You are on page 1of 267

I.

FROMATION EVALUATION OVERVIEW

1. THE SCOPE OF FORMATION EVALUATION


Formation evaluation covers a large variety of measurement and analytic techniques. Although
the emphasis is on wireline logging techniques and log analysis methods, these are far from the
only tools available to the formation evaluator. Well logs are central only in the sense that they
are recorded in practically all well bores and are directly relatable to all the other parameters
available from the associated sciences. For example, a geophysicist needs borehole
measurements to determine a time-depth relationship, and a petrophysicist needs core analysis
to properly define log response, but a thin section or scanning electron microscope (SEM) photo
of a rock sample are of no direct help to the interpretation of a seismic section, nor is a vertical
seismic profile (VSP) of any help in deter-mining relative permeability. However, all the
measurements are pertinent to the complete task of defining a reservoir's limits, storage capacity,
hydrocarbon content, productivity and economic value.

To place the various disciplines in perspective, it is valuable to consider the overall problem of
formation evaluation in terms of orders of magnitude. If one meter is taken as a unit of
measurement, then each formation evaluation technique can be placed in order, as shown in
Table 1 .

Thus, formation evaluation techniques cover at least twelve orders of magnitude. Equally far-
ranging are the physical principles employed to make the basic measurements. An enlightening
way of viewing the vast spread is to consider the frequency employed by the measuring
processes available, as illustrated in Table 2 . Few other sciences require, or use, such a wide
range of measurement techniques over such a wide range of physical dimensions.
2. FORMATION EVALUATION OBJECTIVES

The primary objective of formation evaluation is to determine the size of a reservoir, the quantity
of hydrocarbons in place, and the reservoir's producing capabilities. The initial discovery of a
reservoir lies squarely in the hands of the exploration geologist using seismic, gravity and
magnetic studies, and other geologic tools. Formation evaluation presupposes that a reservoir
has been located and is to be defined by drilling as few wells as possible. Enough data should be
gathered from those wells to extrapolate reservoir parameters field wide and arrive at realistic
figures for both the economic evaluation of the reservoir and the planning of the optimum
recovery method. Formation evaluation offers a way of gathering the data needed for both
economic analysis and production planning.
What, then, are the parameters that the manager, the geologist, the geophysicist, and the
reservoir and production engineers need? Which of these can be provided by seismics, by coring,
by mud logging, by testing, or by conventional wireline logging?
The geophysicist needs to know the time-depth relationship in order to calibrate conventional
seismic and VSP surveys. The geologist needs to know the stratigraphy, the structural and
sedimentary features, and the mineralogy of the formations through which the well was drilled.
The reservoir engineer needs to know the vertical and lateral extent of the reservoir, its porosity
(the nature of the porosity) and permeability, fluid content, and recoverability.
The production engineer needs to know the rock properties, be aware of overpressure if it exists,
be able to assess sanding and associated problems, and recognize the need for secondary
recovery efforts or pressure maintenance. Once the well is in production, he/she also needs to
know the dynamic behavior of the well under production conditions and be able to diagnose
problems as the well ages.
Engineers also need to know formation injectivity and residual water saturation to plan water
flooding and monitor water flood progress when it is operational.
The manager needs to know the vital inputs to an economic study-the original petroleum
hydrocarbons in place, recoverability, cost of development and, based on those factors, the
profitability of producing the reservoir.
Log measurements, when properly calibrated, can give the majority of the parameters required by
all these professionals. Specifically, logs can provide either a direct measurement or a good
indication of

• porosity, both primary and secondary (fractures and vugs)


• permeability
• water saturation and hydrocarbon movability
• hydrocarbon type (oil, gas, or condensate)
• lithology
• formation (bed) dip and strike
• sedimentary environment
• travel times of elastic waves in a formation
From this data, good estimates may be made of the reservoir size and the petroleum
hydrocarbons in place.
Logging techniques in cased holes can provide much of the data needed to monitor primary
production and also to gauge the applicability of water flooding and monitor its progress when
activated.
In producing wells, logging can provide measurements of

• flow rates
• fluid type
• pressure
• residual oil saturation
From these measurements, dynamic well behavior can be understood better, remedial work
planned, and secondary or tertiary recovery proposals evaluated and monitored.
In summary, logging, when properly applied, can answer a great many questions from a wide
spectrum of special interest groups on topics ranging from basic geology to economics.
Of equal importance, however, is the fact that logging by itself cannot provide answers to all
formation evaluation questions. Coring, core analysis, and formation testing are integral parts of
any formation evaluation effort.

2.1 Wireline Logs Objectives


The objective of interpretation of wireline well logs depends very much on the user. Quantitative
analysis of well logs provides the analyst with values for a variety of primary parameters, such as:
• porosity
• water saturation, fluid type (oil/gas/water)
• lithology
• permeability
From these, many corollary parameters can be derived by integration (and other means) to arrive
at values for:
• hydrocarbons-in-place
• reserves (the recoverable fraction of hydrocarbons-in-place)
• mapping reservoir parameters
But not all users of wireline logs have quantitative analysis as their objective. Many of them are
more concerned with the geological and geophysical aspects. These users are interested in
interpretation logs for:
• well-to-well correlation
• facies analysis
• synthetic seismograms
• regional structural and sedimentary history
In quantitative log analysis, the objective is to define
• the type of reservoir (lithology)
• its storage capacity (porosity)
• its hydrocarbon type and content (saturation)
• its producibility (permeability)
As a preliminary to discussing methods of log analysis it is worthwhile to define the terms used.
3. FORMATION EVALUATION METHODS

In practice, the order in which formation evaluation methods are used tends to follow the order-of-
magnitude table, i.e., from the macroscopic to the microscopic. Thus a prospective structure will
first be defined by seismic, gravity, and/or magnetics studies. Most wellbores drilled through such
a structure are mud-logged and/or measured while drilled, from which cores may be cut or
sidewall samples taken. Once the well has reached a prescribed depth, logs are run. Subsequent
to logging, an initial analysis of mud log shows, together with initial log analysis, may indicate
zones that merit examination either by the wireline formation tester or by drillstem testing. Should
such tests prove the formation to be productive, more exhaustive analyses will be made of all
available data, including core analysis? The whole process is summarized in Table 1 .

(Mechanical) Mud Logging

Mud logging, more precisely referred to as hydrocarbon mud logging, is a process whereby the
circulating mud and cuttings in a drilling well are continuously monitored by a variety of sensors.
The combined analysis of all the measurements provides an indication of the rock type and its
fluid content. The sundry measurements are displayed on a log as curves or notations as a
function of depth. Not all wells are logged in this manner. Development wells, for example,
normally are drilled and logged by wireline logging tools only. In contrast, wildcat wells nearly
always are monitored by the mud-logging process. The advantages of mud logging include the
availability on a semi-continuous basis of actual formation cuttings analysis (which, in turn, gives
immediate indications of rock type and hydrocarbon presence) and the ability to predict drilling
problems (such as overpressure) before they become unmanageable.

Coring
A number of methods are in use to cut cores in a wellbore. Conventional cores are cut using a
special core bit and retrieved in a long core barrel. The recovered core sample may undergo
physical changes on its journey from the core depth to the surface, where it can be analyzed.
More sophisticated coring mechanisms now in use conserve either the orientation, the pressure,
or the original fluid saturations of the rock sample gathered. An awareness of these changes and
sampling methods is essential to an understanding of core analysis results.
Other coring methods are available where additional rock samples are required after the well has
been drilled and before it has been cased. These methods require wireline tools that cut core
plugs from the sides of the wellbore.
Many parameters needed to correctly interpret openhole wireline logs can only be determined
from accurate core analysis that presupposes cores have been cut. Thus, coring plays a major
part in field development.

Measurements While Drilling (MWD)


Increasingly, formation properties are being measured by use of special drill collars housing
measuring devices at the time the formation is drilled. These MWD tools are particularly valuable
in deviated offshore wells where wellbore path control is critical and where an immediate
knowledge of formation properties is vital for decision making on such matters as choosing
logging and casing points. Although not as complete as openhole logs, the measurements
obtained by MWD are rapidly becoming just as accurate and usable in log analysis procedures.

Formation Testing
Formation testing is the "proof of the pudding." If the well flows petroleum (oil or gas or both) on a
drillstem test (DST), no amount of logging data or core analysis can deny that a productive zone
has been found. However, a drillstem test provides not only proof that hydrocarbons exist in the
formation and will flow, but also supplies vital data about both the capacity of the reservoir and its
ability to produce in the long term. Correct interpretation of pressure records from drillstem tests
help the overall formation evaluation task immensely.
Wireline formation testers (RFT) complement drillstem tests by their ability to sample the fluid in
many different horizons in the well and also gather detailed formation pressure data that it is
almost impossible to obtain from a DST alone. This detailed pressure information can be used to
calculate fluid contacts, such as the free water level.
Openhole Logging
Openhole logging provides the great meeting place for all the formation evaluation methods. Only
through openhole logging can a continuous record of such formation properties as porosity, water
saturation, and rock type be made, versus depth. In particular, wireline logs can record formation
self potential, electrical resistivity, bulk density, natural and induced radioactivity, hydrogen
content, and elastic properties. Almost without exception, every well drilled for hydrocarbons is
logged with wireline instruments. Unfortunately, the full potential of the logs is not always utilized,
or the logs are incorrectly analyzed, because of a lack of training on the part of the analyst or a
lack of understanding of where wireline logs fit in relation to the other formation evaluation tools.
All too often logs are seen as an end in themselves and are considered in isolation. It is hoped
that this module will encourage the reader to take a broader view of log analysis in the context of
overall formation evaluation.
Figure 1 illustrates the formation evaluation picture and the central role of openhole logging and
log analysis.
II. INTRODUCTION TO WELL LOGGING
1. LOGGING SYSTEMS
1.1 GENERAL DESCRIPTION - Modern Logging Tools
The actual running of a log involves the tool on the end of the logging cable, the cable itself, and
the controlling and recording apparatus on the ground surface. Before discussing downhole tools,
however, the common elements of all logs will be presented. Figure 1 illustrates the basic
components of any logging system. A sensor, incorporated in a downhole measurement
instrument called a sonde, together with its associated electronics, is suspended in the hole by a
multiconductor cable. The sensor is separated from virgin formation by a portion of the mud
column, by mud cake, and, more often than not, by an invaded zone in the surrounding rock. The
signals from the sensor are conditioned by the electronics for transmission up the cable to the
control panel, which in turn conditions the signals for the recorder. As the cable is raised or
lowered, it activates a depth-measuring device-a sheave wheel, for example-which in turn
activates a recording device-either an optical camera (making a film) or a tape deck (making a
digital recording on magnetic tape). The film (or tape) is reproduced to provide a hard copy of the
recorded data.

In general, well-logging jargon distinguishes between a logging survey, a logging tool, and a log,
as well as a curve. There is frequently some confusion about these terms when logging matters
are discussed. A logging survey is provided by a logging service company for a client. During the
course of the survey, the logger may employ several different logging tools, and record several
different logs, on each of which are presented several different curves. The logging tools, in turn,
consist of multiple sensors. Figure 2 illustrates these terms and their interrelationship.
1.2 LOGGING SETUP
Rigging Up to Run a Log

Figure 1 shows a typical setup for a logging job. A logging truck is anchored about 100 to 200 ft
from the well. Two sheave wheels are mounted in the derrick, with one suspended from the
crown block and the other chained down near the rotary table. The logging cable from the truck
winch is then passed over the sheave wheels, attached to the logging tool string, and lowered
into the hole. A more detailed diagram of this hookup is shown in Figure 2 .

Two mechanical details regarding this method of rigging up are worth noting. Between the top
sheave wheel and the elevators a tension device measures strain on the logging cable and
displays it in the logging truck ( Figure 3 ). The tension on the elevators is twice that on the cable.
The elevators should be securely locked and the traveling block braked and chained.

The tie-down chain for the lower sheave is also of great importance. If it breaks or comes untied,
the snap will probably break the cable and catapult the sheave wheel several hundred feet
(Figure 4 ). - Missing
1.3 LOGGING UNITS
Logging Trucks
Logging service companies offer a variety of logging units, each of which has the following
components:
• logging cable
• winch to raise and lower the cable in the well
• self-contained 120-volt AC generator
• set of surface control panels
• set of downhole tools (sondes and cartridges)
• recording mechanism (tape and/or film)
Figure 1 shows a cutaway of a typical logging truck. Land units are mounted on a specially
adapted chassis reinforced to bear the load of a full winch of cable (up to 30,000 ft long). The
instrument and recorder cabs are usually cramped, noisy, too hot or too cold, and sometimes
filled with ammonia fumes from an ozalid copier.

Offshore units are mounted on skids and bolted (or welded) to the deck of the drilling barge,
vessel, or platform.Other units can be disassembled into many small fragments and flown into
remote jungles suspended under helicopters. However, all logging units are basically similar, and
require good mechanical maintenance to avoid problems during logging operations.
Logging Cables
Modern logging cables are of two types: monoconductor and multiconductor. Monoconductor
cables, with a diameter of 1/4 in., are used for completion services, such as shooting perforating
guns, or setting wireline packers and plugs, and for production logging surveys, such as
flowmeters and temperature logs in producing wells. Multiconductor cables, with a diameter of
about 1/2 in, are used by most logging service companies for recording openhole surveys. The
multiconductor cables contain 6 or 7 individual insulated conductors in the core.
The outer sheath is composed of two counterwound layers of steel wire. Such a cable has a
breaking strength of between 14,000 and 18,000 lb and weighs between 300 and 400 lb per 1000
ft. It is quite "elastic" and has a stretch coefficient of around I x l0 -6 ft/1b.
The "Head" and the "Weak point"
The cable ends at the logging "head." The head anchors the cable and attaches to the logging
tool by a threaded ring. Thus, the head provides both the electrical connection between the
individual cable conductors and the various pins in the top of the tool and the mechanical
connection. Built into the head is a "weak point," a short length of aircraft cable designed to break
at a given tension (usually about 6000 lb, but deep-hole weak points are designed to break at
lower tension, e.g., 3500 lb). The weak-point provides a means to free the cable from the tool
when it becomes irrevocably stuck in the wellbore. Several examples follow.
1.4 COMPUTERISED LOGGING UNITS

Available Systems
Major service companies now offer logging services from computerized logging units. The
advantages of using these units are many and their use is encouraged.

Features of Computerized Units


In contrast to conventional logging units, computer-based units offer the following features:
• All logs are directly recorded on digital magnetic tape or onto a hard disk.
• Computer control of the data gathering allows logs to be recorded either
logging up or down with all curves mutually on depth.
• Calibrations are performed under programmed control more quickly and
accurately than in conventional units.
• Logs can be played back from the data tapes on many different scales (both
depth and response scales).
• Wellsite computation of raw data ranges from completion aids (hole volume
integration for cement volumes) to dipmeter computations and complete log
analysis.

Figure 1 is a schematic of a computerized logging system. The logging engineer accesses the
system by keyboard. At his command, the computer loads programs to perform such functions as
calibration, logging, computation, and playback.

Calibration Methods and Tolerances


Conventional logging units require human operation of both sensitivity and zero offset control.
Figure 2 depicts a typical conventional calibration system.
The variable offset resistor is adjusted when the logging sensor is at the low end of its range of
measurement (for example, the caliper tool in a 6-in. ring), and the variable gain resistor is
adjusted when the sensor is at the high end (e.g., the caliper tool in a 12-in. ring).

The computer units eliminate the need for human intervention, other than to place the tool to be
calibrated in the correct environment (e.g., putting the 6-in. ring over the caliper arms). The data-
gathering system accepts the raw uncalibrated readings of the tool and computes a calibration
equation to transform raw uncalibrated data into calibrated data. Figure 3 illustrates this concept.

The important things to check include the agreement between all three numbers with the
specified tolerances listed in Figure 4 . Note that these sets of numbers refer to Schlumberger
logs. Other service company tools use different numbers. Booklets explaining calibration
techniques by each logging service company can be obtained from their sales personnel.

The tolerance table of Figure 4 shows that the near count rates are allowed a variation of ±22 cps
and the far count rates a variation of ±14 cps. Thus, the wellsite calibration in this case can be
considered good.
2. THE BOREHOLE ENVIRONMENT, MUD, MUDCAKE, AND INVASION

After drilling through a permeable formation, generally an invasion process begins. If the pressure
in the mud column exceeds formation pressure, fluid from the mud will move into the formation
(provided it is porous and permeable) and deposit a mud cake on the borehole wall.

It is important to distinguish between the resistivity of the fluid within the pore space and the
resistivity of the rock-fluid system itself. The terms used in Table 1 should be well known to
everyone involved in well log evaluation work.

The flushed zone is important because it affects the readings of some logging tools and because
it forms a reservoir of mud filtrate to be recovered on a drillstem test before formation fluids are
recovered.

Depending on the type of mud used, oil- or water-base, and the relative values of Rmf and Rw, the
invasion process may result in a radial resistivity profile that increases or decreases with distance
from the borehole wall. Figure 1 illustrates what may be expected in a number of cases.
3. LOGGING TOOLS

Logging tools are cylindrical tubes containing sensors and associated electronics that can be
attached to the logging cable at the logging head. Although there are wide variations in sizes and
shapes, a typical logging tool is 3 5/8 in. in diameter and from 10 to 30 ft long. They are built to
withstand pressures up to 20,000 psi and temperatures of 300 to 400 F. The internal sensors and
electronics are ruggedly built to withstand physical abuse. Modern tools are "modularized" to
allow combination tool strings. By appropriate mixing and matching, various logging sensors can
be connected with each other. Among the obvious limitations to this method are the difficulty in
handling very long tools and the limited information-transmitting power of the cable conductors.

Because logging tools have multiple sensors at different points along their axes, their respective
measurements have to be memorized and placed on a common depth reference. Thus, the signal
from the sensor highest on the tool must be "remembered" until the signal from the lowest sensor
arrives from the logging depth being memorized. Figure 1 and Figure 2 illustrate this
characteristic.

The reference point for the logging tool shown in Figure 1 is the sensor A. Higher up the tool,
sensors B and C record other parameters. Without memorization, sensors B and C record curves
off depth that appear on the log to be deeper than sensor A by a distance equal to the spacings
A-B and A-C. It is important, therefore, to ensure that all curves recorded simultaneously are on
depth on the log by means of proper memorization ( Figure 2 , right).

Another associated depth problem arises when several surveys are recorded on different trips
into the hole. Unless care is taken, these surveys may not be on depth with each other. The only
method of ensuring good depth control is to insist on a repeat section that passes a good marker
bed. Each subsequent log should be placed on depth using this repeat section as a depth
reference before the main logging run is made.
Openhole logging tools currently in use are

Formation Fluid Content Indicators


• Induction
• Laterolog
• Microfocused and microresistivity devices
• Dielectric
• Pulsed neutron
• Inelastic gamma
Porosity-Lithology Indicators
• Acoustic (sonic)
• Density and lithologic density
• Neutron
• Natural gamma ray
• Spectral gamma ray
Reservoir Geometry Indicators
• Dipmeter
• Borehole gravimeter
• Ultra-long-spacing electric
Formation Texture Indicators
• Electrical borehole imagers
• Ultrasonic borehole imagers
Formation Productivity Indicators
• Wireline formation tester
• Production logging
These are the basic tools that will answer 90% of the questions about the formation. Omitted from
the list are various types of old logging tools (such as electric logs), and some standard auxiliary
devices, which, although important, do not rate as separate tools since they always piggy-back
along with one of the basic tools. Among those auxiliary tools are the spontaneous potential (SP),
and the caliper.
A discussion of common basic tools follows.
Induction Tools Induct ion tools belong to the resistivity tool family and measure apparent
formation resistivity. They work like mine detectors by inducing electrical currents in the
formation. They may be run simultaneously with a spontaneous potential (SP) or gamma ray
(GR) log (or both) and optionally with various combinations of porosity tools. Curves recorded on
a dual-induction log include deep induction, medium induction, shallow-focused electric, and SP
and/or gamma ray and caliper.
Laterolog Tools Laterolog tools also belong to the resistivity tool family. The most important is the
dual laterolog. This tool can be run with SP, GR, and caliper logs. The curves recorded are
laterolog deep, laterolog shallow and shallow-focused electric.
Microresistivity Devices Microresistivity devices attempt to measure formation resistivity in the
zone very close to the borehole wall where invading mud filtrate has displaced any moveable
formation fluids. They are all variations of a basic microfocused electric log. When certain
constraints on hole conditions are met, these devices produce a measurement of the parameter
Rxo, the resistivity of the flushed zone surrounding the borehole.
Acoustic Tools The modern acoustic log commonly used is known as the borehole compensated
sonic, or BHC. It may be run with a GR, SP, and caliper, or in combination with other porosity
and/or resistivity-measuring devices. Long-spacing sonic tools and tools with multiple transducers
are also in use for special applications.
The curves recorded are D (sonic travel time), GR, SP, and caliper (optional).
Various other acoustical parameters can also be recorded, either simultaneously or on a separate
run. Sonic amplitude logs are used for fracture detection. They may be recorded by various
arrangements of gates for the received wave trains. The tool may also be used to record the
cement bond log (CBL), in which case the recorded curves are D (a single-receiver travel time),
amplitude, and VDL (a variable-density display, or wave trains).
Density Tools Compensated formation density tools are also known as gamma-gamma tools in
some parts of the globe (because their mode of operation is to send gamma rays to the formation
and detect gamma rays coming back).
They record two basic curves b (bulk density) and  (correction) .
Natural gamma ray and caliper tools are normally run simultaneously. Additionally, an apparent
porosity curve can be generated and recorded and, from those data, a formation factor (F) curve
can be generated and recorded as well. A density-derived F is referred to as FD.
A variation to the density tool is known as the lithodensity tool; in addition to measuring bulk
density, it measures the photoelectric factor Pe. Pe is a direct indicator of formation lithology and,
as such, is a valuable adjunct to the basic density measurement.
Neutron Tools There are several types of neutron tools. Today's standard is the compensated
neutron log, which records N, the neutron porosity index, normally recorded for a particular
assumed lithology. Reading the porosity curve requires close attention to the porosity scale and
the assumed matrix. Normally, a natural gamma ray curve is recorded simultaneously with the
neutron log. The standard presentation is a combination density/neutron, where the caliper from
the density survey is also available.
Pulsed Neutron Log The pulsed neutron tool measures the formation capture cross section for
thermal neutrons. The end result is a measurement that helps distinguish oil from salt water in the
formation in cased holes.
The curves appearing on the log are:
 Sigma, the formation capture cross section
T tau, the thermal neutron decay time
 ratio, a porosity-type curve.
Dipmeter Dipmeters come in several versions: four-arm dip-meters, six-arm dipmeters, and eight-
electrode types. High-resolution dipmeters record all the necessary curves for computing
formation dip, hole drift, and azimuth.
Wireline Formation Testers (RFT) There are several types of wireline formation testers available,
which are proving to be a valuable addition to the formation evaluation arsenal. These devices
allow a small sample of formation fluid to be drained from the formation and brought up for
analysis. They also allow multiple formation pressure tests to be conducted during one run into
the hole.
Carbon/Oxygen Logging This relatively new service uses inelastic fast-neutron scattering to
attempt to measure directly the relative abundance of carbon, oxygen, and other elements in a
formation. Its application is in cased holes, and it is a natural candidate in those parts of the world
where fresh formation waters preclude the use of a pulsed neutron-logging survey.
Gamma Ray Spectral Log This service measures the number and energy of naturally occurring
gamma rays in the formation and distinguishes between elements and daughter products of three
main radioactive families: uranium, thorium, and potassium. Since these elements and/or their
decay products are associated with certain distinct types of mineralogy, sedimentology, and
formation waters, the service has obvious appeal.
Borehole Gravimeter The borehole gravimeter measures perturbations in the gravitational
acceleration constant caused by the proximity to the borehole of rock material that is denser or
less dense than normal. Thus, this tool can spot higher porosities, gas, and the like. Its use
requires an exacting set of prerequisites relating to depth, temperature, time, and so forth; it may
not be available or applicable to all wells everywhere.
Dielectric Logging These tools send microwaves along the wall of the wellbore. The speed and
attenuation of these electromagnetic waves are measured and the dielectric constant of the
formation is deduced. Oil and water, having very different dielectric constants, can be
distinguished. The application is in open holes where formation waters are fresh.
Nuclear Magnetic Resonance Measures the precession rate of hydrogen nuclei after the removal
of an intense magnetic field. The measured quantity is related to the free fluid content of the
formation. Recent advances allow determination of formation porosity, permeability, and
irreducible water saturation.
In order to put these tools, surveys, and curves in perspective, the section Logging Tools: Quick
Reference sets out a summary of all the common logging tools, what they measure, and their
uses. Included in this catalog of common wireline logging measurements are some common
interpretive presentations derived from the basic measurements.
Formation "Texture" Indicators Include both electrical and ultrasonic borehole wall imaging
devices that reveal near wellbore sedimentary details as well as wellbore intersections with
fractures and fault planes. These devices are particularly valuable in carbonate formations where
various forms of secondary porosity are imaged.
4. LOGGING PROCEDURES
Choosing a Logging Suite
A logging suite should be selected on the basis of
• type of well--wildcat or development
• hole conditions--depth, deviation from vertical, hole size, mud type
• formation fluid content--(fresh or salt) connate water
• formation type--clastic or carbonate
• economics--rig time, logging dollars, and so forth
Each tool is designed with a specific set of conditions in mind. Outside these limitations, the tool
fails to provide the required measurement and its use is discouraged.
Depth, Pressure, and Temperature Considerations
The majority of logging tools are rated at 20,000 psi and 350 F. These parameters are adequate
for logging most holes. For higher temperatures, special tools are available from the logging
service companies.
Hole Size and Deviation
Six inches is the standard minimum hole size for correct and safe operation of normal logging
tools. Some slim-line, small-diameter tools are available for smaller-diameter holes on a limited
basis. Maximum hole diameter is difficult to define. Most pad contact tools (compensated
formation density logs, microfocused logs, dipmeter, and the like) have spring-loaded,
hydraulically operated arms that push the relevant sensor against the borehole wall. The arms
open to about 20 in., although this limit varies a little from tool to tool. If holes are deviated, good
pad contact may still be obtained, since the tool will "lean" on the low side of the hole. However,
this cannot be guaranteed. Running a pad contact tool in a hole greater than 20 in. in diameter is
risky because the pad may not be able to make contact with the wall of the wellbore. Similarly,
tools that need to be run eccentered--for example, the compensated neutron tool--are less
accurate in enlarged holes. Resistivity devices, such as induction and laterolog, suffer in a
progressive fashion as the borehole gets bigger. Theoretically, there is no fixed limit to the hole
size. Practically, however, there is a limit because borehole corrections to the raw data get so
large that nothing useful can be determined from the logs.
Logging of large-diameter surface holes may thus cause a problem and require logging in a
purposely drilled medium-sized hole that is subsequently underreamed to the desired gauge. In
today’s offshore environment, the deviated hole is the norm rather than the exception. The
greater the angle of deviation from vertical, the greater the difficulties of physically getting a
logging tool to the bottom of the hole. In general, hole deviation greater than 40 degrees causes
problems. A number of techniques have been tried to get logging tools safely to bottom. Among
them are
• keeping the openhole section as short as possible
• removal of centralizers and standoff pads
• use of a "hole finder," a rubber snout on the bottom of the logging tool string
• use of logging tools especially adapted to be run to the bottom of the hole on drillpipe
In difficult situations, the hole may have to be logged through open-ended drillpipe with a slim
logging tool physically pumped down by mud circulation. Using this technique, holes with
deviations as high as 65 have been logged.
Logging Programs
Logging combinations generally consist of one resistivity device and one porosity device.
However, where hydrocarbon reservoirs are more difficult to evaluate, several porosity devices
are needed to provide more accurate porosity data and lithology information. In addition, the
reservoir engineer, the completion engineer, and the geophysicist may need additional
information for evaluation and completion of the well. With the addition of computers to aid in
formation evaluation, such comprehensive logging programs offer greater utilization of the
measurements recorded.
General Logging Program lists recommended logging programs for most logging situations. Mud
resistivity, formation water resistivity, hole conditions, and formation types dictate the type of
devices needed. The extent of the logging program is also a function of the information obtained
in previous wells.
A cross-reference list of tool nomenclature of the various service companies is presented in the
Reference Section under Service Company Terminology.
Influence of the Mud Program
The mud type influences the choice of logging tool, especially the choice of resistivity tool.
Air-drilled holes, which have no conductive fluid in them, must be logged with an induction device.
Likewise, holes drilled with oil can only be logged with an induction log. Where conductive fluids
are in the borehole for logging operations, the choice between induction and laterolog devices is
controlled by the salinities of the mud and the formation water. Fresh muds and salty formation
waters favor the induction log, and salty muds favor the laterolog.
All samples should be protected from excessive fluid losses so that porosity and saturation can
be adequately determined. Bit cuttings can be sufficient to interpret lithology and determine
proper constants for log evaluation formulas. Thus, the mud program should be designed for both
the drilling and the logging operations.
It is possible for a logging program to succeed or fail strictly because of the design of the mud
program. For example, filtrate from a high-water-loss mud can invade a formation so deeply as to
mask the measurement of true resistivity, reduce the amplitude of the spontaneous potential
curve, obscure the detection of the residual hydrocarbons, and result in water recovery on a
drillstem test from zones that would otherwise produce oil. Invasion of oil from oil-base or oil-
emulsion muds can increase the resistivity (Rxo) and decrease the water saturation (S xo) of the
invaded zone. This effect would erroneously indicate the presence of oil in water-bearing
formations, or reduce formation porosity values calculated from microresistivity devices.
The practice of "mudding up" just before reaching the objective zone affects interpretation when
mud filtrate invades the formation beyond the radius of investigation of the resistivity device.
Friable formations, as well, drilled with natural high-water-loss muds are usually badly washed out
and can prevent the logging tools from going down the hole because they hang up on ledges
and/or bridges. Borehole contact devices cannot obtain effective contact with the side of the
borehole in highly rugose holes and will give erroneous measurements. Normally, the extent of
washouts through shale is in proportion to the water loss of the muds (i.e., the higher the water
loss, the larger the washouts). Since many development and semiwild cat wells are drilled with
natural high-water-loss muds through the shallower formations, reliable analysis of logs through
these intervals is most difficult. The decision to drill with natural high-water-loss muds through
shallow formations is normally based on the erroneous assumption that the shallow formations
are of no interest. However, the logs through the shallow formations are invariably consulted later
to find zones for recompletion, to determine prospects for new hydrocarbon-bearing zones in the
area, to locate and evaluate high-pressure zones, and for general correlation work.
Choosing When to Log
Logs should be run just prior to the running and setting of a casing string. Once casing is set,
logging choices are severely limited.
It is recommended that logs be run (1) if hole conditions suggest that a section of hole could be
"lost" (caving, washouts, etc., which would contraindicate the running of a logging tool), (2) if
cuttings indicate that an unexpected formation has been encountered, and/or (3) if one is
otherwise "lost" structurally.
However, one’s enthusiasm for running logs should be tempered somewhat by the economic and
practical realities of service company price lists and fee structures. Each time a logging truck is
called, a setup charge is assessed to cover costs of mobilization. In addition, a depth charge is
assessed per foot of hole from surface to total depth. Finally, a survey charge is assessed over
the actual interval logged. The full cost of a logging operation is thus, more than anything else, a
function of the depth of the well. To log a l00-ft section at 10,000 ft is an expensive proposition,
while a 4000-ft survey at 5000 ft total depth is probably less expensive.
5. ROCK FLUIDS AND PROPERTIES

