

water-contacted pore volume is also a function of Wi Therefore,

Qbeyond

breakthrough

dWi
ibt V E
p A

W i

(Eq. CGM.39)

or changing the limits of integration:

Qbeyond

breakthrough

Wibt Wi / Wibt d WiWibt

V p 1.0
EA

From Eq. CGM.36:

CGM-23

(Eq. CGM.40)

Wibt
Qibt E Abt
Vp

(Eq. CGM.41)

and:

Qbeyond

breakthrough

W / Wibt

Qibt E Abt 1.0i

d Wi / Wibt
EA
(Eq. CGM.42)

Substituting Eq. CGM 42 into Eq. CGM.38:

Qi
W / W d Wi / Wibt
1.0 E Abt 1.0i ibt
Qibt
EA

(Eq. CGM.43)

Finally, from Eq. CGM.30:

E A 0.2749 ln Wi / Wibt E Abt

(Eq. CGM.30)

and:

d Wi / Wibt
Qi
W /W
1.0 E Abt 1.0i ibt
Qibt
E Abt 0.2749 ln Wi / Wibt
(Eq. CGM.44)
A tabular solution of Eq. CGM.44 is presented in Table CGM-3 (or as found in Table
E.9, SPE Monograph III) as a function of

E Abt and Wi / Wibt .

CGM-24

TABLE CGM-3
Values of Qi / Qibt for various Values of Breakthrough Areal Sweep Efficiency
E Abt , percent

Wi / Wibt
1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
3.6
3.8
4.0
4.2
4.4
4.6
4.8
5.0
5.2
5.4
5.6
5.8
6.0
6.2

Values of

50

51

52

53

54

55

56

57

58

59

1.000
1.190
1.365
1.529
1.684
1.832
1.974
2.111
2.244
2.373
2.500
2.623
2.744
2.862
2.978
3.093
3.205
3.316
3.426
3.534
3.641
3.746
3.851
3.954
4.056
4.157
4.257

1.000
1.191
1.366
1.530
1.686
1.834
1.977
2.115
2.249
2.379
2.507
2.631
2.752
2.872
2.989
3.105
3.218
3.330
3.441
3.550
3.657
3.764
3.869
3.973
4.077
4.179

1.000
1.191
1.366
1.531
1.688
1.837
1.981
2.119
2.254
2.385
2.513
2.639
2.761
2.881
3.000
3.116
3.231
3.343
3.455
3.565
3.674
3.781
3.887
3.993
4.097

1.000
1.191
1.367
1.532
1.689
1.839
1.984
2.124
2.259
2.391
2.520
2.646
2.770
2.891
3.010
3.127
3.243
3.357
3.469
3.580
3.689
3.798
3.905
4.011

1.000
1.191
1.368
1.533
1.691
1.842
1.987
2.127
2.264
2.397
2.526
2.653
2.778
2.900
3.020
3.138
3.254
3.369
3.483
3.594
3.705
3.814
3.922

1.000
1.191
1.368
1.535
1.693
1.844
1.990
2.131
2.268
2.402
2.533
2.660
2.786
2.909
3.030
3.149
3.266
3.382
3.496
3.609
3.720
3.830

1.000
1.191
1.369
1.536
1.694
1.846
1.993
2.135
2.273
2.407
2.539
2.667
2.793
2.917
3.039
3.159
3.277
3.394
3.509
3.622
3.735

1.000
1.191
1.369
1.536
1.696
1.849
1.996
2.139
2.277
2.413
2.545
2.674
2.801
2.926
3.048
3.169
3.288
3.406
3.521
3.636

1.000
1.192
1.370
1.537
1.697
1.851
1.999
2.142
2.282
2.418
2.551
2.681
2.808
2.934
3.057
3.179
3.299
3.417
3.534
3.649

1.000
1.192
1.370
1.538
1.699
1.853
2.001
2.146
2.286
2.422
2.556
2.687
2.816
2.942
3.066
3.189
3.309
3.428
3.546

4.608

4.443

Wi / Wibt
6.164

at which

5.944

E A = 100 percent

5.732

5.527

5.330

CGM-25

5.139

4.956

4.779

TABLE CGM-3 (continued)


Values of Qi / Qibt for various Values of Breakthrough Areal Sweep Efficiency
E Abt , percent

Wi / Wibt

60

61

62

63

64

65

66

67

68

69

1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4
2.6
2.8
3.0
3.2
3.4
3.6
3.8
4.0
4.2
4.4

1.000
1.192
1.371
1.539
1.700
1.855
2.004
2.149
2.290
2.427
2.562
2.693
2.823
2.950
3.075
3.198
3.319
3.439

1.000
1.192
1.371
1.540
1.702
1.857
2.007
2.152
2.294
2.432
2.567
2.700
2.830
2.957
3.083
3.207
3.329

1.000
1.192
1.371
1.541
1.703
1.859
2.009
2.155
2.298
2.436
2.572
2.705
2.836
2.965
3.091
3.216

1.000
1.192
1.372
1.542
1.704
1.861
2.012
2.158
2.301
2.441
2.577
2.711
2.843
2.972
3.099
3.225

1.000
1.192
1.372
1.543
1.706
1.862
2.014
2.161
2.305
2.445
2.582
2.717
2.849
2.979
3.107

1.000
1.192
1.373
1.543
1.707
1.864
2.016
2.164
2.308
2.449
2.587
2.723
2.855
2.986

1.000
1.193
1.373
1.544
1.708
1.866
2.019
2.167
2.312
2.453
2.592
2.728
2.862
2.993

1.000
1.193
1.373
1.545
1.709
1.868
2.021
2.170
2.315
2.457
2.597
2.733
2.867

1.000
1.193
1.374
1.546
1.710
1.869
2.023
2.173
2.319
2.461
2.601
2.738
2.873

1.000
1.193
1.374
1.546
1.711
1.871
2.025
2.175
2.322
2.465
2.606
2.744

3.203

3.088

Values of

Wi / Wibt
4.285

at which

4.132

E A = 100 percent

3.984

3.842

3.704

CGM-26

3.572

3.444

3.321

TABLE CGM-3 (continued)


Values of Qi / Qibt for various Values of Breakthrough Areal Sweep Efficiency
E Abt , percent

Wi / Wibt

70

71

72

73

74

75

76

77

78

79

1.0
1.2
1.4
1.6
1.8
2.0
2.2
2.4

1.000
1.193
1.374
1.547
1.713
1.872
2.027
2.178

1.000
1.193
1.375
1.548
1.714
1.874
2.029
2.180

1.000
1.193
1.375
1.548
1.715
1.875
2.031
2.183

1.000
1.193
1.375
1.549
1.716
1.877
2.033
2.185

1.000
1.193
1.376
1.550
1.717
1.878
2.035
2.188

1.000
1.193
1.376
1.550
1.718
1.880
2.037
2.190

1.000
1.193
1.376
1.551
1.719
1.881
2.039
2.192

1.000
1.194
1.377
1.551
1.720
1.882
2.040
2.195

1.000
1.194
1.377
1.552
1.720
1.884
2.042
2.197

1.000
1.194
1.377
1.552
1.721
1.885
2.044

2.6

2.325

2.328

2.331

2.334

2.337

2.340

2.8

2.469

2.473

2.476

2.480

3.0

2.610

2.614

2.394

2.309

2.226

2.147

Values of

Wi / Wibt
2.978

at which
2.872

E A = 100 percent
2.769

2.670

2.575

2.483

TABLE CGM-3 (continued)


Values of Qi / Qibt for various Values of Breakthrough Areal Sweep Efficiency
E Abt , percent

Wi / Wibt
1.2
1.4
1.6
1.8
2.0
2.2

Values of

80

81

82

83

84

85

86

87

88

89

1.000
1.194
1.377
1.553
1.722
1.886
2.045

1.000
1.194
1.378
1.553
1.723
1.887

1.000
1.194
1.378
1.554
1.724
1.888

1.000
1.194
1.378
1.555
1.725
1.890

1.000
1.194
1.378
1.555
1.725

1.000
1.194
1.379
1.555
1.726

1.000
1.194
1.379
1.556
1.727

1.000
1.194
1.379
1.556
1.728

1.000
1.194
1.379
1.557

1.000
1.194
1.379
1.557

Wi / Wibt
2.070

at which

1.996

E A = 100 percent

1.925

1.856

1.790

1.726

CGM-27

1.664

1.605

1.547

1.492

TABLE CGM-3 (continued)


Values of Qi / Qibt for various Values of Breakthrough Areal Sweep Efficiency
E Abt , percent

Wi / Wibt

90

91

92

93

94

95

96

97

98

99

1.0
1.2
1.4
1.6

1.000
1.194
1.380
1.558

1.000
1.195
1.380

1.000
1.195
1.380

1.000
1.195
1.380

1.000
1.195
1.381

1.000
1.195

1.000
1.195

1.000
1.195

1.000
1.195

1.000
1.195

1.115 1.075

1.037

Values of

Wi / Wibt

at which

1.439

Once

Qi / Qibt

1.387

E A = 100 percent
1.338

1.290

1.244

1.199

is determined from this table,

Qi can be calculated and used along

f w2

with the fractional flow curve to define

1.157

Finally,

fo 2 1.0 f w2

and

N ps can be computed using Eq. CGM.34.


The incremental water produced on a one barrel basis,

W ps 1 N ps N pu

W ps is computed as:

(Eq. CGM.45)

The water-oil ratio at reservoir pressure, WOR p is defined as:

WOR p

1 N ps N pu
N ps N pu

(Eq. CGM.46)

The water-oil ratio at surface conditions, WOR , is:

WOR WOR p

Bo
Bw

(Eq. CGM.47)

The oil producing rate in STB/D can be determined as:

CGM-28

qo

iw N ps N pu

(Eq. CGM.48)

Bo

The water producing rate in STB/D is:

qw

iw 1 N ps N pu

(Eq. CGM.49)

Bw

A material balance of the OIP at the start of the waterflood, which neglects the primary
production during the fillup period, leads to the cumulative oil production after fillup,

N p , in STB to be:
Np

V p E A S w S wc S g
Bo

, STB

(Eq. CGM.50)

where:

Sw =

average water saturation in the reservoir at the time of interest

S w S w2 Qi fo 2

(Eq. CGM.51)

Equation CGM.51 is the Buckley-Leverett equation in Chapter 4 whose graphical


interpretation resulted in the tangent line construction to the fractional flow curve to
obtain

S w after water breakthrough.

An alternate means of computing cumulative oil recovery since the start of


waterflooding is with the aid of the following equations:
Incremental waterflood oil production during Stage 4 ,
CGM-29

N p , is:

qo N qo N 1
N p
t
2

and cumulative oil production since the start of waterflooding,

N p , at any time step,

during Stage 4 is:

end of stage 3 N p Stage 4

Np Np

The cumulative water produced, W p in STB can be computed as:

Wp

Wi N p Bo V p S g

(Eq. CGM.52)

Bw

Summary - Stage 4 Calculations


1. Select values of

Wi

from

Wibt

to the economic limit and express as a ratio of

Wi / Wibt .
2. Compute

E A using Eq. CGM.30 for each value of Wi .

3. Determine values of

Qi / Qibt

from Table CGM-3, and compute:

Qi Qibt Qi / Qibt

Qi S wbt S wc Qi / Qibt .
4. Compute the slope of the fractional flow curve,

CGM-30

df w / dS w , using Eq. CGM.35.

5. Use the slope from Step 4 and the fractional flow curve to determine

S w2 .

See

Figure CGM-6 for an illustration of this method.


6. Using

S w2 ,

determine

f w2

from

the

fractional

flow

curve;

then,

fo 2 1.0 f w2 .
7. Compute

S w using Eq. CGM.51.

8. Compute

9. Compute

N pu

10. Compute

using Eq. CGM.32.


using Eq. CGM.29.

N ps using Eq. CGM.34.

11. Compute WOR using Eq. CGM.47.


12. Compute

N p using Eq. CGM.50.

13. Determine the mobility ratio,

M,

(the CGM assumes

increases after water

breakthrough) according to the relationship:

krw S w o
M
kro Swc w
14. Determine

(Eq. CGM.53)

from Figure CGM-5.

15. Compute iw using Eq. CGM.18.


16. Compute incremental and cumulative times associated with each interval.
17. Compute

qo using Eq. CGM.48.

18. Compute

qw using Eq. CGM.49.


CGM-31

19. Compute W p using Eq. CGM.52.


A summary of Stage 4 calculations is presented in Table CGM-4.

CGM-32

Eq. CGM.51

(6)

WOR p

1.0 / (5)

df w / dS w

(13)

(4) x Qibt

Qi

(5)

WOR

(14)

S w2

(7)

fo 2

(8)

Eq. CGM.32 Eq. CGM.29 Eq. CGM.35 Eq. CGM.46 Eq. CGM.47

N ps

N pu

Sw

1.0

(12)

E Abt

Table CGM-3

(11)

1.0

Wibt

Eq. CGM.30

Qi / Qibt

EA

(10)

Wi / Wibt

Wi , bbls

(4)

(3)

(9)

(2)

TABLE CGM-4
SUMMARY OF STAGE 4 CALCULATIONS

(1)

CGM-33

CGM-34

Eq. CGM.53

Eq. CGM.50

Wi

t (20) /(19)

(21)

Np

(20)

(16)

(15)

t t

(22)

qo

(23)

(19)

qw

(24)

Eq. CGM.52

Wp

(25)

(i w ) avg iw n iw n1 / 2

Eq. CGM.49

Eq. CGM.18

iw

(18)

Eq. CGM.48

Figure CGM-5

(17)

TABLE CGM-4 (continued)


SUMMARY OF STAGE 4 CALCULATIONS

VII. Multi-Layer Performance


All of the preceding calculations apply to a reservoir with only one layer. These
predictions can be extended to include other layers if the following assumptions are
made.

No crossflow occurs between layers.

The permeability, thickness, and porosity of the layers can vary. However, the
saturations of oil, gas, and water are assumed to be the same in all layers.

Relative permeability data are the same for all layers.

Injection and production rates associated with each layer are proportional to

kh .
Suppose we predict the performance of Layer 1 using the previously described
calculations. The time required to inject the same number of pore volumes of water
into Layer

n in time, tn , as was injected into Layer 1 during time, t1 , will be:

/ k n
tn t1
/ k 1
If values of

(Eq. CGM.54)

N p1, W p1, Wi1, iw1 , qo1, and qw1 were predicted at time t1 in Layer

1, then at time tn in Layer

n:

h n
N pn N p1
h 1

(Eq. CGM.55)

CGM-35

h n
W pn W p1
h 1
Win Wi1

iwn iw1

(Eq. CGM.56)

h n
h 1

(Eq. CGM.57)

kh n
kh 1

(Eq. CGM.58)

kh n
qon qo1
kh 1

(Eq. CGM.59)

kh n
qwn qw1
kh 1

(Eq. CGM.60)

Procedure
1. Predict the performance of Layer 1 using the method previously described.
2. Plot

N p , W p , Wi , iw , qo , and qw versus time for Layer 1.

3. Obtain values of

/ k , h , and kh for all layers.

4. For a succession of values of


1), determine

(call these values t1 since we are working with Layer

N p1, W p1, Wi1 , qo1,

and

qw1

for Layer 1 from the plots

constructed in Step 2. Table CGM-5 illustrates these results.

CGM-36

TABLE CGM-5
SUMMARY OF LAYER 1 PREDICTIONS
t1

N p1

(t1 )1

( N p1 )1

(t1 ) 2
(t1 )3

W p1

Wi1

iw1

qo1

qw1

(W p1 )1

(Wi1 )1

(iw1 )1

(qo1 )1

(qw1 )1

( N p1 ) 2

(W p1 )2

(Wi1 ) 2

(iw1 ) 2

(qo1 )2

(qw1 )2

( N p1 )3

(W p1 )3

(Wi1 )3

(iw1 )3

(qo1 )3

(qw1 )3

5. Now consider the remaining layers in the reservoir, i.e., consider layer

Corresponding to the times, t1 selected in Step 4 for Layer 1, use Eq. CGM.54 to
compute the times, tn for Layer

required for the same number of pore volumes of

water to be injected into the two layers. These calculations are illustrated by Table
CGM-6.

CGM-37

TABLE CGM-6
EQUIVALENT TIME FOR LAYER n
BASED ON LAYER 1 PREDICTIONS

Layer 1
t1

Layer n
t n (Eq. CGM.55)

(t1)1

(tn )1

(t1)2

(tn )2

(t1)3

(tn )3

6. For each of the values of tn computed in Step 5 for Layer

N pn , W pn , Win , iwn , qon ,

and

qwn

n , compute values of

using Eqs. CGM.55 through CGM.60,

respectively. Repeat these calculations for all remaining layers.


7. Plot

N p , W p , Wi , iw , qo , and qw versus time for all layers.

8. Composite performance at any time

can be obtained from the plots in Step 7 by

summing the performance of individual layers at that time.


Example calculations illustrating the use of the Craig-Geffen-Morse method are presented by
Craig2 and Benton, et al5.

CGM-38

REFERENCES

1. Craig, F.F., Jr., Geffen, T.M. and Morse, R.A., "Oil Recovery Performance of Pattern
Gas or Water Injection Operations from Model Tests," Trans, AIME (1955) 204, pp.
7-15.
2. Craig, F.F., Jr.: The Reservoir Engineering Aspects of Waterflooding, Monograph
Series, SPE, Dallas (1971) 3.
3. Willhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
4. Caudle, B.H. and Witte, M.D.: "Production Potential Changes during Sweep-Out in a
Five-Spot System," Trans., AIME (1959) 216, pp. 446-448.
5. Benton, J.P., Maslanka, P.M. and Smith, R.L.: "Early Implementation of a Full-Scale
Waterflood in the Abo Reef, Terry Co., Texas - A Case History," paper SPE 9475,
presented at the 1980 Annual Technical Conference and Exhibition, Dallas, Sept. 2124.