5.1 INTRODUCTION

Definitions
Porosity is defined as the ratio of the void space in a rock to its bulk volume. There are two
components to a porous rock system: the grain volume VG and the pore volume V p. The sum of
the two gives the bulk volume VB.
VB = VG + VP
Porosity is thus the ratio of VP to VB

It can be measured in a number of ways; for example,

or
Deriving a value of porosity depends on the mechanism of the porosity-measuring device and
knowledge of any two of the three volume fractions.

Saturation is defined as the ratio of the volume of saturating fluid to the volume of the available
storage space (i.e., the pore space). Thus, the water saturation of a porous system is simply
given by

where Vw is the volume of water.

Permeability is defined as the ability of a porous system to allow fluids to flow through it.

Provided flow is laminar, Darcy’s relation can be used to define permeability, k, in terms of flow
rate, Q; area, A; length, L; pressure differential,  P; and fluid viscosity,  such that

If only one fluid is present in the pore system, this relation defines absolute permeability -- i.e., a
rock property independent of the fluid flowing through it.

If Q is in cc/sec, A is in sq cm,  P \ L is in atmospheres/cm, and  is in centipoise, then k is in


darcies. The practical unit is the millidarcy, abbreviated md, equal to 1/1000 of a darcy.

The relationship between permeability and porosity depends on rock type. In general, the log of
permeability is linear with porosity for a given rock type; however, the precise relationship is found
only through direct measurements of representative rock samples. Figure 1 shows some of these
trends.
5.2 POROSITY

Introduction

Petrophysics is the name given to the study of rock-fluid systems. It is particularly important that
the log analyst he aware of the way in which rocks and fluids interact in both static and dynamic
situations.

Although logging measurements are made under static reservoir conditions, the prediction of
reservoir behavior under dynamic flow conditions can only be made if the physics of fluid flow is
understood. The objective of this discussion, therefore, is to equip the formation evaluator with
sufficient information so that log response can be related to reservoir performance, which is what
really counts, rather than to just static reservoir content. Ideally, the reader will come away with a
better understanding of why some reservoirs with low water saturations produce with high water
cut while others with much higher computed water saturations produce water-free hydrocarbons.

The Genesis of Reservoir Rooks

A reservoir rock is one that has both storage capacity and the ability to allow fluids to flow through
it; i.e., to be of practical use it must possess both porosity and permeability. Porosity (void
spaces) can develop between grains of sediments as they are laid down--for example,
intergranular porosity in sandstone reservoirs. Porosity can also develop when chemicals react
with rocks after they have been deposited. Typical of this solution-type porosity are carbonate
reservoirs. Porosity can also develop as fractures induced by the stresses of tectonic movement.
Porosity per se does not guarantee permeability. Swiss cheese, for example, is highly porous but
impermeable, as the void spaces are not connected.

Porosity The porosity developed in sedimentary rocks is a function of many variables--broadly


defined as rock texture--including grain shape, size, orientation, and sorting. If all the grains are of
the same size, sorting is said to be good. If grains of many diverse sizes are mixed together,
sorting is said to be poor.

The packing of the grains (Figure 1 ) determines the porosity. For a given sorting, porosity is
independent of grain size. For example, if spheres of diameter d are packed in a cubic lattice
arrangement, the porosity can be calculated by the following method.
In unit volume n3 spheres are packed n to a side. The total volume is (nd) 3. The volume of any
one sphere is (4/3)(d/2)3, so the volume occupied by n 3 spheres is (4/3)(nd/2)3. Thus the
porosity is

which simplifies to (1 -  /6) or 0.4764. Note that the term d cancels out and is not a determining
factor.

Cubic packing is not an efficient way to store spheres in a box. Nature seeks more compact
packing mechanisms, such as rhombohedral packing, which produces a porosity of 25.95%
(versus 47.64% for cubic packing).

For a given grain size, porosity decreases as sorting gets poorer, since intergranular pores may
be occupied by ever smaller grains.

Quite apart from the mechanics of how sand grains are packed is the question of their
compaction with regard to depth of burial. Porosity decreases with increasing depth in a
predictable manner. A relationship of the sort

generally fits most normally pressured reservoirs; i.e., the log of porosity is linear with depth. For
example, if fo, the porosity at surface, is 45% and depth is in feet, then a typical value of a might
be 12,000, resulting in a porosity of 12.9% at 15,000 ft and 8.5% at 20,000 ft.

Permeability while porosity is a static property of a rock, permeability is a dynamic one.


Permeability is a measure of the ability of a rock to allow fluids to flow through it. Provided the
flow is laminar, Darcy’s relation can be used to define permeability (k) in terms of flow rate (Q),
area (A), length (L), pressure differential (P), and fluid viscosity (µ), such that,

If only one fluid is present in the pore system, then this relation defines absolute permeability; i.e.,
it is a rock property independent of the fluid flowing through it. If Q is in cc/sec, A is in sq cm, P/L
is in atmospheres/cm, and µ is in centipoise, then k is in darcies. The practical unit is the
millidarcy, abbreviated md, equal to one-thousandth of a darcy.
The relationship between permeability and porosity depends on rock type. In general, the log of
permeability is linear with porosity for a given rock type. However, the precise relationship is
found only through direct measurement of representative rock samples. Figure 2 shows some of
these trends.

Over the years various investigators have developed theoretical relationships between
permeability and porosity, taking into account the textural features such as the size, shape, and
distribution of pore channels in the rock. Among these is the Carmen relationship,

k=3 / C(As)2
where C is the Kozeny constant and As is internal surface area per unit bulk volume.

For fracture systems, generalized formulas have been developed relating the permeability to the
square of the fracture width.

In some reservoirs, permeability is a vector; i.e., it takes on directional properties. Depositional


effects may tend to align grains along their long axis, increasing the permeability in that direction.
Vertical permeability may also be different from horizontal permeability. In fractured reservoirs
permeability is likely to be highly directional, depending on the azimuth of the fracture planes.

Summary

The conduction of electric current through a porous rock is conceptually similar to the flow of fluid
through the rock. Thus, measurement by wireline tools of formation conductivity is related to
formation porosity, permeability, and fluid saturation. By combining the basic relationships
established by Archie with the physics of fluid distribution and flow in a reservoir, the analyst may
estimate productivity from the free-water level, the length of the transition zone, and the
irreducible water saturation.

Definitions of Porosity and Effective Porosity

The porosity of a formation is commonly defined as the volume of the pore space divided by the
volume of the rock containing the pore space. This definition of porosity ignores the question of
whether the pores are interconnected or not. Swiss cheese, for example, is quite porous, but is of
very low permeability since the void spaces are not interconnected. Inter granular porosity that is
interconnected is effective porosity. Pores that are blocked in some way (by clay particles, silt,
etc.) are ineffective. Thus a preferred definition gives total porosity (T) as the volume of the pores
divided by the volume of rock, and effective porosity (e) as the volume of interconnected pores
divided by the volume of rock. Figure 3 illustrates this concept.
Porosity is expressed as a fraction of the bulk volume of the rock. The normal convention in
reservoir engineering is to express porosities in percentage units; e.g., a porosity of 0.3 is
referred to as 30% porosity. However, another term frequently used is porosity unit, or P.U. By
using unit rather than percentage, a lot of confusion is avoided, as, for example, in comparing a
20 P.U. sandstone with a 25 P.U. sandstone. The latter is 5 P.U. higher than the former. This
negates the confusion caused by saying one is 5% (or 25%) better than the other.

Types of Porosity

Porosity may develop in a formation by a variety of mechanisms. Where pores are uniformly
distributed throughout the bulk rock, the porosity is referred to as matrix porosity. Where the only
storage space in the rock system is in cracks and fissures in an otherwise zero porosity matrix,
the porosity is referred to as fracture porosity. A third type of porosity may coexist with either of
the other types in the form of vugs, and is referred to as vuggy porosity.

Matrix Porosity

Matrix porosity is common in sandstone and other granular rock formations. The physics of the
porosity measurement is unaffected by the manner in which the void spaces were created; i.e., it
is not important whether the porosity was originally created by sedimentation of individual grains
or by leaching by acidic solution after deposition. Thus, individual logging tools cannot tell directly
the type or origin of the matrix porosity in a rock sample. Petrographic analysis of cores is
required for that kind of information.

Fracture Porosity

Fracture porosity is unevenly distributed throughout the rock. It appears normally as near-vertical
cracks, or fractures, whose orientation depends on the azimuth of the stresses in the formation.
Not all logging tools respond to fracture and/or vuggy porosity in the same manner. Thus, it is
sometimes possible to distinguish fracture and/or vuggy porosity from matrix porosity with
judicious use of a combination of porosity-measuring devices and careful analysis of the results.
5.3 RELATIVE PERMEABILITY

If only one fluid is present in a pore system, fluid flow is nicely governed by Darcy’s law. If two or
more fluids are present together in a pore system, the dynamic behavior of the individual phases
is not quite so straightforward. Consider the case of oil and water together in a pore system. The
effective permeability is defined as the permeability to a particular phase at a particular
saturation. Thus, if, under a given pressure gradient, oil and water flow through a pore system
together, we find that

and

Furthermore, we find that the total flow rate Q t = (Qo + Qw) is less than the flow rate either phase
would have if it were at 100% saturation. Thus it appears as though the two phases interfere with
each other’s progress through the pore system.

A useful way to quantify this phenomenon is to define the relative permeability, k r. This is the ratio
of the effective permeability of the rock to one phase divided by the absolute permeability, and it
is quoted at some particular saturation value.

kro = kr/k

krw = kw/k

Figure 1 shows typical relative-permeability curves. Several things are worth noting. Relative
permeability to oil at irreducible water saturation is 100% or 1, and as water saturation increases,
kro decreases until it effectively reaches zero at some high water saturation corresponding to S or,
the residual oil saturation.
Relative permeability to water, on the other hand, commences effectively at zero when the rock is
at Swi and thereafter increases as Sw increases. It should also be noted that in an oil-wet system
kro is always less, at a given Sw, than in a water-wet system. Conversely, krw is always greater in
an oil-wet system than in a water-wet one.

A common way of representing these differences is a plot of the relative permeability ratio, k rw/kro,
versus water saturation, Sw. Figure 2 shows that in water-wet systems the relation is such that if
the krw/kro ratio is on a log scale and Sw on a linear one, a straight line is obtained. In an oil-
wet system an S-shaped line is observed.
When plotting relative permeability curves, the distinction is usually made between two possible
scenarios: imbibition and drainage. Imbibition refers to the case in which the wetting fluid is
increasing in saturation. For example, in a water-wet reservoir a rise in the water table subjects
the transition zone to imbibition of water. Drainage refers to the case in which the wetting fluid
saturation is decreasing, as, for example, when oil first migrates into a water-wet rock. The
difference between the two sets of relative permeability curves reflects the saturation history and
the trapping of the nonwetting phase that occurs after it has been imbibed. Figure 3 illustrates
these different cases.

Many workers in this field have proposed generalized empirical equations to relate k ro and krw to
Sw, Swi, and Sor Of particular note are those cited in Honarpour, Koederitz, and Harvey 1982,
Molina 1983, and Pirson, Boatman, and Nettle 1964. A commonly used approximation gives

If a well is completed above the transition zone where the reservoir is at irreducible water
saturation (i.e., krw = 0), water cannot be produced. However, if completion is contemplated in the
transition zone, it is helpful to know in advance what water cut may he expected. Fortunately this
can be calculated as follows:
The oil-flow rate is Qo = ko • P • A/µo • L

The water-flow rate is Qw = kw • P • A/µw • L

Thus the water-oil ratio is given by


WOR = kw µo/ko µw
The ratio kw/ko is numerically equivalent to krw/kro, which can be deduced from measured relative
permeability ratios or estimated from one of the generalized correlations.

The actual water cut of the production into the wellbore is given by

WC = Qw/(Qw + Qo)
which is equivalent to
WC = WOR/(l + WOR)
Surface water cut is a function of the formation volume factors of the oil and water, so the
complete expression is
WC = WOR o/(w + WOR • o)
5.4 FLUID DISTRIBUTION IN THE RESERVOIR

Initially sediments are laid down in water--either in river and lake beds (continental) in deltas and
along shore lines (transitional), or on the continental shelves (marine), as illustrated in Figure 1 .
Eolian dune sediments, initially deposited in a water-free environment, are the exception to
this rule.

Later in geologic time, after the reservoir rock has been buried, hydrocarbons migrate into the
reservoir. Because of gravity segregation, gas accumulates above oil, and oil over-lies water. In
the absence of any rock, gas, oil, and water form distinct layers with sharp contacts between
each phase. In the reservoir, however, the contact lines between gas, oil, and water become
blurred. To understand why this occurs, consider the simple case of a reservoir containing oil and
water.

Figure 2 shows such a reservoir. It is divided into three sections. The section at the top is
mainly oil, the section at the bottom is all water, and the section in the middle has ever-increasing
amounts of water as depth increases. Plotted on the right-hand side of the Figure is a curve of
water saturation, together with a plot of the pressure of the fluids in the pore space.

In order to understand the shape of the water saturation curve in the transition zone, consider the
classical experiment of a small glass tube held in a beaker of water ( Figure 3 ). A capillary
tube of radius r is found to support a column of water of height h. If the density of the air is a and
the density of the water is w, then the pressure differential at the air-water contact is simply (w -
a)h. This pressure differential acting across the cross-sectional area of the capillary is exactly
counterbalanced by the surface tension, T, of the water film acting around the inner
circumference of the capillary tube. If, at the water-glass interface, the contact angle is, then at
equilibrium we have

2rT cos = (w - a)h • r2

Force = Pressure • Area

By simplifying and rearranging this expression we have

We see that the smaller r gets, the larger h gets.

Translating this laboratory observation into terms of reservoir fluids, we can see that water can be
drawn up into what would otherwise be a 100% oil column by the capillary effect of the small
pores present in the rock system. We can equate the air in our experiment with oil, water with
water, and the tube with pore throats. Thus the maximum height, h, to which water can he raised
is controlled by the following factors:

• the surface tension, T, between the two phases (here oil and water)

• the contact angle,  , between the wetting fluid (water) and the rock

• the radius of the pore throats (r)

• the density difference between the phases (w - o in this case)

Given these factors, it is simple to predict the length of a transition zone in a reservoir. Reservoirs
with large pore throats and high permeability have short transition zones, and the transition zone
at a gas-oil contact is shorter than that at an oil-water contact simply because of the interphase
density differences involved ( Figure 4 ).

Since a pore system is made up of a variety of pore sizes and shapes, no single value of r can be
assigned to one reservoir. Depending on the size and distribution of the pore throats, certain
available pore channels will raise water above the free-water level. The water saturation above
the top of the transition zone will thus he a function of porosity and pore-size distribution.

In a water-wet system, water wets the surface of each grain or lines the walls of the capillary
tubes. At the time oil migrates into the reservoir the capillary pressure effects are such that the
downward progress of oil in the reservoir is most strongly resisted in the smallest capillaries. A
distinct limit is reached to the amount of oil that can be expected to fill the pores. Large-diameter
pores offer little resistance (Pc is low because r is big). Small-diameter pores offer greater
resistance (Pc is high because r is small). For a given reservoir, o and w determine the pressure
differential that an oil-water meniscus can support. Thus, the maximum oil saturation possible is
controlled by the relative number of small and large capillaries or pore throats. This maximum
possible oil saturation, if looked at in terms of water saturation, translates into a minimum
possible water saturation, and this is referred to as the irreducible water saturation, Swi Shaly,
silty, low-permeability rocks with their attendant small pore throats tend to have very high
irreducible water saturations. Clean sands of high permeability have very low irreducible water
saturations. Figure 5 illustrates this important concept by comparing capillary pressure curves for
four rock systems of different porosity and permeability.
5.5 LABARATORY MEASURMENTS
Measuring Porosity

Porosity may be measured by a variety of methods, including

• borehole gravimetrics
• wireline logging
• core analysis

Each method investigates a different volume of the formation. The borehole gravimeter samples
very large volumes in the order of 103 to 106 cu ft. Wireline logging tools investigate a much
smaller volume, on the order of 1 to 10 cu ft, depending on the specific porosity device used.
Core analysis investigates much smaller volumes, ranging from 10 -3 to 10-1 cu ft. From one
extreme to the other lie nine orders of magnitude, so we should not be surprised to learn that
porosity estimates using different tools and techniques do not always agree.

Measuring Permeability

There are many ways to estimate permeability, including

• pressure buildup from drillstem tests


• pressure drawdown and buildup from wireline repeat formation testers
• log analysis
• core analysis

Again, each method investigates an effective radius different from the others by several orders of
magnitude. In increasing order, these are:

Method Approximate radius (ft.)


DST 102 - 104

RFT buildup 10 - 102

RFT drawdown 10-2 - 100

log analysis 10-1 - 5

core analysis 8 x 10-2 - 3 x 10-1

It should come as no surprise that these different methods of measurement occasionally


produce disparate results, especially in a heterogeneous reservoir. Where the drilling process
has caused clay swelling in the invaded zone, it is also to be expected that measurements
made near the wellbore (logs, RFT tests) will reflect permeabilities that are lower than true
permeabilities. Care must be taken in using the results of permeability measurements made
on cores; they vary, depending on the type of fluid (air or brine) used for the measurement
and the pressure and temperature of the sample at the time of the measurement (standard or
reservoir conditions).

Many investigators have attempted to correlate rock permeability to measurements made by


wireline logging tools. These measurements fall into two categories--those that apply above the
transition zone and those that apply only in the transition zone--among which are:

k = 8581 • 4.4 • Swi -2 (Timur)


k = [250 • 3 • Swi-1]2 -oils (Schlumberger) or
k = [79 •  3 • Swi-1]2 -gas (after Wyllie and Rose)

for O/W (Raymer and Freeman 1984) or

for G/W

where:

Swi is fractional irreducible water saturation

 is fractional porosity
h is height in feet from free-water level to the top of the
transition zone

w is the water density in gm/cc

o is the oil density in gm/cc

g is the gas density in gm/cc

The first three equations (Timur, Schlumberger, Wyllie/Rose) apply to points above the transition
zone, since that is the only Place that Swi can be measured. The fourth equation applies to the
oil/water transition zone, and the fifth to the gas/water transition zone.

In the transition zone, the resistivity gradient is usually linear; i.e., a resistivity log on a linear scale
shows a straight line in a transition zone. The resistivity gradient may then be related to k,
provided the density difference between the wetting and nonwetting phases is known.

Figure 1 gives a graphical solution to the equation

where

and c = 20.

rw and rhy are in gm/cc, DR/DD is expressed in ohms/ft and k is in md.

Measuring Saturation

Fluid saturations for the most part are adequately measured by log analysis techniques, provided
formations are clean and connate waters are saline. Problems arise with shaly formations and
fresh-water-bearing formations. Other methods of saturation determination are available from
proper coring and subsequent core analysis techniques. Mud logging can provide a qualitative
measure of oil and gas saturations.

Many similarities exist between the flow of fluids through a rock and the flow of electric current
through a rock. The permeability to water, for example, can be equated with the electrical
conductivity of a porous system, since both depend on interconnected pores. In cases in which
both oil and water are present in a pore system, a parallel also exists between relative
permeability to water and electrical conductivity of oil-bearing sand.

Investigation of the electrical properties of water- and oil-hearing rocks was pioneered by Archie
(1942). To follow the development of his experimental observations, let us examine the electrical
behavior of electrolytes and water-filled rocks.
Water Resistivity (Rw) Connate waters range in resistivity from about 1/100 of an ohm-m up to
several ohm-m, depending on the salinity and temperature of the solution. To find water
saturation by quantitative analysis of porosity and resistivity logs, a value of R w is required.

For our basic concept, we only need to understand that the ability of a rock to conduct electricity
is due entirely to the ions in the water found in the pore spaces. Figure 2 shows a cube of rock
with a system of cylindrical tubes drilled through it. If the cylindrical "pores" are filled with
water of resistivity Rw, their total area is A, and their length is L, we can estimate that the
resistivity of the total rock system is proportional to Rw • L/A.

If the area A is small, there is a small conductive path per length L and the resistivity of the rock
system is high. Conversely, if A is large, the resistivity is low. The resistivity of a rock 100%
saturated with water is referred to as Ro. It can be seen that A is proportional to the porosity itself.
Thus we may write

Ro = f (Rw,f)

or that Ro is related to Rw by some formation factor, F, such that

Ro=F • Rw

Electrical Formation Factor The method Arch is used to arrive at this conclusion was simple. He
took a number of cores of different porosity and saturated each one with a variety of brines. He
could then measure, at each brine salinity, the resistivity of the water R w, and the resistivity of the
100% water saturation rock system, Ro. when the results were plotted, he found a series of
straight lines of slope F, as shown in Figure 3 .

Archie conducted many experiments that showed that the formation factor is related to porosity in
a predictable manner. Our simple tubular model bears little relationship to the tortuous paths that
pores actually take. The factor L (the length of the tubular pore) grows larger as the tortuosity of
the pore system increases. Figure 4 illustrates the concept.

Note that by definition the formation factor is the ratio of Ro/Rw i.e., the resistivity of a rock sample
100% saturated with water to the resistivity of the water itself. Archie found that laboratory-
measured values of F could also be related to the porosity of the rock by an equation of the form

F = a / fm
where a and m are experimentally determined constants. In porous formations, a is usually close
to 1 and m is usually close to 2 ( Figure 5 ).

Three commonly used formation-factor-to-porosity relations are:

F = 1/ 2 (compacted formations
and chalky rocks)

F = 0.62/ 2.15 (Humble formula-soft


formations and sucrosic
rocks)

F = 0.81/ 2 (simplified Humble formula


for sands)

For a wet formation, we may combine the F to f relationship with the definition of F, and arrive at
the equation

Rw = Ro/F = Ro • fm/a

Saturation Archie's experiments showed that the saturation of a clean formation could he
expressed in terms of its true resistivity as Swn = Ro/Rt.

Since Ro = F • Rw, the water saturation equation can also be written

Swn = F • Rw/Rt,

where n is the saturation exponent and is usually set to 2 ( Figure 6 ).

Archie's classical relationships work well in clean formations, but not in shaly formations and
where connate water is fresh. Archie's model considers the electrolytes in the pores as the only
conductive path. As we shall see, an additional conductive path exists as the result of surface
conductance effects which only become noticeable when they begin to provide a substantial
percentage of the total rock system conductivity.

Permeability Estimates

As with porosity measurement, there are many ways to estimate permeability. These include

• pressure buildup from drillstem tests


• pressure drawdown and buildup from wireline repeat formation testers
• log analysis
• core analysis

Different methods of measurement may produce dissimilar results because there are many
orders of magnitude that separate the effective radius of investigation of each method. In
increasing order these are

Method Approximate radius of


investigation, ft

DST 102 - 104

RFT buildup 10 - 102

RFT drawdown 10-2 - 100

log analysis 10-1 - 5

core analysis 8 x 10-2 - 3 x 10-1

In a heterogeneous reservoir, such dissimilarities are to be expected. Where the drilling process
has caused clay swelling in the invaded zone, measurements made near the wellbore (logs,
cores, RFT tests) usually reflect permeabilities lower than actual permeabilities. Careful review of
the results of permeability measurements made on cores is necessary to distinguish between the
type of fluid (air or brine and its salt concentration) used for the measurement, and the pressure
and temperature of the sample at the time of the measurement (standard or reservoir conditions).

Many investigators have attempted to correlate rock permeability to measurements made by


wireline logging tools. These attempts fall into two categories: those that apply above the
transition zone and those that apply only in the transition zone. Among them are

Timur (1968)

(Oils)

(Gas)
where:

k = permeability (in md)

 = fractional porosity

Swi = fractional irreducible water


saturation

Figure 7 shows a graphical representation of the Wyllie and Rose relationship where f and S wi are
crossplotted to yield values of k.

In the transition zone a common observation is that the resistivity gradient is linear; i.e., a
resistivity log on a linear scale will show a straight line in a transition zone. The resistivity gradient
(Tixier 1949) then may be related to k, provided the density difference between the wetting and
nonwetting phases is known.

Figure 8 gives a graphical solution to the equation

and

C = 20

where rw and rh are densities of water and hydrocarbon in gm/cc; DR/DD is expressed in ohms/ft
and k is in md.

Laboratory Measurement of Porosity

Rock samples suitable for laboratory analysis may come from a variety of sources, such as
cuttings, sidewall cores and/or plugs, and conventional cores.

Depending on the source of the sample, the type of analysis made may be more or less
sophisticated. At worst, a good idea of the rock type and porosity can be obtained, and, at best, a
vast range of rock and fluid properties can be measured, including

• porosity
• fluid saturation
• permeability
• relative permeability
• wettability
• capillary pressure
• pore throat distribution
• grain size distribution
• grain density
• mineral composition
• electrical properties
• effects of overburden stress
• sensitivity to fluids
• hydrocarbon analysis

Since most of these rock/fluid system properties are of vital interest to the formation evaluator, it
is helpful to learn more about core analysis methods and the application of their results to log
analysis in particular and formation evaluation in general.

Sidewall cores are usually shot at preselected depths determined from wireline logs. Cuttings are
usually collected at 2-, 5-, 10-, or 20-ft intervals, and should be properly tagged and identified as
to source depth range. Whole cores are usually cut to driller's depth, which may be at odds with
wire-line logging depths; thus, a good starting point for whole core analysis is a gamma ray scan
of the core. The core is laid out in its shipping container and moved relative to a gamma ray
counter, which records a graph of radioactivity versus distance traveled along the core. This
record may then be compared directly with a wireline gamma ray log. Figure 9 shows an
example.

When cutting conventional cores, it is wrong to assume that the only formations of interest are the
clean reservoir rocks. Useful data may be extracted from shales as well, and the temptation to
high-grade the core at the wellsite by throwing shale sections into the outer darkness should be
resisted. The core gamma scan, for example, would be useless if the radioactive section of the
core were missing.

Depending on the particular analysis to which the core is to be submitted, either a plug is cut or
the whole core itself is used. Plugs are 1 to 1-1/2 in. in diameter and 1 to 3 in. long. Whole cores
are normally 5 in. in diameter and up to 60 ft long.

Fresh cores are cores cut with water- or oil-base muds preserved without cleaning. Native state
cores are those cut with lease crude as the coring fluid to minimize changes in rock wettability.
Restored cores are cleaned and dried prior to testing; their wettability and fluid distributions are
changed.

Cores may be epoxy coated, jacketed in heat-shrinkable tubing or metal (for unconsolidated
cores), or molded in lucite.

A porous rock system has two components: the grain volume and the pore volume. The sum of
the two gives the bulk volume:

VB = VG + VP
The porosity is defined as the ratio of the pore volume to the bulk volume, for example,

Thus porosity can be measured in a number of ways, such as

or

Provided any two of the three entities are measured, porosity can be deduced. Among the
commonly employed methods of deduction are

• summation of fluids method, in which the volumes of water, oil, and gas are
independently measured and then summed to give Vp. VB is deduced from the
dimensions of the core.
• Boyle's law method, in which the core is cleaned and dried. Air, or other gas, is allowed
to fill the pore space. When the sample is connected to another chamber filled with gas at
a different pressure, the gas in the core pore space expands.
• The final pressure in the system allows deduction of VP by use of Boyle's (Charles's) law.
Again, VB is deduced from the dimensions of the core.
• Washburn-Bunting vacuum extraction and collection of gas in the pore space, which is
somewhat similar to the Boyle's law method and measures VP.
• liquid restoration, which involves simply filling the pore space with a liquid of known
density and measuring the weight increase. VP is then deduced by dividing the weight
increase by the known liquid density.
• grain density methods, which require that both bulk volume and dry weight of the sample
be determined first. The sample is then crushed to grain-size particles and VG is
measured. VP is then deduced as the difference between VB and VG. A side benefit is
the estimation of grain density from the knowledge of dry weight and grain volume. The
disadvantage of the method is the physical destruction of the core sample.

Accuracy of core porosity measurements are claimed to be within a half P.U. Methods requiring
that the core be cleaned and dried are subject to error if the core contains hydrated clay material.
Heating such a sample in a retort may drive off water of hydration; the porosity measured thus
may be larger than effective porosity. Use of a humidity-controlled oven to dry samples alleviates
this problem.

Porosity measurements from sidewall cores may produce a value that is slightly different from the
average over the zone. Therefore, sidewall-core porosity values should be used only in addition
to other methods.

Porosity measurements should be "weighted" prior to their use, based on (1) the method used to
obtain them, (2) their anticipated application, and (3) the homogeneity of the reservoir. A core
plug is very localized and core plug porosities may be higher than whole core porosities if the
whole core includes portions of rock of a lower porosity. The selection of places to plug a whole
core is somewhat subjective and usually the "best-looking spot" is picked. This tendency should
be resisted, lest the core data become so high-graded that they no longer tie to the log. Regularly
spaced core plugs should be taken, regardless of the lithology. These data can then be
processed in a rolling average to mimic the wireline logging tool response, and thus produce
correlatable results.

Porosity measurements made from sidewall core plugs can be either higher or lower than true
porosities, depending on the porosity range. This condition is illustrated in Figure 10 by the
plotting of sidewall core porosity against conventional core porosity. In general, low-porosity
formations tend to fracture when sidewall core bullets strike them and hence induce additional
pore volume.

Porosity measurements may also be made on sample cuttings as small as a cubic centimeter or
less. Such measurements compare well to core plug porosities within one P.U.

Effects of Overburden Pressure

Although most basic rock properties are measured at atmospheric conditions, in some cases it is
important to make the measurements (especially those of porosity and permeability) at simulated
reservoir conditions. Both rock matrix and pore fluids are compressible, although matrix
compressibility is generally very low. Thus, measurements at standard conditions give overly
optimistic values for f and k. Figure 11 illustrates the effects of net overburden pressure on
permeability.

Depending on the type of material tested, reduction of up to 80% can be expected. Note that
permeability is an intrinsic rock property, independent of the fluid in the pore space.

When it comes to discussing porosity reductions that are the result of changes in pressure,
however, the total rock/fluid system must be considered, since both the rock matrix and the fluids
in the pore space are compressible. Thus, both overburden and pore pressure come into play.
Compressibilities are expressed in vol/vol/psi. It would not be uncommon to find a 6% reduction in
porosity caused by compressibility; e.g., if a sample in the lab measured 20%, the porosity at
depth might be 18.8%. Lab measurements of oil, water, and rock compressibility can be made
and the exact pore reduction factor deduced if reservoir pore and overburden stress are known.

Other pressure-sensitive parameters include acoustic velocity and formation factor. The effects of
pressure are to raise the lab-reported values by 20 to 30%.
5.6 POROSITY AND FORMATION FACTOR
Porosity may be measured by a variety of methods, including
• core analysis
• wireline logging
• borehole gravimetrics
Each method investigates a different volume of the formation. Core analysis investigates very
small volumes, ranging from l0-3 to 10-1 cu ft. Wireline logging tools investigate volumes on the
order of 1 to 10 cu ft, depending on the specific porosity device used. The borehole gravimeter
evaluates very large volumes in the order of 103 to 106 cubic feet. With nine orders of magnitude
(10-3 to 106), one should not be surprised to learn that porosity estimates using different tools and
techniques do not always agree.
If formations are clean and connate waters are saline, fluid saturations are usually measured by
log analysis techniques. Problems arise in shaly formations and where formation waters are
fresh. Other methods of saturation determination are possible through core analysis techniques.
Mud logging can qualitatively measure the presence of oil and gas, but is not available on all
wells.
Many similarities exist between the flow of fluids through a rock and the flow of electric current
through a rock. The permeability to water, for example, can be likened to the electrical
conductivity of a porous system, since both depend on interconnected pores. In cases in which
both oil and water are present in a pore system, there is a parallel between relative permeability
to water and electrical conductivity of an oil-bearing sand.
Investigation of the electrical properties of water- and oil-bearing rocks was pioneered by Archie.
A good starting point to follow the development of his experimental observations is the behavior
of electrolytes.
5.7 INTERPRETING PETROPHYSICAL DATA

When a formation is above the transition zone, i.e., at irreducible water saturation, the product of
 and Sw is a constant. variations of porosity are normal on a local scale, caused both by changes
in the depositional environment and by subsequent diagenesis. If porosity is reduced locally,
either a greater proportion of the pore throats are small or there are simply fewer pore throats.
Either way, the mean radius r is smaller; thus Pc is larger and more water can be held in the pore
maintaining the constant

 • Swi product .
This has a practical application. After a zone has been analyzed on a foot-by-foot basis for
porosity and water saturation, a plot of  versus Sw reveals the presence or absence of a
transition zone.

Figure 1 shows a plot on log-log paper. Here, the points at irreducible saturation plot on a
straight line and the points in the transition zone plot to the right of the irreducible line.

Reservoirs may be characterized by the  • Swi product, and this knowledge used as a basis for
predicting production characteristics. For points not at irreducible saturation, some water
production is to be expected, depending on the mobility ratio, (k wµo/koµw), for the particular fluids
present. Note that in a low-porosity, low-permeability formation, surprisingly high water
saturations can be tolerated without fear of water production. Conversely, in others with good
porosity and permeability, even with moderate values of Sw, water production can be expected.
This salient fact is all too often overlooked.
5.8 CONDUCTION OF ELECTRIC CURRENT IN POROUS ROCKS

Ohm’s law states that the potential difference, V, between two points on a conductor is equal to
the product of the current flowing in the conductor, I, and the resistance of the conductor, R.
Practical units of measurement are, respectively, the volt, the amp, and the ohm. Expressed as
an equation, the relationship is

V=I•R

Volts = Amps • Ohms

Of more interest in logging is the resistivity, rather than the resistance, of a rock. Resistivity is
defined as the resistance of a specific amount of a substance. It is further defined as the voltage
required to cause one amp to pass through a cube of face area one meter square. Figure 1
illustrates this concept. The unit of resistivity is the ohm-meter2/meter and abbreviated as
½m2/m or ½m.

When discussing formation resistivities, it is common to say "this is a 25-ohm sand" rather than to
say "this sand has a resistivity of 25 ohms meters squared per meter." So the field jargon, when
talking about resistivity logs, is to say "ohm" when "ohm m2/m" is really meant.

Why the need to know the resistivity rather than the resistance? Because resistance is a function
not only of the resistivity measured, but also of the geometry of the body of material on which the
measurement is being made. The geometry of the body is not of prime interest. The
measurement that characterizes the rock, as far as fluid content is concerned, is the resistivity,
not the resistance. The resistance of a wire stretching across the Pacific Ocean could be high, but
the resistivity of the wire itself could be very low.

Conductivity is the reciprocal of resistivity. A substance with infinite resistivity (empty space) has
a conductivity of zero, and a substance with low resistivity has high conductivity. Common units of
conductivity are milliohms, or 1/1000th of a reciprocal ohm-m.

conductivity (C) (in milliohms) = 1000/resistivity (in ohm-m)

Typical formation resistivities range from 0.2 ohm-m to 1000 ohm-m. Soft formations (shaly
sands) range from 0.2 ohm-m to about 50 ohm-m. Hard formations (carbonates) range from 100
ohm-m to 1000 ohm-m. Evaporites (salt, anhydrite) may exhibit resistivities of several thousand
ohm-m. Formation water, by contrast, ranges from a few hundredths of an ohm-m (brines) to
several ohm-m (fresh water). Sea water has a resistivity of 0.35 ohm-m at 75 F.
Types of Resistivity Measurements

Given an infinite isotropic homogeneous medium with a spherical electrode implanted in it


emitting a current I radially in a spherical distribution (see Figure 2 ), the voltage drop between
any two concentric spherical shells with radii  and  + d can be determined in the following
manner:

dV = I • dr
where dV is the voltage drop, I is the current, and dr is the resistance between the two shells. If
the resistivity of the medium is R, then
dr = R • d / 4 ¹ and,

dV = I • R • d / 4 ¹

Integrating this equation from  = A to  = M, the equation to determine the value for Vm becomes

where Vm is the measured voltage at some point a distance M from the current electrode A, and R
is the formation resistivity .

This ideal derivation does not fit the real world for two reasons. First, a borehole is required in
order to introduce an electrode into the formation, and, second, no formation is infinite and
homogeneous. Over the years, many improvements have been made to this simple, but
inadequate, method of measuring formation resistivity .

Resistivity

Electricity can pass through a formation only because of the conductive water it contains. With a
few rare exceptions, such as metallic sulfide or graphite, dry rock is a good electrical insulator.
But perfectly dry rocks are very seldom encountered; water is in their pores or absorbed in their
interstitial clay, therefore subsurface formations have finite, measurable resistivities.

The resistivity (R) of a formation depends on:


• the resistivity of the formation water

• the amount of water present

• the pore structure geometry

Resistivity, a key parameter in determining hydrocarbon saturation, is defined as the specific


resistance of a substance, i.e., the resistance of a specific amount of it. It is numerically
equivalent to the voltage required to cause one amp to pass through a cube of face area one
meter square. Figure 3 illustrates this concept. The unit of resistivity is the ohm-meter2/meter,
abbreviated as m2/m, or simply ohm-meter (ohm-m).

Conductivity is the reciprocal of resistivity and is expressed in mhos per meter (mho/m). A
substance with infinite resistivity (empty space) has a conductivity of zero and a substance with
low resistivity has high conductivity.

Conductivity is usually expressed in millimhos per meter (mmho/m), where 1000 mmho/m = 1
mho/m:

Typical formation resistivities range from 0.2 ohm-m to 1000 ohm-m. Soft formations (shaly
sands) range from 0.2 ohm-m to about 50 ohm-m. Hard formations (carbonates) range from 100
ohm-m to 1000 ohm-m. Evaporates (salt, anhydrite) may exhibit resistivities of several thousand
ohm-m. Formation water, by contrast, ranges from a few hundredths of an ohm-m (brines) to
several ohm-m (fresh water). Seawater has a resistivity of 0.35 ohm-m at 75 F.

Water Resistivity (Rw)

Connate waters range in resistivity from about 1/100 of an ohm-m up to several ohm-m,
depending on the salinity and temperature of the solution. A value of R w is required to determine
water saturation by quantitative analysis of porosity and resistivity logs.
The ability of a rock to conduct electricity is due entirely to the ions in the water found in its pore
spaces. Figure 4 shows a cube of rock with a system of cylindrical tubes drilled through it. If
the cylindrical "pores" are filled with water of resistivity R w, their total area is A, and their
individual length is L, we can estimate that the resistivity of the total rock system is proportional to
Rw L/A. By definition the resistivity of water-bearing porous rock systems is R0.

If the area A is small, there is a small conductive path per length L and the resistivity of the rock
system is high. Conversely, if A is large the resistivity is low. It can be seen that A is proportional
to the porosity itself. Thus we may write:

Ro = f(Rw,)
The resistivity of the formation water, Rw is an intrinsic property of the water and is a function of
its salinity and temperature. The higher these two variables, the more conductive the water and
the lower its resistivity.

Electrical Formation Factor

The method Archie used to arrive at the functional form of the relationship was simple. He took a
number of cores of different porosities and saturated each one with a variety of brines. He could
measure, at each brine salinity, the resistivity of the water, R w, and the resistivity of the 100%
water-saturated rock system, Ro. When the results were plotted, he found a series of straight
lines of slope F, as shown in Figure 5 . Archie determined that the relationship between Ro
and Rw is as follows:
Ro = F RW

Archie conducted many experiments that showed that the formation factor is related to porosity in
a predictable manner. Our simple tubular model ( Figure 6 ) bears little relationship to the tortuous
paths that pores actually take. The factor L, the length of the tubular pore, grows longer as the
tortuosity of the pore system increases.

By definition, the formation factor is the ratio of Ro/Rw; i.e., the ratio of the resistivity of a rock
sample 100% saturated with water to the resistivity of the water itself. Archie found that
laboratory-measured values for F could also be related to the porosity of the rock by an equation
of the form

where a is the cementation factor and m is the cementation exponent. The values of a and m are
experimentally determined constraints; a is usually close to 1 and m is usually close to 2 in
porous formations ( Figure 7 ).

Two commonly used formation-factor-to-porosity relations are

(carbonates)
(Humble formula – sands)

To eliminate the fractional cementation exponent, the Humble formula is sometimes simplified to

The exponent m can be as high as 3 in some severely ooliclastic packs.

In a wet formation we may therefore combine the F to 4 relationship with the definition of F and
arrive at the equation

Resistivity Index and Water Saturation

Archie’s experiments showed that the saturation of a core could be related to its resistivity. He
found that the fractional water saturation, Sw, was equal to the square root of the ratio of the wet
formation resistivity, Ro to the formation resistivity, Rt. That is,

In a more generalized form this equation can be written as

where n is the saturation exponent ( Figure 8 ). Laboratory experiments have shown n = 2.0 in
the average case.

These classic Archie’s relationships work well in clean formations. In shaly formations and where
connate water is fresh they do not work as well. Archie’s model considers the electrolyte in the
pores as the only conductive path. However, a second conductive path exists, due to surface
conductance effects.
5.8 ELASTIC WAVES PROPAGATION

Introduction

Many disciplines meet at a common point when formation evaluation is discussed from the point
of view of elastic waves. Elastic formation properties control the transmittal of elastic waves
through subsurface formations; indeed, the whole science of seismic evaluation is based on the
physics of rock elasticity. Acoustic logging is a localized, downhole branch of geophysics. By
properly combining measurements both from surface and downhole, a wealth of information can
be gathered concerning formation properties. For example,

• acoustic logs and check shot surveys can be used to "calibrate" seismic surveys

• combined acoustic and density logs can provide "synthetic seismic" traces
• combined acoustic and density logs can deduce formation mechanical
properties, used in turn to deduce pore pressure, rock compressibility, fracture
gradients, sanding problems, etc.
• acoustic logs, used in conjunction with other logs, can deduce porosity, lithology,
and fluid saturations
• in borehole measurements, acoustic logs can produce vertical seismic profiles
(VSP) that "see" below the bottom of the well
• acoustic tools may be used for cement bond logging in cased holes

Since the elasticity of subsurface formations is basic to all of these measurements and
interpreted answers, a good starting point is the study of elastic wave propagation through a
medium.

Propagation of Elastic Waves

Two types of sound waves are propagated in an infinite medium:


Compressional Waves. Compressional (or pressure) waves are longitudinal, that is, the
direction of propagation is parallel to the direction of particle displacement ( Figure 1 ). Gases and
liquids, as well as solids, tend to oppose compression, therefore compressional waves can be
propagated through them.

Shear Waves. Shear waves are transverse; that is, the direction of propagation is perpendicular
to the direction of particle displacement ( Figure 2 ). Shear waves can be propagated through
solids, owing o their rigidity. On the other hand, gases (and liquids having negligible viscosity)
cannot oppose shearing, and shear waves cannot be propagated through them. In practice,
viscous fluids do permit some propagation of shear waves, though they become highly
attenuated.

In a finite medium (e.g., a borehole) other types of waves are propagated. These are guided
waves, which include:

Rayleigh Waves. Rayleigh waves occur at the mud/formation interface and are a combination of
two displacements, one parallel and the other perpendicular to the interface. Their speed is
slightly less than the shear wave velocity (VRayleigh is 86% to 96% of VShear). When energy leaks
away from the interface as compressional waves are set up in the mud, the waves are then
referred to as pseudo-Rayleigh waves.

Stoneley Waves. Stoneley waves ("tube waves") can travel in the mud by interaction between
the mud and the formation. The amplitude of these low-frequency waves decays exponentially in
both the mud and the formation away from the borehole boundary. Stoneley wave velocity is
lower than the mud compressional velocity.

Proper interpretation of any measurement made using elastic wave data requires an
understanding of the elastic properties of a medium.
Elastic Constants

The properties derived from testing rock samples in the laboratory, such as measuring the strain
fore a given applied stress, are static elastic constants. Dynamic elastic constants are determined
by measuring elastic wave velocities in the material. Acoustic logging and waveform analysis
provide the means for obtaining continuous velocity measurements and, thus, knowledge of the
mechanical properties of the rock in situ.

The speed at which a wave travels through a medium may be expressed in two ways.
Geophysicists think in terms of velocity, i.e., distance traveled per unit of time. Subsurface
formation velocities range from 6000 to 25,000 ft/second. Log analysts think in terms of time, i.e.,
the time taken to travel one unit of distance. A convenient unit of measurement is the
microsecond per foot (µ sec/ft) given the symbol t. With these definitions in mind, the dynamic
elastic constants of a medium can be expressed as a function of bulk density ( b) and travel time
fore compressional and shear waves, tc and ts, respectively, as shown in Table 1 .
6. OTHER INFORMATION
NOMENCLATURE

Mud and Borehole Terms

dh Diameter of hole (in.).

hmc Mudcake thickness (in.).