CGM-39

PROBLEM CGM:1
Use the Craig-Geffen-Morse method to calculate the performance of the five-spot pattern
waterflood described below:
Pattern area,

A = 40 acres

Oil formation volume factor,

Bo = 1.056 RB/STB

Water formation volume factor,


Oil viscosity,

Bw

= 1.0 RB/STB

o = 0.853 cp

Water viscosity,

w = 0.375 cp

Injection pressure = 3200 psig


Average reservoir pressure at start of waterflood = 100 psig
Producing well pressure after the end of Stage 2 = 100 psig
Wellbore radius, rw = 0.5 ft
Injection well skin factor = 0
Production well skin factor = 0
Formation thickness,
Permeability,

= 1.5 ft

ko Swir

= 20 md

Porosity, = 0.16
Oil saturation at beginning of flood,

So = 0.70

Gas saturation at beginning of flood,

Sg

Water saturation at beginning of flood,

= 0.10

S wc

= 0.20

Oil production rate at beginning of flood = 1.0 BOPD

CGM - 40

Relative permeability data for the reservoir and calculations to determine the fractional
flow curve are summarized in Problem CGM:1 - Table 1; the fractional flow curve is
presented in Problem CGM:1 Figure 1. The derivative of the fractional flow curve is
presented in Problem CGM:1 Figure 2.
Calculate time, cumulative water injected and cumulative produced fluids at:
A. Interference
B. Fillup
C. Water breakthrough
D. Economic limit

qo 2.0 STB/D
Problem CGM:1 - Table 1: Relative
permeability data and fractional flow
calculations

Sw

kro

krw

kro w
krw o

fw

0.20
0.35
0.40
0.45
0.50
0.55
0.60
0.65
0.70

1.0000
0.4120
0.2720
0.1770
0.1090
0.0627
0.0317
0.0111
0.0000

0.0000
0.0678
0.1040
0.1300
0.1630
0.2030
0.2540
0.3180
0.3970

2.6714
1.1498
0.5986
0.2940
0.1358
0.0549
0.0153
0.0000

0.000
0.272
0.465
0.626
0.773
0.880
0.948
0.985
1.000

CGM - 41

PROBLEM CGM:1 FIGURE 1


FRACTIONAL FLOW CURVE
1.0

S wbt 0.585

0.9
0.8
0.7

S wf 0.472

0.6

fw

0.5
0.4
0.3
0.2
0.1
0.0
0.1

0.2

0.3

0.4

0.5

0.6

Water Saturation, fraction

CGM - 42

0.7

0.8

PROBLEM CGM:1 FIGURE 2


DERIVATIVE OF FRACTIONAL FLOW CURVE

3.0
2.8
2.6
2.4
2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
40 42 44 46 48 50 52 54 56 58 60 62 64 66 68 70
Water Saturation, fraction

CGM - 43

WORKSHEET: CRAIG-GEFFEN-MORSE METHOD


INITIAL CALCULATIONS

1. Pore Volume, V p

V p 7758 Ah
V p 7758

V p ___________ bbls
2. Stock Tank Oil at Start of Waterflood

No
No

V p So
Bo

No _________ STBs
3. Mobility Ratio Prior to Breakthrough

krw S wbt o
M
kro Swc w
S wbt _____________

krw S wbt

___________
CGM - 44

kro Swc
M

____________

_________ _________
_________ _________

M _____________

4. Sweep Efficiency at Breakthrough

E Abt _____________
5. Water Injected at Time of Interference, Wii

Wii

Wii

rei2 h S g
5.615

2
3.14 ________ ________ ________ ________

5.615

Wii _____________ bbls


6. Cumulative Water Injected at Gas Fillup, Wif

Wif V p S g
Wif _________ _________

Wif _____________ bbls


CGM - 45

7. Water Injected at Time of Water Breakthrough,

Wibt

Wibt V p E Abt S wbt S wc


Wibt _________ _________ _________

Wibt _____________ bbls

CGM - 46

STAGE 1: PERFORMANCE PRIOR TO INTERFERENCE


Table 1: Summary of Stage 1 Calculations
(2)
(3)
(4)
(5)

(1)

Wi , bbls

re2

re , ft

r , ft

krw

ln

r
rw

(6)

kro

ln

re
r

Wii

(7)
(5) + (6)

Table 1: Summary of Stage 1 Calculations


(8)
(9)
(10)

iw , bbl/D iw avg

(11)

t t , days
Wi
, days
iw avg

CGM - 47

STAGE 2: PERFORMANCE FROM INTERFERENCE TO FILLUP

1.

Wii __________________bbls

2.

Wif _________________bbls

3.

E A at Fillup

EA

EA

Wif

V p S wbt S wc

E A __________________
4. Mobility Ratio = _____________________
5. From Figure CGM-5,

6.

ibase

0.003541 ko S

o ln

ibase

___________
wir

hp

d
0.619 0.5s p 0.5si
rw

0.003541

ibase ____________bbls/day

CGM - 48

i
iwf base

iwf

iwf _____________bbls/day
7. Time Interval,

t , of Stage 2

Wif Wii

0.5 iwi iwf

t _____________days
8. Cumulative Time to Fillup

_______+_______=_______days

CGM - 49

STAGE 3: PERFORMANCE FROM FILLUP TO WATER


BREAKTHROUGH

(1)

Wi , bbls

Table 2: Summary of Stage 3 Calculations


(2)
(3)
(4)
E A (Eq. CGM.22) (Fig.CGM-5) iw (Eq.CGM.18)

(5)

iw avg

Wif

Wibt

(6)

Table 2: Summary of Stage 3 Calculations


(7)
(8)
(9)

(10)

t t ,days qo ,STB/D 1 Wif ,bbls N p ,STB


Wi
,days
iw avg
4 / Bo
9 / Bo

CGM - 50

STAGE 4: PERFORMANCE AFTER BREAKTHROUGH


Table 3: Summary of Stage 4 Calculations
(2)
(3)
(4)

(1)

Wi ,bbls

Wi / Wibt

EA

Qi / Qibt

(Eq. CGM.30)

(Table CGM-3)

(5)

Qi 4 Qibt

Wibt

Table 3: Summary of Stage 4 Calculations


(6)

(7)

df w / dS w S w 2
1.0/(5)

(8)

(9)

fo2

Sw

(10)

(11)

N pu

Eq.CGM.51 Eq.CGM.32 Eq.CGM.29

CGM - 51

(12)

N ps
Eq. CGM.34

Table 3: Summary of Stage 4 Calclulations


(13)

(14)

WOR

WOR p
Eq.CGM.46

(16)

Np

(17)

Eq.CGM.47 Eq.CGM.50 Eq.CGM.53 Fig.CGM-5

(19)

iw avg

(15)

Table 3: Summary of Stage 4 Calculations


(20)
(21)
(22)

iw n iw n1
2

Wi ,bbls

CGM - 52

20 t t
19

(18)

iw
Eq.CGM.18

(23)

qo ,STB/D
Eq. CGM.48

Table 3: Summary of Stage 4 Calculations


(24)
(25)

qw ,bbl/D

W p , bbl

Eq. CGM.49)

Eq. CGM.52)

CGM - 53

PROBLEM CGM:2

The table on the next page summarizes rate and cumulative volume data previously
calculated for a single layer five spot pattern described in Problem CGM:1. Listed below
are the rock properties for layer one. Extend the calculations to a second layer with the
properties given below. Assume that fluid saturations

S wc , So , and S g

water relative permeability data are identical for both layers.


PARAMETER

LAYER 1

LAYER 2

20.00

10.00

Thickness, ft.

1.50

1.00

Porosity, fraction

0.16

0.13

Permeability ko S

wir

, md

CGM - 65

and oil-

CRAIG-GEFFEN-MORSE WATERFLOOD PERFORMANCE LAYER 1


TIME,
DAYS
7
14
28
43
59
75
88
118
152
222
293
366
468
578
685
492
899
1,006
1,113
1,223
1,446

Wi ,BW N p , BO W p ,BW iw ,BWPD qo ,BOPD qw ,BWPD


500
1,000
2,000
3,000
4,000
5,000
5,846
7,450
9,000
12,000
15,000
18,000
22,085
26,500
30,900
35,400
39,800
44,200
48,600
53,000
61,900

0
0
0
0
0
0
0
0
1,460
4,309
7,150
9,991
13,859
16,375
18,124
19,740
21,223
22,442
23,816
24,410
24,551

0
0
0
0
0
0
0
0
0
0
0
0
0
1,758
4,311
7,105
9,938
12,946
16,000
19,773
28,524

CGM - 66

75
71
67
65
64
62
62
46
44
43
42
41
40
41
41
41
41
41
40
40
40

0
0
0
0
0
0
0
43
41
41
40
39
38
14
11
10
9
8
7
6
2

0
0
0
0
0
0
0
0
0
0
0
0
0
27
29
31
32
33
33
34
38

WATERFLOOD SURVEILLANCE

I.

Introduction
Engineering aspects of a waterflood do not end with the completion of an initial
engineering and geological report, an economic evaluation of project profitability, or
management approval of an AFE.

Engineering aspects of waterflooding are an

ongoing activity. A large number of technical papers have been presented in the
literature which describes waterflood prediction techniques.

Unfortunately, the

literature has been to a great extent silent on methods and procedures which the
operating engineer can use to monitor performance of actual reservoirs.

In the

beginning, the operations engineer usually has a rate and reserve forecast. Actual
performance usually does not agree with predicted performance. These differences
can frequently be attributable to the use of "average data" rather than data which is
specific to a particular geological layer or a particular area of the field. Forecasts also
differ from actual production performance due to the lack of accurate fluid saturations

So , Swc , S g , Sor , rock properties ( kro , krw , wettability), and geological


descriptions (stratification, permeability distribution,

V,

rock continuity). Further,

even if forecasts are made utilizing an accurate data base, the production forecasts can
differ from actual behavior due to operational considerations. It is normal for the
operation and timing of installations to be different from that projected.

Well

conversions are delayed (especially high production rate oil wells), drilling and rework of existing wells do not always meet schedules due to oil price changes,
corporate budget constraints, or changes in management practices. Nonetheless, if
forecasts are made using predictive models which incorporate reasonably complete and
accurate rock, fluid, and geological data, estimates of ultimate waterflood recovery
should be reasonably correct. Unfortunately, the timing of actual events such as fillup,
initial water breakthrough, and peak production is more difficult to achieve.

9-1

The primary objectives of this chapter are to provide tools, techniques, and procedures
which will supplement prediction techniques so as to assist in the surveillance of
ongoing activities and to aid in the alteration of the initial waterflood design. In
addition, procedures presented should assist in quantifying the flood performance.
Waterflood surveillance includes close monitoring and professional management of the
entire waterflood operation including production wells, injection wells, injection
facilities, water quality, and metering capabilities.

As a general objective, a

surveillance program should allow for the maximum oil recovery to be achieved at the
lowest water-oil ratio and operating cost. From a reservoir viewpoint, this can be
achieved by maximizing the waterflood recovery factors which primarily consist of

E A , EV , and ED with a minimum amount of injected water.


II. Production and Injection Test Analyses
A. Maps
A sound waterflood surveillance program requires that high quality and easily
readable maps be prepared for each important geological flow unit. It is essential
that one large scale map showing only bottomhole locations for each active and
inactive injector and producer be prepared. Extra information such as wellbore
trajectories is usually not helpful. Such extra information can be displayed on
other maps. These maps do not need to show wellbore trajectories and shallower
or deeper zone completions. Also, it is desirable to have pore volume maps for
each flow unit with pore volume computed on a pattern by pattern basis for each
flow unit. The net pay cutoff parameters should be clearly delineated. Good
quality structure maps can be helpful in showing faults, fluid contacts, and the
productive area of a field. Lastly, a finely gridded cross-section can be helpful in
defining vertical flow barriers and lateral discontinuities.

9-2

B. Production Well Testing Procedures


Production and injection graphs are valuable tools for monitoring and detecting
changes in both well and reservoir performance. Production and injection trend
analysis is usually based on allocated oil, gas, and water production or injection
data which comes from well tests.

Production trend analysis (decline curve

analysis) is an accepted industry procedure for evaluating remaining recoverable


reserves. Production or injection trends frequently form the basis of working over
wells, stimulating wells, setting bridge plugs, installing larger equipment, or
drilling of new wells.

It is recognized that if well tests are not reliable,

unnecessary well work may be carried out or production opportunities may be


missed. Decisions to make capital expenditures on individual wells or a larger
portion of the field are tied to production and injection trend analysis. Clearly,
there is a need for frequent and reliable well tests. Each production well should be
tested monthly for 8 to 12 hours or more. High volume or strategic production
wells should be tested at least twice a month. Changes in production by more than
a pre-determined percentage for these strategic wells should automatically result
in a re-test with the information being brought to the field reservoir or production
managers attention.
Production test procedures and test equipment used early in the life of the field
(during primary depletion) when water cuts are low, may need to be changed
under waterflooding when total fluid production is greater and water cut is much
higher. Test equipment and flow meters should be calibrated on a regular basis to
insure that oil, water, and gas measurements are reliable. A reliable and regular
well testing program is an integral part of any type of reservoir monitoring and
reservoir management program.
C. Production and Injection Trend Analysis
Production and injection trend analysis is a valuable tool in monitoring and
detecting changes in well and reservoir performance. Traditionally, oil, water, and
9-3

gas production rates are plotted versus time for individual wells, groups of wells,
and entire fields. Most engineers agree the most reliable trend analysis occurs
when well evaluations are conducted using accurate well test data. A well-by-well
review allows the analyst to incorporate recent changes in production behavior
resulting from reservoir performance or well work activities. Production trend
analysis for groups of wells, while commonly used in the industry, may be less
accurate because it is difficult to account for individual well workovers,
mechanical changes to existing wells, the shut-in of old wells, or the addition of
new wells. The grouping of wells to evaluate production trends tends to combine
the good, the bad, and the ugly wells and may give an unrealistic picture of
reservoir behavior.
1.

Production Wells
Various graphs of production data can provide valuable insight into the
performance of an ongoing waterflood. Before discussing some of the more
common graphs used in waterflooding, it should be noted that many of these
graphs represent a simple carryover from primary depletion to waterflooding.
Because the reservoir drive mechanism in waterflooding is much different
than in primary depletion, the production trend analysis tools which may be
applicable for primary depletion may not be applicable for water injection.
For example, a considerable number of technical papers have discussed the
conditions under which exponential, harmonic, or hyperbolic decline curves
develop during primary depletion. Reference 1 summarizes these methods.
Significantly, none of the declines may be applicable in waterflooding. As
mentioned in Eqs. 4.30 and 4.31 in Chapter 4, oil and water production rates
are directly related to the water injection rate. Consequently, changes in the
rate of water injection into existing wells or the drilling of new injectors may
alter the rate of oil production and the decline rate but may not result in
changes in remaining reserves.

9-4

Decline curves in waterflood analysis have limited theoretical justification.


Nevertheless, when production data are plotted on various types of graphs,
linear trends frequently develop and may be used for evaluating and
forecasting purposes. As a result, many decline curves represent empirical
evaluations. Listed below are some of the common plots used to evaluate
production and injection performance.
a. Coordinate Graph
A coordinate graph showing oil production rate, water production rate,
and gas/oil ratio (GOR) (not gas production) versus time can be one of the
most valuable diagnostic plots in assessing waterflood behavior.

By

using a linear scale, as shown in Figure 9-1, it is easy to see small but
important changes in the data trends.

Figure 9-1
Single Pattern Production and Injection Data vs Time
1000
900

4500

Water Inj
GOR

800

4000

700

3500

600

3000

500

2500

400

2000

300

1500

200

1000

100

500

0
0

10

15

# of Years

9-5

20

25

30

GOR, SCF/BO

Production Rate, Bbls/Day

5000

Oil
Water

If a semi-log graph of oil rate versus time is used, frequently the log scale
tends to mask or obscure small changes in production. The GOR graph
can provide great insight into gas fillup. If there is free gas saturation at
the start of water injection, the GOR is likely to be much greater than the
initial solution GOR. During the gas fillup period of the waterflood, the
GOR declines. This can be a valuable indicator that reservoir pressure is
increasing, gas is being dissolved in the oil, and oil response should be
anticipated in the near future. As gas fillup occurs in all of the important
geological flow units, the GOR will decline to and level out at the
solution GOR existing at the start of water injection.
b. Exponential Decline Curves (and Hyperbolic and Harmonic)
In primary depletion there are certain conditions in which it can be
mathematically demonstrated that oil (or gas) production rate will decline
exponentially (or hyperbolic or harmonic). In an exponential decline, oil
production rate is plotted versus time on semi-log graph paper.

In

waterflood operations, if oil production rate forms a straight line, it is an


empirical decline which is exponential. Figure 9-2 is an example of oil
production rate versus time characterized by exponential decline
behavior.

9-6

Figure 9-2
Oil Rate vs Time
Center Production Well in a 5-Spot Pattern

Oil Rate, BO/Month

10,000

1,000

100

10
0

10

12

14

16

18

20

22

24

26

28

# of Years

Eq. 9-1 describes exponential decline.

qt qi e Dt

(Eq. 9.1)

where:

qi

= initial production rate at start of decline where t = 0.0, BOPD

= exponential decline, fraction/year

qt

= production rate at future time t , BOPD

= time since the start of decline, years

Integration of Eq. 9.1 with respect to time yields:


9-7

30

q qt
N pt i
D

* 365

(Eq. 9.2)

where:

N pt

= cumulative oil production, STBO

By rearranging Eq. 9.2, it can be seen that:

qt qi

D N pt
365

(Eq. 9.3)

Eq. 9.3 describes a straight line. That is, if the oil production rate is
declining exponentially, then oil production rate ( qt ) plotted versus
cumulative oil production ( N pt ) on coordinate graph paper is also a
straight line. This graph can be extrapolated to the economic limit and
ultimate production can be estimated. Eq. 9.3 removes the time factor,
however, from a mathematical perspective, Eq. 9.3 is equivalent to Eq.
9.1.
The advantage of using the oil rate versus cumulative production graph is
that it is easier to observe changes in oil production rate which may be
masked on a log scale. Figure 9-3 is a graph of oil rate versus cumulative
production.

9-8

Figure 9-3
Oil Rate vs Cumulative Oil Produced
Center Production Well in a 5-Spot Pattern
4.0

Oil Rate, MBO/Month

3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0.0

50.0

100.0

150.0 200.0 250.0 300.0


Cumulative Oil, MBO

350.0

400.0

The data used in Figure 9-3 are also the data used to construct Figure 9-2.
Any graph, such as shown in Figures 9-2 and 9-3, in which historical
oil production rate trends are extrapolated for estimating future
production should be used with caution. It should be recalled from
previous discussions that changes in future water injection rates (into
existing or new wells) would be expected to have an effect on future oil
production rates but do not necessarily change ultimate recoverable
reserves. Future changes in injection may only accelerate recovery.
2.

Injection wells
For each injection well, a linear graph should be maintained which shows
daily water injection rate, average daily wellhead injection pressure and
estimates of average reservoir pressure versus time. Average pressure points
9-9

should be gathered at least every two years, particularly during the early
years of the waterflood. These pressure measurements provide insight into
the voidage replacement ratio and may be helpful in evaluating Hall plots
which are discussed in a later section.
3.

Patterns
A pattern (or cluster of wells) analysis is helpful when evaluating reservoir
performance in localized areas.

Pattern graphs are similar to those

constructed for individual wells except that pattern graphs represent


allocated oil, water, and GOR data. For example, in a perfect injection
centered five-spot pattern 25 percent of the oil, water and GOR data for each
of the four producing wells would be assigned to the production graph.
A principal difficulty of performing pattern analysis is production allocation
for injection centered patterns (or injection allocation for production well
centered patterns). This is complicated due to the fact that waterfloods are
usually characterized by irregular patterns, out of balance injection and
production rates, directional permeability trends, or wellbore operational
issues. Figure 9-4 shows a production centered five-spot pattern. Figures 95 and 9-6 show production and allocated injection versus time.

9 - 10

Figure 9-4
Production Centered 5-Spot Pattern
N-Well

80 Acres

W-Well

E-Well
C-Well

S-Well

Figure 9-5
Production and Allocated Injection for a
Production Centered 5-Spot Pattern
8.0

32000
Oil

28000

Wtr
Wtr Inj

6.0

24000

GOR

5.0

20000

4.0

16000

3.0

12000

2.0

8000

1.0

4000

0.0

10

12

14

# of Years

9 - 11

16

18

20

22

24

GOR, SCF/BO

Production Rate, MB/Day

7.0

Figure 9-6
Annotated Production Centered 5-Spot Pattern
8.0

32000

7.0

S-i

W-i

C-p

Oil
Wtr

6.0

24000

GOR

5.0

20000

4.0

16000

3.0

12000

2.0

W-p

E-i

N-p

8000

N-i
S-p

1.0

GOR, SCF/BO

Rate, MB/Day

28000

Wtr Inj

4000

E-p

0.0

10

12

14

16

18

20

22

24

# of Years

Figure 9-6 is annotated and indicates when the five wells which create the
five-spot in Figure 9-4 were initially drilled as producers and later converted
to injection.
4.