Rm Resistivity of mud (ohm-meters).

Rmc Resistivity of mudcake (ohm-meters).

Rmf Resistivity of mud filtrate (ohm-meters).

Formation Terms

h Formation thickness (ft).

 Porosity = Fraction of formation volume that is pore space.

h Permeability (millidarcies) - Fluid flow characteristic of formation.

d1 Resistivity of connate water (ohm-meters).

Rw Diameter of invasion (in.).

Rwa Apparent resistivity of formation water (ohm-meters).

Ro Resistivity of uninvaded formation with pores completely filled with connate


water (ohm-meters).

Rt Resistivity of uninvaded (deep ) formation with pores containing both


connate water and hydrocarbon.

Ri An average resistivity of invaded zone; pores may contain a mixture of mud


filtrate, connate water, and hydrocarbons; a nebulous term, not commonly
used.

Rxo Resistivity of shallow zone completely flushed by mud filtrate; pores may
contain residual hydrocarbon as well as mud filtrate.

F Formation factor=Ro/Rw or Rxo/Rmf by definition. = 1/2(ls) or 0.81/2 (ss), by


experiment.

C Conductivity (mmho) = 1,000/resistivity (ohm-meters).

Sw Water saturation = Fraction of pore space in uninvaded zone containing


water = (Ro/Rt)1/2=

Sh Hydrocarbon saturation = Fraction of pore space in uninvaded zone


containing hydrocarbons = (1 - Sw).

So Oil saturation = Fraction of pore space in uninvaded zone containing oil.


Sg Gas saturation = Fraction of pore space in uninvaded zone containing gas;
(So + Sg = Sh).

Sxo Fraction of pore space in flushed zone containing water = (F R mf/Rxo)1/2

Shr Residual hydrocarbon saturation = Fraction of pore space in flushed zone


containing hydrocarbons.

Soxo Fraction of pore space in flushed zone containing oil.

Sgxo Fraction of pore space in flushed zone containing gas. (Soxo + Sgxo=Shr).

Terms Related to Logging Measurements

SP

SSP Static spontaneous potential = SP deflection (millivolts from shale line ) in


PSP thick, clean formation.

PSP Pseudo-static spontaneous potential = SP deflection (mV ) in thin or shaly


formation.

Resistivity - Deep
RILd or RID Resistivity measured by deep induction.

RLLD Resistivity measured by deep laterolog.

Rs Adjacent bed resistivity read by deep induction (for bed thickness


correction).

Resistivity - Medium

RILM or RIM Resistivity measured by medium induction.

RLLS Resistivity measured by shallow laterolog.

Rs Adjacent bed resistivity read by deep induction (or deep laterolog) .

Resistivity - Shallow

RSN or R16 Resistivity measured by short normal.

RLL8 Resistivity measured by laterolog-8.

RSFL Resistivity measured by spherically focused.

Resistivity - Flushed Zone (Pad Measurements)

RMLL Resistivity measured by microlaterolog.

RPL Resistivity measured by proximity tool.

RMSFL Resistivity measured by micro- spherically-focused logging tool.

Resistivity - Mudcake Controlled


R1"x1" Resistivity measured by extremely shallow microlog curve.

R2" Resistivity measured by slightly deeper microlog curve.

Sonic

t (or t) Travel time measured by sonic tool (sec/ft).

tma (or tma) Travel time of solid rock matrix.

tf (or tf) Travel time of 100% pore fluid.

Cp Compaction correction (³1).

s Sonic-derived porosity.

Density

b Bulk density measured by density tool (gm/cc).

ma Density of solid rock matrix (grain density).

g Density of pore fluid.

D Density-derived porosity (ss or ls matrix specified).

Neutron

N Neutron-derived porosity (ss or ls matrix specified).


SERVICE COMPANY NOMENCLATURE

Schlumberger Computalog

1. Electrical Log (ES) Electrical Log

2. Induction Electric Log Induction Electrical Log

3. Induction Spherically –

Focused Log –

4. Dual Induction Spherically Focused Dual Induction

Log Laterolog

5. Laterolog-3 Laterolog-3

6. Dual Laterolog Dual Laterolog

7. Microlog Micro-Electrical Log

8. Microlaterolog Microlaterolog

9. Proximity Log –

10. Microspherically Focused Log –

11. Borehole Compensated Sonic Log Borehole Compensated Sonic

12. Long Spaced Sonic Log Acoustilog

13. Cement Bond/Variable Density Log Sonic Cement Bond System

14. Gamma Ray Neutron Gamma Ray Neutron

15. Sidewall Neutron Sidewall Neutron

Porosity Log Porosity Log

16. Compensated Neutron Log Compensated Neutron Log

17. Thermal Neutron Decay Time Log –

18. Formation Density Log Compensated Density Log

19. Litho-Density Log –

20. High Resolution Dipmeter Four-Electrode Dipmeter

21. Formation Interval Tester –

22. Repeat Formation Tester Selective Formation Tester

23. Sidewall Sampler Sidewall Core Gun

24. Electromagnetic Propagation Log Dielectric Constant Log


25. Borehole Geometry Tool X-Y Caliper

26. Ultra Long Spacing Electric Log –

27. Natural Gamma Ray Spectrometry –

28. General Spectroscopy Tool –

29. Well Seismic Tool –

30. Fracture Identification Log Fracture Detection Log

Western Atlas Halliburton Logging Services

1. Electrolog Electric Log

2. Induction Electrolog Induction Electric Log

3. 4. Dual Induction Induction Log Dual Induction Log

5. Focused Log Guard Log

6. Dual Laterolog Dual Guard Log

7. Minilog Contact Log

8. Microlaterolog FoRxo Log

9. Proximity Log

11. Borehole Compensated Acoustic Velocity Log

10.- Acoustilog

12. Long Spacing BHC Acoustilog

13. Acoustic Cement Bond Log Microseismogram

14. Gamma Ray Neutron Gamma Ray Neutron

15. Sidewall Epithermal Neutron Log Sidewall Neutron Log

16. Compensated Neutron Log Dual Spaced Neutron

17. Neutron Lifetime Log Thermal Multigate Decay

18. Compensated Densilog Density Log

19. Z Densilog –

20. Diplog Diplog

21. Formation Tester Formation Test

22. Formation Multi Tester Multiset Tester


23. Corgun Sidewall Coring

24. Dielectric Log Dielectric Constant Log

25. Caliper Log Caliper

26. –

27. Spectralog Compensated

28. Carbon/Oxygen Log

29. Borehole Seismic Record

30. –
LOGGING TOOLS: QUICK REFERENCE

Openhole logs, logging tools, and what they measure

Generic name of log: Induction

Type of tool: Resistivity

When run: Primary log in fresh or oil- base mud where invasion is shallow

Purpose: Measures formation resistivity, Rt

Limitations: Behaves badly in salt muds and/or large boreholes and in formation with
high resistivities

Often combined with: Porosity tools

Operating principle: 20 kHz coil induces current in formation

Curves recorded: Deep induction, conductivity, shallow-focused electric log SP and/or GR

Generic name of log: Dual induction

Type of tool: Resistivity

When run: Primary log in fresh mud where invasion is moderate or deep

Purpose: Measures formation resistivity, Rt

Limitations: Behaves badly in salt muds and/or large boreholes

Often combined with: Porosity tools

Operating principle: 20 kHz coil induces current in formation

Curves recorded: Deep induction, medium induction, shallow-focused electric log SP


and/or GR

Generic name of log: Dual laterolog

Type of tool: Resistivity

When run: Openhole

Purpose: Measures formation resistivity

Limitations: Works best in salt muds. Cannot be used in oil-base muds

Often combined with: Porosity tools

Operating principle: Horizontal bean of current sent to formation

Curves recorded: Laterolog deep, laterolog shallow, microfocused electric log


Generic name of log: Microlog

Type of tool: Microresistivity

When run: Openhole

Purpose: Sand count, permeability indication, invaded-zone resistivity Rxo

Limitations: Water-base muds required

Often combined with: Other microresistivity devices

Operating principle: Measures two shallow investigation resistivities

Curves recorded: Micro (normal), micro (inverse or lateral)

Generic name of log:: Compensated density

Type of tool: Radioactivity/porosity

When run: Openhole, primary porosity device

Purpose: Measures porosity, indication of lithology

Limitations: Requires smooth borehole wall

Often combined with: Other porosity and/or resistivity tools

Operating principle: Gamma rays from source scatter in formation

Curves recorded:
; Bulk density, , correction, apparent porosity, (for selected
1ithology)

Generic name of log: Compensated neutron

Type of tool: Radioactivity/porosity

When run: Open or cased hole

Purpose: Measures porosity/lithology

Limitations: Requires liquid-filled hole

Often combined with: Other porosity and/or resistivity tools

Operating principle: Fast neutrons thermalized by hydrogen atoms in formation

Curves recorded:
(for selected lithology), GR, etc.

Generic name of log: Acoustic

Type of tool: Sonic/porosity


When run: Openhole

Purpose: Measures porosity, lithology

Limitations: Some hole size limitations depending on value of formation

Often combined with: Other porosity and/or resistivity tools

Operating principle: Measures travel time of compressional waves in formation. If wave trains
are recorded, the travel time of shear waves in the formation can also be
deduced.

Curves recorded:
, wave train

Generic name of log: Gamma ray

Type of tool: Radioactivity/lithology

When run: Open or cased hole

Purpose: Sand/shale discriminator

Limitations: None

Often combined with: Any and all open- and cased-hole tools

Operating principle: Scintillation detector measures natural formation gamma ray activity

Curves recorded: GR

Generic name of log: Gamma ray spectral log

Type of tool: Radioactivity/lithology

When run: Open or cased hole

Purpose: Measures K, U, and Th concentrations in formation

Limitations: Slow logging speed

Often combined with: Neutron/density

Operating principle: Gamma ray energy spectrum characterizes source of gamma rays

Curves recorded: Total counts, uranium,

potassium, and thorium

Generic name of log: Dipmeter

Type of tool: Resistivity correlation/ orientation

When run: Openhole

Purpose: Measures formation dip, detects fractures


Limitations: Does not perform well in oil-base muds

Often combined with: SP, GR

Operating principle: 4, 6, or 8 independent pad electrodes record correlation curves

Curves recorded: Hole deviation, azimuth of Pad 1, relative bearing, correlation, caliper

Generic name of log: Sidewall sampler

Type of tool: Percussion core cutter

When run: Openhole

Purpose: Retrieves samples of the formation

Limitations: Limited number of cores (typically 30 or 60) per trip in hole

Often combined with: SP for depth control

Operating principle: Explosive charge propels hollow cylinder into formation

Curves recorded: SP, only for depth control; recovery report

Generic name of log: F-overlay

Type of tool: Quick-look log analysis

When run: Openhole

Purpose: Generates Ro curve

Limitations: Works best in clean formations of constant lithology

Often generated with: Density log or neutron/ density when run with deep induction or deep
laterolog

Operating principle: If Ro < Rt then Sw < 100

Curves recorded: F or rescaled as an Ro curve

Generic name of log: Rwa

Type of tool: Quick-look log analysis

When run: Openhole

Purpose: Distinguishes water- and oil-bearing rocks, finds Rw

Limitations: Needs careful attention to detail in gas-bearing formations and/or shales,


depending on porosity tool used

Often generated with: Sonic, porosity/resistivity combination tools

Operating principle: Apparent porosity and Rt combined to give apparent water resistivity
Curves recorded: Rwa

Generic name of log: Rxo/Rt versus SP

Type of tool: Quick-look log analysis

When run: Openhole, fresh muds

Purpose: Hydrocarbon indicator

Limitations: Requires an SP log. Used primarily when no porosity tool is available

Often generated with: Induction log

Operating principle: Scaled RxoRxo/Rt ratio compared to SP

Curves recorded: Rxo/Rt, SP


GENERAL RECOMMENDED LOGGING PROGRAM

Condition-- Data Desired


Correlation and lithology in sand/shale

Fresh mud Recommended Services


Dual Induction--SFL--GR/SP; Induction Electrical; Dual Laterolog-- GR/SP
(low porosity and/or high resistivities); Sidewall Cores

Remarks
The dual induction and dual laterolog devices are superior to single
induction and single laterolog in all cases Sidewall cores give lithology in
sand/ shale

Data Desired
Porosity; water saturation; lithology in carbonates and evaporates;
hydrocarbon type

Recommended Services
Density and/or Neutron and /or Sonic and GR; Formation Tester

Remarks
For Shaly sands or simple mixed lithologies, density--neutron or neutron--
sonic. For complex mixed lithologies use all three. For hydrocarbon type as
above and or Formation Tester

Data Desired
Producible hydrocarbons and permeability indicators

Recommended Services
Proximity-Microlog; Microlaterolog or Micro-SFL; Dual Resistivity device;
Sidewall Cores; Formation Tester

Remarks
Proximity log used for thick mudcake; Microlaterolog for thin mudcake;
Micro-SFL available with Dual Laterolog. Sidewall cores give permeability
estimate in shaly sands

Data Desired
Hydrocarbon indication at the wellsite

Recommended Services
Wellsite Computer Products, including apparent formation water resistivity,
Rwa; formation factor/resistivity overlay; and Rxo/Rt vs. SP overlay

Remarks
Available with computer units only. Others available with both units.

Data Desired
Formation dip magnitude and directional

Recommended Services
Dipmeter
Remarks
Wellbore directional information is also available

Condition-- Data Desired


Correlation and lithology in sand/shale

Salt mud Recommended Services


Dual Laterolog--Micro-SFL--GR

Remarks
In some areas of low porosity and/or mixed lithology the neutron or density--
neutron is usable as a correlation log

Data Desired
Porosity, water saturation, and lithology in carbonates and evaporates;
hydrocarbon type

Recommended Services
Density and/or Neutron and/or Sonic and GR; Formation Tester

Remarks
See Remarks under fresh mud

Data Desired
Producible hydrocarbons and permeability indicators

Recommended Services
Proximity--Microlog; Microlaterolog-Microlog; Dual Laterolog--Micro-SFL;
Sidewall Core; Formation Tester

Remarks
In very salty muds or when Rxo is less than Rmc, the microlog results may
be unsatisfactory. See Remarks under fresh mud

Data Desired
Hydrocarbon indication at the wellsite

Recommended Services
Wellsite Computer Products; Rwa, Ro Overlay

Remarks
See Remarks under fresh mud. Apparent formation water resistivity may not
be satisfactory in low and/or mixed lithologies

Data Desired
Formation dip magnitude and directional

Recommended Services
Dipmeter

Remarks
Wellbore directional information is also available
Condition-- Oil-base Data Desired
mud Correlation and lithology in sand/shale

Recommended Services
Dual Induction--GR; Induction--GR; Sidewall Cores

Remarks
High temperature (350 to 400&deg;F) will restrict the use of the dual
induction and sidewall cores

Data Desired

Porosity, water saturation, lithology in carbonates and evaporites;


hydrocarbon type

Recommended Services
Density and/or Neutron and/or Sonic; Formation Tester

Remarks
See Remarks under fresh mud. High temperature (350&deg;F) and small
holes(6 in.) restrict the use of the formation tester

Data Desired
Producible hydrocarbons and permeability indicators

Recommended Services
Dual Induction; Sidewall Cores; Formation Tester

Remarks
Sidewall cores give permeability estimates in shaly sands

Data Desired
Hydrocarbon indication at the wellsite

Recommended Services
Wellsite Computer Products; Rwa, Ro Overlay

Remarks
See Remarks under fresh mud

Data Desired
Formation dip magnitude and directional

Recommended Services
Dipmeter

Remarks
See Remarks under fresh mud

Condition-- Air- or Data Desired


gas-filled hole Resistivity, correlation, and lithology in sand/shale

Recommended Services
Dual Induction--GR; Induction--GR

Remarks
None

Data Desired
Porosity, water saturation, gas saturation, estimated lithology in carbonated
and evaporates

Recommended Services
Density and/or Neutron

Remarks
Dual spacing neutron CNL cannot be used; must use gamma ray-
neutron(GRN) or preferable sidewall neutron porosity (SNP)

Data Desired
Permeability Indicator

Recommended Services
Temperature Log; Noise Log

Remarks
Gas entry

Data Desired
Hydrocarbon indication at the wellsite

Recommended Services
Wellsite Computer Products; Ro Overlay, Rwa

Remarks
See Remarks under fresh mud

Condition-- Fresh or Data Desired


unknown formation Resistivity; correlation and estimated lithology in sand/shale
water

Recommended Services
See fresh mud and salt mud

Remarks
See fresh mud and salt mud

Data Desired
Porosity water saturation; estimated lithology in carbonates and evaporates;
hydrocarbon type

Recommended Services
Density and/or Neutron and/or Sonic and GR; Electromagnetic Propagation
Tool; Inelastic Neutron Scattering and Capture Gamma Ray Spectroscopy;
Formation Tester

Remarks
Under conditions of no invasion the electromagnetic propagation tool will
yield water saturation directly. See Remarks under fresh mud

Data Desired
Producible hydrocarbons and permeability indicators
Recommended Services
Sidewall Cores; Formation Tester

Remarks
None

Data Desired
Hydrocarbon indication at the wellsite

Recommended Services
Wellsite Computer Products.

Remarks
See Remarks under fresh mud. May use electromagnetic propagation tool

Condition-- Cased Data Desired


hole Fluid type and lithology

Recommended Services
Pulsed Neutron Log; Inelastic Neutron Scattering and Capture Gamma Ray
Spectroscopy; GR and Natural Gamma Ray Spectroscopy

Remarks
None

Data Desired
Porosity and hydrocarbon type

Recommended Services
Pulsed Neutron Log; Gamma Ray-Neutron; Density Formation Tester;
Inelastic Neutron Scattering and Capture Gamma Ray Spectroscopy

Remarks
Dual spacing neutron (CNL) cannot be used if hole is gas filled. Under
favorable conditions the density may be used for porosity

Data Desired
Permeability

Recommended Services
Formation Tester

Remarks
None
REFERENCES
Allaud, L., and N. Martin. 1977. Schlumberger, the history of a technique. New York: Wiley and
Sons.

Bateman, R. M. 1984. Log quality control. Boston: IHRDC.

___________. 1985. Openhole log analysis and formation evaluation. Boston: IHRDC.

Bateman, R. N., and E. E. Konen. 1977. The log analyst and the programmable pocket calculator.
The Log Analyst (SPWLA) Sept. -Oct.

Burke, J. A., R. L. Campbell, and A. W. Schmidt. 1969. The litho-porosity crossplot. SPWLA
Symposium, May.

___________. 1969. The litho-porosity crossplot. The Log Analyst (SPWLA) Nov. -Dec.

Burke, J. A., M. R. Curtis, and J. T. Cox. 1966. Computer processing of log data enables better
production in Chaveroo field. SPE 1576, presented at the 41st Annual Meeting. October, Dallas.

Cox, J. W., and L. L. Raymer. 1976. The effect of potassium salt muds on gamma ray and
spontaneous potential measurements. SPWLA 17th Annual Logging Symposium.

Doll, H. G. 1949. The SP log: Theoretical analysis and principles of interpretation. Trans. AIME
179:146.

_________. 1949. The SP log in shaly sands. J. Pet. Tech. 2912.

__________. 1955. The invasion process in high permeability sands. Pet. Engr. (January).

Dresser Industries Inc. 1974. Log Review 1. Dresser Atlas.

_________. 1979. Gamma ray spectral data assists in complex formation evaluation. Dresser
Atlas (February).

_________. 1980. Formation multi-tester interpretation manual. Dresser Atlas (June) 9404.

_________. 1981. Spectralog. Dresser Atlas.

_________. 1984. Wireline service catalog. Dresser Atlas.

_________. 1984. Services catalog. Dresser Atlas.

Eck, M. E., and D. E. Powell. 1983. Application of electromagnetic propagation logging in the
Permian Basin of West Texas. SPE 12183, presented at the 58th Annual Technical Conference
and Exhibition. October, San Francisco.

Ellis, D., C. Flaum, C. Roulet, E. Marienbach, and B. Seeman. 1983. Litho-density tool calibration.
SPE paper 12048, presented at the Annual Technical Conference and Exhibition. October, San
Francisco.
Evers, J. F., and B. G. Iyer. 1975. A statistical study of the SP log in fresh water formations of
northern Wyoming. 16th Annual Logging Symposium of the SPWLA.

Fertl, W. H., and E. Frost, Jr. 1982. Experiences with natural gamma ray spectral logging in North
America. SPE paper 11145, presented at the 57th Annual Technical Conference and Exhibition.
September, New Orleans.

Fertl, W. H., W. L. Stapp, D. B. Vaello, and W. C. Vercellino. 1980. Spectral gamma ray logging in
the Texas Austin Chalk trend. J. Pet. Tech. (March).

Fitzgerald, D. D. 1980. Obtaining valid dipmeter surveys in deviated wells. World Oil (November).

Garner, J. S., and J. L. Dumanoir. 1980. Litho-density log interpretation. Paper N, Trans., SPWLA
21st Annual Logging Symposium. July, Lafayette, LA.

Gaymard, R., and A. Poupon. 1968. Response of neutron and formation density logs in
hydrocarbon-bearing formations. The Log Analyst (SPWLA) Sept. -Oct.

Gilreath, J. A., and R. W. Stephens. 1975. Interpretation of log response in a deltaic environment.
Paper presented at the AAPG Marine Geology Workshop, April, Dallas.

Gondouin, M., M. P. Tixier, and G. L. Simard. 1957. An experimental study on the influence of the
chemical composition of electrolytes on the SP curve. Trans. AIME 210:58.

Hassan M., A. Hossin, and A. Combaz. l976. Fundamentals of the differential gamma ray log-
interpretation technique. Paper presented at the SPWLA 17th Annual Logging Symposium. June,
Denver.

Hilchie, D. W. 1984. A new water resistivity versus temperature equation. The Log Analyst
(SPWLA) July-August: p. 20.

Kokish, F. P. 1951. Gamma ray logging. Oil and Gas Journal, July 26.

Marett, G., P. Chevalier, P. Souhaite, and J. Suau. 1976. Shaly sand evaluation using gamma ray
spectrometry applied to the North Sea Jurassic. SPWLA 17th Annual Symposium. June.

Neinast, G. S., and C. C. Knox. 1973. Normalization of well log data. SPWLA 14th Annual
Symposium Transactions, Paper I.

Nugent, W. H., G. R. Coates, and R. P Peebler. 1978. A new approach to carbonate analysis.
SPWLA 19th Annual Logging Symposium, June.

Overton, H. L., and L. B. Lipson. 1958. A correlation of the electrical properties of drilling fluids
with solids content. Trans. AIME, Vol. 213.

Poupon, A., R. W. Hoyle, and A. W. Schmidt. 1971. Log analysis in formations with complex
lithologies. J. Pet. Tech. (August).

Quirein, J. A., J. S. Gardner, and J. T. Watson. 1982. Combined natural gamma ray spectral/litho-
density measurements applied to complex lithologies. SPE paper 11143, presented at the 57th
Annual Technical Conference and Exhibition. September, New Orleans.
Raymer, L. L., and W. P. Biggs. 1963. Matrix characteristics defined by porosity computations.
Trans., SPWLA 4th Annual Logging Symposium.

Raymer, L. L., W. R. Hoyle, M. P. Tixier. 1962. Formation density log applications in liquid-filled
holes. J. Pet. Tech. (March).

Raymer, L. L., and E. R. Hunt. 1980. An improved sonic transit time-to-porosity transform.
SPWLA 21st Annual Logging Symposium. July, Lafayette, LA.

Savre, W. C. 1963. Determination of a more accurate porosity and mineral composition in


complex lithologies with the use of sonic, neutron, and density surveys. J. Pet. Tech. (September)
945-959.

Schmidt, A. W., A. G. Land, J D. Yunker, and E. C. Kilgore. 1971. Applications of the Coriband
technique to complex lithologies. SPWLA 12th Annual Logging Symposium (May), paper Z.

Segesman, F. 1962. New SP correction charts. Geophysics, 27(6): Part 1.

Segesman, F. and M. P. Tixier. 1958. Some effects of invasion on the SP curve. SPE Annual Fall
Meeting, October.

Sherman, H., and S. Locke. 1975. Effect of porosity on depth of investigation of neutron and
density sondes. Paper SPE 5510, presented at SPE Annual Fall Meeting, September-October,
Dallas.

Silva, P., and z. Bassiouni. 1981. A new approach to the determination of formation water
resistivity from the SP log. SPWLA 22nd Annual Logging Symposium.

Smith, H. D., Jr., C. A. Robbins, D. M. Arnold, and J. G. Deaton. 1983. A multi-function


compensated spectral natural gamma ray logging system. SPE paper 12050, presented at the
58th Annual Technical Conference and Exhibition. October, San Francisco.

Smolen, J., and L. Litsey. 1977. Formation evaluation using wireline formation tester pressure
data. SPE paper 6822, presented at the 18th Annual Technical Conference and Exhibition.
October, Denver.

Stewart, G., and M. J. Wittmann. 1981. The application of the repeat formation tester to the
analysis of naturally fractured reservoirs. SPE paper 10181, presented at the 22nd Annual
Technical Conference and Exhibition. October, San Antonio.