Voidage Replacement Ratio (VRR) (Monthly and Cumulative)


The VRR is defined as being the sum of the water injection and natural water
influx divided by reservoir voidage, measured at reservoir conditions. VRR
is usually evaluated on a monthly basis. In many instances, natural water
influx is negligible; however, in some instances it may be important.
Monthly reservoir voidage, evaluated at reservoir conditions, is calculated
using Eq. 9.4:
Voidage =

qo Bo qw Bw qo (GOR Rs ) Bg

Where
9 - 12

(Eq. 9.4)

qo

= oil rate, STBOM

qw

= water rate, STBWM

GOR = current monthly producing GOR, SCF/STBO

Rs

= solution GOR, SCF/STBO

Bw

= water formation volume factor, RB/STBW

Bo

= oil formation volume factor, RB/STBO

Bg

= gas formation volume factor, RB/SCF

If the VRR for a given month is equal to or greater than 1.0, the reservoir
pressure is being maintained or increased for the month. If the ratio is less
than 1.0, reservoir pressure declines for the month. When computing the
reservoir voidage, it should not be assumed the free gas term is negligible
without making appropriate calculations.
Due to leaking faults, poor cement casing bond, discontinuous sands depleted
prior to water injection, a gas cap, or an inactive aquifer, it is common to lose
some of the injected water to areas outside of the floodable pore volume.
The volume of water lost varies from reservoir to reservoir, but this writers
experience is that 10 to 50 percent of the injected water is lost. Therefore for
computing effective injection volumes, the actual injection needs to be
reduced by the estimated percentage of water lost out of zone. In the absence
of other information, it can be assumed that 20 to 30 percent is lost.
Cumulative VRR is the cumulative injection and natural water influx since
the start of injection divided by cumulative reservoir voidage since the start
of injection. As long as this cumulative VRR is equal to or greater than 1.0,
9 - 13

after taking into account injection losses, reservoir pressure since the start of
water injection will be maintained or increased. Many engineers compute
cumulative VRR by using cumulative reservoir voidage since the date of
initial production from the reservoir with the idea that waterflood response
will not occur until VRR is 1.0. This is not correct. Waterflood response
within the most permeable layers will occur when gas fillup is achieved as
discussed in Chapter 4. When the cumulative VRR computed since initial
production reaches 1.0, reservoir pressure will have increased back to near
the original reservoir pressure.
Figure 9-7 presents a graph of monthly and cumulative VRR since the start
of injection.

Figure 9-7
Monthly and Cumulative VRR
Since the Start of Injection
3.0

Monthly
VRR
Cum
VRR

2.5

VRR

2.0

1.5

1.0

0.5

Cum VRR Since Initial Production = 0.44


0.0
0

10

11

12

13

14

15

16

Time, Months

5.

Spaghetti Graph
The reservoir analyst should not try to include too much data on a single
graph. Frequently, analysts attempt to plot five or six production variables
9 - 14

such as oil, water, and gas production rate, GOR, water injection rate, water
cut, and well count on the same graph. Consequently, the graph becomes
cluttered and important production trends may not be identified. Figure 9-8
is an example of a spaghetti graph.

Figure 9-8
Spaghetti Graph for a Production Well
1000

1.4
BOPD
MCFPD
1.2

WCUT
GOR

1.0

100
0.8
GOR

Oil (BOPD); Water(BWPD); Gas(MCFPD); WC%

BWPD

0.6

10
0.4

0.2

0.0

10

12

14

16

18

20

22

24

26

Years

It was only after each production parameter was plotted on separate graphs as
illustrated in Figures 9-9, 9-10, 9-11,

9-12, and 9-13 that important

production trends were observed leading to improved reservoir management


strategies. The message is to construct simple graphs until sudden changes
or gradual trends can be seen. At the time significant changes occur, graphs
may be combined to help identify causes and effects of field activities.

9 - 15

Figure 9-9
Single String of Spaghetti Oil Rate vs Time
140
130

BOPD

120

Oil Production, BOPD

110
100
90
80
70
60
50
40
30
20
10
0
0

10

12

14

16

18

20

22

24

26

Years

Figure 9-10
Two Strings of Spaghetti Oil & Water Rate vs Time
140
BOPD

130

BWPD

Oil and Water Production, BPD

120
110
100
90
80
70
60
50
40
30
20
10
0
0

10

12

14
Years

9 - 16

16

18

20

22

24

26

Figure 9-11
Two Strings of Spaghetti Oil Rate & GOR vs Time
140

1.4

130

BOPD

120

GOR

1.2

110
1.0

90
80

0.8

70
60

0.6

50
40

GOR, MSCF/BO

Oil Production, BOPD

100

0.4

30
20

0.2

10
0

0.0

10

12

14

16

18

20

22

24

26

Years

Figure 9-12
Spaghetti String Exponential Decline

Oil Production (BOPD)

1000

BOPD

100

10

1
0

10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44
Years

9 - 17

Figure 9-13
Spaghetti String Exponential Decline
100
90

Oil Production (BOPD)

80
70
60
50

EUR @ 10 BOPD = 625MBO

40
30
20
10
0
0

100

200

300

400

500

600

700

800

Cum Oil - MBO

6.

Water/Oil Ratio Plot


A common tool used to evaluate reservoir performance and forecast ultimate
recovery is a semi-log graph of WOR versus cumulative oil production.
WOR is plotted on the log scale and cumulative oil production ( N p ) is on
the linear axis. This graph has been used for many years in the petroleum
industry for evaluating waterflood behavior but very little technical data has
been published on its limits of application. References2,3 suggest that a graph
of WOR versus

Np

for WOR values exceeding 1.0 to 2.0 can be

extrapolated to higher WOR values to estimate ultimate recovery. Again,


like other production plots, the WOR versus

Np

graph is most dependable

when it is applied on a well-by-well basis so that the good, bad and ugly
(low, medium, and high) WOR wells are not combined. Finally, at very high
WOR values, the data points bend upward and form a vertical line as the
9 - 18

reservoir is completely swept to a residual oil saturation and maximum


waterflood recovery is achieved.
It is recalled from a previous discussion in this chapter and from discussions
in Chapter 4 that oil and water production rates are influenced by water
injection rate. Significantly, when applied on a well-by-well basis, the WOR
versus

N p plot is independent of injection rates.


WOR

Eq. 4.28 shows that:

fw
B
o
1.0 f w Bw

(Eq. 4.28)

Figure 9-14 is an example of a WOR graph for one well in a waterflood


pattern. The data used in this graph were used to construct Figures 9-2 and
9-3.

100.0

Figure 9-14
WOR vs Np
Center Production Well in 5-Spot Pattern

WOR

10.0

1.0

0.1
100

150

200

250

300

350

Cumulative Oil, MBO

9 - 19

400

450

500

Figure 9-15 is a WOR plot for a group of production wells. Multi-well


graphs are usually more difficult to analyze than single wells due to different
conditions in the various wells.

Figure 9-15
WOR vs Np Group of Production Wells
100.0

WOR

10.0

1.0

0.1
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Cumulative Oil, MBO

7.

Oil Cut
A decline curve which has been successfully used in many waterflood
projects is a graph of the fraction of oil in the total produced well stream (oil
cut) versus cumulative production. As seen in Eq. 4.47, WOR is related to
water cut and consequently, oil cut ( f o

1 f w ).

The oil cut is plotted

on a log scale and cumulative production is plotted on a linear scale. The oil
cut graph is similar to but different from the WOR plot. Figure 9-16 is an
example of the oil cut plot.

9 - 20

FIGURE 9-16
Oil Cut (Fieldwide) vs Cumulative Oil Production

Oil Cut, %

100

10

1
0

10000

20000

30000

40000

50000

60000

Np, MBO

While there are no specific theoretical guidelines, but because oil cut is
related to WOR, it would be reasonable to expect the limitations of WOR
plots extend to oil cut plots. Since the WOR plot is most reliable when
applied to individual wells for WOR values greater than 1.0 to 2.0, oil cut
plots should be applied to individual wells using production data where the
oil cut is less than about 70 percent. When applicable, the graph can be
extrapolated to the oil cut economic limit to estimate ultimate recovery.
8.

Plot

Ershaghi and Omoregie4 suggested an alternate method of analyzing


waterflood performance. Their method is a graphical technique which is
referred to as an

plot. They recommend plotting

production where:

9 - 21

versus cumulative

1.0
1.0
1.0
X ln
fw
fw

(Eq. 9.5)

and:

fw

qw
qo qw

References4,5,6 indicate the

(Eq. 9.6)

plot is a more accurate method of plotting

waterflood data for extrapolation purposes. For actual fractional flow,

fw ,

> 1.0 ), a linear relationship between X

and

data greater than 0.5 (WOR

Np

may develop when applied on a well-by-well basis. Extrapolation of

these data permit an estimate of future recovery as a function of


Figure 9-17 is an example of the

9 - 22

plot.

and

fw .

FIGURE 9-17

X vs CUMULATIVE PRODUCTION

5
Economic Limit

Oil
Recovery
at
Economic
Limit

2
0

Cumulative Oil Recovery, MMBO

The

plot is not the answer to all problems. The basic theory behind this

graph assumes a 1-D Buckley-Leverett model without consideration of (1)


stratification or (2) changes in areal sweep efficiency after water

9 - 23

breakthrough. Like other production plots such as the oil cut and WOR
methods, the
9.

plot methodology should be used with caution.

Oil Cut versus Cumulative Production (Coordinate Graph)


Several experienced engineers have recommended (without theoretical
justification) the use of a simple coordinate graph of oil cut (or water cut)
versus cumulative oil production as an excellent method for forecasting
production in mature waterfloods (where the oil cut is less than about 30
percent). In several confidential consulting projects, it has been observed
that this graph yields results which appear to be very consistent with results
from other empirical plots. Figure 9-18 is an example of a coordinate graph
of oil cut versus cumulative oil production for a mature stratified, carbonate
waterflood. This figure presents actual field data and shows an apparent
increase of 20 MMSTBO in ultimate oil recovery attributable to a pattern
realignment and pattern regularization which occurred at a cumulative
production of about 290 MMSTBO. A conventional semi-log graph of oil
production rate versus time gives similar results.

9 - 24

FIGURE 9-18
Oil Cut vs Cumulative Oil Production Showing Effects of
Pattern Regularization
40

Continued Waterflood Forecast


(Ultimate = 414 MMBO at 5.0% Oil Cut)
Pattern Regularization Forecast
(Ultimate = 434 MMBO at 5.0% Oil Cut)

Oil Cut, percent

30

20

Waterflood Pattern
Regularization

10
5
0

200

250

300

350

400

450

500

Cumulative Oil Production, MMBO

10. Recovery Factor versus Hydrocarbon Pore Volumes Injected


A useful tool in waterflood analysis is a coordinate graph of oil recovery
factor versus hydrocarbon pore volumes of water injected. Hydrocarbon
pore volume, HPV, is simply net pore volume multiplied by the initial oil
saturation at the time of field discovery. That is:

HPV V p * (1 S wc )
where:

HPV

= hydrocarbon pore volume, RB

Vp

= net floodable pore volume, RB

9 - 25

(Eq. 9-7)

S wc

= connate water saturation ( S wc


fraction

Figure 9-19 is a graph of

RF

versus

Wi / HPV

irreducible water),

(hydrocarbon pore

volumes of water injected) where Wi is cumulative injection.

Recovery Factor, percent

FIGURE 9-19
RECOVERY FACTOR vs HYDROCARBON PORE
VOLUMES INJECTED
40

Better Than Average


Field Average,
Analog, or Simulation

30

20

Below Average
10

0
0.0

1.0

2.0

3.0

W i /HPV, Hydrocarbon Pore Volumes Injected

An important function of this type of graph is to compare how individual


patterns or clusters of wells compare to theoretical values, field average
values, or similar plots from other fields.

Efforts should be made to

determine why actual performance for a group of wells is similar to or


different from the theoretical or field average plots.
Plotting

RF

versus Wi

/ HPV

removes the time element from the graph

and allows for a comparison of actual performance to a base performance


where actual injection rates are different from the original injection rates.
9 - 26

The base performance graph can also be obtained from reservoir simulation
or from analogous fields.

When making comparisons between adjacent

fields or between patterns, it is assumed that factors such as (1) permeability


variation, (2) fluid saturation at the start of injection, (3) fluid properties, and
(4) relative permeability are similar.
11. Multiple Trend Forecasting with Field Production Constraints
In 1998, Wilson, et al.7 presented several interesting ideas related to
combining multiple production trends to improve predictions of the future
performance of the single primary phase when production constraints exist.
Two methods are described in detail which specifically account for
mature waterflood performance and gas-handling constrained production.
The waterflood analysis accounts for waterflood response when total
constrained liquid production is sometimes limited by the ability to process
and lift large volumes of water along with decreasing amounts of oil. The
method forecasts total liquid rate and water-oil ratio versus cumulative oil
production. Several field examples are discussed.
12. Summary of Production Graphs
Engineers have developed and utilized many different types of production
plots for monitoring and predicting waterflood performance. Baker8,9 does a
good job in summarizing many of these plots. Nevertheless, it is clear that
no complete theoretical basis for constructing many of the production plots
exists. The various graphs should generally be considered as empirical plots
which historically have been found to be helpful in monitoring performance.
The graphs may be influenced by a number of factors such as injection rate,
infill drilling, pattern geometry, the degree of pattern balancing, relative
permeability characteristics, fluid properties, fluid saturations, directional
permeability, and reservoir stratification. Yet these graphs, when applied on
an individual well basis, can be helpful in estimating future production with a
9 - 27

continuation of historical injection, production, and operational procedures.


It is also clear that if operations are changed (infill drilling, pattern realignment, increased injection, well stimulation, etc.), these production plots
may show the incremental benefits derived from the changes. Finally, when
used correctly on a well-by-well basis, the graphs can provide valuable
insight into existing waterflood behavior and how current operations and oil
production can be improved.
III. Pressure Transient Testing
The performance of injection wells and production wells is of major importance in
waterflooding operations. It is important that a complete testing program be used
before and during injection to determine the condition of the reservoir. A number of
parameters including formation flow capacity, formation damage, average reservoir
pressure, and formation parting pressure can be determined from a properly designed
and analyzed test. In general, the major objectives of a testing program are to monitor
changes in formation damage, reservoir pressure, and parting pressure gradients.
Robertson and Kelm10 have descried a testing program used by one company to
accomplish these objectives.
The pressure tests which are most frequently conducted in a waterflood surveillance
program include:

buildup

falloff

step rate

Hall plot

In major projects, a pulse test or interference test is occasionally run to determine


pressure communication between wells and to estimate inter-well rock properties.

9 - 28

Anytime a production or injection well is performing below expectations, a pressure


buildup or falloff test should be considered.
A. Pressure Buildup and Pressure Falloff Testing
A complete waterflood surveillance program should include a regular schedule for
conducting pressure buildup and pressure falloff tests. The general objectives of
these tests are to determine flow capacity,
and average reservoir pressure,

p.

kh , wellbore flow restriction (skin),

It would be desirable (in a perfect world) to

conduct these tests on each production and injection well on an annual basis.
Unfortunately, it is not always possible to live in the ideal world. As a result, it is
recommended that 20 to 25 percent of the key wells (both producers and injectors)
be tested on an annual basis.
Change in average reservoir pressure

is one of the most useful diagnostic tools

available in a surveillance program. Average pressure values can be posted on


maps and iso-pressure contours can be constructed. It is recognized that while
these maps are not theoretically correct, they provide considerable reservoir
insight regarding the degree of reservoir fillup, injection efficiency, and the
Voidage Replacement Ratio in various segments of the field.
A number of technical books have been published on conventional well testing.
References 11, 12, and 13 summarize the important design and interpretation
features of pressure tests. It is beyond the scope of this section to discuss pressure
testing in more detail.
Several technical papers have discussed the use of pressure falloff tests to estimate
the distance to various flood fronts such as water fronts and oil banks. Due to
saturation gradients which exist in the water zone, reservoir stratification, and the
lack of a properly balanced pattern, it is the experience of the writers that flood
front distance calculations are difficult at best. Further, confidence in a calculated
distance would likely be very low.
9 - 29

B. Step Rate Test


To realize maximum injectivity without fracturing the reservoir, it is necessary to
have a reliable method of determining the fracture pressure. Fracture gradients are
generally known for most areas, but the actual fracture pressure in a given
reservoir can vary from well to well. Further, the fracture pressure can vary with
time and operating conditions within a given well largely due to reservoir
pressure.

As reservoir pressure declines, formation parting pressure (FPP)

decreases.

Conversely, as reservoir pressure increases during fillup, the FPP

increases. In several instances14 where substantial primary depletion has occurred,


the FPP at the start of waterflooding has been found to be only slightly greater
than a hydrostatic column of salt water. The parting pressure will often increase
with injection during the early life of a flood. An approximate fracture pressure
can also be obtained from the instantaneous shut-in pressure following a
stimulation treatment; this is approximate, however, and as noted previously, the
pressure can change for a number of reasons.
The best method available to determine safe injection pressures is the step rate
test15. Singh, et al.16 and Gidley, et al.17 (Chapter 15, and Appendix 1) presented a
comprehensive review of the variables affecting the step rate test (SRT).
Generally, the test well is either shut-in or stabilized at a reduced but constant
injection rate prior to the start of the SRT. Ideally, the shut-in period should be
long enough so that the bottom-hole pressure stabilizes near the static formation
pressure. Alternatively, if the well is stabilized at a reduced injection rate, the
stabilization period should be long enough to achieve a steady-state or a pseudo
steady-state condition. The SRT that follows consists of a series of constant-rate
injections with rates increasing from low to high in a stepwise fashion. Each
constant-rate step is normally of equal time length. Injection rates and pressures
are recorded for each step and analyzed to determine the FPP.

9 - 30

The conventional analysis method assumes a steady-state Darcy flow into an


injection well and is based on Eq. 9.8.

r
pwi p 141.2 ln e S iw
kh rw

(Eq. 9.8)

If a steady-state condition is achieved during each injection step,


constant. It is further assumed that during the test,

re

will be

can be assumed constant.

A linear relationship will therefore exist between the injection pressure

pwi

at

the end of each constant-rate injection step and the corresponding injection rate

iw .
For most SRTs, step time lengths are seldom long enough to reach a steady-state
condition. In this case, re can be replaced by rd (the drainage radius as defined
by Aronofsky and Jenkins18). If time step size is constant, a linear relationship
should still exist between

pwi

and iw for the test data below parting pressure.

This is because the logarithmic term

ln rd / rw

will be constant for each

injection rate during the test.


A Cartesian plot of

pwi

versus iw is made for the SRT data. When the injection

pressure exceeds parting pressure the resulting fracture acts as an additional fluid
conductor which changes (reduces) the slope of the
accordingly. This is illustrated by the
as shown in Figure 9-20.

9 - 31

pwi

pwi

versus

iw

curve

versus iw plot of the simulated data

FIGURE 9-20

Injection Pressure, psi

BASIC PLOT FOR STEP RATE TEST

Fracture Pressure

Injection Rate, BPD

The pre-parting data falls on a straight line and is governed by Eq. 9.8. Normally,
another straight line is drawn through the points above the parting pressure as
shown in Figure 9-20. The pressure corresponding to the point where the two
lines intersect is interpreted as the parting pressure. This method provides an
approximate estimate of parting pressure. Since the fracture length continues to
increase above the FPP, there is no theoretical basis for drawing a second straight
line through the points above the parting pressure. Accordingly, on the vertical
axis, this line extrapolates back to a pressure point which is much higher than the
pretest pressure,

p.

The value of the intercept with respect to the pretest pressure

provides a qualitative indication that the pressure points corresponding to the


second straight line are above the parting pressure.
Many engineers are concerned that this test will fracture and permanently damage
the formation. This should not be a concern if the test is properly conducted.
When using this test for the first time in a field, rates should be increased slowly
9 - 32

until a fracture pressure is definitely established. Tests in other wells can then be
designed so that pressures will not exceed this pressure for appreciable lengths of
time.
Multi-rate pressure transient analysis techniques can be used in some cases to
analyze the results of a step rate test for permeability and skin. However, rates
must be carefully controlled during each flow period for this analysis to be
accurate.
It is recommended that Reference 17 be studied in more detail to review other
factors which may affect SRT analysis such as injection time step size, wellbore
storage, skin damage, and rate increment size.
________________________________________________________________
EXAMPLE 9:1
A 2800-foot water injection well located in a developed flood in the Permian
Basin of West Texas has undergone a step rate test. The test data were measured
as shown. Prior to the test, the well was shut-in for seven days. Analyze the data
and determine:
(a) the fracture pressure of the reservoir,
(b) the fracture gradient, and
(c) the optimum water injection rate.

9 - 33

TABLE 1
Date
3/28/99
4/ 4 /99
4/ 4/99
4/ 5/99
4/ 6/99
4/ 7/99
4/ 8/99
4/ 9/99
4/10/99
4/11/99
4/12/99
4/13/99
4/14/99

iw , bbls / day

Bottom-Hole Pressure, psi

631.0
0.0
100.0
175.0
250.0
325.0
400.0
475.0
--550.0
625.0
700.0
775.0

1,940.0
1,615.0
1,683.0
1,725.0
1,760.0
1,812.0
1,824.0
1,857.0
--1,870.0
1,880.0
1,901.0
1,920.0

SOLUTION
The step rate data are plotted in Figure 9-21. It is observed that two straight lines
are defined by the data; the intersection of these lines defines the reservoir fracture
pressure as 1825 psi and the optimum water injection rate as 370 bbls per day.

9 - 34

FIGURE 9-21 STEP RATE TEST


(EXAMPLE9:1)

B ottom -H ole Injection P ressure, psi

2000

1900

1800
Approximate Fracture Pressure
1825 psi at iw = 370 BWPD

1700

1600
0

200

400

600

800

1000

Injection Rate, BWPD

Based upon this information:


Fracture gradient = 1825 psi / 2800 feet or 0.65 psi/ft
_________________________________________________________________
9 - 35

An issue which frequently is discussed in waterflood operations is the impact of


vertically induced fractures on waterflood recovery. These fractures can result
from either conventional hydraulic fracturing or by injecting above the FPP for
extended periods of time. The effect of such fractures can be helpful, harmful, or
have negligible effect on recovery. In general, if the direction of the fracture is
such that it is parallel to a line connecting two adjacent injectors, overall areal
sweep efficiency will be enhanced and recovery should increase when compared
to the non-fractured reservoir. In this case, vertical fractures may be helpful. On
the other hand, if the direction of the vertical fracture is such that it is parallel to a
line connecting two adjacent producers, overall areal sweep efficiency may be
reduced and oil recovery will decrease when compared to the non-fractured case.
Fracture orientation is not always known. As a result, Reference 19 presents a
review of the technical literature as well as information which considers the effect
of fracture length on sweep efficiency as a function of mobility ratio and fracture
orientation. In general, if the fracture length is less than approximately ten percent
of the distance between the injector and producer, areal sweep efficiency will not
be significantly affected.
In tight or low permeability waterfloods, it is frequently necessary to inject above
the fracture pressure to inject meaningful quantities of water. In these instances, it
is desirable to create a short hydraulic fracture possessing a high conductivity.
Subsequent water injection should take place at injection pressures which are
below fracture pressures so as to prevent fracture growth as a function of time.
Even so, injection above the fracture pressure or into hydraulically fractured wells
makes it very difficult to control the vertical distribution of the injected water
which could be harmful to vertical sweep efficiency.
Finally, it should be noted that based on fracture theory and the work of Dyes et
al20 as reviewed in Reference 17, injection rates in reservoirs with short fractures
(ten percent of the distance between injectors and producers) show about a 2.5 to
3.0 increase over injection rates for comparable non-fractured systems. If actual
9 - 36

injection rates show more than a threefold increase, it could be that skin damage
was removed, a much longer fracture length (greater than ten percent) is present,
or that injected water is being lost out of zone.
C. Hall Method of Analyzing Injection Well Behavior
In 1963, Hall21 presented a technique for interpreting routinely collected injection
well data to draw conclusions regarding skin effects at water injection wells.
Buell et al22 studied the application of the Hall Plot for injection well analysis for
both waterfloods and polymer injection projects.

The required data includes

average monthly bottom-hole injection pressure (wellhead injection pressures can


be used if they are correctly converted to bottom-hole pressures by accounting for
the hydraulic head and friction pressure losses in the tubing), average reservoir
pressure

( p ) , monthly injection volumes, and injection days for the month.

The

procedure assumes gas fillup has occurred (GOR has collapsed), the mobility ratio
is 1.0 and steady-state injection is present such that the injection rate can be
expressed as:

iw

0.00707 kh( pwi p )


r

ln e S
rw

At this point, it is assumed that k , h, , re , rw , and

(Eq. 9.9)

are constant. Thus, Eq.

9.9 reduces to:

iw C ( pwi p )

(Eq. 9.10)

where:

9 - 37

0.00707kh
r

ln e S
rw

(Eq. 9.11)

Rearranging Eq. 9.10 results in:

( pwi p )

iw
C

(Eq. 9.12)

Integration of both sides of Eq. 9.12 with respect to time gives:

t
o

pwi p dt

1 t
o iwdt
C

(Eq. 9.13)

The integral on the right-hand side is cumulative water injected. Hence, Eq. 9.13
becomes:

t
o pwi p dt

Wi
C

(Eq. 9.14)

where:

Wi = cumulative volume of water injected at time t , bbls


Examination of Eq. 9.14 indicates that a coordinate graph of the integral term
versus Wi should form a straight line with a slope of 1/ C . This type of graph is
called a Hall Plot. If
value of
change,

k , h, , re , rw and S

are constant, then from Eq. 9.11, the

C is constant and the slope is constant. However, if the parameters


C will also change and thus the slope of the Hall plot will change.

Herein lies the value of the method. Changes in injection conditions may be noted
from the Hall plot. For example, if wellbore plugging or other restrictions to
injection are gradually occurring, the net effect is a gradual increase in the skin
9 - 38

C decreases and, thus, the slope of the Hall Plot


increases. Conversely, if S is decreasing as would be the case when injection
exceeds fracture pressures causing fracture growth, then C will decrease and the
factor.

As

increases,

slope of the Hall Plot will decrease. Figure 9-22 is a Hall plot for various injection
conditions.

FIGURE 9-22
HALL PLOT
FOR VARIOUS INJECTION WELL CONDITIONS

300
A

250

200

150
100
50
Gas Fillup

(A) No Change with MR=1.0


(B) Abrupt + Skin
(C) Gradual + Skin and/or MR> 1.0
(D) Gradual Skin and/or MR>1.0
(E) Abrupt - Skin

0
0

100

200

300

400

Cumulative Water Injection, MB

9 - 39

500

The most difficult part of developing a Hall Plot is the evaluation of the pressureintegral function on the left side of Eq. 9.14. Fortunately, the integral can be
easily solved. Consider Figure 9-23 which shows a graph of monthly bottomhole
injection pressure,

pwi , and periodic estimates of average reservoir pressure, p .


FIGURE 9-23

BOTTOM-HOLE INJECTION AND RESERVOIR


PRESSURE vs TIME
1500

Pressure, psi

Bottom-hole Injection Pressure

1000

500

Average Reservoir Pressure

0
0

Time, Years

Usually

can be estimated at regular intervals, such as one per year. A straight

line connecting annual

estimates will provide a basis for estimating

less

frequently, such as on a monthly basis. Consider the following example.


________________________________________________________________________

9 - 40

Example 9:2
During the last six months, the data listed below have been estimated for an
injection well. Compute the information necessary to prepare six points on a Hall
Plot.

TABLE 2
Month

pwi ,psi

p , psi

t , days

iw , BPM

1.0

2,240.0

1,300.0

11.0

11,000.0

2.0

2,270.0

1,302.0

28.0

28,000.0

3.0

2,285.0

1,304.0

30.0

30,000.0

4.0

2,290.0

1,306.0

16.0

16,000.0

5.0

2,298.0

1,308.0

31.0

31,000.0

6.0

2,312.0

1,310.0

28.0

28,000.0

Solution
If it can be assumed that the tabulated values of

pwi

and

month, then:

t
o

pwi p dt p x t

where:

p pwi p
t

= number of injection days for the month


9 - 41

are average for the

TABLE 3

Month

p x t

iw

Wi

days

1.0

psi
940

11.0

psi-days x 103
10.3

BWPM
11,000.0

BW
11,000

2.0

968

28.0

37.4

28,000.0

39,000

3.0

981

30.0

66.8

30,000.0

69,000

4.0

980

16.0

82.6

16,000.0

85,000

5.0

990

31.0

113.3

31,000.0

116,000

6.0

1,002

28.0

141.4

28,000.0

144,000

These data make up a portion of the data presented in Figure 9-24.

9 - 42

FIGURE 9-24
HALL PLOT FOR WATER
INJECTION WELL
300

250

200

150

100

50

50

100

150

200

250

300

350

400

Cumulative Water Injection, MB

__________________________________________________________________
Changes in the Hall plot will occur gradually and, as in production decline curve
analysis, several months (6 or longer) of injection history may be required to reach
conclusions about injection behavior. Finally, several words of caution are in
order because changes in the slope of the Hall plot can be the result of other
factors. Early in the life of an injection well before gas fillup, the radii of the
water and oil zones will increase with cumulative injection and cause the value of

to increase resulting in the Hall plot being concave upward. Therefore, during

the gas fillup period, the increasing radii create the effect of an increasing skin
9 - 43

factor. Also, the Hall plot assumes a mobility ratio of 1.0. If the actual mobility
ratio is greater than 1.0, then after gas fillup the Hall plot tends to behave like
curve D in Figure 9-22; if the mobility ratio is less than 1.0, the Hall plot will tend
to follow curve C. Finally, as average water saturation in the reservoir increases
over time,

kw increases and may affect the slope.

After gas fillup, if it can be assumed that the difference between

pwi

and

p,

does not materially change, then preparation of the Hall plot is greatly simplified.
In this situation,
constant, and

can be neglected. If the difference between

pwi

and

is

is ignored, the Hall plot is only shifted on the vertical scale

without changing the slope or its interpretative procedures. Under this condition,
the bottom-hole injection pressure,

pwi ,

is simply the wellhead injection

pressure plus a hydrostatic gradient minus a friction loss term. These two terms
can usually be assumed constant and neglected. As a result the left hand side of
Eq. 9.14 simply becomes the integral of the wellhead injection pressure. On the
other hand, if the difference between
include

pwi and p

is not constant, the failure to

can result in changes in the slope leading the reservoir analyst to

incorrect conclusions regarding changes in skin factor. The consequences of


ignoring

p is discussed in more detail in Reference 22.

To determine whether average reservoir pressure is changing, it is necessary to


conduct regular pressure buildup/falloff tests and to monitor monthly VRR plots.
The objective of the Hall plot is to detect changes in the injection well skin factor.
It is not a perfect tool but can, under certain conditions, provide reasonable insight
on skin changes. The best tool for quantifying injection wellbore skin damage is a
properly designed, well executed, and fully analyzed pressure falloff test.

9 - 44

IV. Pattern Balancing


In a large multi-pattern waterflood, significant improvements in the efficiency of the
waterflooding process can be achieved through careful management of the flood
injection rates in each waterflood pattern. Balancing injection and production rates
within and between patterns should substantially reduce produced water handling
requirements, improve long-term production rates, and enhance ultimate recovery.
In balanced patterns, important events such as fillup or water breakthrough for the
various patterns will occur at the same time. To illustrate the principle of pattern
balancing, consider two five spot patterns which are a part of a developed five spot
system as depicted in Figure 9-25.

FIGURE 9-25
TWO 5-SPOT PATTERNS FROM A DEVELOPED 5-SPOT
SYSTEM WITH STREAMLINES A1 AND A2

A2

A1

Pattern 1

Pattern 2

Note the two analogous stream lines A1 and A2. Consider the time to initial water
breakthrough for patterns 1 and 2. For the patterns to be balanced, water breakthrough
at the producing well on stream lines A1 and A2 will occur at the same time. That is:

tbt1 tbt 2

(Eq. 9.15)

9 - 45

tbt1

= time to water breakthrough in pattern 1, days

tbt 2

= time to water breakthrough in pattern 2, days

Also, the time to water breakthrough is the cumulative water injected to breakthrough
divided by the average daily injection rate as shown in Eq.9.16.

Wibt1
iw1

tbt1

(Eq. 9.16)

and from Eq. 5.11 of Chapter 5:

Wibt V p1E Abt1 ( S wbt S wc )1, BW

(Eq. 5.11)

iw daily water injection rate, BW/D


tbt1

V p1E Abt1 ( S wbt S wc )1

(Eq. 9.17)

iw1

Referring to Eqs. 9.15 and 9.17, it is clear that

V p1E Abt1 ( S wbt S wc )1


iw1

V p 2 E Abt 2 ( S wbt S wc ) 2
iw2

(Eq. 9.18)

If the two patterns are similar such that variables including the fractional flow graph,
mobility ratio, average permeability, directional permeability, and stratification are
similar, then Eq. 9.18 reduces to:

V p1
iw1

Vp2

(Eq. 9.19)

iw2

or
9 - 46

iw2 V p 2

iw1 V p1

(Eq. 9.20)

Equation 9.20 indicates that water injection into various patterns or segments of the
field should be in proportion to the pore volume. For example, if a pattern contains 10
percent of the reservoir pore volume, then 10 percent of the total injection should be
into that pattern.
Several writers (References 23, 24, and 25) appear to indicate that a slight modification
to Eq. 9.20 is appropriate. These authors suggest that injection rate should be in
proportion to the displaceable hydrocarbon pore volume, VD . That is:

iw2 VD 2

iw1 VD1

(Eq. 9.21)

where

VD V p (1 S wc Sor )

(Eq. 9.22)

and

Vp

S wc

= connate water saturation ( S wc

Sor

= residual oil saturation to water injection, fraction

pattern pore volume, bbls

Swir ) , fraction

In summary, to achieve a balanced waterflood project, the fraction of total field


injection into an individual pattern should be in proportion to the pattern displaceable
hydrocarbon pore volume,

VDp , relative to the field displaceable hydrocarbon pore

volume,VDf . In other words:


9 - 47

(iw )ith pattern

VDp
VDf

iwf

(Eq. 9.23)

where

iwf

= target field injection rate

(If the connate water and waterflood residual oil saturations between the patterns are
similar, Eqs. 9.20 and 9.21 are equivalent.)
Moreover, for an ideally balanced pattern, total oil production from each pattern should
be in proportion to the displaceable hydrocarbon pore volume. Consider the following
example.
____________________________________________________________________

Example 9:3
Six five-spot patterns (A, B, C, D, E, and F) are shown in Figures 9-26 and 9-27. The
solid lines represent no-flow boundaries caused by the reservoir pressure distribution
and the stream lines for ideally balanced patterns. The displaceable hydrocarbon pore
volume,

VD , is tabulated below.

If field-wide injection rate is to be maintained at

6,000 BWPD, determine the necessary injection and production rates to maintain a
balanced five-spot pattern in the field.
TABLE 4
Pattern

VD , Mbbls

A
B
C
D
E
F
Total

3,000
2,500
1,500
1,000
1,500
500
10,000
9 - 48

FIGURE 9-26
DEVELOPED FIVE-SPOT PATTERN
1

10

11

9 - 49

12

FIGURE 9-27
THEORETICAL WATER INJECTION RATES
DURING AND AFTER FILLUP
TOTAL FLUID PRODUCTION RATES AFTER FILLUP
FOR BALANCED FIVE-SPOT PATTERNS
2

1
450

450
A

1500
450

450
225

225

375
150

900
225

225
225

225

150

150

150

75

75
F

900
225

600

10

375

1800

375

375

300
225

11

75

75

12

Solution
Balanced patterns require that injection and production rates into and out of each
pattern be in proportion to the pattern displaceable pore volume.
9 - 50

TABLE 5
Percent of
Pattern

VD , Mbbls

Total VD

iw , Mbbls

3,000

30

1,800

2,500

25

1,500

1,500

15

900

1,000

10

600

1,500

15

900

500

300

TABLE 6
Production Volumes From Each Pattern at Various Wells Based on Pattern Injection,
Pattern
A
B
C
D
E
F
Totals

1.0
450
0
0
0
0
0
450

2.0
450
375
0
0
0
0
825

3.0
0
375
0
0
0
0
375

4.0
450
0
225
0
0
0
675

5.0
450
375
225
150
0
0
1,200

6.0
0
375
0
150
0
0
525

7.0
0
0
225
0
225
0
450

8.0
0
0
225
150
225
75
675

9.0
0
0
0
150
0
75
225

10.0
0
0
0
0
225
0
225

11.0 12.
0
0
0
0
0
0
0
0
225
0
75 75
300 75

Not only must the production rates be correct, the allocation of production from each
pattern must be in the manner shown. Achieving this sort of injection and withdrawal
is at best very difficult and is most likely impossible. Consequently, precise pattern
balancing is usually not possible in actual waterflood operations.
____________________________________________________________________

9 - 51

The concept of pattern balancing is further complicated by the presence of actual


reservoir flow boundaries (pinchouts, sand discontinuity, fluid contacts, etc.). One
example is shown in Figure 9-28.

This diagram shows that well locations and

reservoir boundaries affect the pattern shape.

FIGURE 9-28
FIVE-SPOT PATTERN WELL ARRANGEMENT
WITH EFFECTS OF RESERVOIR LIMITS

RESERVOIR
LIMITS

Consider the problem discussed in Example 9:3 and whose injection and production
values are shown in Figure 9-27 under balanced conditions. Let us assume that due to
9 - 52

areal heterogeneities, injection well damage, variations in flow capacity or reservoir


pressure, it is not possible to physically inject (or produce) at the desired rates.
Further, assume that a decision is made to inject at a rate of 1,000 BWPD per well (we
assume this can be achieved without exceeding formation parting pressures). Further,
assume the presence of reservoir limits. Because of the aforementioned reservoir
complexities, the no-flow boundaries (see Fig. 9-29) will, in practice, result in irregular
patterns which will probably not be accurately known.