Tilly, H. 0., B. J. Gallagher, and T. D. Taylor. 1982. Methods for correcting porosity data in a
gypsum-bearing carbonate reservoir. J. Pet. Tech. (October) 2449-2454.

Tittman, J. 1956. Radiation logging lecture 1: Physical principle, and lecture 2: Applications.
Petroleum engineering conference on the fundamental theory and quantitative analysis of electric
and radioactivity logs. University of Kansas.

Tittman, J., and J. S. WahI. 1965. The physical foundations of formation density logging (gamma-
gamma). Geophysics (April).

Tixier, M. P., and R. P. A1ger. 1968. Log evaluation of nonmetallic mineral deposits. SPWLA 9th
Annual Logging Symposium, New Orleans.
Truman, R. B., R. P. Alger, J. G. Connell, and R. L. Smith. 1972. Progress report on interpretation
of the dual-spacing neutron log (CNL) in the U.S. SPWLA Trans., 13th Annual Logging
Symposium, Tulsa.

Wah1, J. S., J. Tittman, C. W. Johnstone. 1964. The dual spacing formation density log. J. Pet.
Tech. (December) 17.

Watson C. C. 1983. Numerical simulation of the litho-density tool lithology response. SPE paper
12051, presented at the Annual Technical Conference and Exhibition. October, San Francisco.

Welex, a Halliburton Company. (undated). Open-hole services. Catalog G 6003.

Williams, H., and H. F. Dunlap. 1984. Short term variations in drilling parameters, their
measurement and implications. The Log Analyst. (SPWIA) Sept.- Oct. :3-9.
III. WELL LOGGGING TOOLS AND TECHNIQUES

1. RESISTIVITY LOGS

1.1 Definitions of Resistivity and Dielectric Constant

The first logging device ever designed measured formation resistivity. It was a modification of a
method previously used to detect underground resistivity anomalies associated with either
geologic features or concentrations of metallic ores. Figure 1 illustrates this old surface surveying
method.

A voltage source sent a current through the ground between two widely spaced electrodes. The
voltage drop between two other more closely spaced electrodes was used as a measure of the
ground resistivity. By moving the whole electrode array across the countryside, it was possible to
"map" underground features, as shown in Figure 2 . By rotating the whole setup through a 90
angle and lowering it into a borehole, the electric log was born.

Since those early days, a series of improvements have resulted in five main families of resistivity
tools--electric logs, induction logs, laterologs, microresistivity devices, and dielectric logs.
Responses of Resistivity Logs

Although the original electric logging principles were sound, their practical embodiment left much
to be desired. Efforts to improve the measurement of formation resistivity have been busily
pursued for the last fifty years at least. As a result, three main branches of resistivity logging have
evolved: focused electric logs, induction logs, and microwave devices.
Focused electric logs were a logical step up from the early (unfocused) electric logs. By adding
focusing electrodes to the basic four-electrode array, the current could be "steered" in the right
direction. Its modern descendant, the laterolog, uses a multiplicity of focusing electrodes both to
direct current into the formation and to eliminate most of the detrimental borehole effects on
electric logs. Spherical focused logs also rely on focusing electrodes.
Induction logs broke the tradition of using current and voltage electrodes by introducing a system
of focused coils that induce the flow of currents in the formation away from the disturbing
influence of the borehole and the invaded zone.
More recently, microwave devices have been built to measure the dielectric constant of the
formation. Strictly speaking, they do not measure formation resistivity; however, they are usually
classified as resistivity devices since their end use is the same as for resistivity devices, i.e.,
determination of formation fluid saturation.
Surface Surveys
The first logging device ever designed was used to measure formation resistivity. It was a
modification of a method for detecting underground resistivity anomalies associated with either
geologic features or concentrations of metallic ores.
Borehole Measurements
By rotating the whole setup 90 and lowering it into a borehole, the electric log was born. The first
electric log ever run was in the pechelbronn field in France in 1927.
Since those early days, continued improvements have resulted in the development of five main
families of resistivity tools:

• electric logs
• induction logs
• laterologs
• microresistivity devices
• dielectric logs
In addition to these, a multitude of other sensors have been developed to measure and record
formation porosity, fluid content, and other physical properties of the formation.

Conventional Resistivity Measurements

Conventional Resistivity (ES)

During the first thirty years of well logging, the only resistivity logs available were the conventional
electrical surveys (sometimes called old E-logs). Thousands of them were run each year in holes
drilled all over the world. Currents were passed through the formation by means of current
electrodes, and voltages were measured between measure electrodes.
The measured voltages provided the resistivity determinations for each device, as follows: In
Figure 3 , a current I flows between electrode A and electrode N in a homogeneous, isotropic
medium. The corresponding equipotential surfaces surrounding the current emitting electrode
A would be spheres. The voltage on electrode N situated on one of these spheres is proportional
to the resistivity of the formation, and the measured voltage can be scaled in resistivity units.

Although the original electric logging principles were sound, their practical embodiments left much
to be desired. Efforts to improve the measurement of formation resistivity have been busily
pursued for over 60 years. As a result, three main branches of resistivity logging have evolved.
They are focused electric logs, induction logs, and microwave devices.

Focused Electric Logs The responses of conventional electrical logging systems can be greatly
affected by the borehole and adjacent formations. These influences are minimized by a family of
resistivity tools that use focusing currents to control the path taken by the measure current. These
currents are emitted from special electrodes on the sondes.

The focusing electrode tools include the laterolog and spherically focused devices (SFL). These
tools are much superior to the conventional electrical logs (ES) because they eliminate many of
the detrimental borehole effects. They are also better for resolution of thin beds. Focusing
electrode systems are available with deep, medium, and shallow depths of investigation.

Induction Logs The induction logging tool was originally developed to measure formation
resistivity in boreholes containing oil-base muds and in air drilled boreholes. Electrode devices
did not work in nonconductive muds.

Experience soon demonstrated that the induction log had many advantages over the
conventional ES log when used for logging wells drilled with water-base muds. Designed for deep
investigation, induction logs can be focused in order to minimize the influences of the borehole,
the surrounding formations, and the invaded zone.

Microwave Devices Recently, microwave devices (also called electromagnetic propagation


logging) have been designed to measure the dielectric constant and the conductivity of the
formation. Strictly speaking, they do not measure formation resistivity. However, they are
sometimes classified as resistivity devices since the end use of their measurement is the same as
for resistivity tools, i.e., determination of formation-fluid saturation.

Unfocused Electric Logs The original conventional electrical logs are still used occasionally for
special applications.

Spontaneous Potential Very early in the development of electric logging, the spontaneous
potential (SP) was discovered and put to good use.
1.2 Philosophy of Measuring Formation Resistivity

Whatever device is used to measure formation resistivity, there are common factors that conspire
to confound these efforts. Although modern resistivity-measuring devices represent a
considerable improvement over the original unfocused electric log (commonly called the old E-
log), there is still plenty of room for improvement. In addition to measuring the resistivity of the
undisturbed zone, Rt, the tool, by its design, is influenced by the resistivities of the mud in the
borehole, the adjacent beds, and the invaded zone ( Figure 1 ). Thus, we cannot assume that
the reading from a resistivity log represents R t. Depending on the device used, the particular
circumstances of the well, and the formations logged, the actual reading nay be greater or less
than Rt.

We discuss below how to recognize those cases where resistivity measurements depart radically
from Rt. For now, use this rule: a big contrast between the resistivity of the bed of interest and the
resistivity of either the mud column or the adjacent bed is a danger signal that calls for the use of
correction charts. In this context, "big" means a factor of 10 or more. Of particular note are
conditions where the bed of interest is thin (say, 15 ft or less) and/or invasion is deep ( d i greater
than 40 in).

To summarize, assume that a deep-resistivity device measures Rt unless

• Rt/Rm is greater than 10

• Rt/Rs is greater than 10

• hole size is greater than 12 in.

• the bed is thinner than 15 ft

• invasion is greater than 40 in.

If any of these adverse conditions exists, refer to the appropriate correction chart. As will become
apparent, induction logs and focused electric logs (laterologs) behave differently when faced with
these problems; in many cases, what may adversely affect an induction tool can be an advantage
to a laterolog, and vice versa.

Whatever device is used to measure the resistivity of the undisturbed zone (R t), three elements,
individually or collectively, make measurement more difficult. They are the borehole itself, the
adjacent beds, and mud filtrate invasion.
Although modern resistivity-measuring devices represent a considerable improvement over the
original unfocused electric log, there is still plenty of room for improvement. When using a
resistivity log, the analyst must remember that the device is not perfect, and the measurement
displayed is a composite of the four items in Figure 2 .

Note that in addition to Rt, the resistivity of the undisturbed zone (which is what we are trying to
measure), the tool, by its design, is influenced by the resistivity of the mud in the borehole, the
adjacent beds, and the invaded zone, if present.

It is unwise to assume that the reading from a resistivity log represents R t. Depending on the
device used, the particular circumstances of the well, and the formations logged, the actual
reading may be greater or less than Rt. In the sections that follow, we shall learn how to
recognize those cases where resistivity measurements depart radically from R t. As a rule of
thumb, a large contrast between the resistivity of the bed of interest and that of the mud column
or the adjacent bed is a danger signal that calls for the use of correction charts. In this context a
"large" contrast could be classified as a factor of 10 or more. Of particular note are conditions
where the bed of interest is thin (say, 15 ft or less) and/or invasion is deep (di is greater than 40
in.).

By way of summary, assume that a deep resistivity device measures Rt unless:

• Rt/Rm is greater than 10

• Rt/Rs is greater than 10

• hole size is greater than 12 in.

• the bed is thinner than 15 ft

• invasion is greater than 40 in.

If any of these adverse conditions exists, then the appropriate correction chart is called for. As will
become apparent later, induction logs and focused electric logs (laterologs) behave differently
when faced with these problems. What is poison to an induction tool is often an advantage to a
laterolog, and vice versa.
1.3 Conventional Electric Logs
Introduction

The basic electric logging system consists of two current electrodes A and B (the ground return)
and two voltage measuring electrodes M and N. These can be arranged in a variety of
configurations and spacings to suit particular requirements, such as bed resolution or deep
investigation. Some of these arrangements became industry standards, such as the normal and
the lateral electrode spacings.

Normal Devices

Figure 1 illustrates a normal device. Constant current is passed between electrodes A and B.
The measure voltage appears between electrodes N and N. The distance AN is called the
spacing. Thus, the 16-in. short-normal device has electrode A separated from electrode M by 16
in.

Lateral Devices

Figure 2 illustrates a lateral device. A constant current is sent between the A and B electrodes
and the measure voltage appears between the M and N electrodes.

Shortcomings of conventional Devices

All of these old electric devices, though used for many years, were plagued with inherent
shortcomings related to borehole and adjacent bed effects. Their idiosyncrasies are numerous.
Students requiring further details to complete a study of old electric logs are directed to the
References section at the back of the manual.

The only survivors from the early electric log are the microlog and a device known as the ULSEL,
the ultra-long spacing electric log, which is a normal device with AM spacings from 100 to 1000 ft.
It is used for remote sensing of one borehole from another (blowout control) or for remote sensing
of resistive anomalies such as salt domes.
1 d. Later logs
Introduction

In the 1920s, Conrad Schlumberger put forward the idea of a "guarded electrode" in an attempt to
improve on the electrical logs of the time that had undesirable borehole effects. His idea was not
put into practice until H. G. Doll designed a working guard electrode system. From this starting
point, laterologs evolved in a number of ways. The laterolog 7, which used small guard
electrodes, operated on the same principle of a constant survey current (io) being "forced" into
the information by bucking currents from the guard electrodes. By monitoring the voltage required
to maintain the fixed current io, the formation resistivity was measured. later the laterolog 3, which
used long guard electrodes, was placed in service. It was known as the conductivity laterolog,
and maintained a constant voltage on the measure electrode so that current variations monitored
the formation conductivity.

Today, the laterolog tool most commonly used is the simultaneous dual laterolog. It is neither a
conductivity nor a resistivity laterolog, but rather a hybrid using a constant product of current and
voltage (constant power). The design of this tool solved many problems associated with earlier
laterologs and it is now the standard basic resistivity log for salt mud environments.

When to Use a Laterolog

Laterologs should be used when the following conditions exist:

• There is seawater or brine mud in the hole.

• The Rmf/Rw ratio is less than 3.

• Hole size is less than 16 in.

Furthermore, the laterolog is superior to the induction log when R t exceeds

150 Wm2/m. It also gives a better estimate of Rt than the induction log when bed thickness is less
than 10 ft. Figure 1 provides specifics about when to run a laterolog. This Figure shows a plot
of the Rmf /Rw ratio versus porosity (). The laterolog is preferred for use when the crossplot of R mf
/Rw versus  falls on the left side of the chart.
1.5 The Dual Laterolog Tool
The dual laterolog tool makes two resistivity measurements: the laterolog deep (LLd) and the
laterolog shallow (LLs). A microspherically focused log (MSFL) may be run in conjunction with the
laterolog measurements. In addition to these resistivity measurements, auxiliary curves, such as
caliper, gamma ray, and spontaneous potential curves, may be recorded. The resistivity curves
are presented on a standard four-decade logarithmic scale ( Figure 1 ).

Figure 2 shows one version of the tool with its associated measure electrodes.
The mechanics of measuring both a deep and shallow laterolog from a single set of electrodes
are handled by circuitry inside the tool. Figure 3 shows the respective current paths for the lid and
LLs devices. The lid uses long-focusing electrodes and a distant return electrode, while the
shallow laterolog uses short focusing electrodes and a near return electrode.

Figure 4 shows the current paths for the MSFL, which has five rectangular electrodes mounted on
a pad carried on one of the caliper arms.

Under the normal conditions found when using a dual laterolog, the radial profile of resistivities is
as shown in Figure 5 ; i.e., Rt > Rxo > Rm. Between the invaded zone and the undisturbed
formation is a transition zone with a resistivity value between R t and Rxo.

If a horizontal slice were made through the tool and its surrounding formation and examined in
plan view, the image in Figure 6 would be seen. Here the current is flowing radially outward
from the tool and has to pass through the mud, the invaded zone, and the undisturbed formation
before arriving at the return electrode. The current, if held constant, thus develops a series of
voltage drops across each zone encountered. The relationship between these voltages may be
simplistically expressed as

Vtotal = Vmud + Vinvaded + Vundisturbed


Each voltage drop is proportional to the product of the current, the resistivity of the zone, and
some geometrical constant, depending on the size of the zone.
Dual Laterolog "Fingerprints"

The characteristic behavior of the DLL tool in zones with movable hydrocarbons makes quick-
look interpretation very simple. The golden rule is that the pattern in which R LLD > RLLS > RMSFL is a
good indication that hydrocarbons are present, and conversely, the pattern in which R MSFL > RLLS >
RLLD is a good indication that the zone is wet ( Figure 7 )

Any relative ordering of the curves other than the two cases above suggests little or no invasion
and indicates that the zone is impermeable ( Figure 8 ).

Log Quality Control

Deep and shallow laterolog curves should read the same in impermeable formations (shales and
evaporites). In porous and permeable zones, some separation between the two laterolog curves
is to be expected, depending on the invasion diameter and the ratio of R xo to Rt .
1.6 Later log Corrections
Borehole and Invasion Corrections

Borehole corrections to the raw data may be necessary. Charts are available from wireline
service companies to make such corrections.

The MSFL, a pad contact device, is sensitive to mudcake thickness (h mc) and mudcake resistivity
(Rmc).

In the range of normal interest, when laterolog readings lie in the range of 10 < (R LL/Rm) < 100, all
corrections are within ±10%. where hole diameters are large, however, the LLs correction can
become intolerably large.

Once raw log readings have been corrected for borehole effects, they may be corrected for
invasion effects, using what is commonly known as a "butterfly chart" ( Figure 1 ). This chart
plots the ratio of RLLD/RLLS against the ratio of RLLD/Rxo. There are three families of lines on the
chart. They are constant values of Rt/RLLD constant values of Rt/Rxo, and constant values of di.

In order to use the chart, it is first assumed that (R MSFL)cor is equal to Rxo. A point is then located on
the chart at the coordinates RLLD/RLLS and RLLD/RMSFL. This point uniquely defines the three
unknowns: Rt, Rxo, and di.

The lower left portion of the chart corresponds to the invasion pattern R MSFL > RLLS > RLLD , which
usually occurs in water-saturated zones where Rmf > Rw.
1.7 Anomalous Laterolog Behavior
The early laterologs were prone to various types of anomalous behavior, which are chronicled
here to give some insight into the few anomalies that can still occur, even with the dual laterolog.
The Delaware Effect

In the early 1950s in the Permian basin, logging engineers found that laterologs behaved
anomalously when approaching a thick resistive bed, such as the massive anhydrite and salt that
overlies the Delaware sand. The effect manifested itself by a gradual increase in apparent
resistivity, starting when the bridle entered the highly resistive bed. Apparent resistivities would
climb to as much as 10 times the value of Rt before the sonde itself entered the highly resistive
bed. The solution for the laterolog 7 was to place the B return electrode at the surface. For the
conductivity laterolog, the solution was not so easy, since these devices were using a 280 Hz
survey current generated in the cartridge. Having the return at the surface did not solve the
problem, since skin effect restricted the return current to a sheath around the borehole, thus
resulting in the effective return electrode as the lower part of the cable ( Figure 1 ).
Compensation for this effect with the Laterolog 3 involved a messy setup, with two sondes, one
on each side of a cartridge, and a B return electrode on the bottom for Delaware situations.
However, for all practical purposes, the laterolog 3 remains susceptible to the Delaware effect.
The Anti-Delaware Effect
In an attempt to improve on the situation and provide a dual spacing laterolog, a tool was
introduced with both deep and shallow devices. However, this device also behaved anomalously
beneath highly resistive beds. The deep laterolog showed a gradient of decreasing resistivity, the
exact opposite of the Delaware effect. With the B electrode at surface (effectively at zero
potential), the N electrode acted as the takeoff point of a potential divider formed by the borehole
below and above N; thus the approaching sonde, at some positive potential, would cause N to
raise its potential. The anti-Delaware effect would at worst cause a 50% reduction in the deep
laterolog and would only be noticeable within 35 ft of the resistive bed. In fact, the effect had been
present on the earlier B electrode at surface, Delaware-free laterologs, but it had not been
noticed since there was no shallow laterolog with which to compare the deep laterolog.

The dual laterologs in use today have incorporated features that assure virtual freedom from
Delaware and anti-Delaware effects. However, a new effect has been observed on the dual
laterolog, again associated with highly resistive beds.
The Groningen Effect

The Groningen effect, first observed in the course of logging gas wells in Holland, manifests itself
as the lid reading too high when the N electrode enters a highly resistive bed. From a distance of
AN below the bed boundary (about 102 ft), the LLd will rise over a short distance to an
anomalously high value, which it will then maintain until the bed is entered. Experiments have
indicated that the effect depends on the operating frequency, and is only trouble-some in low-
resistivity formations immediately below a massive salt or anhydrite bed. Modern laterolog
devices can detect and correct for the Groningen effect.

The Groningen effect appears (if at all) within 102 ft (31 m) of a resistive bed and will be of
interpretive importance only where Rt in the underlying bed is less than 10 m2/m. It can appear
even if casing is set to the bottom of the resistive bed.

Dual Laterolog "Normal" Anomalies

Dual laterologs experience environmental effects, even if resistive bed effects do not occur. A tool
has not yet been designed that is entirely free of the disturbing effects of the borehole and
adjacent beds, although progress has been made in reducing these effects. For interpretive work,
these environmental effects must be taken into account. The hole size and invasion effects have
already been covered in the previous discussions, and another set of corrections is worth noting.

Shoulder Bed Corrections--Squeeze and Antisqueeze

When the sonde is in front of a bed with a resistive shoulder on either side, current tends to
concentrate in the least resistive path; in other words, it is "squeezed" between the resistive
shoulders into the formation of interest. Charts are available to correct for this effect. The
correction factor to be applied to the borehole corrected log reading is a function of bed thickness
and the contrast between the apparent reading and the shoulder resistivity R a /Rs. where this
factor is less than one, a squeeze situation exists and the apparent log reading is too high. where
Ra /Rs is greater than one, the bed is surrounded by a conductive shoulder and the current tends
to fan out into the path of least resistance--the conductive shoulders. Since this is the reverse of
"squeezed," it is called "antisqueeze." The apparent log readings are too low in this situation.

The LLd is much more affected by squeeze and antisqueeze than is the LLs even in what might
be considered thick beds (50 ft or more). When making detailed interpretations, one should use
the Shoulder Bed Correction Charts for lid after borehole correction and before any other step.

Invasion corrections may then be made. A word of caution is in order. In general, an ideal
laterolog has a depth of investigation response that behaves logarithmically with respect to
invasion diameter, but it is also a function of the contrast between R xo and Rt .

Furthermore, the effect of a hole larger than 8 in. (20.32 cm) is to replace part of the R xo zone by
mud, thus changing the effective position of the origin on the invasion correction chart.
1.8 Induction Tools

Introduction

Logging systems used before the introduction of induction logging depended on the presence of
an electrically conductive fluid in the borehole to transmit electric current to the formation. In most
rotary drilled wells, the drilling fluid is a water-base mud that conducts electricity. However, some
wells are drilled with nonconductive fluids, such as oil-base muds, air, and gas. Under such
conditions, it is impossible to obtain a satisfactory electrical log using conventional electric logging
tools.

Induction logging does not depend upon physical contact between the walls of the wellbore and
the logging tool. The induction logging tool acts like a transformer: the transmitter coil is
energized with alternating current, which induces in the formation a secondary current that is
proportional to the electrical conductivity of the formation and to the cross-sectional area affected
by the energizing coil. The higher the conductivity of the formation, the lower the resistivity, and
the larger the formation current will be. This current in turn induces a signal into a receiver coil,
the intensity of which is proportional to the formation current and conductivity. The signal detected
by the receiver coil is amplified and recorded at the surface.

The direct measurement is therefore one of conductivity. Both the conductivity and reciprocated
conductivity (resistivity) curves are shown on the log. The deflections of these curves are
proportional to formation conductivity. Formations having resistivities of 10, 100, or 1000 ohm-m
would have conductivities of 100, 10, and 1 mmho/m, respectively.

Induction logging equipment provides a record of the formation conductivity over a wide range.
The accuracy is excellent for conductivity values higher than 20 mmho/m (resistivity values less
than 50 ohm-m) and is acceptable in lower conductivity ranges (down to 5 mmho/m). Beyond this
limit, the induction log continues to respond to formation conductivity variations, but with
diminished accuracy. There is a small uncertainty of about ±1 mmho/m on the zero of the present
equipment.

When to Use an Induction Log

Induction logs are recommended for use when:

• the hole to be logged is filled with fresh water or oil-base mud

• the hole to be logged was air drilled

• the Rmf/Rw ratio is greater than 3

• the Rt is less than 150 ½m2/m

The induction log is the only resistivity device that works in oil-base mud (where oil is the
continuous phase) or air-filled holes. The laterolog measurement is preferred when Rmf/Rw falls
to the left of the vertical dashed line and to the left of the solid line for the appropriate value of Rw
( Figure 1 ). The induction log is preferred above the appropriate Rw line. To the right of the
dashed line and below the appropriate Rw curve, either or both logs may be required for an
accurate interpretation.
Tool Types

Two commonly used induction tools are the single- and dual-induction devices. Each of these
tools can be combined with the other sensors, thereby allowing both porosity and resistivity logs
to be recorded simultaneously. Figure 2 shows a typical tool string.
Presentations and Scales
Induction logs and combination induction logs are recorded on a variety of scales and
presentations. The primary measurements of conductivity are always recorded on a linear scale
when presented. In contrast, resistivity can be plotted on a linear or logarithmic scale. when
porosity data are presented, a split grid is usually employed. Figure 3 , Figure 4 , and
Figure 5 illustrate the various possibilities.
1.9 Theory of Induction Devices

Today’s induction tools have many transmitter and receiver coils. The principle can be
understood clearly by considering a sonde with only one transmitter coil and one receiver coil (
Figure 1 ).

A high-frequency alternating current of constant intensity is sent through a transmitter coil.


The alternating magnetic field created induces currents into the formation surrounding the
borehole. These currents flow in circular ground loops coaxial with the transmitter coil and create,
in turn, a magnetic field that induces a voltage in the receiver coil.

Because the alternating current in the transmitter coil is of constant frequency and amplitude, the
ground loop currents are directly proportional to the formation conductivity. The voltage induced
in the receiver coil is proportional to the ground loop currents and, therefore, to the conductivity of
the formation.

The induction tool works best when the borehole fluid is an insulator, such as air or gas. The tool
also works well when the borehole contains conductive mud, unless the mud is too salty, the
formation too resistive (above 150 ohm-m), or the borehole diameter too large.

Calibration of the system is a two-step process. First, the tool is suspended high off the ground
away from any conductive materials. Since the tool is in a zero-conductivity environment, it is
adjusted to read zero conductivity (infinite resistivity). Second, a circular loop or ring of a known
conductivity (known resistivity of either 1 or 2 ohm-m) is placed on the tool. The tool response is
now adjusted to measure this "calibration" value.

Now, with the two end points defined and measured (the high end simulated by the tool in air and
the low end simulated by the test loop), the tool is capable of measuring most normally
encountered oilfield resistivity or conductivity values.