9 - 53

FIGURE 9-29
RESERVOIR PATTERN DEVELOPMENT WITH FIVE-SPOT
WELL LOCATIONS, OUTER BOUNDARY EFFECTS,
AND UNBALANCED INJECTION AND PRODUCTION

A
i

= 1000
W

= 1000
W

5
C

= 1000

= 1000
W

E
i

F
i

= 1000
W

= 1000
W

11
10

12

Nonetheless, the pore volume in each pattern (whatever the shape) can be
approximated by the percentage of the total injection rate into the pattern. Hence, for
equal injection rates, each pattern in Figure 9-29 contains approximately 16.6 percent
(1/6) of the total pore volume. It is clear that while injection and production wells may
be physically located on a five-spot (or any other arrangement), the flow patterns
which develop within the reservoir may be vastly different depending on (1) the
9 - 54

location and type (no-flow or constant pressure) of outer boundary and (2) injection
and production rates from individual wells.
The above discussion allows the determination of target injection and production rates
in terms of the field-wide injection rate, iwf At some low value of iwf , all the wells
in the project will be able to achieve the target rate, satisfying the equalities expressed
in Eq. 9.23 and leading to an ideally balanced flood. Unfortunately, the field-wide
injection rate at which all patterns can be balanced is usually below the optimum
economic injection rate. Consequently, as iwf is progressively increased, fewer and
fewer wells are able to achieve their target rate and the project becomes increasingly
unbalanced. Nevertheless, if the optimum balanced injection rates could be achieved,
then given a sufficiently long period of time, the ideally balanced flood will produce
the most recovery for the least amount of water production. For additional discussion
on this topic, please see Reference 24.
Finally, the actual pattern which may develop within each geological zone is
dependent on the number of wells completed within the zone. Consider Figure 9-30.
Initial inspection of the well location map indicates a possible developed five-spot
pattern. A more complete analysis of this flood shows the injection wells to be
completed in different zones. For example, Figure 9-31 shows the injection and
producing wells completed in Zone 1 and a possible pattern configuration. Figure 9-32
shows wells competed in Zone 2.

9 - 55

FIGURE 9-30
APPARENT FIVE-SPOT PATTERN

FIGURE 9-31
ACTUAL PATTERN IN ZONE 1

9 - 56

FIGURE 9-32
ACTUAL PATTERN IN ZONE 2

Waterflood patterns do not start injection at the same time. It is recognized that
even in a multi-patterned field, the start of injection in the different patterns takes place
at different times. Therefore, even in the ideal pattern case, events (gas fillup, water
breakthrough, etc.) will occur at different times for the different patterns. However, if
injection into each pattern and each geological layer is in proportion to the pore
volume of that pattern and geological layer, the no-flow boundaries that develop
would be similar to the no-flow boundaries for the ideal system. Of course, it is
recognized that the ideal pattern would only occur if such factors as permeability,
skin factors, and pressure drop are identical. Since this does not occur in the field, it is
likely that ideal patterns never exist within the reservoir even though the well bore
locations may give the appearance of an ideal pattern.

9 - 57

In summary, it is very difficult to predict the exact shape of a flood pattern. Many
factors determine the final shape including well location (and completion intervals),
injection and production rates, reservoir heterogeneities, and reservoir boundary
locations. Because of these complexities, some companies have elected to subdivide
the reservoir into several areal segments (which may contain several patterns) and to
simply balance overall injection and total reservoir withdrawal within each areal
segment.
V. Volumetric Sweep Efficiency
Calculation of volumetric sweep efficiency in a mature waterflood is important. It
provides an indication of the fraction of the reservoir which has been swept or not
swept by the injected water. Additional oil recovery potential exists in the unswept
portion of the reservoir. Volumetric sweep efficiency of the injected water,

Evw , is a

combination of areal sweep and vertical sweep efficiencies as illustrated in Eq. 9.24.

Evw E A Ev

(Eq.9.24)

In most waterflood analyses, volumetric sweep calculations are computed using


estimates of net injected water volume. In many reservoirs, net injection volume
cannot be computed due to poor injection records or the loss of injection water out of
zone.
Cobb and Marek26 describe a technique for computing volumetric sweep efficiency
using only oil production data. The method is based upon a volumetric material
balance and standard waterflood principles. It can be applied to those oil reservoirs
possessing a free gas saturation at the start of injection or when no free gas is present.
The procedure, however, is only valid after free gas fillup, and has application in
many mature waterfloods. The method is also applicable for water-drive reservoirs or
reservoirs which have both water injection and aquifer influx.

9 - 58

Results of the analysis are dependent upon floodable pore volume, oil saturation at the
start of the waterflood, connate water saturation, and average water saturation in the
water swept portion of the reservoir. The procedure can be applied to both regular and
irregular patterns.
The method is illustrated with the use of a five-spot pattern; however, the method has
direct application for any regular or irregular pattern waterfloods after gas fillup. If a
free gas saturation does not exist at the start of injection (implementation of water
injection when reservoir pressure is above the initial bubble point pressure), the
technique may be applicable from the start of initial injection.
Figure 9-33 shows waterflood saturations early in the life of a waterflood prior to gas
fillup. The conventional water zone, oil zone, and unaffected part of the reservoir are
shown. These regions and saturations have been discussed in detail in Chapters 4 and
5 and References 27 and 28.

( S w S wbt

The oil saturation in the water zone is

before and at breakthrough) where

Sw

is estimated from waterflood

fractional flow theory. The oil saturation in the oil bank is

1 S wc

saturation in the unaltered region can be calculated using Eq. 3.9.

9 - 59

1 Sw

and the oil

Figure 9-33
Water and Oil Banks and Unaltered Zone
Prior to Fillup in a Single Layer of a 5-Spot Pattern

Producer
Unaltered Zone

Oil Bank

Injected
Water Bank
Injector

So 1.0 S wbt

So 1.0 S wc

So 1.0 S g S wc

When the leading edge of the oil zone (or oil bank) reaches the producing well, gas
fillup is achieved as depicted in Figure 9-34.

9 - 60

Figure 9-34
Water Bank and Oil Bank
At Gas Fillup in a Single Layer
of a 5-Spot Pattern
Producer
Oil Bank

Injected
Water Bank

Injector

So 1.0 S wbt

So 1.0 S wc

Figure 9-35 depicts the position of the water zone and oil zone after water
breakthrough. Note that after gas fillup, the oil remaining in the reservoir resides in
the water zone and oil zone.

This observation allows for the development of a

volumetric material balance equation on the reservoir oil which can be solved to
determine volumetric sweep efficiency of the injected water.

9 - 61

Figure 9-35
Oil and Water Banks in a Single Layer
after Water Breakthrough
Producer
Oil Bank

Water Bank

Injector

S o= 1.0 - Sw

S o= 1.0 - Swc

Figure 9-36 depicts a multi-layered reservoir prior to reservoir fillup. In Figure 9-36
certain layers have reached gas fillup but the reservoir as a whole has not achieved
fillup.

9 - 62

Figure 9-36
Stratified Reservoir Showing Flood Front Banks
before Reservoir Fillup

PRODUCER

INJECTOR
Oil
Bank

Unaltered
Zone
Injected
Water
Bank

With continued injection, gas fillup is obtained in all significant layers possessing
permeability and porosity values greater than the net pay cutoff values as depicted in
Figure 9-37. (Reservoir fillup can be characterized by a period of stable producing
GOR after falling from a higher GOR value early in the life of the waterflood project.)
Figure 9-37 shows that water breakthrough has occurred in some layers and water
breakthrough has not occurred in other layers.

The volumetric sweep efficiency

technique proposed in the Cobb and Marek paper assumes that gas fillup has been
9 - 63

achieved in all layers. Also, the oil remaining in the reservoir is located in the water
swept portion of the reservoir and the oil bank portion of the reservoir as illustrated in
Figure 9-37.

Figure 9-37
Stratified Reservoir after Reservoir
Fillup
Oil
Bank

INJECTOR

PRODUCER

Injected
Water
Bank

With the above mentioned assumptions, it is possible to write the following material
balance:

9 - 64

(Oil in place at start of waterflooding) = (Produced oil since the start of injection)
+ (Oil currently in reservoir)

(Eq. 9.25)

where:

(Oil in place at start of waterflooding) =

V p So

(Produced oil since the start of injection) =

, STBO

(Eq. 9.26)

N p , STBO

(Eq. 9.27)

Bo

Following reservoir gas fillup as illustrated in Figure 9-37, it is possible to show that:
(Oil currently in reservoir) = (Oil in water bank) + (Oil in oil bank)
and

(Oil in water bank) =

(Oil in oil bank) =

V p Evw (1.0 S w )
Bo

, STBO

V p (1.0 Evw )(1.0 S wc )


Bo

(Eq. 9.28)

, STBO

Substituting Eqs. 9.26, 9.27, 9.28, and 9.29 into Eq. 9.25 and solving for

N p Bo
Evw
If

Vp

(Eq. 9.29)

Evw leads to:

1.0 So S wc
(Eq. 9.30)

S w S wc

S w can be assumed constant as in the case of piston-like displacement, Eq. 9.30

becomes the equation of a straight line relating


9 - 65

Evw to N p .

That is:

Evw G HN p

(Eq. 9.31)

where

Sg
1.0 S wc So

S w S wc
S w S wc

(Eq. 9.32)

Bo
V p ( S w S wc )

(Eq. 9.33)

and

Equation 9.31 shows that the volumetric sweep of the injected water,
function of the oil production, N p . Further,

Evw can be computed at any time after

Bo , Swc , S w , So , and V p can be estimated.

gas fillup if

negligible changes in

Bo

mobility ratio waterfloods,


be approximated by

Sor

Sw can be approximated by two methods.

S wbt

breakthrough and increases from

Sw

First,

Sw

can

obtained from waterflood fractional flow theory as

S w can be approximated by 1 Sor

is the waterflood residual oil saturation. In fact,

approximate

The technique assumes

during the life of the waterflood. For moderate to low

described in Figure 4-20 of Chapter 4. Second,


where

Evw , is a linear

Swbt

to

1.0 Sor .

Sw

changes after water

Also, it may be possible to

S w by the following relationship:


S wbt (1.0 Sor )
2

(Eq. 9.34)

9 - 66

For these conditions, after gas fillup,

Evw can be graphed as a function of N p using

historical production data and Eq. 9-30.

If ultimate oil recovery under current

waterflood operations can be estimated from decline curve analysis, the volumetric
sweep efficiency plot, or Eq. 9.30, can be used to compute ultimate volumetric sweep
efficiency.

Further, the production benefits resulting from increasing volumetric

sweep efficiency through alternate operations can be quantified if the increase in


volumetric sweep can be estimated. The volumetric sweep procedure is illustrated
with the aid of an example.
_____________________________________________________________________

Example 9:4
A large multi-layered sandstone reservoir is under waterflood. At the start of injection,
the reservoir pressure was below the initial bubble point pressure. However, the
waterflood has responded favorably. Cumulative oil recovery to date since the start of
water injection is 40,000 MSTBO.

Production decline curve analysis indicates

remaining oil recovery under current operations is 5,000 MSTBO. Key reservoir fluid
saturations and oil properties at the start of water injection, floodable pore volume, and
production values since the start of water injection are summarized below.

9 - 67

TABLE 7

VOLUMETRIC SWEEP ANALYSIS


Conditions at Start of Waterflood
Connate Water Saturation

= 22 percent

Gas Saturation

= 8 percent

Oil Saturation

= 70 percent

Oil Viscosity

= 0.3 cp

Residual Oil Saturation

= 31 percent

Oil Formation Volume Factor

= 1.57
Total Unit

Pore Volume

= 350,000 MB

Cumulative Oil Production since Start of Injection

= 40,000

Current Volumetric Sweep Efficiency (Eq. 9.30)

= 0.552

Remaining Oil Production under Current Operations


Estimated Waterflood Ultimate Recovery

= 5,000 MB
= 45,000

Ultimate Volumetric
Operations (Eq. 9.30)

Sweep

Efficiency

under

Current = 0.600

The operator of the project desires to know if there is an opportunity to increase


recovery beyond that projected by decline curves through an infill drilling or an
injection well realignment program. Application of Eq. 9.30 provides possible answers
as to the remaining recovery potential. Using historical production data obtained after
gas fillup, volumetric sweep efficiency of the reservoir was computed. The volumetric
sweep efficiency values are summarized in the following table and are presented in
Figure 9-38.

9 - 68

TABLE 8
HISTORICAL PRODUCTION AND VOLUMETRIC SWEEP EFFICIENCY
Cumulative Oil Production, MSTB

Volumetric Sweep Efficiency

30000

0.46

35000

0.50

40000

0.55

Figure 9-38

Volumetric Sweep Efficiency for Waterflood Project


(Pore Volume Based on 6.0% Porosity Cutoff)
1

26.0 MMSTB

Evw 0.85
0.8

0.6

0.4

Cumulative Oil Production


= 40.0 MMSTB
Remaining Oil Production
= 5.0 MMSTB
Estimated Ultimate Recovery = 45.0 MMSTB

0.2
0

10

20

30

40

50

60

70

80

Production Since Start of Waterflood, Np, MMSTB

9 - 69

The calculated current volumetric sweep is 55.2 percent and the calculated ultimate
volumetric sweep efficiency under current operations is approximately 60 percent. This
means that at the economic limit under current operations, 40 percent of the reservoir
will not have been swept by the injected water. Moreover, conventional waterflooding
theory27,28 indicates that the oil saturation in this unswept part of the reservoir will be
equal to

1.0 Swc .

To determine the potential for a re-engineered waterflood, including infill drilling


and/or pattern realignment, a waterflood model was used to estimate the volumetric
sweep efficiency of a well-managed project. Results of the model work indicated
ultimate volumetric sweep efficiency values of approximately 85 percent could be
achieved at the WOR economic limit. Extrapolation of the line in Figure 9-38 (or using
Eq. 9.30) to 85 percent volumetric sweep indicates ultimate recovery could be increased
from an expected value of 45,000 MSTBO to approximately 71,000 MSTBO, an
increase in recovery of 26,000 MSTBO.
Engineering and geological judgment caused the operator to believe that the additional
26,000 MSTBO was too high. Therefore, Eq. 9.30 was evaluated in more detail. In
particular, the accuracy of the factors used in Eq. 9.30 such as floodable pore volume
and average water saturation,

S w , were reassessed.

Sensitivity Analysis
Net pay determination. The floodable pore volume of 350,000 MBO as shown in the
Table 7 was originally computed using a six percent porosity cutoff. While this value
may accurately depict the oil-in-place that contributes to primary production, the six
percent porosity cutoff value was deemed too low to compute net pay for the
waterflood project. A reevaluation of net pay under water flood operations using the
water-cut methodology described in Chapter 3, indicated a permeability cutoff which
translated to a porosity cutoff of 10 percent. Using this higher cutoff value, revised
floodable pore volume calculations resulted in approximately a 25 percent reduction in
9 - 70

pore volume.

With the revised pore volume, new estimates of volumetric sweep

efficiency were computed for this sandstone reservoir. Figure 9-39 presents these
revised calculations.

It indicates that volumetric sweep efficiency under current

operations to recover 45,000 MSTBO is projected to be 74 percent. Therefore, from


Figure 9-30, it can be determined that if infill drilling or pattern realignment can be
implemented to increase sweep efficiency to 85 percent, an additional 8,400 MSTBO
could be obtained giving a remaining ultimate recovery of 13,4000 MSTBO (8,400 +
5,000).

9 - 71

Figure 9-39

Volumetric Sweep Efficiency for Waterflood Project


(Pore Volume Based on 6.0% and 10.0% Porosity Cutoff)
1
26.0 MMSTB
8.4
MMSTB

Evw 0.85
0.8

0.6
10% Porosity Cutoff

6% Porosity Cutoff

0.4

Cumulative Oil Production


= 40.0 MMSTB
Remaining Oil Production
= 5.0 MMSTB
Estimated Ultimate Recovery = 45.0 MMSTB

0.2
0

10

20

30

40

50

60

70

80

Production Since Start of Waterflood, Np, MMSTB

Average Water Saturation.

The oil viscosity for this reservoir at the start of

waterflood (and throughout the life) is 0.3 cp. This low oil viscosity results in a
favorable mobility ratio and a fractional flow graph which gives piston-like
displacement.

Therefore the average water saturation,

estimated to be

1.0 Sor .

Sw ,

for the analysis is

Core data indicate the residual oil saturation is 31


9 - 72

percent, yielding an average water saturation in the water swept portion of the reservoir
of 69 percent. This 69 percent value is used in Figures 9-38 and 9-39. However, some
limited data indicated the residual oil saturation could be as high as 36 percent, giving
an average water saturation in the swept zone of only 64 percent.
Using a floodable pore volume which has been reduced by 25 percent (based on the 10
percent porosity cutoff) and an average water saturation of 64 percent in the swept
zone,

Evw values were recomputed from Eq. 9.30 and results are presented in Figure

9-40.

Figure 9-40

Volumetric Sweep Efficiency for Waterflood Project


(Pore Volume Based on 10.0% Porosity Cutoff)
1

8.4
MMSTB

Evw 0.85
0.8

0.6
Residual Oil
Saturation = 36%
Residual Oil
Saturation = 31%

0.4

Cumulative Oil Production


= 40.0 MMSTB
Remaining Oil Production
= 5.0 MMSTB
Estimated Ultimate Recovery = 45.0 MMSTB

0.2
0

10

20

30

40

50

60

Production Since Start of Waterflood, Np, MMSTB

9 - 73

For these new conditions, the current volumetric sweep efficiency is 76 percent and that
the sweep efficiency at a recovery of 45,000 MSTBO will be 84 percent. If the
maximum practical volumetric sweep efficiency is only 85 percent, there appears to be
very little recovery to be gained through infill work.
This example illustrates an application of the method on a field wide basis. However,
the method could be applied to a smaller area of the field such as a single pattern or a
group of patterns.
______________________________________________________________________
For unfavorable fractional flow graphs in which piston displacement is not valid,
volumetric sweep efficiency is not a linear function of cumulative oil production as
assumed in the previous example. For non-piston displacement, the average water
saturation in Eq. 9.30 is a function of cumulative water injected (and therefore time). In
this situation, it is more difficult to determine average water saturation.

As an

approximation, it may be possible to take the water cut from the pattern or area of the
field under study and relate that to an average water saturation from the fractional flow
curve using the tangent line methodology discussed in Chapter 4. If average water
saturation can be computed as a function of water cut (or WOR), then volumetric sweep
can be computed versus WOR. Further, if future oil recovery can be estimated from
WOR versus

Np

plots, future values of volumetric sweep can be computed at future

WOR values.
The following observations can be made regarding this volumetric sweep method.
A procedure is presented that allows for the determination of volumetric sweep
efficiency of a waterflood based upon waterflood principles after gas fillup for
both regular and irregular patterns, including the effects of water influx..
The technique utilizes data that can be estimated from oil production records,
geological information, core analysis, and fluid property data.
9 - 74

The method provides a basis for computing current and ultimate volumetric sweep
efficiency under existing waterflood operations using only oil production data.
The procedure requires reliable estimates of floodable pore volume and average
water saturation in the water swept zone.

It is likely that in most field

applications, floodable pore volume (as opposed to primary depletion pore


volume) is the greatest source of potential error for this analysis.
VI. Injection Profile Testing
It has been shown in previous chapters that oil reservoirs are generally stratified.
Seldom does nature provide a reservoir which can be characterized as a single
homogenous layer.

Stratification can create major problems in waterflooding

operations and lead to low vertical sweep efficiency.


Waterflood prediction techniques assume all injection water enters the various layers
in proportion to the flow capacity (actual injection is related to layer skin factor,
radius to water and oil, areal sweep of the injected water, and mobility ratio as well as
flow capacity) of each layer. In other words:

(Injection Rate)i-Layer

(kh)i-Layer
(kh)All-Layers

Unfortunately, this natural injection behavior creates a problem.

(Eq. 9.35)

Thin, high

permeability intervals serve as highly conductive streaks or channels for the injected
water. These channels prevent efficient flooding of the other zones. Core and log
analysis may give valuable clues as to the probability of this occurrence, but in many
instances, injection history is required to verify channeling. Accordingly, injection
water seeks the zones of highest permeability, but to flood all layers simultaneously,
water should be injected into each layer based upon the displaceable hydrocarbon
pore volume,

hShc .

For ideal injection behavior, layer injection rate should

conform to:
9 - 75

(Injection Rate)i-Layer

( hShc )i-Layer
( hShc )All-Layers

(Eq. 9.36)

where

= effective layer porosity, fraction

= effective layer thickness, feet

Shc

1 Swc Sor

For many reservoirs, the

Shc

value can be assumed constant for each layer.

Exceptions include thick reservoirs of low permeability which possess large water
saturation transition zones as governed by capillary pressure relationships or zones
with significantly different values of

Sor .

If

Shc

can be assumed constant for all

layers under investigation, Eq. 9.36 reduces to:

(Injection Rate)i-Layer

( h)i-Layer
( h)All-Layers

(Eq. 9.37)

A comparison of Eqs. 9.35 and 9.37 more clearly explains the stratification problem.
Injection rates tend to be governed by Eq. 9.35, but to achieve uniform flood front
movement in all layers and to maximize vertical sweep efficiency at low WOR
values, injection rates should conform to Eq. 9.37. Figures 9-41 and 9-42 depict
waterflood front position when injection conforms to flow capacity and displaceable
pore volume, respectively.

It is relatively clear that maximum vertical sweep

efficiency and recovery efficiency are achieved when the injection rate into each zone
is in proportion to hShc of the zone.

9 - 76

FIGURE 9-41
WATER INJECTION INTO A STRATIFIED SYSTEM
BASED ON ZONE kh
kh
percent of total
40

30

15

10

FIGURE 9-42
WATER INJECTION INTO A STRATIFIED SYSTEM
BASED ON ZONE PORE VOLUME
hShc
percent of total
14

26

25

18

17

9 - 77

VII. Interval Selection For Waterflood Monitoring


When monitoring a field flood, it is desirable to identify several major porosity
intervals within a given injection well which are correlatable to adjacent producing
wells. In theory, it may be possible to identify a large number of these intervals.
From a surveillance consideration, it is usually satisfactory to combine these layers so
as to treat the well as if it contains only five or six layers. If a thin, high permeability
streak is present, it may be desirable to treat this as a single layer. Normally, the five
or six zones are selected using natural geological zone boundaries such as shales or
dense non-pay intervals. Figure 9-43 presents a porosity log for a thick, dolomite San
Andres interval of West Texas. Six intervals identified as M-1 through M-6 are
shown. It is assumed that

Shc

is constant for all zones. The average

each layer is shown.

9 - 78

for

43

9 - 79

VIII. Injection Profiles


One of the most important measurements conducted in a water injection well is an
injection profile. Figure 9-44, with the perforations and profiles, permits a visual
assessment of the distribution of the injected fluid between perforation sets as well as
within a set. Injection profiles identify perforation sets that are not taking injected
water, possibly because the perforations are plugged. Also, profiles reveal contrasts
of injectivity within a perforation set. High concentration of injected water over small
intervals indicates zones of high permeability which are frequently called thief zones.
Failure to correct or alter profiles within thief zones will result in early water
breakthrough and will usually result in poor vertical sweep, lower oil recovery, and
undesirable surface water production.

FIGURE 9-44
TYPICAL INJECTION PROFILE
Injection Rate Percentage Analysis
4900

0%

25%

50%

25%
50%
10%
15%

5000
0%

9 - 80

75%

100%

Figures 9-45, 9-46, and 9-47 present useful bar charts which allow the operations
engineer a visual means of comparing actual injection profiles with ideal profiles
based on hShc .

Figure 9-45
INJECTION PROFILE HISTORY FOR EXAMPLE WELL
IDEAL PROFILE

Z-1

20%

Z-2

30%

Z-3

18%

32%

Z-4
0
Initial Inj.
7-1-98
Date ______________

25
iw = ___________

50
Pwi = ___________psi

IDEAL Based on hS hc
Inj. Profile ___________________________________________________

9 - 81

Figure 9-46
INJECTION PROFILE HISTORY FOR EXAMPLE WELL
ACTUAL vs IDEAL 7/1/01
12

Z-1

20%

Z-2

44%

Z-3

34%

10
32%

Z-4
0
7-1-01
Date _______

25

50

1000
800
625
iw = ______B/D
Pwi = ______psi
Cum Inj. ________
MBW

0
Previous Inj. Survey _____________
Cum Inj @ Last Survey _______
MBW
1000
Cum Inj. Since Last Survey __________MBW

9 - 82

Figure 9-47
INJECTION PROFILE HISTORY FOR EXAMPLE WELL
ACTUAL vs IDEAL 10/1/04
16

Z-1

20%
26

Z-2

30%

Z-3

34%

32%

Z-4
0
Date 10-1-04
_______

25

50

1900
600
770
iw = ______B/D
Pwi = ______psi
Cum Inj. ________
MBW

7/1/01
1000 MBW
Previous Inj. Survey _____________
Cum Inj @ Last Survey _______
900
Cum Inj. Since Last Survey __________MBW

Various types of profile measuring devices exist, from radioactive injection


techniques to continuous spinner surveys. These services are provided by many
service companies. The cost per survey depends on such factors as location, well
depth, and injection pressure, but the cost is usually not expensive.

9 - 83

When monitoring a waterflood, it is recommended that an injection profile be


conducted every six months during the first two years of a wells injection life. After
two years, surveys should be conducted annually.

After remedial work on an

injection well, it is recommended that a survey be conducted 30 to 60 days after


completion of the work program. The major objective of an injection survey is to
provide a means of monitoring the injection water and to ensure injection distribution
conforms with zonal

hShc .

Also, by measuring the injection volume into each

zone, flood front maps (bubble maps) can easily be constructed. Injection profile tests
will pay substantial dividends as they lead to increased vertical sweep and ultimate
recovery. Perhaps the single most valuable tool for monitoring waterflood operations
is the injection survey. It tells us where the injected water is entering the formation.
Without this knowledge, it becomes very difficult (impossible?) to perform
meaningful production or reservoir engineering calculations using either analytical or
numerical simulation procedures.
IX. Alteration of Injection Profiles
It is beyond the scope of this course to discuss all the methods of altering or changing
injection profiles. Most service companies are capable of providing technical expertise
on this issue. Some of the techniques which have been successfully used in the industry
include:
selective perforating
low pressure squeeze cementing
acidizing
thief zone blockage through injection of fine grain sand
thief zone blockage through injection of certain polymers

9 - 84

Publications by Seright, et al29,30,31,32,33 provide an excellent discussion on the use of


polymers for injection profile management.
It is important to note that the benefits of profile modification may not be observed for
several months after the work. The real benefit must be determined at the production
well. The time required to see changes in WOR trends or water production rate at
producing wells is usually measured in months or years rather than days. Nevertheless,
it should be a daily goal of the surveillance engineer to ensure that the long term
injection strategy is to inject into each layer in proportion to each layers hShc .
Finally, action steps implemented during a waterflood surveillance and management
program such as injection profile modifications, should be considered in view of how
they may affect alternate recovery operations such as

CO2 injection.

X. Flood Front (Bubble) Maps


A waterflood surveillance tool which can be useful in monitoring flood behavior is a
flood front map or bubble map. The bubble map is a pictorial display showing the
location of the various oil and water flood fronts. The bubble maps allow visual
differentiation between areas of the reservoir that have been swept by injected water
and those areas that have not been swept.
Early in the life of an injection well, it can usually be assumed that water and oil banks
are radial in shape around the injector as shown in Figure 9-48

9 - 85

FIGURE 9-48
RADIAL WATER AND OIL BANKS

OIL

re

WATER

The outer radius of the oil bank can be computed from Eq. 4.37.

1
5.615Wi Einj 2

re

h S g

where:

re

= radius of oil bank, feet

= porosity, fraction

= interval thickness, feet


9 - 86

(Eq. 4.37)

The

Wi

thickness

Sg

= gas saturation at start of injection, fraction

Wi

= cumulative water injected, bbls

Einj

= layer injection efficiency

represents the apparent cumulative water injection into a geological layer

h.

The apparent volume of water allocated to each layer is based on

injection surveys and/or rock properties of the layer such as permeability and porosity.
Further, based on numerical simulation history matches of actual waterflood
performance and from field observations, it is generally recognized that actual
production performance is less than anticipated performance.

This difference is

frequently attributed to a deficiency in injection. In short, actual performance behaves


as if a portion of the injected water is lost. The layer injection efficiency,

Einj , takes

into account this lost water phenomenon. Injection efficiency values generally range
from 0.5 to 0.9 but they are difficult to precisely quantify. In thick, heterogeneous, low
permeability formations,

Einj

tends to be low. Personal experience suggests that a

value of 0.50 to 0.75 may be expected in carbonate reservoirs, whereas for thin, clean
and continuous sands, a value of 0.75 to 0.90 could be expected. Further, it has been
observed that injection efficiency into the pay zones may increase with time. For
example, early in the life of a waterflood, the injection efficiency may be in the order of
40 to 50 percent whereas by the time fillup occurs in most zones, the overall efficiency
may increase to 70 to 80 percent. The injection efficiency will be dependent on
reservoir conditions including the presence of natural or induced fractures, injection
above fracture pressures, presence of leaking faults, gas cap injection, loss to inactive
aquifers, the presence of porous but laterally discontinuous zones, and mechanical
integrity of the injection well bores.

9 - 87

If all of the injection water is assumed to remain within the water bank, the water bank
radius can be estimated from Eq. 4.41:

1
2

Sg

r re

S wbt S wc

(Eq. 4.41)

where:

= radius of water bank, feet

Swbt

= average water saturation behind front, fraction

S wc

= connate water saturation, fraction

Inspection of both Eqs. 4.37 and 4.41 indicates that distance to the various fronts is
dependent on the gas saturation,

Sg .

Unfortunately, the gas saturation in some

reservoirs cannot be accurately calculated. Yet, even in these situations, it is possible to


compute

using the following water material balance relationship present in Eq. 4.38.

1
2

5.615 Wi Einj
r

h
S
S

wbt
wc

For approximate calculations,

(Eq. 4.38)

S wbt 1.0 Sor


1
2

5.615 Wi Einj

h 1.0 Sor S wc
where:

9 - 88

and:

(Eq. 9.38)

Sor

= residual oil saturation to water injection, fraction

Bubble maps should be updated each six months for the first two years of a flood and
on an annual basis thereafter. Further, bubble maps should be constructed on each of
the major identifiable and correlatable intervals as previously discussed.
The use of bubble maps provides certain obvious benefits. First, they graphically
delineate portions of the reservoir that need additional injection or curtailed injection.
Secondly, bubble maps help identify areas within the reservoir that would support infill
drilling opportunities. After water breakthrough, front shapes are more difficult to
construct.
Frequently, bubble maps are manually constructed. However, in larger fields, the
procedures are computerized. In some instances, bubble maps are generated using
numerical or front tracking simulation models. Simulation models provide certain
advantages over the conventional manual graphs because they can more easily account
for variations in variables which effect fluid movement including porosity, thickness,
directional permeability, irregular well locations, and pressure differentials between
wells, incorporate the effects of offset wells, and account for flood front movement
after water breakthrough in offset wells.
The utility of profile management and bubble map monitoring early in the life of an
injection well is illustrated with an example.
______________________________________________________________________

Example 9:5
Bubble Map Construction
An injection well with injection profiles described in Figures 9-45, 9-46, and 9-47 has
been subdivided into four zones (Z-1, Z-2, Z-3 and Z-4) for surveillance purposes.
Given the following rock properties and fluid saturations, estimate the radius to the
water and oil banks as of 7/1/01 when 1,000 MB of water had been injected.
9 - 89

Sg

= 14 percent

S wc

= 25 percent

Swbt

= 55 percent

Einj

= 80 percent

Layer Porosity, percent Net Thickness, feet


Z-1
Z-2
Z-3
Z-4

24.4
26.4
18.0
15.6

18.0
25.0
22.0
45.0

Date: 7/1/01
Layer
Z-1
Z-2
Z-3
Z-4

Injection

Water Injection

12.0
44.0
34.0
10.0

120.0
440.0
340.0
100.0

re*(Eq. 4.37)
feet
528.0
825.0
936.0
382.0

r(Eq. 4.41)
360.0
564.0
639.0
261.0

* Einj = 80%
The flood fronts for each layer are presented graphically in Figure 9-49

9 - 90

FIGURE 9-49
FLOOD FRONTS FOR FOUR GEOLOGICAL ZONES
AS OF 7/1/01
1000 FEET

re

Z-1

528

Z-2

825

Z-3

936

Z-4

382
0

1,000

______________________________________________________________________
XI. Injection Analysis
This section provides a general method23 for analyzing water injection behavior in a
mature waterflood. The displaceable hydrocarbon saturation,

DHS , is defined as:

DHS 1 S wc Sor

(Eq. 9.39)

where

Swc Swir , fraction

S wc

= connate water saturation,

Sor

= residual oil saturation to waterflooding, fraction

Further, the displaceable hydrocarbon pore volume,

9 - 91

DHPV , is defined as:

DHPV V p (1 S wc Sor )

(Eq. 9.40)

where

Vp

= floodable pore volume, barrels

______________________________________________________________________

EXAMPLE 9:6
At the start of waterflooding, the oil, gas, and connate water saturations are 55 percent,
10 percent, and 35 percent respectively. The irreducible water saturation is 22 percent
and the residual oil saturation is 38 percent. The floodable pore volume is 20 million
barrels. Compute the

DHPV .

DHPV V p (1 S wc Sor )
Substituting:

DHPV 20MMB * (1 0.35 0.38)

DHPV 5.4MMB
______________________________________________________________________
Consider a bounded or closed reservoir or pattern in which:
there is no influx or efflux out of the pore volume
there is no loss of injected water to non-oil pay zones
there is negligible change in the oil formation volume factor.
For these assumptions at reservoir conditions, one net barrel of injected water results in
one net barrel of produced hydrocarbons. Moreover, the maximum net water which
9 - 92

can be injected during the waterflood operation is equal to the displaceable hydrocarbon
pore volume. That is:

(Wi W p Bw ) MAX DHPV

(Eq. 9.41)

where

Wi

= cumulative injected water, BW

Wp

= cumulative produced water since the start of injection, BW

Bw

= produced water formation volume factor, RB/STBW

Define the net displaceable hydrocarbon pore volume injected, VD , as the ratio of net
water injected to the displaceable hydrocarbon pore volume. That is:

VD

net water injected, BBL


displaceable hydrocarbon pore volume, BBL
(Eq. 9.42)

or

VD

Wi W p Bw
V p (1 S wc Sor )

A. Analysis Without Free Gas

(Eq. 9.43)

( S g 0)

At the start of injection (when Wi = 0), the produced oil at that instant is the result
of a primary drive mechanism and is denoted by a primary recovery factor,

( RF ) p .
9 - 93

Also at the start of injection

( E A EV 1.0)

VD

is 0.0. Conversely, when the entire reservoir

incremental recovery factor attributable to injection is denoted by

( RF ) I

Sor ,

is waterflooded to the residual oil saturation,

( RF )i

So Sor
So

the

where

(Eq. 9.44)

where

So = oil saturation at the start of waterflooding, fraction (1.0 S wc if


S g 0.0 )
When the reservoir is waterflooded to residual oil, the value of
Therefore, during a waterflood,

VD

is 1.0.

RF ranges from ( RF ) p to ( RF ) F where

( RF ) F ( RF ) p ( RF ) I

(Eq. 9.45)

Because each net barrel of injected water results on one barrel of produced oil when
measured at reservoir conditions, a coordinate graph of

RF

versus VD will yield

a straight line. This type of graph has been called a conformance plot in References
8 and 9. References 24 and 25 refer to these graphs as a Staggs plot because such
plots were first presented in Reference 23.

9 - 94

_____________________________________________________________________

Example 9:7
An oil reservoir whose reservoir pressure is above the initial bubble point pressure

( S g 0)
injection,

is to be waterflooded. The primary oil recovery factor at the start of the

( RF ) p , is 4 percent.

The connate water saturation is 35 percent (the

irreducible saturation is 25 percent) and the waterflood residual oil saturation is 30


percent. Assume that

Bo

is essentially constant during the waterflood. Plot

RF

versus VD for this reservoir.


The final (or maximum) waterflood recovery factor,
Eq. 9.45 as :

( RF ) F ( RF ) p ( RF ) I
from Eq. 9.44

( RF ) I

1 S wc Sor
1 S wc

( RF ) I

1 0.35 0.30
1 0.35

and

( RF ) I 0.54
Finally:

( RF ) F 0.04 0.54
9 - 95

( RF ) F , can be computed from

( RF ) F 0.58
For

( RF ) p of 0.04 (four percent), the value of VD

= 0 and for

( RF ) F of 0.58,

the value of VD = 1.0. Results are shown in Figure 9-50.

Figure 9-50
Ideal Reservoir Recovery Factor
vs
Net Displaceable Pore Volumes Injected
( RF ) F

0.6
0.5
0.4
0.3
0.2
0.1

( RF ) p

0
0

0.2

VD

0.4

0.6

0.8

Wi W p Bw
V p (1 S wc Sor )

_____________________________________________________________________
9 - 96

Once the

RF

versus VD plot has been constructed, actual production and injection

data can be plotted on the same graph and a comparison with the theoretical
performance can be made. Reference 23 indicates that when field data are plotted on
the theoretical graph, the field data frequently deviate from the theoretical line. If it is
assumed that the theoretical line is correct (that is if

V p , S wc , Sor , and W p

are

correct), the departure of the field data from the straight line can be used to infer (1)
the loss of injection out of zone, (2) water influx, or (3) inter-pattern flow as
illustrated in Figure 9-51.