Of course, when the tool is at the bottom of a l0,000-ft well, there is no way a test loop can be
placed around the sonde, so an internal calibrator is included in the tool. The calibrator will have a
nominal value of 1 or 2 ohm-m; its precise value is determined monthly by reference to the test
loop. These internal calibrators shift with age but behave reasonably well under normal use. A
check of the zero conductivity point when the tool is in the hole is accomplished by simply
opening the receiving coil. Any extraneous signal is canceled out by a zero adjustment.
Skin Effect

Linkage of each ground loop with its own magnetic field (a ground loop has self-inductance), and
with the magnetic fields of the other nearby ground loops creates a cross-coupled system and the
resultant eddy currents will not be quite as predictable on the basis of the theory already
discussed. That is, it cannot be assumed that the individual ground loops are independent of one
another. It can be predicted that, with increasing distance from the source (i.e., transmitter coil),
there will be attenuation in the amount of transmitted power because

1. The dissipation of energy by the flow of eddy currents in the region near the source
decreases the energy available for transmission to regions farther out.

2. Regions far from the source are shielded from the magnetic field of the transmitter coil
by the annulling effect of magnetic fields of opposite sign from the eddy currents in the
conductive medium closer to the transmitter. In a sense, the "shielding" of the outer
regions is equivalent to a reflection of the energy back toward the source.

As a consequence of these interactions, there is a reduction in the receiver-coil signal; i.e., a


reduction in high conductivity. This reduction is commonly called a "skin effect."

Thus, if g is the conductivity reading observable in a given configuration of media without skin
effect and a is the conductivity actually observed, then the difference, s, is the "skin effect."

s = g - a

An amount s is added to the observed reading by means of a skin effect compensating network.
It is nonlinear and can best be illustrated by Figure 2 . In practical terms, the tool reads a
resistivity that is too high unless the skin effect compensation is working properly.
Environmental Effects

In addition to the transmitting and receiving coils of the simple two-coil device ( Figure 1 ), a
practical field tool also includes additional focusing coils ( Figure 3 ). These focusing coils make
the current ground loop flow as far away from the borehole as possible to eliminate borehole and
drilling-mud-filtrate invasion effects.

Bed Thickness Corrections

Unfortunately, a tradeoff has to be made when designing an induction tool. Good bed resolution
can only be obtained with closely spaced transmitter-receiver coil arrangements, but this close
spacing results in a relatively shallow radial depth of investigation. Conventional induction
devices, designed for deep investigation, have poor vertical bed resolution. Effectively, the signal
received is a mixture of signals from points both above and below the horizon being measured.

The surface control equipment offsets the poor bed resolution characteristic by emphasizing the
zone of interest and playing down the measurement made on either side of the horizon ( Figure 4

The electronic circuitry used in this tool can manipulate three measurements in such a way
that the reading recorded on the log is equal to a "weight" value A times the value of the interval
being measured plus B times the values at points 78 inches above and below the point being
measured. The values for A and B should be chosen so that A - 2B = 1. This is logical: in a
homogeneous formation where all three measurements are the same, the net effect is similar to
the gross effect. This scheme assists in correcting the log for the effects of adjacent beds.

The more modern phaser processing of the induction tool signals allows for enhanced bed
thickness response.
Induction Current Paths
Current loops flow around the borehole in a horizontal plane. The measured signal includes
signals from the mud, the filtrate invaded zone, and the undisturbed zone. The tool "sees" three
resistances in parallel ( Figure 5 ).

Hole Size Corrections


The borehole effects due to the current loop in the mud can be corrected by using a special chart.
The size of the correction is insignificant in fresh, resistive muds, but quite significant in salty,
conductive muds.

If an SFL (spherically focused log) is run in conjunction with an induction log, a hole size
correction is also needed. Figure 6 is provided for this purpose. The RSFL/Rm ratio is plotted
against the ratio of (RSFL) cor to RSFL. The lines on the chart are for different hole sizes.
Invasion Affects
The radial response of the induction tool is described by the "integrated radial geometric factor,"
or "G." The G factor reveals which fraction of the measured signal comes from which radial
distance from the tool. Mathematically, it can be described by the equation

If di (the diameter of invasion) is small, then G is small and all the signal comes from the
undisturbed zone; in this case, RID is equal to Rt. If di is large, then G also is large and a large
pert of the total signal comes from the filtrate invaded zone. In this case, R ID reads somewhere
between Rt and Rxo.
Figure 7 shows G as a function of di for the deep induction tool. This plot can be used to solve
the following case. Suppose di is 80 in., Rxo = 20, and Rt = 10. What will the induction tool read?
From Figure 7 , G for a di of 80 in. is 0.4. Therefore, the equation given above can be written as:

RID = 1/0.08 = 12.5 Thus, RID reads greater than Rt.

Figure 8 illustrates a typical invasion pattern with high filtrate saturation in the invaded zone and
low connate water saturation in the undisturbed zone.

We should realized that this treatment of the invasion problem is the reverse of what is
encountered in the field; i.e., in practice Rxo, Rt, and di are not known in advance. The objective is
to find Rxo, Rt, and di from the measured log values. In fact, if only the value of R ID is known, there
is no solution to the problem. If three unknowns exist, then three known quantities are needed to
solve the problem. The solution is to use the dual induction SFL combination logging tool. Since
the geometric factor for the medium induction log (G’) is different from the geometric factor for the
deep induction log (G) at the same di, the following three equations can be solved
simultaneously:

RSFL = f (Rxo)
Tool Calibration
Induction tool calibration can be performed on land at any time. The sonde is placed in a zero
conductivity environment. This is normally done by raising the sonde up in the air well away from
metallic objects. This defines a zero point. A calibration loop is then placed around the sonde to
give a known conductivity signal, usually 500 mmhos. This calibration is performed monthly. It is
almost impossible to perform on an offshore rig because of the surrounding metal structure. In
cases where it is not possible to set the zero point under controlled conditions at the surface, it is
permissible to set it with the tool in the hole opposite a thick, very highly resistive zone (salt,
anhydrite, dense low-porosity carbonate, etc.) if one exists. The sonde and its associated
electronic cartridge form a matched set and should always be used together.
The High-Resolution Induction
Because of the design limitations of conventional induction tools the resulting measurements of
formation resistivity are distorted by the adjacent beds, by the invaded zone, and to some extent
by the borehole. Recent (mid 1980s) advances in signal processing have led to improved output
from the standard dual induction hardware. These "new" Induction tools are referred to as the
high-resolution induction (HRI) and the phasor induction. These improvements stem from the use
of both the conventional (indirect) EMF’s induced in the receiver coil and the directly coupled, out
of phase, ones known as the "X signals."
The result of the additional data is a measurement of formation resistivity, which is less affected
by adjacent beds and allows far better precision in correcting for invasion effects. Modern
software routines allow real-time deconvolution in the logging truck and hence output of R t, Rxo
and di directly on the log.

Figure 9 shows the same formation logged with (a) a conventional dual induction and (b) an HRI
log. Note the improvement in bed resolution between (a) and (b).
1.9 MICRO RESISTIVITY TOOLS
Introduction
Microresistivity tools make measurements that have a variety of applications in well-to-well
correlation, and are used to determine the following:
o flushed zone saturation, Sxo Sxo
o residual oil saturation, (ROS)
o hydrocarbon movability
o hydrocarbon density, hy
o invasion diameter, di
o invasion corrections to deep resistivity devices
Microresistivity Tools
A variety of tools, old and new, are available. Each tool has its own special characteristics. The
following list covers the majority of the microresistivity devices that are now, or have been, used
extensively. These tools can be divided into two main groups: the mandrel tools and the pad
contact tools.
Mandrel Tools: 16-in. SN Short Normal
LL8 Laterolog 8
SFL Spherically Focused Log
Pad Contact Tools: MLL Microlaterolog
PL Proximity Log
MSFL Microspherically Focused Log
ML Microlog
The mandrel tools have electrodes placed on a cylindrical mandrel. Such tools do not require
physical contact with the formation. In contrast, the pad contact tools have their electrodes
embedded in an insulating pad carried on a caliper arm that is forced against the borehole wall.
The microlog is worthy of special mention as an underrated device that should be run more
frequently than it is. It was one of the first microresistivity devices on the market and has had a
spectacular career. Originally, it was used as a pseudo-porosity device. When that function was
improved with modern porosity devices, the microlog was relegated to the pile of has-beens by
many people in the logging industry. It is still a valuable tool because it offers a superb visual
identification of porous and permeable zones. Figure 1 shows a microlog and proximity log
presentation. The presence of permeability is indicated wherever the microinverse curve
reads higher than the microinverse curve, and the microinverse curve reads close to R mc.
The microlog records two resistivity curves with shallow depths of investigation. The microlog
looks for a resistivity contrast between the mudcake and the flushed zone. If no porosity or
permeability are present in the formation, there is no filtrate invasion, and thus no mudcake
buildup. Hence, there is no positive separation between the two resistivity curves.
Depth of Investigation
Each microresistivity tool has its characteristic depth of investigation. It is important to know these
values for each of the tools in order to select the one with the right characteristics for the job. A
tool with a shallow depth of investigation is needed if invasion is shallow and the tool is to read
Rxo without undue influence from Rt. Conversely, in situations where deep invasion exists, a deep
investigation tool will ensure a reading of Rxo free from any effects of Rmc.

As with other tools, no single value for the depth of investigation can be used. Rather, a
pseudogeometric factor must be used. This factor indicates how much of the total tool signal is
received from an annular formation volume represented by distances (expressed in inches) from
the borehole wall ( Figure 2 ).
Bed Resolution
Just as each of the microresistivity tools has its characteristic depth of investigation, so too does
each tool have its own characteristic bed resolution; i.e., some tools are better than others at
distinguishing thin beds. Tools with coarse bed resolution values are "blind" to thin shale and/or
sandstone layers. For example, 3-in. shale streaks will not be "seen" by a short normal log but
may easily be delineated by a microlog. By way of contrast, a shallow-focused log, depending on
its electrode spacing, may be able to resolve beds 1 or 2 ft thick at best.
Environmental Corrections
Microresistivity devices of the mandrel type are subject to aberrations caused by the size of the
wellbore. These effects can be quite severe. The pad contact tools, however, are only affected by
excessive mudcake buildup, hole rugosity, and fractures.
The mudcake corrections can be made by using appropriate charts, available from the wireline
service companies that relate a correction factor to the mudcake thickness and resistivity.
The mudcake thickness is calculated as half the difference between the bit size and the
measured caliper reading when the caliper reads less than bit size.
The mandrel-type tool corrections can be made by use of service company charts that relate the
log reading, the mud resistivity (Rm), and the hole size.
Sxo and Hydrocarbon Movability
The water (filtrate) saturation in the flushed zone (S xo) may be estimated by using Archie’s
equation: (Sxo) n = F Rmf/Rxo where F = a/ m
To solve this equation, the values of a, m, n, Rmf, and Rxo must be known. Rmf should be
corrected for formation temperature.

The value of Sxo may not reveal much about the amount of oil in place, but it will reveal a great
deal about whether the oil that is in place is likely to flow or not. The invasion process acts like a
miniature waterflood. Invading filtrate displaces not only connate water, but also any movable
hydrocarbons. In the undisturbed state at initial reservoir conditions, the fractional pore volume
occupied by oil is (1 - Sw). After filtrate invasion has taken place, the fractional pore volume
occupied by oil is (l - Sxo). The difference between these two values is the fractional pore volume
that contained movable oil. Figure 3 shows this process.
The pore volume fraction of movable oil is determined by the relationship (S xo - Sw). The fraction
of the original oil in place that has moved is determined by:
(Sxo - Sw) / (1 - Sw)
This index can then be used as a measure of the quality of the pay. In formations where the
relative permeability to oil is low, Sxo is likely to be close to Sw and the index will be low. This
same formation will not be as productive as another with the same value of S w but better relative
permeability to oil and hence a higher value of Sxo
Hydrocarbon Density
The computation of the hydrocarbon density in a pay zone can be critical when there is doubt
about the type of hydrocarbon present; i.e., does the formation contain oil, light oil, condensate,
or gas? Since the porosity tools make their measurements in the flushed zone, they "see" a bulk
volume of hydrocarbon equal to (1-Sxo). This leads to the interesting paradox that where
hydrocarbons are movable they will have been flushed away from the zone where they can be
seen. Thus large hydrocarbon effects on porosity tools may be misleading and really only indicate
large volumes of residual hydrocarbons. Lack of pronounced hydrocarbon effects could mean
either that movable hydrocarbons are present or the formation is wet. Either way, a good value of
Sxo is essential for correct prediction of hydrocarbon density and hence the type of hydrocarbon
present in the formation.
Quality Control
Quality control for these devices can be summarized by the following maxims:
o Beware of washed-out holes because (a) pad contact tools lose contact with the
formation and "float" in the mud column and (b) mandrel tools give severely
inaccurate readings.
o Beware of thick mudcakes because pad contact tools require large corrections.
o If hole conditions are bad, forget about trying to measure Rxo, because either the tool
will stick or the pad will tear up. Either way, no usable log reading will be obtained.
It should also be noted that pad contact resistivity devices do not measure accurately in oil-base mud.
THIS SECTION IS INCOMPLETE. OTHER LOG
TOOLS AND TECHNIQUES HAVE TO FILLED IN
IV. WELL LOG INTERPRETATION
1. QUICK LOOK ANALYSIS
1.1 INTRODUCTION
Quick-look log analysis refers to a number of techniques for plotting log data in a reasonably
effortless and simple way that reveals either the formation content or the formation lithology.
These methods are widely used by log analysts for wellsite evaluations. Their great appeal lies in
their simplicity and subtlety. Most quick-look methods can be applied without any special
equipment and produce quite acceptable results. Broadly speaking, there are three branches to
quick-look analysis:

• compatible overlays of curves


• crossplots of selected curve readings
• simple algorithms for calculators
In general, compatible overlays manage to eliminate some unknown quantity by taking a ratio,
while revealing another quantity that is of interest; for example, water saturation.
Crossplots are indispensable for computer-generated analysis, but also offer a quick and
convenient means of determining endpoints such as Rw.
Algorithms offer a quick and simple means of calculating items of interest, such as porosity and
water saturation. They are widely incorporated into log analysis routines on both programmable
calculators and computers.
1.2 OVERLAYS
Theory
Compatibly scaled overlays entail comparison of two or more log curves. In practice, one curve is
overlaid on the other and a tracing made (either by hand or via a computer graphics display) so
that the end result is a composite with two curves. The relative deflection between the two curves
usually indicates a formation property of interest. The basis for all overlays is that the curves
being compared be compatibly scaled, i.e., both must be in the same system of units. Apples
must be compared with apples and not oranges. Many of the quick-look overlays in use today are
quite subtle in that they eliminate some quantity that is either unknown or of no interest. The most
commonly used are:

• for hydrocarbon detection: SP with Rxo/Rt Ro with Rt dielectric log-derived


porosity with porosity

• for lithology, porosity, and hydrocarbon typing: neutron porosity with density
porosity density porosity with sonic porosity
SP with Rxo/Rt
This overlay is by no means universally applicable, since it requires conditions that result in
healthy SP developments. However, where those conditions are met (i.e., in wells drilled with
fresh mud having salty connate water) it is an elegant way to detect hydrocarbons without the
need to know porosity. A natural candidate for wells where only an induction-SP log is available, it
is usually produced at the time the log is run and requires some manipulation of the raw data in
the logging service company’s surface equipment. Inputs required are an SP, a deep resistivity
measurement (usually a deep induction), and a shallow resistivity measurement (usually a
shallow-focused electric log). The theory behind the method depends only on Archie’s equation
and the SP relationship, which we may write as:
• SP = -K log (Rmf/Rw)
Note that for the purposes of this overlay Rmf and Rw are used rather than their "equivalent"
values Rmfe and Rwe.
Archie’s equation, if written for both the invaded and undisturbed zones, allows a ratio to be taken
which eliminates F, the porosity-dependent formation factor:
•Swn=FRw/Rt
Sxon = F Rmf/Rxo

If an assumption is made that Sxo is related to Sw--, for example, that


• Sxo = (Sw)1/5
--then the quantity (Sw/Sxo)n can be replaced by (Sw)8/5 if "n" is assumed equal to 2.
Thus, the term Rmf/Rw can be replaced by (Rxo/Rt) (Sw) 8/5 and the SP equation rewritten as:
• -SP = K log [(Rxo/Rt) (Sw)5/8] or:
• -SP = K (log (Rxo/Rt)) + K log (Sw5/8)
In a water-bearing zone with Sw=1, the term K log (Sw) 5/8 is equal to 0, and thus the term K log
(Rxo/Rt) is numerically equal to -SP. However, in an oil-bearing zone with Sw less than 1, the
term K log ((Sw)5/8) will be less than 1, and hence the term K log (Rxo/Rt) will be numerically less
than -SP. Provided that no substantial SP reduction is due to the presence of hydrocarbons
(usually the case in all but very shallowly invaded formations), a comparison of the actual SP with
the quantity K log (Rxo/Rt) will have the following characteristics:
• In wet zones, the two curves will "track."
• In hydrocarbon-bearing zones, the K log (Rxo/Rt) curve will separate from the SP curve (
Figure 1 ). (Note that in the lower sand, which is wet, the SP and the massaged
Rxo/Rt ratio closely coincide, and in the upper sand which is hydrocarbon bearing, the
two curves separate.)
• In shales, since the Rxo/Rt ratio is close to 1, the massaged Rxo/Rt is effectively zero.

In practice some experimentation is usually required in order to obtain a valid overlay with the two
traces overlaying in both shales and wet zones. This requires using the correct K value for the
formation temperature in question and correct offsetting of either the SP baseline or the
massaged Rxo/Rt curve.

To sum up, use this method, by all means, if you have sand-shale sequences, good SP
development (high Rmf/Rw ratio), and you do not envisage recording a porosity log. This method
cannot be used with oil-base mud. An unusual variation of this method uses the SP to generate a
pseudo Rxo/Rt ratio, which may then be compared with the actual Rxo/Rt ratio. A refinement of
the method allows a more accurate computation of the real Rxo/ Rt ratio by taking into account
invasion effects. In deeply invaded formations, for example, the ratio of RSFL to RILd is not as
large as the real Rxo/Rt ratio. This refinement requires a solution to the "tornado" chart to be
carried in the surface equipment software package.
Ro versus Rt -- The F Overlay
Another popular and extremely effective overlay is the F overlay, which effectively compares Ro
with Rt.
In clean formations of known lithology this is a simple matter for the logging service company’s
surface equipment. It merely requires values of ma, f, a, and m to be dialed in, and F appears as
an output. In practice, this F curve is recorded on a logarithmic scale as a separate log. The
analyst then uses the logarithmic F curve in conjunction with a logarithmic R t curve (a deep
induction or deep laterolog) in the manner of a slide rule to normalize the two curves so that they
overlay each other in clean wet zones. By so doing, the log (F) curve has in fact been shifted by
an amount equal to log (Rw) . Since the product of F and Rw is Ro, the net effect is an overlay
that compares Ro to Rt. Wherever the two curves separate, with Rt greater than Ro, the
implication is that Sw is less than 1 and therefore hydrocarbons are expected.
Figure 2 shows an F overlay. Note that in the water-bearing section the normalized F curve
(i.e., Ro) coincides with the Rt curve and that in the hydrocarbon-bearing section the Rt curve
separates from the Ro curve. This separation can be quantified in Sw units by using an
appropriate logarithmic scaler such as that illustrated (not to scale) in Figure 3 . The scaler is
placed across the overlay so that its long axis parallels the depth lines on the log. The S w - 100
mark is placed on the normalized F curve, and the Sw value is read at the Rt curve.
Another useful feature of this overlay is that Rw need not be known. The act of normalizing the F
curve to the deep resistivity trace in a wet zone effectively computes R w for you. If the F curve is
on top of the resistivity log when the normalization is made, locate the 100 line on the F scale at
the top of the F log and read the resistivity value on the resistivity scale that lies immediately
beneath it. This is numerically equal to 100 times R w. For example, after normalizing in a wet
zone, F = 100 lies over the Rt scale line corresponding to 20 Ohm-m. This means that Rw is 0.2
Ohm-m.
The Logarithmic Movable Oil Plot
An extension of the F overlay is the logarithmic movable oil plot, which requires that two overlays
be made, one to indicate Sw and a second to indicate Sxo.
In practice, both are made on a single log and the result is a movable oil plot (MOP) on a
logarithmic scale. The production of a MOP proceeds in two stages. First, the F curve is
normalized to Rt in a wet zone and the resulting Ro curve traced onto the resistivity log (or, if
preferred, the Rt curve may be traced onto the normalized F log). Second, the F curve is
normalized to an Rxo trace (such as a focused, or microelectric log) in a wet zone, and the
resulting Rxoo curve (Rxoo = Rxo in a rock 100% saturated with a fluid of resistivity R mf) is
traced onto the resistivity log (or, as noted, the Rxo curve may be traced onto the normalized F
log) . The result of such tracings will be as shown in Figure 4 . Note that the area between the
Ro and Rt curves is representative of the hydrocarbon-filled pore space. The MOP adds a further
dimension by subdividing that space into two parts, one containing residual oil and the other
movable oil. This presentation can help assess the most productive intervals (i.e., those with high
movable oil).
Consider the bulk volume model shown in Figure 5 . Note that in the undisturbed zone the
saturating fluids are connate water and oil and that the water saturation is S w. In the invaded
zone, however, the saturating fluids are mud filtrate and residual oil and the water saturation is
Sxo. If Sxo is high (i.e., residual oil saturation, ROS, is low) then a large fraction of the oil in
place has been moved by the mini-waterflood effected by the drilling process. If, however, Sxo is
close to Sw, then very little of the oil in place has been moved, and in such cases a formation
may be less productive. Thus, on the MOP, the area between the R o curve and the Rxoo curve
is a measure of the residual oil saturation and the area between the R xoo curve and the RT
curve is a measure of the movable oil saturation.

In practice, once a tracing of this sort has been prepared, it is common to color in the appropriate
areas by use of high-lighters of different colors, e.g., red for residual oil, orange for movable oil,
and blue for water. Note that zero porosity point never appears on such a logarithmic plot, since F
or Rt corresponding to zero porosity is infinity. Thus some arbitrary line is chosen as the right-
hand limit for the purpose of coloring in water with a blue highlighter.
The Neutron-Density
Overlay The neutron-density overlay is probably the most commonly used presentation. Its
popularity stems from (a) the fact that the two logs are nearly always run together and (b) the
enormous wealth of information that can be gathered from such a presentation, even with a
cursory inspection.
Two distinct types of overlay are considered here: the compatible sandstone-scaled overlay and
the compatible limestone-scaled overlay. In sand-shale sequences the former is needed and in
carbonate-evaporite sequences the latter.
The compatibly scaled sandstone presentation requires that the neutron log (and here the dual-
spacing thermal neutron log is the norm) be recorded on a sandstone matrix setting and
displayed on a 0 to 60% scale. Likewise the bulk density recording, b should be converted to an
apparent sandstone porosity curve by choice of an appropriate value for ma (2.65 or 2.68 gm/cc,
for example) and displayed on the same scale as the neutron log. The choice of a scale, be it 0 to
60% or 0 to 50%, is immaterial provided both curves are on the same scale and the shale
readings do not exceed the high end of the scale (i.e., a presentation on a 0 to 30% scale is
inappropriate if N in the shales is 45% and thus off-scale) .
Figure 6 shows such an overlay. Note that in shales, N is much larger than D (dotted curve
to the left of the solid curve) . Note that in water-bearing intervals, N equals D where the
formation is clean. Note that in gas-bearing zones, .N reads a lower apparent porosity than does
D and that the traces. tend to be mirror images of each other. Even in oil-bearing sections it is
common to note N reading slightly less than D.
Thus, on one plot it is possible to quickly distinguish porous and permeable sections from shales
and, within the potential reservoir rocks so defined, to distinguish gas, oil, and water.
Since the purpose is slightly different in carbonate reservoirs, a slightly different presentation is
used. Here the norm is not sandstone or shale but limestone and possibly evaporites. Porosities
in general are lower than in sandstone reservoirs and dolomite may be present as well. The
compatible limestone presentation, therefore, calls for the neutron log on a limestone matrix
setting, with the density log on an apparent limestone porosity basis as well. Since dolomite and
anhydrite may compute apparent porosities less than zero on a limestone basis (remember ma
for dolomite is higher than ma for limestone), an appropriate scale for this type of overlay is 45%
to -15% (left to right). In some locations 30% to -10% is favored. However, the 45% to -15% scale
has an added advantage. If the bulk density curve has not been converted to an apparent
porosity curve then all Is not lost. By appropriate shifting, the neutron porosity and bulk density
curves may be overlain so that on the neutron log 2.95 gm/cc coincides with -15% and 1.95
gm/cc coincides with 45%. The reader may satisfy himself that 1.95 gm/cc translates to
approximately 45% porosity on a limestone scale, and 2.95 gm/cc to -15%.