9 - 97

Figure 9-51
Injection Analysis Showing Effect of
Eflux/Influx into Pore Volume
0.6

0.5
Inflow of Oil or Water
from Outside of V p

0.4

Theoretical

0.3
Loss of Oil or Injected
Water Outside of V p

0.2

0.1

0
0

0.2

0.4

VD

0.6

0.8

Wi W p Bw
V p (1 S wc Sor )

_____________________________________________________________________

EXAMPLE 9:8
A waterflood project was initiated when the reservoir pressure was above the bubble
point pressure. Pressure has been maintained such that changes in fluid properties
9 - 98

( Bo

and

o ) are negligible.

At the present time, the cumulative injected water is

24.5 MMB and the cumulative produced water is 16.0 MMB. Compute the water

Eing ,

injection efficiency,

into the pore volume,

Vp .

Other reservoir data are

tabulated below.

Bw

= 1.0 RB/STB

Sor

= 36 percent

Bo

= 1.35 RB/STB

( RF ) p

= 6 percent

S wc

= 32 percent

( RF ) I

= 47.1 percent

So

= 68 percent

( RF ) F

= 53.1 percent

Vp

= 31,250 MBW

Using actual injection and production data, the calculated values of

RF

and VD are

tabulated below:
TABLE 9

The theoretical
values of

RF

RF
and

VD

RF

VD

RF

0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40

0.060
0.072
0.093
0.115
0.137
0.154
0.164
0.173
0.191

0.45
0.50
0.55
0.60
0.65
0.70
0.75
0.80
0.85

0.207
0.219
0.230
0.242
0.254
0.269
0.275
0.290
0.300

versus

VD

VD

straight line is seen in Figure 9-52. The actual

are also plotted on Figure 9-52. Note the actual points fall

below the theoretical line. By comparing Figure 9-52 with Figure 9-51, it appears
that injected water is being lost from the pore volume,
9 - 99

Vp .

Consider the last field

data point. For a recovery factor of 0.30, the ideal value of

VD

is 0.51 whereas the

calculated value of VD is 0.85. If it is assumed the error is related to the cumulative


injection, Wi , then it is possible to write:

VD

Eing * Wi W p Bw
V p (1 S wc Sor )

or

0.51

Eing (24,000MB 16,000MB)


31, 250MB(1 0.32 0.36)

or

Eing 0.861
This analysis illustrates that 14 percent of the injected water is being lost outside of
the pore volume, V p .

9 - 100

Figure 9-52
Recovery Factor
vs
Net Displaceable Pore Volumes Injected
0.6

0.5

0.4

Theoretical
0.3

0.2

Actual

0.1

0
0

0.2

0.4

VD

0.6

0.8

Wi W p Bw
V p (1 S wc Sor )

______________________________________________________________________
B. Analysis With Free Gas

( S g 0)

The previous paragraphs illustrate application of the

RF

versus

VD

graph when

there is negligible gas. As mentioned throughout this manual, it is likely that at the
start of injection most projects possess a free gas saturation. Accordingly, this
9 - 101

generalized injection analysis graph needs to be modified to account for free gas. If
the free gas saturation at the start of injection is

S g , the net volume of injected

water necessary to reach fillup is governed by Eq. 4.33 that states:

Wif V p S g

(Eq. 4.33)

Therefore, for a homogeneous reservoir, the value of

VD

at gas fillup is obtained

by substituting Eq. 4.52nto the numerator of Eq. 9.43 which gives:

VDF

VpSg

(Eq. 9.46)

V p (1 S wc Sor )

or

VDF

Sg

(Eq. 9.47)

1 S wc Sor

Assume that oil production during the fillup period is negligibly small. During
reservoir fillup,

RF

VD

increases from 0 to

is constant and is equal to

VDF

( RF ) p .

and with negligible oil production,

Once fillup is obtained, then for each net

barrel of injection, there will be one barrel of oil produced when measured at
reservoir conditions. As a result, after fillup, there is a linear relationship between

RF
value,

and

VD .

Finally, when

VD

becomes unity (1.0), the

( RF ) F , which is governed by Eq. 9.45.

9 - 102

RF

reaches its final

_____________________________________________________________________

Example 9:9
For the reservoir system described in Example 9:8, assume that at the start of
injection, a gas saturation of 10 percent is present. Construct a
graph. Assume that before and after fillup, the change in
At the start of injection,

( RF ) p is 6 percent and

So 1 S wc S g

So 1 0.32 0.10
So 0.58
The final recovery factor from Eq. 9.45 is:

( RF ) F ( RF ) p ( RF ) I
from Eq. 9.44

( RF ) I

So Sor
So

( RF ) I

0.58 0.36
0.58

or

( RF ) I 0.379
9 - 103

RF

versus

Bo is negligible.

VD

Substituting into Eq. 9.45:

( RF ) F 0.060 0.379
( RF ) F 0.439 or 43.9%
[As a side note, observe that the

( RF ) F

( RF ) F

with a gas saturation is less than the

without gas (43.9% versus 53.1%).]

Next, compute VD at fillup from Eq. 9.47.

VDF
VDF

Sg
1 S wc Sor
0.10
1 0.32 0.36

VDF 0.312
Figure 9-53 is a graph of

RF

versus

VD .

For comparison, the

RF

line for the case of no gas saturation is shown on the upper dashed line.

9 - 104

versus

VD

FIGURE 9-53
Recovery Factor
vs
Net Displaceable Pore Volumes Injected
0.6

0.5

0.4

Sg 0

0.3

0.2

S g 10%

0.1

0.312

0
0

0.2

VD

0.4

0.6

0.8

Wi W p Bw
V p (1 S wc Sor )

_____________________________________________________________________
The solid line of Figure 9-53 assumes a homogeneous reservoir.

In practice,

reservoirs are comprised of multiple layers of varying pore volumes and permeability.
As a result, fillup of the most permeable layer occurs first with fillup of the tight
zones occurring last. As a result, the
reservoir fillup.

RF

typically increases prior to reaching full

The following actual case study illustrates the effects of

stratification.

9 - 105

_____________________________________________________________________

EXAMPLE 9:10
Injection Efficiency Calculation
An oil reservoir has been waterflooded for 20 years. Given the tabulated data,
construct the RF versus VD graph, and compute the injection efficiency.
OOIP

49,638 MSTBO (at bubble point)

S wc

20 percent

So

64 percent (start of waterflood)

Sg

16 percent

Sor

36 percent

Bo

1.15 RB/STB (start of waterflood)

Bob

1.27 RB/STB

N pp

5,771 MSTBO (start of waterflood)

( RF ) p

11.6 percent (start of waterflood)

9 - 106

TABLE 10
YR

Wi ,MBW

W p ,MBW

N p ,STBO

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

603
2,264
3,952
5,515
7,353
9,238
12,081
14,791
17,235
19,918
22,667
25,347
27,740
29,882
32,760
35,68
37,587
40,071
42,593
44,210

0
1
1
8
22
56
93
176
382
721
1,205
1,742
2,327
3,083
4,046
4,858
5,715
6,861
8,080
9,213

5,805
5,956
6,110
6,314
6,747
7,363
8,045
8,956
9,835
10,620
11,433
12,04
13,115
14,105
15,157
16,035
16,801
17,397
17,900
18,312

The

( RF ) F

from Eq. 9.45 is:

( RF ) F ( RF ) p ( RF ) I
( RF ) F 0.116

0.64 0.36
0.64

( RF ) F 0.554
At gas fillup, VDF from Eq. 9.47 is:

9 - 107

VDF
VDF

Sg
1 S wc Sor
0.16
1 0.20 0.36

VDF 0.364
These calculations can be summarized as follows:

RF ,%

VD

0.000
0.364
1.000

11.6
11.6
55.4

These three points define the horizontal and diagonal parts of the injection efficiency
graph shown in Figure 9-54.

9 - 108

FIGURE 9-54
Recovery Factor
vs
Net Displaceable Pore Volumes Injected
0.6

Recovery Factor, fraction

0.5

0.4

0.3

0.2

0.1

VDF 0.364

0
0.00

0.25

VD

Next,

RF

versus

VD

0.50

0.75

1.00

Wi W p Bw
V p (1 S wc Sor )

are computed from production records, shown below, and

plotted on the theoretical graph as illustrated in Figure 9-55

9 - 109

TABLE 11
YR

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

RF

VD

0.117
0.120
0.124
0.128
0.136
0.149
0.163
0.181
0.199
0.214
0.231
0.246
0.265
0.285
0.306
0.324
0.339
0.351
0.361
0.369

Wi W p Bw
V p (1 S wc Sor )
0.018
0.66
0.114
0.159
0.212
0.265
0.346
0.422
0.487
0.554
0.620
0.681
0.733
0.773
0.829
0.886
0.920
0.958
0.996
1.010

9 - 110

FIGURE 9-55
Recovery Factor
vs
Net Displaceable Pore Volumes Injected
0.6

Recovery Factor, fraction

0.5
0.4
0.3
0.2
0.1

VDF 0.364

0
0.00

0.25

VD

0.50

0.75

1.00

Wi W p Bw
V p (1 S wc Sor )

The data points tend to lie to the right of the diagonal line. After 20 years the actual

RF

is 0.369. It is expected that for this

RF , a value of VD

diagonal should be 0.73. The actual value of

VD

from the theoretical

is 1.010. If it is assumed the

discrepancy is attributable to injection water losses, it is possible to compute an


injection efficiency. Adjust the last data point in the following manner:

9 - 111

VD

Wi Einj W p Bw
V p (1 S wc Sor )

0.73 (for RF 0.369)

At 20 years:

44, 210 Einj 9, 213


78,800(1 0.20 0.36)

0.73

and

Einj 0.78
From this analysis, it can be concluded that at least 22 percent of the injected water is
lost to thief zones.
injection efficiency,

Re-compute

VD

by adjusting each value of

E inj , factor of 0.78 as illustrated below.

9 - 112

Wi

by the

TABLE 12
YR

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

RF

VD

0.78Wi W p Bw
V p (1 S wc Sor )

0.117
.120
0.124
0.28
0.136
0.149
0.163
0.181
0.199
0.214
0.231
0.246
0.265
0.285
0.306
0.324
0.339
0.351
0.361
0.369

0.014
0.051
0.089
0.124
0.165
0.207
0.270
0.328
0.377
0.428
0.476
0.520
0.557
0.584
0.621
0.661
0.681
0.704
0.725
0.729

Re-plot the adjusted data on the theoretical graph as shown in Figure 9-56. Finally, if
a straight line is constructed which connects the end point

RF s, an obtuse triangle

is created within which the data points should fall. It is noted in Figure 9-56 that the

RF

values begin to increase immediately. That is due to (1) primary recovery

during fillup not being negligible or (2) effects of stratification.

9 - 113

FIGURE 9-56
Recovery Factor
vs
Net Displaceable Pore Volumes Injected

Recovery Factor, fraction

0.6
0.5
0.4
0.3
0.2
0.1

VDF 0.364

0
0.00

0.25

VD

0.50

0.75

1.00

Wi W p Bw
V p (1 S wc Sor )

_____________________________________________________________________
C. Numerical Simulation
A theoretical graph of

RF

versus

VD

could be constructed using numerical

simulation which takes into account variables such as stratification, primary


production during fillup, and variations in S wc ,
9 - 114

Sg ,

and

Sor

between layers.

This plot would be nonlinear and concave upward until reservoir fillup but would
become a straight line after fillup.
XII. Water Testing Program
In the early days of waterflooding, only the quantity of water was considered, not
quality. As waterflooding projects increased in number and as waterfloods matured, the
petroleum industry began to notice that higher injection pressures were required to
maintain desired injection rates and that corrosion problems increased. More recently,
it has become clearly evident that water quality is as important as water
quantity34,35,36,37.

The absence of water quality control can prove disastrous to

waterfloods and can result in engineering and economic failure.


After a water source is identified, a water analysis is required to determine the
following.
compatibility with reservoir water
compatibility with swelling clays in the reservoir
whether a closed or open injection facility is needed
the treatment program necessary to have an acceptable water for the reservoir
and to minimize corrosion of the equipment
Water analysis in the injection system should be conducted on a regular basis, such as
quarterly or semi-annually. These analyses should be to determine the presence of
dissolved gases, minerals, microbiological growth, and dissolved solids (all undesirable
constituents of water).
A. Dissolved Gases
Dissolved gases such as hydrogen sulfide, carbon dioxide, and oxygen promote
corrosion.

Carbon dioxide and oxygen also contribute to the formation of

precipitates which cause reservoir plugging and thus lowers injection rates.
9 - 115

B. Microbiological Growth
The control of colonies of one-celled animals and plants is of concern to operators
attempting to maintain suitable water for injection. Aerobic, anaerobic, fungi, and
algae growths will cause reservoir and equipment plugging. Major emphasis is
required to control sulfate-reducing bacteria. The presence of the sulfate ion is
essential for the growth and reproduction of these colonies which in turn cause
plugging. The sulfate ion reacts with the bacteria to create a sulfide ion which in
turn reacts with iron. Iron sulfide is a serious plugging material. A by-product of
the reaction is

H 2 S , an extremely corrosive agent.

C. Minerals
Water should be checked for mineral content. Some form of iron is one of the most
common plugging agents encountered in injection wells.

Low iron content is

desirable in any water. Sulfates are also of interest; they lead to substantial amounts
of deposition. Chlorides are the primary indication of salinity or the strength of the
brine or the presence of fresh water. Chloride tests can be used to track the progress
of a waterflood. Also, barium is a difficult and troublesome mineral. Barium
sulfate is difficult to remove.

In general, it is undesirable to mix water with

significant amounts of barium with water containing high sulfates.


D. Total Solids
Total solids means the combination of dissolved solids and suspended solids and
should be minimized. Suspended solids are the suspended or precipitated material
in the water. They can lead to plugging of injection wells. Dissolved solids can
precipitate at the injection well face or in the formation and lead to plugging which
reduces injectivity.

9 - 116

E. Produced Water
Water sample analysis is also important in production wells and can assist in a
waterflood surveillance program. Samples should be collected and analyzed on a
regular basis (semi-annually).

Initially, produced water probably represents

formation water and when analyzed provides excellent estimation of water


resistivity,

Rw , for log analysis.

Water sample analysis can be used as an indicator

of water breakthrough and possibly the source of the produced water.


XIII. Pie Charts
A tool that is helpful in monitoring multi-zone waterfloods is a pie chart completion
map. It provides a quick visual presentation of the current perforated producing and
injection pay intervals. For multi-zone waterfloods, this is valuable help. In the event
new perforations are added or old perforations are closed, the pie map should be
updated. Figure 9-57 is a simple example of a pie chart showing the current perforated
intervals for a four-layered reservoir.

9 - 117

FIGURE 9-57
EXAMPLE PIE CHART SHOWING EXISTING
PERFORATIONS

Z-1

Z-4

Z-2

Z-3

Lease
Lines

Open Perforations
Closed or No Perforations

It is observed that the center injection well and three of the corner producers are
completed in Zones 1, 3, and 4. The producer in the lower right corner is completed
only in Zones 2 and 4. Pie charts such as these when updated on an annual basis
provide a quick reference to the existing and historical perforations of a well and the
entire field.
9 - 118

XIV. INTEGRATED WATERFLOOD MONITORING


Integrated monitoring is a process. Integrated monitoring is not a single test or
measurement or a reservoir study done at a single point in time, but it involves
bringing together a number of tests and measurements in which the reservoir
manager utilizes to evaluate the overall efficiency of the injection project on a real
time and ongoing basis. No single measurement or tool answers all of the questions.
Listed below is a summary of important tests and plots which, in the aggregate, can
be used to provide insight as to the current and future management efficiency of a
waterflood project. These steps enable problems and inefficiencies to be recognized
as soon as possible and the results of the corrective action to be incorporated into
the overall process.
1. Maintain a clean well location map showing only bottom hole locations for
each important geological flow unit. Use crisp symbols identifying current
producers, current injectors, and shut-in or abandoned wells.

The wells

should be clearly marked so that when the map is placed on the wall, it can be
easily read by someone standing 10 to 15 feet away. Avoid trying to put too
much data on this map; i.e., well trajectory, injection or production rates,
shallow or deep well completions. The idea is to have a simple clean map
showing the well locations without any clutter.
2. Re-evaluate net pay, permeability, porosity, water saturation, and shale
volume cutoffs and update the geological model with pore volume being
computed for small areas of the field, such as patterns, for each flow unit.
This should be done on a regular basis (at least every two years) and
incorporate the data collected as part of the monitoring discussed throughout
this section. Any new data (log, core, pressures) collected as part of a drilling
program should be immediately incorporated into the geologic model.
3. Conduct production well tests on a regular basis. Ideally, well tests should be
conducted two times per month with each test lasting at least six to eight
hours. Calibrate test equipment, as needed, to ensure test results are reliable.
9 - 119

Well tests provide the information necessary to allocate field-wide production


to individual wells. Individual well data are analyzed using various types of
decline curves and form the technical basis to evaluate waterflood
performance for that specific well.

Care should be taken to accurately

measure all fluids, oil, gas and water. Free gas production, in particular, is
often overlooked but can contribute significantly to reservoir voidage
problems.
4. Keep the fluid levels in the producing wells in a pumped off condition. This
is a Golden Rule in waterflooding. By keeping the fluid levels pumped down,
bottom hole producing pressure can be minimized, throughput rates can be
increased, waterflood recovery is accelerated, primary production from
discontinuous or low permeability sands not responding to injection can be
maximized, and decline curve analysis becomes more reliable. Fluid levels
should be recorded at the same time as the well is being tested so that this
information is compatible.
5. Prepare production rate vs. time graphs for each well. These plots should be
simple with the results being graphed on Cartesian paper. The idea is to plot
the data such that it can be viewed and not distorted by a logarithmic scale.
The key data to be shown on the plot is oil production rate, water production
rate, GOR, and reservoir voidage. Reservoir voidage should take into account
any free gas production. Individual well graphs require frequent and reliable
production tests.
6. Prepare graphs of oil production rate versus cumulative oil production for
each well on a linear scale. This rate versus cumulative plot is equivalent to
rate versus time on semi log paper; it eliminates the log scale, and allows the
analyst to more clearly note changes in oil production. When using this graph
for extrapolating purposes to estimate future oil recovery, considerable care
must be used because in a reservoir in which gas fillup has been achieved, oil
production decline rates can be altered by changing water injection rates.
9 - 120

That is to say, the oil rate and water rate are directly related to the injection
rate.
7. The GOR (not gas rate) versus time plot is one of the best tools for identifying
when reservoir gas fill-up has occurred. At gas fill-up, the GOR will have
declined to the solution GOR evaluated at the reservoir pressure at the start of
injection.
8. Plot WOR versus cumulative oil production on a semi-log graph for each
well. To determine ultimate recovery at different WOR values, utilize and
extrapolate those data points in which the WOR is greater than approximately
1.5 to 2.0. After gas fillup, changes in water injection rate impact both the oil
and water production rates in the same percentage. Consequently, changes in
the injection rate from established injectors should not alter estimates of
future oil recovery obtained by the WOR plot. However, if new injection
wells are placed in service, they may distort the volumetric sweep pattern of
the prior injectors. If this occurs, estimates of future oil production at a given
water oil ratio may change.
9. Annotate all production and injection plots for each well by noting significant
events that can affect production trends including pump changes, well
stimulations, perforation changes, drilling of offset production or injection
wells, or changing offset well injection rates.
10. Prepare a linear graph showing recovery factor (RF) since the start of water
injection versus hydrocarbon pore volumes injected (HPVI). Compare actual
RF versus HPVI from small segments of the field, such as a pattern, with the
RF versus HPVI for the entire field. This graph provides a comparison of
how the small segment of the field is performing relative to the total field. If
a numerical simulation model for the field exists, compare field RF versus
HPVI with RF versus HPVI developed in the simulator.

9 - 121

11. Prepare a voidage replacement ratio (VRR) plot versus time for each pattern,
each flow unit, and for the entire field. Remember, pattern totals may contain
fractions of certain wells contribution if an individual well is in more than
one pattern. In preparing this graph, the analyst should account for free gas
production and recognized that water injection efficiency will probably be no
more than about 80 percent. When accounting for free gas production and
water injection efficiency, if the resulting VRR is greater than 1.0, reservoir
pressure is increasing. If the VRR is less than 1.0, reservoir pressure is
declining unless there is water influx from the aquifer. Clearly, it should be
realized that, when looking at the field as a whole, the reservoir pressure
could be increasing or decreasing, but in localized areas, reservoir pressure
may be just the reverse.
12. If the injection pressure is constant during gas fillup, the water injection rate
will decline over time for all values of the mobility ratio.
13. If pressure between the injector and producer well is constant after gas fillup,
injection rate will decline, remain constant, or will increase if the mobility
ratio (MR) is less than 1.0, is equal to 1.0, or is greater than 1.0 respectively.
Changes in water injection rate directly affect oil production rate and water
production rate. Caution is urged in the use of conventional decline curves
when projecting future recoverable oil and gas volumes because the rate of
decline is directly affected by the injection rate and injection rate is directly
related to mobility ratio.
14. Maintain a graph of water injection rate and well head injection pressure for
each well in the field. If equipment is not in place to monitor at the wellhead
on a daily basis, well tests should be performed at least twice a month using
the production well testing standards discussed in point 3. These tests should
be used to allocate monthly injection volumes.

9 - 122

15. Conduct pressure buildup (PBU) and pressure falloff (PFT) tests on 1/4th to
1/5th of the producing and injection wells each year. This means that each
well is tested at least once every 4 to 5 years. These pressure tests provide the
best estimates of average reservoir pressure and indicate how pressure is
changing with time. These tests also give the best indicators of wellbore skin
damage and wellbore permeability.
16. Perform injection profile surveys (PLTs) on new injection wells after three
months, six months, and 12 months of injection. Thereafter, run surveys on
an annual basis for the next three years and then at least one survey every two
years. These surveys provide critical information on where the injection
water is going. If the engineer or geoscientist does not know how much water
is entering the various flow units, it is very difficult for that person to perform
reliable reservoir management calculations. Frequent surveys are needed due
to the fact that the injection profiles change over time if the skin factor
changes, if new wells are drilled, or if MR is not equal to 1.0.
17. Immediately following the PLTs in the injection wells, shut those wells in for
five to seven days and obtain a pressure falloff test. Following the pressure
falloff test, conduct a step rate test starting with low injection rates and
gradually increasing to higher values. This test gives the best opportunity to
determine the reservoir pressure at that time.
18. Maintain a Hall plot on each injection well to identify possible changes in the
injection well skin factor. Remember, the routinely constructed Hall plot
assumes gas fillup has been completed, average reservoir pressure is constant,
and MR is 1.0. If reservoir pressure is gradually increasing, this change in
pressure creates the effect of a positive skin.

9 - 123

Variable

Hall Plot Results

Gas fillup not yet achieved

Skin factor increasing

Average reservoir pressure increasing

Skin factor increasing

Average pressure is decreasing

Skin factor decreasing

MR greater than 1.0

Skin factor decreasing

MR less than 1.0

Skin factor increases

19. In a pattern waterflood, inject into each pattern in proportion to the pore
volume. (Unfortunately nature honors KH and not pore volume).
20. Construct injection efficiency plots as described in Section XI. These plots
provide an indication as to the amount of injected water that goes out of zone.
Results are strongly dependent on the accuracy of pore volume, residual oil
saturation to waterflooding, and initial connate water saturation.
XV. Project Review
The most important part of a reservoir monitoring program in primary or waterflood
operations is a well review conference.

These conferences, normally held on a

quarterly basis, include the reservoir engineer, operations engineer, development


geologist, and several representatives from the field office.

The purpose of the

meetings is to review injection and production behavior for individual wells and the
field as a whole. Another purpose is to ensure free exchange of information from all
parties, discuss the status of previous well recommendations, propose future well
work, and address general injection, production, and well testing issues. Each well in
the project should be reviewed at least one time per year. Problem wells should be
reviewed more frequently. These well reviews should help to ensure that each well is
operating in the most efficient manner.
9 - 124

The project review builds and maintains a good working relationship between the
engineers and field personnel. This close relationship serves to promote teamwork,
smooth operations, and improved recovery.

9 - 125

CHAPTER 9 REFERENCES
1. Fetkovich, M.J.: Decline Curve Analysis Using Type Curves, Journal of Petroleum
Technology, (June 1980), pp. 1965-1077.
2. Lo, K.K., Warner, H.R., and Johnson, J.B.: A Study of Post-Breakthrough
Characteristics of Waterfloods, SPE Paper 20064 presented at 60th California
Regional Meeting, Ventura, California, April 1990.
3. Yortsas, Y.C., Choi, Y., Yang, Z, and Shah, P.C.: Analysis and Interpretation of
Water/Oil Ratio in Waterfloods, SPE Journal, (December 1999), p. 413.
4. Ershaghi, I. and Omoregie, O.: A Method for Extrapolation of Cut vs. Recovery
Curves, Journal of Petroleum Technology, (February 1978), p. 203.
5. Ershaghi, I. and Abdassah, D.: A Prediction Technique for Immiscible Processes
Using Field Performance Data, Journal of Petroleum Technology, (April 1984), p.
664.
6. Ershaghi, I., L.L., and Hamdi, M.: Application of the X-Plot Technique to the Study
of Water Influx in the Sidi El-Itayem Reservoir, Tunisia, Journal of Petroleum
Technology, (September 1987), p. 1127.
7. Wilson, S.J., Miller, M.R., Frazer, L.C., and Digert, S.A.: Multiple Trend Forecasting
Accounts For Field Constraints, SPE Paper 39929 presented at Rocky Mountain
Regional Meeting, Denver, Colorado, April 1998.
8. Baker, R.: Reservoir Management for Waterfloods Part I, The Journal of
Canadian Petroleum Technology, (April 1997), p. 20.
9. Baker, R.: Reservoir Management for Waterfloods Part II, The Journal of
Canadian Petroleum Technology, (January 1998), p.12.
10. Robertson, D.C., and Kelm, C.H.: Injection Well Testing to Optimize Waterflood
Performance, Journal of Petroleum Technology, (November 1975), p. 1337.
11. Matthews, C.X. and Russell, D.G.: Pressure Buildup and Flow Tests in Wells,
Monograph Series, SPE, Richardson, TX (1967) 1.
12. Earlougher, R.C., Jr.: Advances in Well Test Analysis, Monograph Series, SPE,
Richardson, TX (1977) 5.
13. Lee, W.J.: Well Testing, Textbook Series, SPE, Richardson, TX (1982) 1.

9 - 126

14. Felsenthal, M. and Ferrell, H.: Fracturing Gradients in Waterfloods of Lowpermeability, Partially Depleted Zones, Journal of Petroleum Technology, (June
1971), p. 727.
15. Felsenthal, M.: Step Rate Tests Determine Safe Injection Pressures in Floods, Oil &
Gas Journal, (October 28, 1974), p. 49-54.
16. Singh, P.K., Agarawal, R.G., and Drase, L.D.: Systematic Design and Analysis of
Step-Rate Tests to Determine Formation Parting Pressure, paper SPE 16798 presented
at the 1987 SPE Annual Technical Conference and Exhibition, Dallas, Sept. 27-30.
17. Gidley, J.S., Holditch, S.A., Nierode, D.E., and Veatch, R.W., Jr.: Recent Advances in
Hydraulic Fracturing, Monograph Series, SPE, Richardson, TX (1989) 12.
18. Aronofsky, J.S. and Jenkins, R.: A Simplified Analysis of Unsteady Radial Gas
Flow, Trans., AIME, (1954) 201, pp. 149-154.
19. Bargas, C.L. and Yanosik, J.L.: The Effects of Vertical Fractures on Areal Sweep
Efficiency in Adverse Mobility Ratio Floods, paper SPE 17609 presented at the 1988
SPE International Meeting, Tianjin, China, November.
20. Dyes, A.B., Kemp, C.E., and Caudle, B.H.: Effect of Fractures on Sweepout
Patterns, Trans., AIME, (1958) 213, pp. 245-249.
21. Hall, H.N.: How to Analyze Injection Well Performance, World Oil, (October
1963), pp. 128-130.
22. Buell, R.S., Kazemi, H., and Poettman, F.H.: Analyzing Injectivity of Polymer
Solutions with the Hall Plot, paper SPE 16963 presented at the 1987 SPE Annual
Technical Conference and Exhibition, Dallas, Sept. 27-30.
23. Staggs, H.M.: An Objective Approach to Analyzing Waterflood Performance, paper
SPE presented at Southwest Petroleum Short Course (Spring 1980 Lubbock, Texas.
24. Anthony, J.L. and Meggs A.J.M.: An Approach for Determination of Economically
Optimum Injection and Production Rates in a Large Multi-Pattern Waterflood, paper
SPE 16957 presented at the 1987 SPE Annual Technical Conference and Exhibition,
Dallas, Sept. 27-30.
25. Chapman, L.R. and Thompson, R.R.: Waterflood Surveillance in the Kuparuk River
Unit with Computerized Pattern Analysis, Journal of Petroleum Technology, (March
1989), p. 277.
26. Cobb, W.M., and Marek, F.J.: Determination of Volumetric Sweep Efficiency in
Mature Waterfloods Using Oil Production Data, SPE Paper 38902 presented at the
1997 SPE Annual Technical Conference and Exhibition, San Antonio, Oct. 5-8.

9 - 127

27. Craig, F.F., Jr.: The Engineering Aspects of Waterflooding, Monograph Series, SPE,
Dallas, TX (1971) 3.
28. Willhite, G.P.: Waterflooding, Textbook Series, SPE, Dallas (1986) 3.
29. Seright, R.S.: Placement of Gels To Modify Injection Profiles, paper SPE 17332
presented at 1988 Enhanced Oil Recovery Symposium, Tulsa, April 16-21.
30. Seright, R.S.: A Review of Gel Placement Concepts, New Mexico Petroleum
Recovery Center, New Mexico Tech, Socorro, New Mexico 87801, PRRC 96-21, July,
1996.
31. Liang, Jenn-Tai, Seright, R.S.: Further Investigations of Why Gels Reduce Water
Permeability More Than Oil Permeability, SPEPF (November 1997) 225-230.
32. Seright, R.S., Lane, R.H. and Sydansk, R.D.: A Strategy for Attacking Excess Water
Production. SPEPF (August 2003) 158-169.
33. Seright, R.S.: Clean Up of Oil Zones after a Gel Treatment, paper SPE 92772
presented at 2005 SPE International Symposium on Oilfield Chemistry, The
Woodland, Texas, Feb. 2-4.
34. Frick, T.C.: Petroleum Production Handbook, Volume II, Reservoir Engineering,
Society of Petroleum Engineers of AIME, Dallas, 1962.
35. Vetter, O.J.: Oilfield Scale Can We Handle It?, Journal of Petroleum Technology,
(December 1976), p. 1402.
36. Vetter, O.J., Kandarpa, V., and Harauka, A.: Prediction of Scale Problems Due to
Injection of Incompatible Waters, Journal of Petroleum Technology, (February 1982),
p. 273.
37. Patton, C.C.: Oilfield Water Systems, Campbell Petroleum Series, 1215 Crossroads
Blvd., Norman, OK (1981).

9 - 128

INTENTIONALLY LEFT BLANK

9 - 129

PROBLEM 9:1
WOR, CUMULATIVE OIL PRODUCTION, AND VOLUMETRIC SWEEP
EFFICIENCIES
Figure 9:1-1 is a graph of WOR versus cumulative oil production for a producing well
located in the center of a five-spot pattern which is surrounded by four injection wells.
Figure 9-4 illustrates the well location. Cumulative oil production to date since the start
of water injection is 552 MBO. The pattern consists of the following data:

Vp

= 6,056 MBBL

Sw

= 68% (Assume constant)

Sg

= 5%

Bo

= 1.20 RB/STBO

S wc

= 32%

Evw , using Eq. 9.30..

1.

Compute the current pattern volumetric sweep efficiency,

2.

If the WOR economic limit is 25:1, what is the expected ultimate oil recovery at

the economic limit under current operations.

Evw at the economic limit?

3.

What is the expected

4.

If the reservoir is infill drilled and the volumetric sweep efficiency is increased by

20 percent, what is the anticipated incremental oil recovery?

9 - 130

INTENTIONALLY LEFT BLANK

9 - 131

9 - 132

0.1

1.0

10.0

100.0

WOR

0.0

0.1

0.2

0.3

0.5

0.6
Cumulative Oil, MMBO

0.4

0.7

0.8

Figure 9:1-1
WOR, Np, and Evw
For a Producer Centered 5-Spot Pattern

0.9

1.0

INTENTIONALLY LEFT BLANK

9 - 133

WHAT RESERVOIR PARAMETERS DO I NEED TO CONSIDER WHEN


IMPLEMENTING A NEW WATERFLOOD?
I.

Remember the four driving forces which collectively control waterflood recovery:

N D N E A EV ED

N D = Displaceable Oil (Ultimately Recoverable), BBL

N = Oil in Place at Start of Waterflood, BBL


E A = Areal Sweep Efficiency, Fraction

EV

Where

= Vertical Sweep Efficiency, Fraction

(Also

And

ED

E A EV

= Volumetric Sweep Efficiency)

= Displacement Efficiency of Oil in Volumetrically Swept Portion of


Reservoir, Fraction

A. What's the Oil in Place at start of waterflood,


Saturation, Net Pay, and Porosity?
Oil Saturation,

Compute

N?

Did you properly consider Oil

So

So at start of waterflood

Did you use reliable estimates of S wc ,


factor?

Bo , BOBP , and primary recovery

So compare with Sor ?

How does

Is

What's maximum recoverable waterflood oil in reservoir at start of


waterflood if E A and EV = 100%?

So sufficient to create an oil bank?

N MAX

Pore Volume (So Sor )


Bo

Did you evaluate Bo at reservoir pressure at start of waterflooding? Why


not? That would be a good estimate!

Net Pay,

How did you select net pay cutoffs?

Did you use local "rules of thumb" or actual special core analysis?

Remember - air permeability vs. porosity plots are not reliable! Plot oil
permeability measured at immobile water!

Did you distinguish between net pay for primary production and net pay for
secondary injection? If you don't, you could be making a huge mistake!

Secondary net pay is usually less than primary net pay!

Did you consider lateral continuity?

Prepare detail cross sections

Porosity,

Did you rely on logs?

How does log porosity compare with core porosity?

Remember a 1-2 unit porosity difference in low porosity reservoirs could


impact waterflood pore volume and ultimate waterflood recovery by 15 20%.

B. How did you quantify E A ? How did you consider such factors as Pattern,
Directional Permeability, Mobility Ratio, and Volumes Injected?
What is the basis for your pattern?
Does your pattern require new wells? Injectors or producers?
Does it attempt to utilize existing wells? Remember, old, fracture-stimulated
wells can be poor injectors! Injected water cannot be controlled and
production will ultimately suffer!

Did you take advantage of directional permeability trends with proper injection
well orientation to maximize areal sweep efficiency? Failure to do so will
result in poor recovery.
Waterflood

MR "rule of thumb" M 0.333 oil viscosity !

Remember, as reservoir pressure increases, little or no free gas re-dissolves in


oil! It's been produced! Thus Bo decreases and oil viscosity increases with
increasing pressure.
How did you compute MR ? Is krw evaluated in water region and kro in
displaced oil bank? Did you evaluate viscosity at reservoir temperature and
pressure existing at start of waterflood?
In high gas saturation reservoirs, does water breakthrough occur before fillup?
Did you consider that in reservoirs with existing free gas saturations, the
advancing oil bank moves into areas not easily swept by injected water? This
is a resaturation effect. No, this is not a "gas cap" issue!
Important message: Areal sweep is governed by the pressure distribution,
MR , and directional permeability trends between the injector and producer!
Check this with CUSA Chi Tube streamline model. Even for ideal well
patterns, these three factors can cause the actual reservoir flow pattern and
areal sweep to be vastly different from that predicted using a pattern based on
simple well locations.
C. Did you consider vertical sweep efficiency? Did you take into account stratification,
MR , and gravity segregation?
Did you select "actual" geological layers or did you develop layers using a
statistical approach such as with the Dykstra-Parsons co-efficient? If actual
layers, did you consider sub-layering?
Did your forecasting possess sufficient layers to replicate actual reservoir
stratification or did you use a minimum number to simplify calculations!
In selecting number of layers, did you consider F. F. Craigs JPT paper on
reservoir description in October 1979?
Does the model Dykstra-Parson permeability variation agree with the actual
reservoir?
Did you account for vertical crossflow? This can be important when vertical
permeability is moderately high (1-5 md or greater) due to the large cross

section area available for vertical flow. Crossflow is usually more important in
sandstones without interbedded shales.
Reservoirs with high vertical permeability can result in poor vertical sweep due
to gravity segregation of water and oil! In these cases, closer well spacing and
increased injection rate helps!
Interlayer crossflow in favorable
recovery.

MR

(less than unity) can increase oil

D. Did you consider the efficiency with which the injected water displaced the oil
within the volumetrically swept pore volume, ED .
Remember that
swept portion.

ED

is an indicator of the decrease in oil saturation in the

Did you consider that the oil saturation in the water swept portion of the
reservoir is:
a.

greater than

Sor

k
b. governed by fractional flow ( ro and
krw

c.

decreases as injection throughput increases

Remember if
and
II.

o
)
w

ED

So

at start of waterflood is low, then little moveable oil exists

will be low.

Water Injection Rate Calculations


For constant bottomhole (essentially constant wellhead) injection pressure,
injection rate will decline sharply (10-30%) until fillup regardless of

MR !

After fillup (and with constant bottomhole injection well and producing well
pressure), injection rate will gradually increase

MR 1.0 depending on the MR .

k
Many variables affect injection rate such as o ,

MR 1.0

kw ,

or decrease

oil/water viscosity;

pattern, MR , skin (injector and producer), bottom hole injection pressure,


producing pressure (high fluid levels also decrease injection), distance between
wells, and pattern. The next time injectivity is a problem, don't automatically
assume skin damage!

Injecting above fracture pressure will increase injection by only a 2-3 fold
factor. Larger increases indicate (1) wellbore skin or (2) water going out of
zone! I bet you thought you fraced into "new pay"!
Is actual vertical distribution of water in proportion to kh ? No! Now you
wonder why actual performance does not agree with predicted performance.
Second important message: Actual injection should usually exceed projected
injection by 15-20% or more to account for thief (gas cap, natural fractures,
aquifer, poor cement, porous but discontinuous zones, etc.) zone losses!

You might also like