Figure 7 illustrates a neutron density compatible limestone overlay. It is constructed for a


formation porosity of 15%. Note that in limestone the two traces coincide. In dolomite the
apparent neutron porosity is higher than the apparent density porosity and in sandstone the
reverse is true. This sandstone "crossover" is due entirely to the matrix effects on the two porosity
devices and is the result of the limestone porosity scaling used. It should not be confused with
crossover produced by gas. Note that in sandstone intervals, both curves have a similar character
although separated from one another. The gas effect "mirror imaging" is not evident. If gas were
present it would manifest itself by an additional separation, with neutron even lower than density.
The evaporites salt and anhydrite are also shown along with shale. Such a presentation is
extremely valuable for a quick-look guide to the rock types in the column drilled and is a very
good starting point for other more detailed analysis.
Density-Sonic

Overlay Although not as widely used as a neutron-density overlay, the density-sonic overlay is
particularly useful for detection of secondary porosity, i.e., vugs and fractures. Figure 8 shows a
formation with both matrix (intergranular) porosity and fracture porosity. Provided the matrix is
known, the density porosity will be equal to the total porosity:

• = = matrix + secondary
The sonic tool, however, responding as it does to the first compressional wave arrival, will only
respond to the matrix porosity. Compressional waves traveling through a vertical, fluid-filled
fracture, for example, will travel more slowly than those traveling through the matrix system.
Since the sonic tool triggers on the first (faster traveling) arrival, the later arrivals passing through
the fracture system will be ignored. Thus:

• S= matrix

An overlay of S and D, has the interesting property that in fractured or vuggy intervals the two
curves separate. Figure 9 illustrates such an overlay. A number of scaling options are
available. Compatible limestone scaling on a 45% to -15% scale (left to right) is adequate for
most applications. If bulk density and A are to be overlaid, then compatible scaling is 1.95 to 2.95
gm/cc (left to right) for b and 108 to 28 µsec/ft (left to right) for D, equivalent to 40 µsec/ft per
track. In metric units a scaling of 350 to 90 µsec/meter is adequate (equivalent to 65 µsec/meter
per track) .
Dielectric Porosity Overlay

The compatible porosity overlay ( Figure 10 ) of a dielectric log-derived porosity with another
porosity curve is useful for quick-look hydrocarbon detection. The theory behind the overlay is
simple. A dielectric log-derived "porosity" is effectively the water-filled pore space, whereas a
density porosity, for example, is effectively the total porosity. If the formation is wet, traces will
overlay. In hydrocarbon-bearing zones separation may be expected. The usual cautions are
appropriate: clean formations and the correct choice of matrix parameters are required when
computing the two porosities.

1.3 CROSSPLOTS

Theory
Crossplots of log data are indispensable tools for the analyst. A crossplot is a two-dimensional
representation of the variation of data with respect to two or more properties. Although widely
used in conjunction with computer processing of log data, crossplots may, of course, be prepared
by hand using only limited amounts of data that have been screened by the analyst. The end
result of using a crossplot is finding a parameter needed for quantitative log analysis, for
example, Sw and Rw. Figure 1 , a graphical representation of the Archie equation, shows a
crossplot of porosity versus resistivity. Within a given formation, an Ro line is determined that
is representative of 100% water saturation. Then, an arbitrary R o line can be determined to
examine points that have water saturations less than a given value. Points below this line could
be subjected to further analysis. Crossplots come in many other forms; here follows an overall
view of crossplotting techniques.
Trend Analysis and Groupings

In logging terms, crossplots can be used to discern trends or groups. Suppose a well is
extensively cored and core analyses are made to find porosity and permeability throughout the
section. A plot of permeability against porosity might look like the plot shown in Figure 2 . Note
that three distinct rock types are revealed by the crossplot.
Rock type A is of low porosity and permeability (perhaps a carbonate), rock type B is of moderate
porosity and high permeability (perhaps a sandstone), and rock type C is of high porosity and
moderate permeability. Thus, crossplots visually show us trends and groupings of data points
that, in turn, help us understand the nature of the population being plotted.
Extrapolations
Another useful feature of the crossplot is in making extrapolations. The extrapolation of patterns
and trends on crossplots is an extremely valuable tool for predicting such key analysis
parameters as porosity. In the example shown in Figure 3 , D-ls = 15 and N-ls = 21. This
defines point P, lying between the limestone and dolomite curves and falling near a line
connecting the 18% porosity graduations on the two curves. Assuming a matrix of limestone and
dolomite and proportioning the distance between the two curves, the point corresponds to a
volumetric proportion of about 40% dolomite and 60% limestone; porosity is 18%.
Frequency Plots

Computers are particularly adept at manipulating log data to produce frequency plots. These are
crossplots on which the number of occurrences of a particular pair of data points is printed at a
particular map location. Figure 4 shows a neutron-density frequency crossplot for data from a
section logged through a sand-shale sequence. For the purpose of the plot, each axis is
divided into one-porosity-unit cells. There are thus 2500 (50 x 50) cells on the plot. Each data
point on the log is inspected and its N and D coordinates used to place a point on the plot.
Where many points fall in the same cell their total number is accumulated, and this number (or
frequency) is the item printed on the final plot at that particular cell address.
Where no data are found the plot remains blank. Where the majority of the points lie gives a good
visual image of the distribution of the formation properties in the logged section.
Such plots are used extensively for log data normalization (i.e., recalibration shifts), shale picks,
etc.
Z Plots
The Z plot is a companion plot to the conventional frequency crossplot (or X-Y plot) described
above. The purpose of the Z plot is to map the distribution of three variables instead of two. Given
the limitations of a two-dimensional sheet of paper, this task is accomplished by printing the
scaled average of the third variable (Z) at any given X-Y cell address.

For example, in Figure 4 there were 8 occurrences of the condition N (PHIN-the X axis) = 26%
and D (PHID-the Y axis) = 12%. The same cell on the Z plot of Figure 5 shows a scaled average
of a third parameter, the gamma ray. At the chosen address the scaled average of the 8
values of gamma ray occurring at N = 26% and D = 12% is seen to be 8. On this particular plot
the scaling of the Z axis was accomplished by applying the formula


and z-min and z-max were chosen so that the least radioactive levels "scored" 0 and the most
radioactive levels 10. Note that two-digit numbers (10 or more) are represented by symbols such
as "*."
Reviewing the Z plot of Figure 5 , it is evident that the shaliest points lie to the southeastern
corner of the plot (high z values of 8, 9, and *), and the least radioactive points lie to the
northwestern part of the plot (low z values of 1, 2, and 3). Z plots are useful for data analysis
where three variables are of interest.
1.4 Histograms

Histograms of raw log data and computed results of log analysis also provide useful tools for the
analyst. Figure 1 illustrates a histogram. The Y axis of such a plot is either scaled in actual
number of occurrences or in percent of the total number of data points analyzed. The X axis is
scaled over the useful range of the data analyzed. Among the better known uses for histograms
are picking minima and maxima, re-scaling logs, and checking the validity of computed results.
Illustrations of particularly useful histograms will be given in the appropriate modules of this
logging series when discussing particular aspects of log analysis.
1.5 Quick-Look Algorithms
With the advent of the hand-held calculator, especially the programmable type, log analysts have
tended to condense their wisdom into algebraic manipulations to ease their daily toil. Today much
of what used to be done by reference to charts is done via pocket computer. This has advantages
and disadvantages. On the plus side, obviously more data points can be analyzed in a shorter
time and with greater precision. On the minus side, however, the analyst no longer has the
latitude to shift points on a chart to get an immediate feel for the sensitivity of a computed answer
to a change in the raw data values used.
Algorithms for solving many of the common manipulations that log analysts use will be given in
the appropriate modules as the particular analysis method is discussed.
A word of caution is in order. Although the calculator/ computer is a useful device for the
crunching of numbers and relieves the analyst of that chore, it also places a greater burden on
the analyst, that of ensuring (a) that the model fits the circumstances and (b) that legitimate data
is submitted to be processed. A more precise answer should not be confused with a more
accurate one.
2. FLUID SATURATION
2.1 ARCHIE’S EQUATION
Saturation Determination
This section reviews in detail different methods of determining the water saturation of a rock
system. Once determined, the water saturation can be used to determine whether hydrocarbon
production is probable and to calculate the volume of oil in a unit volume of rock (e.g., how many
barrels/acre).

Water saturation is defined as the fraction of the pore volume occupied by water. This is shown
schematically in Figure 1 . The actual bulk volume of water in a unit volume of porous rock is
given by the product  Sw and the bulk volume of hydrocarbons is given by  (−Sw). This latter
value is also referred to as the hydrocarbon pore volume, or HCPV.
Two available methods for determining the water saturation of a rock sample are core analysis
and log analysis. Although core analysis is not covered in detail here, it is worthwhile to mention
that, in determining water saturation, only native state cores can be relied upon for representative
Sw values. As regards the role of measurements of permeability from cores, many low-resistive
pay sections with very high apparent water saturations do in fact produce water-free
hydrocarbons. Since these are generally formations of low permeability, core analysis can help
evaluate such troublesome formations.
Archie's Equation
Formation Factor, F
Archie combined three measurable observations into one equation. By saturating a rock sample
with salt solutions of different salinities, he found that the resistivity of the "wet" (water-saturated)
rock (Ro) was related to the resistivity of the saturating water (R w) by the relation
Ro = FRw
where F is called the formation factor. This formation factor was found to vary predictably as the
rock porosity changed according to

where a is a constant and m is known as the cementation exponent.


Lastly, rocks with less than 100% water saturation and resistivity R t were found to obey the rule
where n is the saturation exponent.
Combining these relationships gives Archie's equation:

The values of a, m, and n can be set to generally accepted values when they are unknown. For
serious petrophysical analysis, they should be determined by core analysis.
Generally accepted values for these constants and formation-factor-to-porosity relationships are
for sandstones: a = 0.81 m = 2 n = 2 or a = 0.62 m = 2.15 n = 2
for carbonates: a = 1 m = 2 n = 2
Thus, three commonly used versions of Archie's equation are, for sandstones:

or
for carbonates:

The analyst should avoid the mind set that locks the conventional values for a, m, and n into any
and all log analysis evaluations. Some crossplot methods allow one or another of these
parameters to be deduced; however, there is no substitute for rigorous core analysis to pin down
the precise values required for each and every reservoir unit. While n is usually assumed to be 2,
its exact value may vary depending on rock wettability. Oil-wet systems usually exhibit an n
greater than 2. Some water-wet systems show an n less than 2. To find n, a core sample is
prepared at a number of different water saturations and for each saturation the ratio of R o to Rt is
measured. By taking the logarithm of Archie's equation, it is transposed into

Therefore, a plot of log (Sw) versus log (Ro/Rt) should produce a straight line of slope n. Figure 2
illustrates the method.
It is worthwhile to explore the sensitivity of Sw to the value of n. Suppose a calculation of Sw is
made using n = 2 and the resulting value of Sw is 0.5, or 50%. Then we can deduce

hence
Having found that F Rw/Rt is 0.25, we can now play a "what if" game and see the results of raising
n to, say, 2.2:

or 53%
Lowering n to, say, 1.7 gives
Sw = (0.25)1/1.7 = 0.44 or 44%
In summary, other factors remaining the same, raising n raises S w; likewise, lowering n lowers Sw.
Each rock type has its characteristic formation-factor-to-porosity relationship. Core
measurements to determine this relationship require a range of core samples of different
porosities. Then at each porosity, a measurement is made of Ro, the resistivity of the rock at
100% water saturation. If the value of Rw is known, the formation factor (F) can be deduced by
using the definition F = Ro/Rw. Note that the solution used to saturate the core should be of the
same NaCl concentration as that of the connate water found in the formation. By taking the
logarithm of both sides of the equation that relates formation factor to porosity, the result is

Thus, a plot of log (F) versus log () should give a line of slope -m and an intercept at  = 1 of a.
Figure 3 illustrates this method.
2.2 INVADED ZONE SATURATION (Sxo)
Rmf and Invasion Diameter
The reliability of any estimate of Sxo depends on familiarity with Rmf and on the invasion diameter.
If the mud system is such that invasion is deep (generally greater than d i = 60 in.) then the
annular volume of formation measured by the Rxo devices is sufficiently large for a good
measurement of Rxo to be made. On the other hand, if invasion is shallow, it is possible that the
measurement of Rxo will be inaccurate, since part of the tool's measurement is in the uninvaded
zone.
Rmf changes when the mud composition in a drilling well is changed. As a consequence, mud
filtrate salinity (hence Rmf) may vary with depth.
Rxo Determination
Where only single Rxo measuring devices are used, there is no way of confirming the validity of
the measurement. A better estimate of Rxo can be made when multiple resistivity devices are
used. Since the use of the dual induction tool in conjunction with the tornado chart outputs the
ratio of Rxo to Rt, it is thus possible to back-calculate a good representative value for Rxo within a
fairly wide range of invasion diameters.
Sxo and the Movable oil Plot
The water (filtrate) saturation in the flushed zone (Sxo) may be estimated by using Archie's
equation:

where
To solve this equation, the values of a, m, n, , Rmf (at formation temperature), and Rxo must be
known.

The value of Sxo may not reveal much about the amount of oil in place, but it does reveal a great
deal about whether the oil in place is likely to flow or not. The invasion process acts like a mini-
waterflood. Invading filtrate displaces not only connate water, but also any movable
hydrocarbons. In the undisturbed state at initial reservoir conditions, the fractional pore volume
occupied by oil is  (1 - Sw). After filtrate invasion has taken place, the fractional pore volume
occupied by oil is  (1 - Sxo). The difference between these two values is the fractional pore
volume that contained movable oil. Figure 1 gives a pictorial representation of this process.
The pore volume fraction of movable oil is determined by the relationship  (Sxo - Sw). The fraction
of the original oil in place that has moved is determined by
This index can then be used as a measure of the quality of the pay. In formations where the
relative permeability to oil is low, Sxo is likely to be close to Sw and the index will be low. This
same formation will not be as productive as another with the same value of S w, but will have
better relative permeability to oil and hence a higher value of Sxo.
2.3 EFFECTS OF DIFFERENT MUD SYSTEMS ON SATURATION MEASURMENTS
Water-base Muds
In general, water-base muds produce a moderate invasion diameter and thus a measurable
invaded zone. In cases where the water loss is excessively high or the mud column highly
overbalanced, invasion diameters can be quite large. In such instances the measurement of R xo
is easy but that of Rt is more difficult, and the corresponding value for S w is less reliable.
Oil-base Muds
Oil-base mud systems generally produce an oil filtrate. In water-bearing formations, the relative
permeability to oil is low and invasion is slight. In oil-bearing formations, reservoir oil may be
replaced by invading oil filtrate, but only a small portion of the connate water is likely to be
displaced. Consequently, Sxo is likely to be equal to or lower than Sw when oil-base muds are
used. For the same reasons, Rxo may be equal to or greater than Rt.
2.4 EFFECTS OF CLAYS ON CALCULATED SATURATIONS
Surface Conductance
The log analyst faces formidable problems when confronted by shaly sands. Porosity,
hydrocarbon content, permeability, and productivity must be defined from logs (and any other
available data) for situations where neither an adequate model exists nor the fundamental
electrical behavior of the rock system is understood. Complications arise whenever clay material
is present with a quartz matrix and formation waters are fresh, in which case conventional log
analysis overestimates water saturation; or whenever formation waters are relatively salty and
abnormally low formation resistivities are observed in pay zones. Both situations lead to
bypassed production, unless adequate steps are taken to identify these difficult zones.

The problem is that the classical Archie relationship fails to hold up under conditions encountered
in shaly sands. Here the postulate that Co (wet rock conductivity) is linearly related to C w (water
conductivity) is found to be false. Figure 1 shows that as formation water conductivity increases,
Co increases through a nonlinear zone until it reaches a plateau at some conductivity offset (X)
above the clean sand line.
The form of the Co versus Cw plot is a function of the shaliness of the sand. Figure 2 shows that
the apparent formation factor, Fa (= Cw/Co) decreases as shale content increases. Note that
at high values of Cw (salty formation waters) Fa approaches the classical Archie formation factor.
An explanation is therefore sought for the source of the "excess" conductivity, X.

In general the approach first models the behavior of the 100% wet shaly sand system and then
transposes the model to include nonconductive hydrocarbons as well. A large body of published
material is available on the subject; the reader may check the SPWLA reprint volume Shaly Sand
(SPWLA 1982).
Clay Distribution and Type
In discussions of shaly sand, the distinction is usually made between the terms clay and shale. In
this discussion, the term clay or dry clay is used to refer to dehydrated shale and shale is used to
refer to rehydrated dry-clay materials.
Shale effects depend on
• distribution of clay material
• type of clay material
• amount of clay material formation water salinity
• water saturation
Figure 3 illustrates three different states in which clay materials may be found.
1. Shale can exist in the form of laminae separated by layers of sand. Laminar shale
does not affect the porosity or permeability of the sand streaks themselves. As laminar
shale increases, however, and the porous medium decreases, the overall average
effective porosity is reduced in proportion.
2. Shale can exist as grains or nodules in the formation matrix. This matrix or structural
shale is usually considered to have properties similar to those of laminar shale and
nearby massive shales.
3. Shale material can be dispersed throughout the sand, partially filling the intergranular
interstices. This dispersed shale may adhere to or coat the sand grains, or partially fill
smaller pore channels, thus markedly reducing the permeability of the formation.

All these forms of shale can, of course, occur simultaneously in the same formation.
Clay Types There are two ways of defining shales: defining grain size and describing mineral
content. Standard definitions for grain-size diameter are:
• sand 0.05 to 2 mm (50 to 2000 micron)
• silt 0.004 to 0.05 mm (4 to 50 micron)
• clay less than 0.004 mm (less than 4 micron)
Mineralogical analysis defines the common clay minerals as montmorillonite (smectite), illite,
kaolinite, and chlorite. These are various molecular arrangements of alumino-silicates with
various quantities of quartz and feldspar.
These clays are further subclassified by their origin. Detrital clays are deposited with the
sandstone at the time the sediments are laid down. Authigenic clays appear as precipitates from
solution at a later time. A schematic of the molecular building blocks and their various
arrangements to form clay crystals is shown in Figure 4 .
The most important effect of these minerals is their ability to hold adsorbed water on their grain
surfaces. Table 1 lists the specific areas of porous formations; Figure 5 generalizes the
relationship between grain size and grain surface area.
The Archie water saturation equation (4.1), relating rock resistivity to water saturation, assumes
that formation water is the only electrically conductive material in the formation. The presence,
then, of another conductive material (i.e., shale) requires either that the Archie equation be
modified to accommodate the existence of another conductive material or that a new model be
developed to relate rock resistivity to water saturation in shaly formations. The presence of clay
also complicates the definition of rock porosity. Although the layer of closely bound surface water
on the clay particle can represent a very significant amount of porosity, this porosity is not
available as a potential reservoir for hydrocarbons. Thus, a shale or shaly formation may exhibit a
high total porosity yet a low effective porosity as a potential hydrocarbon reservoir.

Microporosity
As grain diameter decreases, so too does pore throat radius. This decrease, accompanied by an
increase in capillary pressure, increases the amount of water that can be imbibed into the system.
When silt- and clay-sized particles are present, "microporosity" accounts for a large percentage of
the total porosity. Under these circumstances irreducible water saturation can be very high.

To give a visual understanding of where and how the micro pore system develops, Figure 6
shows SEM photos of the major clay types.

Cation Exchange Capacity


Crystal surfaces have what are known as "exchange sites," where ions can temporally reside as
the result of the charge imbalance on the external surface of the clay's molecular building blocks.
These exchange sites effectively offer an electrical path through the clay by means of surface
conductance. A dry clay is an insulator; a wet clay is not. To quantify the conductivity of wet clays,
the cation exchange capacity (CEC) can be measured. It will come as no surprise to find that in
general the larger the specific surface area, the larger the CEC. The clay type per se is not
important, only its specific surface area. Figure 7 shows the relationship between CEC and
specific surface area. Chlorite, one of the principal clay minerals, may not follow the trend
shown in Figure 7 quite as well as the other clay minerals. Even quartz sand grains, if sufficiently
small, exhibit surface conductance and can be ascribed a CEC value.

Unfortunately, log measurements do not allow a direct measurement of either CEC or specific
surface area. Table 2 lists the properties of, and log responses to, the common clay minerals.

Notice in Table 2 that there is no general correlation between the CEC or specific area numbers
and any one clay indicator. Perhaps the best hope for determining clay type from logs is to note
that the least electrically active clays have a large neutron/density difference and the most active
a small one.

Summary of Shaly Sand Models


Presence of shale in reservoir rock is an extremely troublesome factor in formation evaluation.
Most sands contain some clay or shale, which reduces the effective porosity, lowers the
permeability, and alters the resistivity predicted by the Archie equation.

Over the years, a large number of models relating resistivity and fluid saturations have been
proposed. Many have been developed assuming shale exists in a specific geometric form (i.e.,
laminar, dispersed, structural) in shaly sand. All these models are composed of a clean sand
term, described by the Archie water saturation, plus a shale term. The shale term, whether fairly
simple or quite complex, may be relatively independent of, or interact with, the clean sand term.
All models reduce to the Archie water saturation equation when the fraction of shale is zero; for
relatively small amounts of shaliness, most models and methods yield similar results.

The key to successful log analysis in these difficult formations lies in the experimental sciences.
The observations are to hand. What is lacking is an adequate theory to explain the observations
and adequate logging tools to make the necessary measurements. Figure 8 shows the key
observation, that the Cw/Co versus Cw plot is dependent on both Cw and "shaliness,"
expressed via CEC or specific surface area.
3. FRACTURES

Fractured Formations
Analysis of fractured formations may be subdivided into

• detection of fractures
• calculation of the relative magnitude of matrix porosity and fracture porosity through
quantitative analysis
• determination of the orientation of fractures in the formation, including the dip and strike
of the fracture plane
• evaluation of water saturation (especially when the saturation in the matrix porosity may
differ from that of the fractures)
Detection of Fracture
Fracture detection is based on a number of distinct principles, which include

• comparison of neighboring dipmeter microresistivity readings


• comparison of density and sonic porosity behavior of microcaliper and/or density
correction curve relative readings of K, U, and Th on gamma ray spectra log
• sonic wave train analysis to allow comparison of compressional and shear wave travel
times
• comparison of apparent resistivity values from multielectrode devices
Seldom does any one method suffice to detect all fractures. It is common to use several fracture
indicators and go with the "majority vote."
Quantitative Analysis

The density-sonic overlay is particularly useful for detecting secondary porosity, i.e., vugs and
fractures. Figure 1 shows a formation with both matrix (intergranular) porosity and fracture
porosity. Provided the matrix is known, the density porosity is equal to the total porosity:

The sonic tool, however, responds only to the first compressional wave arrival, i.e., the matrix
porosity. Compressional waves traveling through a vertical, fluid-filled fracture, for example, travel
more slowly than those traveling through the matrix system. Since the sonic tool triggers on the
first (faster traveling) arrival, later arrivals passing through the fracture system are ignored. Thus:
An overlay of  Sand  D ( Figure 2 ) therefore has the interesting property of curve separation in
fracture or vuggy intervals.

A number of scaling options are available. Compatible limestone scaling on a 45% to -15% (left to
right) scale is adequate for most applications. If bulk density and are to be overlaid, then
compatible scaling is 1.95 to 2.95 gm/cc (left to right) for  b and 108 to 28 µsec/ft (left to right) for
t, equivalent to 40 µsec/ft per track. In metric units a scaling of 366 to 86 µsec/m is adequate
(equivalent to 70 µsec/m per track).
4. POROSITY

4.1 Density Porosity


Since the density of a mixture of components is a linear function of the densities of its individual
constituents, it is a simple matter to calculate the porosity of a porous rock. Fore this purpose it is
useful to consider the bulk volume model of a clean formation with water-filled pore space (Fig 1).

Unit volume of porous rock consists of a fraction  made up of water and a fraction (1 - ) made
up of solid rock matrix.
Thus the bulk density of the sample can be written as:
b = ma (1 - ) + f
where ma refers to the matrix density and f refers to the fluid density. Simple rearrangement of
the terms leads to an expression for porosity given by:
D = (ma - b) / (ma - f)
The same concept is illustrated graphically in Figure 2 , where b is plotted against porosity.
Note that points falling on the line connecting the matrix point (ma,  = 0%) and the water point
(f,  = 100%) represent all possible cases, extending from zero porosity rock matrix up to 100%
porosity. Any intermediate value of b corresponds to some porosity .

Note that, strictly speaking, porosity is a decimal fraction and that the assumption is made that b
is equal to a, the tool reading. Density porosity, then, can be found either by using a calculator or
the chart ( Figure 3 ), which generalizes the concept from Figure 2 and builds in response
lines fore a number of different matrix materials and pore-filling fluids.
This derivation of porosity assumes a clean matrix of known density and water-filled pore space.
These computed values of D are wrong when (1) the lithology is mixed or unknown, (2) shale is
present, (3) gas or light hydrocarbons are present in the flushed zone, or (4) pad contact with the
formation is lost in washed-out holes.
Since the tool makes its reading in the zone adjacent to the borehole wall where mud filtrate has
flushed away most of the original formation fluids, the choice fore a value of f is dictated by the
mud filtrate density, which in turn is a function of its salinity, pressure, and temperature.
Figure 4 is useful for estimating f in salt-water-filled formations.

Shale Effects
If, in addition to matrix and fluid, a third component, shale, is introduced, the same principle can
be applied as before. Figure 5 illustrates a bulk volume model of a shaly formation. The
volume fraction of shale is referred to as Vsh, hence that of matrix is (1 -  - Vsh).
As above, an equation may be written for the bulk density of such a mixture
b = (1 - Vsh -  ) ma + Vsh (sh) + f
Hence,

or  = D - Vsh •Dsh
where Dsh is the apparent density porosity of the shale.
Note that, since most shale densities are lower than common reservoir rock matrix densities, D in
a shaly formation is always greater than the true effective porosity. It is common practice among
log analysts to correct density readings fore shale effects by the simple expedient of estimating
shale content (Vsh) from a gamma ray or SP log and reading the density tool response in a shale
bed ( Dsh). Although the procedure is not strictly valid, since clay materials disseminated in a
sand may differ from those deposited in a pure shale, it is still widely used as a "quick-look," and
usually gives satisfactory results.
In this context, it is worth remembering that shales are in fact porous formations, albeit with
effectively zero permeability and hence zero effective porosity. If the dry clay minerals in a shale
have a matrix density of 3 gm/cc (a good average), then a hydrated (wet) shale with a bulk
density of 2.5 gm/cc has a total porosity of 25%. The same shale logged on an apparent
sandstone porosity scale would appear to have a porosity of about 9%.
Gas Effects

Figure 6 shows a graphical situation where a clean (non shaly) formation under investigation has
a certain amount of its pore space filled with a gas. A density tool response based on the total
ore bulk volume model for such a case would be
b = ma(1 - ) + f Sw + g  (1 - Sw)
In this situation the three unknowns of the equation,  (porosity), Sw (water saturation), and g
(gas density) make analysis difficult without either additional formation measurements ore good
estimates of the unknown values.
In another form, the total or bulk volume equation can be rewritten fore the case of the invaded
portion of the formation, as:
b = ma(1 - ) + mf Sxo + g (1 - Sxo)
and rearranging the equation to solve for , we obtain:
 = (ma - b)/[ ma - g - Sxo(mf - g)]
Again, we are confronted with the case where , g , and Sxo are not known, but with the aid of
some simple Archie equations, gas density approximations, and some mathematical
substitutions, a fair value of  (porosity) can be obtained.
We start with
Sxo2 = F Rmf/Rxo
Utilizing the simple case of F = 1/2, we can substitute and obtain:
Sxo = (Rmf / Rxo)1/2 1/
 Sxo = [Rmf / Rxo]1/2
and substituting [Rmf / Rxo]1/2 for Sxo we obtain:
 = (ma - b) + (Rmf/Rxo)1/2 (mf - g)/( ma - g)
Now, except for g , all other measurements can be obtained from either a density log ore a
resistivity log, ore are already known.
An approximation for g can be obtained by using:
g = 0.18/([7644/depth] + 0.22)
where depth is in feet and g is in gm/cc. Once g is obtained, a correction fore the (2Z/A) effect
of density tool calibration results in
g (apparent) = 1.325 g - 0.188
Now, along with values of ma , g from the log, Rmf, Rxo from a log, mf , and g (apparent), a value
of  (porosity) can be obtained.
Depth of Investigation

The depth of investigation of the density tool is quite shallow. Figure 7 illustrates the depth of
investigation of the density tool as compared with the MSFL and compensated-neutron tools fore
average conditions. It should be pointed out that the exact shape of the curve depicted in
Figure 7 is also a function of formation density and the salinity of the formation fluids. Most of the
density tool signal comes from a region less than 8 in. from the borehole wall. The CNL tool, by
contrast, gathers most of its signal from the region within 12 in. of the borehole wall. If the
invasion process has left Sxo higher in the zone sensed by the density tool than in the zone
sensed by the neutron tool, then the density tool is less affected by light hydrocarbons than the
neutron tool. Where deep invasion has occurred, there may be very little hydrocarbon effect on
either tool.
Density Log Quality Control

Practical calibration of the density tool is accomplished by a series of standards. The primary
standard ( Figure 8 ) is made by using laboratory formations. Since these cannot be easily
transported, a set of secondary standards is available at logging service company bases in the
form of aluminum, magnesium, and/ore sulfur blocks of accurately known density and geometry.
These blocks, which weigh up to 400 pounds, are not easily transportable either. So a field
calibrator containing two small gamma ray sources is used to reproduce the same count rates as
those found in the blocks.
The wellsite calibration should be performed before and after each log is run; the shop calibration
should be run at least every 60 days, and a copy of it attached to the main log. It is important to
note that the field calibrator, the skid with the detectors, and the source form a matched set. If any
of the three does not match the serial numbers on the master calibration, the log should be
rejected.
Natural benchmarks fore checking the validity of a density log are salt, which has a a of 2.032
gm/cc, and anhydrite, which has a a of 2.977 gm/cc. These minerals may not appear in the
wellbore being logged; even if they do, they may not be 100% pure, and should be used with
caution. In general, density logs are either well calibrated (and therefore correct) or very
noticeably bed.
Apart from the natural benchmarks already discussed, the next best quality check is a review of
the  curve. If the short-spacing detector fails, then the whole compensation mechanism is
thrown out of kilter. So if  is roughly within the limits of ± 0.05 gm/cc, the log may be assumed
to be correct. If in light muds the  is negative, however, something is wrong. Likewise, positive
values for  in heavy (barite) muds is a danger signal.

4.2 Porosity from Neutron, Density, and Sonic Logs


Porosity tools include compensated neutron, compensated formation density (with or without
photoelectric factor measurement), and acoustic logs. These tools are sensitive to both rock
matrix and to the fluid filling the pore space; consequently, their measurements reflect not only
porosity, but also the type of rock, the clay content, and the fluid type. As yet, no tool has been
invented that reads only porosity.

Clean (shale/clay-free) water-bearing formations of known lithology represent the simplest


environments for porosity determination, because the effects of mixed lithology, clay, and
hydrocarbons do not confuse the issue. Before discussing these analysis techniques, however, a
review of the scales used on porosity logs and the procedures and techniques used to read them
will be presented.

Three factors must be determined before reading the log:


1.The type of curve recorded -- e.g., bulk density, density porosity, neutron porosity,
acoustic travel time -- and (in the case of porosity curves) the lithology on which it is
based.
2.The log scale and the number of units per log division.
3.Local knowledge (whenever available), such as lithology and formation fluid content.
Once this information is obtained, true porosity can be determined in those intervals where the
particular porosity device under consideration can reasonably be expected to work; e.g., with a
pad contact device such as the density tool, readings should not be attempted in rugose
boreholes.

Reading the Density Log Normal scaling for the density log is from 2.00 to 3.00 gm/cc from left to
right across two film tracks, or 20 divisions. Each division thus represents 0.05 gm/cc.

On this basis, it is a simple matter to find the bulk density, b in gm/cc. The range of bulk
densities encountered in sandstone, for example, is from 1.00 gm/cc (water) in 100% porosity to
2.65 gm/cc at zero porosity. Thus, a spread of 1.65 gm/cc represents a spread of 100 P.U.s, and
each of the twenty log-scaling divisions represents 0.05 gm/cc. The number of P.U.s in a scale
division can be determined as follows:

1.65 gm/cc = 100 P.U.s


0.05 gm/cc = (0.05/1.65) x 100 P.U.s
= 3.03 P.U.s
Thus, for all practical purposes, the log scale can be marked off in increments of 3 P.U.s, starting
with 0 P.U. at 2.65 gm/ cc. The error at 30 P.U.s is only 0.3 P.U., or 1%.

Combinations of Porosity Tools Often a combination of porosity devices is available that can be
run into the wellbore together. Usually these combination tools can define porosity better than any
single porosity device by itself. The most commonly available pair is the density/neutron
combination.

When working with well logs generated by such a tool combination, the log scales should be
checked carefully before conclusions are reached. Probably more confusion exists about
density/neutron combinations and their presentations than about any other log on the market
today.

Density/Neutron Combinations
The density log can be recorded either as a bulk density curve (b) in gm/cc or as a porosity
curve D if certain assumptions about ma and b are made. The neutron can be recorded in one
of three matrix settings. The important items to note on each porosity log are (a) the zero porosity
point on each log and (b) the porosity scale for each log.
If these parameters have been chosen correctly, comparing the two logs is like comparing apples
with apples. Occasionally, however, density/neutron combinations are like comparing apples with
pears, because porosity scales are different or the zero points do not coincide.
A common rule is to record both logs on compatible limestone scales; i.e., both logs read the true
porosity if the lithology is limestone. To accomplish this, the neutron is run using a limestone
matrix setting, and the density is recorded as a bulk density curve and shifted so that the zero of
the neutron scale coincides with a b value of 2.70 gm/cc. Alternatively, and a little more
accurately, b is converted into a porosity curve by using ma = 2.71 and using an appropriate
value of b, depending on the salinity of the mud filtrate.
For predominantly sandstone reservoirs, two methods can be used:
1. Run the neutron on a sandstone matrix using a porosity scale of 60 to 0%. Run the
density as a b curve on a scale 1.65 to 2.65 gm/cc. The zero for both is now at the
extreme right of Track 3.
2. Alternatively, run b as a D curve and scale it as for the neutron on a 60 to 0% porosity
scale.
For predominantly carbonate reservoirs, or where lithology is not well known, the best method is
to use compatible limestone scales; i.e., run the neutron on a limestone matrix using a 45% to -
15% porosity scale, and run the density using a b scale of 1.95 to 2.95 gm/cc. In compacted
formations, a scale of 30% to -10% is frequently used.
In analyzing combination density/neutron logs, some unusual combinations of scales may be
encountered. If this happens, reason back to the zero point for each log and begin the analysis
there.
Uses of Combination Porosity Logs Three parameters can be determined by combining two
porosity logs. They are porosity, lithology, and fluid content of the pore space. Regardless of the
actual lithology, a rough approximation of porosity can be made by combining the neutron
limestone porosity with the density limestone porosity in the following manner:

where x is referred to as the crossplot (Xplot) porosity. This approximation is somewhat


inaccurate in low-porosity dolomite, and its use in such cases is not recommended.
If both logs are recorded on limestone porosity scales and the lithology is limestone, both logs will
read the same porosity value.
If the lithology is sandstone, then DLM will be too high and NLM will be too low, but the combined
average will be close to the actual porosity value. If gas or light hydrocarbons are present, use of
the average porosity formula results in porosity values that are only good to a first order
approximation. Better values of porosity, when light hydrocarbons are present, can be determined
by the Gaymard method.

4.3 POROSITY FROM RESISTIVITY


In clean formations of known lithology, porosity may be determined from all three common
porosity tools. When lithology is unknown, the safest quick-look porosity is the neutron/density
crossplot porosity.
When light hydrocarbons are present, the Gaymard quick-look formula should be used.
In shaly formations a simultaneous solution of any two of the three porosity tool response
equations gives an adequate "effective porosity" for quick-look purposes.
To determine reliable porosity information, the minimum tool combination required is the
neutron/density. All three common porosity tools – neutron, density, and sonic -- should be run to
obtain the best possible results.
In complex cases where mixed lithology, clays, and light hydrocarbons all coexist, a more
sophisticated approach is required that combines the methods already discussed with some
newer ones.
Use of Ro and Rw
Archie’s equation for clean formations relates porosity, resistivity, and water saturation by the
familiar relationship

By rearrangement of this equation, an expression for porosity may be written:

If it is assumed that the formation of interest is at 100% water saturation (S w = 1), then the
expression for porosity simplifies to

Note that Ro has been substituted for Rt since Sw is assumed = 1. Thus an estimate of porosity
can be obtained from formation resistivity, provided the values of a, R w, Ro, and m are known and
the formation is 100% water bearing.
Use of Rxo and Rmf
In the flushed zone, Archie’s equation for clean formations can be written

where Sxo is the water saturation in the flushed zone.


Rmf can be measured from a sample of filtrate pressed from the mud. Rxo can generally be found
from the readings of the Microlog, provided the depth of invasion is not too small.
A reasonable assumption has to be made concerning the value of S xo. The residual oil saturation
(ROS) seems to vary between 15 and 25% in most cases, with an average value of 20%.
Therefore, Sxo can be assumed equal to 0.8 without too great an error.
Limitations
Obviously there are limitations to the use and applicability of these methods of porosity
estimation. In shaly formations, for example, the porosity estimate will be overly optimistic and in
hydrocarbon-bearing zones overly pessimistic.
Porosity determinations from resistivity measurements are, at best, crude approximations. They
should only be attempted in the absence of an independent porosity device such as an acoustic
log, a density log, or a neutron log. Interpreters of old electrical logs make good use of these
techniques.
4.4 Clean Formations
Density Porosity The simple relation

can be represented graphically, as in Figure 1 . The matrix density in normal reservoir rocks
varies between 2.87 gm/cc (dolomite) and 2.65 gm/cc (sandstone). The fluid density in normal
brines ranges from 1 to 1.1 gm/cc.

Neutron Porosity Compensated neutron tools are run on a matrix setting chosen by the logging
engineer or the company witness. If the actual lithology coincides with the chosen matrix setting,
porosities may be read directly from the log. This, however, is seldom the case. If the matrix is
something other than that used in running the log, the porosity reading from the log will not be
correct. In most instances, too, the lithology is not known prior to logging the well. Therefore, a
standard matrix setting is normally used.
A convenient standard for the neutron is the "Neutron Porosity Index (Limestone)." This is the
same value that the tool would have read had it been recorded on a limestone scale. Figure 2
plots the porosity measured by one service company’s neutron tool using a limestone matrix
against the true porosity for the indicated lines of constant lithology. This chart can be used to
determine the true porosity if the actual lithology is known.
Acoustic Porosity Acoustic porosity may be estimated from the Wyllie time average equation:

This equation is represented graphically in Figure 3 . The interval transit time, t in


microseconds/ft, is plotted against porosity. Lines on the chart correspond to various matrix
velocities in ft/sec. Common reservoir rocks have matrix velocities in the range of 18,000 ft/sec
(sandstone) to 26,000 ft/sec (dolomite). Fluid velocities depend on fluid type (oil, gas, or water)
and the temperature, pressure, and salinity of the fluid, but a working range is 5000 to 5300 ft/
sec. The Wyllie equation is only accurate when the fluid in the pore space is water or dead oil and
when the porosity is intergranular. In gas-bearing formations and vuggy pore systems, the
equation is not even close.
Unconsolidated formations do not conform to the simple Wyllie time average equation. A
correction factor is used in these cases to extend the usefulness of Figure 3 . This correction
factor is the term Bcp, also known as the compaction factor. The value of B cp is usually
determined for a given formation from core analysis and existing acoustic logs. The Hunt
transform (Raymer and Hunt 1980) may be used as an alternative.
Neutron Density Cross plot In cases where lithology is unknown, no single porosity tool can
uniquely determine the porosity. A porosity value can be estimated by combining two porosity
logs, such as the neutron and the density logs.

Figure 4 shows the neutron porosity index plotted against the bulk density. The lines of
constant porosity are substantially straight lines that are parallel to the NW-SE axis. Thus,
regardless of the lithology, a reading of b and  N is sufficient to define a porosity. One limitation
of the chart is that it assumes water-filled porosity; if the porosity is filled with hydrocarbons, use
of the chart becomes complicated.
A simple rule of thumb for finding a neutron-density crossplot porosity is to express  b as an
apparent limestone porosity, add this value to neutron limestone porosity, and divide by 2:

Neutron Acoustic Crossplot On many occasions, a density log may be unavailable (or rendered
useless) by a washed-out hole or other operational problems. An acoustic log may then be
combined with a neutron log, as plotted in Figure 5 . Again, constant porosity lines proceed
generally in the NW-SE direction. However, the trend is not as predictable as the trend in Figure 4
and may become unreliable in gas-bearing formations.
4.5 Shaly Formations
Depending on the perspective of the analyst, a sample of shaly sand can be treated in several
different ways. For example, a core sample may be analyzed for porosity and/or grain density.
Ideally, the application of the density porosity equation to such a sample bears out both lab
measurements and bulk volume theory; i.e., if the sample is of bulk density  and the core
analysis yields a grain density of gr then

whence

If the same shaly sand sample had been logged by a density tool, the derivation of porosity would
be more troublesome, since the value of gr would be unknown. The matrix may be assumed to
be a known substance, such as quartz sand, and a default value used for its density ma.
However, the shale "impurities" in the sand sample change the effective grain density. Thus, the
application of the bulk density/porosity relationship requires care and attention to detail.
The value of gr may be estimated if the fraction of clay solids and sand solids are known.

where Vclay is the bulk volume of clay solids in unit volume of the shaly sand sample. Thus, in
clean sands Pgr is numerically equal to Psand (or Pma) at close to 2.65 gm/cc, and in shales
Pgr is numerically equal to Pclay at close to 3 gm/cc. It should be recognized that shale densities
are considerably less than 3 gm/cc, since a shale is composed of dry clay and water-filled pore
space. For example, a 30% porous shale has a bulk density of 2.4 gm/cc (.3 · 1 + .7 · 3). Only a
very minute fraction of the pores are interconnected and so the shale has effectively zero
permeability and zero useful storage space (zero effective porosity).
Log analysts have evolved many methods to handle the appropriate transforms from logging
measurements to porosity in such cases.
The following development is somewhat generalized.
If a porosity device P reads p on the log and psh in the foreign material (clay fillers or shale
laminae), then the response of the device can be written as

where Vsh is the bulk volume of foreign material, (1 -  - Vsh) is the bulk volume matrix, p matrix is
the tool’s response in pure matrix material, and  is the actual porosity.
By definition, p matrix is zero. Thus, the relationship reduces to

and, by rearrangement,

This generalized equation may be applied to each of the porosity devices. For the density tool, for
example,

Typical values of Dsh range from 20% to -10%, depending on the bulk density of the extraneous
materials. For example, if a density log reads 20% for the assumed matrix, and an independent
indicator (such as a gamma ray) reveals that 30% of the bulk volume is occupied by shale (Vsh =
30%), which has an apparent density porosity of 15%, the equation can be written

In most cases, shaly formations appear more porous than they actually are.
For the neutron log, the equation may be written as

and for the sonic log, it may be written as

Typical values for Nsh lie between 15 and 50% and typical values for Ssh are in the same range.
Again, the combination of two devices can eliminate unknown qualities. For example, a
simultaneous solution of the neutron and density equations for porosity in shaly formations leads
to a useful result:

Eliminating Vsh, these equations can be combined to give

Eliminating , the following solution is found for Vsh:


4.6 Complex Lithology and Secondary Porosity

Neutron and density logs respond to total porosity regardless of its form or distribution. Acoustic
logs, however, tend to ignore irregular porosity (such as vugs) and fracture porosity.
Compressional waves passing through the formation can be thought of as finding a faster path
through the rock matrix around the water-filled vugs and fractures than through the water in them,
although the true mechanism is more complex. Since the sonic tool registers the first arrival time
of the compressional wave, it follows that s represents the primary, or intergranular, porosity.
This effect may be turned to good advantage. A comparison of D and s should indicate whether
or not secondary porosity is present. If D is greater than s we are likely to find vugs and/or
fractures in the formation.

This poses the problem of computing s and D in complex lithology when neither ma nor tma is
accurately known.

4.7 Hydrocarbon Effects

In order to quantify the effects of hydrocarbons on porosity devices, the basic measurement
mechanics of the tools involved should be reviewed. Neutron tools respond to the hydrogen
index. Gas and light oils have a low hydrogen index and thus make neutron readings lower than
they would be had water occupied the pore space instead of hydrocarbons. Density tools respond
to electron density and, because of hydrogen’s unique Z/A ratio, gas and light oils (which contain
hydrogen) appear less dense than water. Sonic tools measure the transit time of compressional
waves, which travel slowly through gas.

Thus, in the presence of gas and light hydrocarbons,

• the neutron log indicates less than true porosity

• the density log indicates more than true porosity

• the sonic log indicates more than true porosity


To quantify the effects of light hydrocarbons, two items must be known: (1) the volume of the pore
space containing hydrocarbons in the volume of the formation in which the tool makes its
measurement, and (2) the response of the tool to hydrocarbons. In general, porosity tools are
shallow investigation devices; i.e., the majority of the measurement is made in the filtrate-invaded
zone ( Figure 1 ), where saturation is Sxo. Thus, the bulk volume of hydrocarbon seen by the tool
is (l - Sxo).

The equivalent porosity of the hydrocarbon depends on its density; this is shown in Figure 2 for
neutron and density tools. Fair approximations for these functions are

where the subscript "hy" refers to hydrocarbon.

The response of the density tool can be described by

whence

In gas-bearing formations,  Dhy is much larger than  T hence  D is larger than


 T For example, if  Dhy = 1.7,  T = 30%, and Sxo = 75%, the density log reading
can be calculated as
 D = 0.3 (0.75 + 1.7 (1 - 0.75)) = 0.3525 or

35.25% porosity
A neutron log under the same conditions would have a  Nhy of zero and
 N = 0.3 (0.75 + 0.(1 - 0.75)) = 0.225 or 22.5%
The log analyst has the problem of deducing porosity from a measurement distorted by the fluid
contained in the pore space. A quick-look method is provided by the Gaymard relationship:

In the previous example with  N = 22.5% and D = 35.25%, the Gaymard formula gives

(or 29.57%)

which is a close approximation to the true porosity of 30%.

More rigorous solutions to the problem can be made via iterative procedures that refine the value
of porosity and Sxo, but these methods require complex processing.

Figure 3 offers a quick solution to the determination of hy. The ratio of  N/ D is plotted against
Sxo. The curved lines are for values of constant hy.
5. LITHOLOGY
5.1 LITHOLOGY INDICATORS

Logs can be used to interpret lithology. The most useful logs for this purpose are

• gamma ray: natural gamma and gamma ray spectra

• density: D and b

• neutron: N

• acoustic: t

In some cases, resistivity or conductivity logs can be of value.

With the exception of the photoelectric factor measurement, e, and the natural gamma ray
spectra log, no single porosity tool measurement gives, by itself, an indication of lithology.
However, much useful information can be gathered by using combinations of porosity tool
measurements.

The most useful combinations are

• crossplots such as b versus N, e versus t, and t versus N

• the M-N plot

• the MID plot

• combinations of the above with e and/or the gamma ray spectra


measurements of K, U, and Th

In many cases, it is possible to scale porosity logs in such a way that two curves, when overlaid,
give an immediate visual indication of the rock type. These methods are to be encouraged. A very
good picture of a geologic column may be gained by simply color-coding an appropriately scaled
combination porosity log suite. In mixed lithology it is essential to identify the rock type in order to
pick correctly the parameters needed to perform other log analysis calculations for porosity and
water saturation. Correct identification of lithology also assists in well-to-well correlation.
5.2 ESTIMATING SHALE CONTENT FROM GAMMA RAY LOGS

Since it is common to find radioactive Materials associated with clay minerals that constitute
shales, it is also common practice to use the relative gamma ray deflection as a shale volume
indicator. The simplest procedure is to rescale the gamma ray between its minimum and
maximum values (in one consistent geologic zone consisting of both sands and shales) from 0%
to 100% shale. A number of studies have shown that this is not necessarily the best method, and
alternative relationships have been proposed. To explain these methods in detail, the gamma ray
index is defined as a linear rescaling of the OR between GRmin and GRmax such that

If this index is called X, then the alternative relationships can be stated in terms of X as follows:

Relationship Equation

Linear Vshale = X

Clavier Vshale = 1.7 - (3.38 - (X + .7)2)1/2

Steiber

Bateman Vshale = X(X + GR factor)

In the Bateman equation, the GR factor is a number chosen to force the result to imitate the
behavior of either the Clavier or the Steiber relationship. Figure 1 illustrates the difference
between these alternative relationships.
5.3 GAMMA RAY

Figure 1 illustrates a gamma ray spectral log. Note that in the left-hand track both total gamma
ray activity (SGR) and a "uranium-free" (CGR) version of the total activity are displayed. Units are
in API. In the two right-hand tracks the concentrations of U, Th, and K are displayed. Depending
on the logging service company, the units may be in counts/sec, ppm, or percentage.

Interpretation of the gamma ray spectra log is a relatively new and developing art. One approach
is to take ratios of elemental concentrations as indicators of formation properties or of formation
type.

The thorium/uranium (Th/U) ratio, for example, varies with depositional environment; it is highest
for a continental, oxidizing environment and lowest for a marine, reducing one that produces, for
example, black shales. It can thus be used in gauging the distance to ancient shorelines or the
location of rapid uplift during the time of deposition. Similarly, stratigraphic correlations of
transgressions or regressions are possible using the Th/U ratio.

The uranium/potassium (U/K) ratio has been found to be a good indicator of the potential of
source rocks in argillaceous sediments. The ratio also correlates well with vugs and natural
fracture systems and, at times, with hydrocarbon shows in both clastic and carbonate reservoirs.

The thorium/potassium (Th/K) ratio has proved useful in rock typing. It is particularly useful in clay
typing, because it increases from glauconite through to bauxite ( Figure 2 ).
Alternatively, the uranium versus potassium crossplot ( Figure 3 ) may be referred to as a
guide to rock type. If additional data are available -- for example, the photoelectric absorption
coefficient ( e) -- then additional plots of the sort shown in Figure 4 and Figure 5 can be made.
Field presentations of gamma ray spectral logs can assist the analyst in the task of mineral
identification by offering curve plots with the ratios of the three components (U, Th, and K)
already computed. Figure 6 gives an example of one such presentation. The left-hand track
shows total gamma ray together with a uranium-free curve. The middle track gives three ratios:
U/K, Th/U, and Th/K. The right-hand track gives a coded display on which the coded area
represents the formations with both the highest potassium and the highest thorium content,
generally the non-reservoir rocks, such as shale.
5.4 USE OF POROSITY LOGS FOR LITHOLOGY IDENTIFICATION

Logs can be used as indicators of lithology. The most useful curves for this purpose are

density - b and Pe
neutron - N
acoustic - 
gamma ray - natural gamma and gamma ray spectra
With the exception of the photoelectric factor measurement, P e, and the natural gamma ray-
spectra log, no single porosity tool measurement will, by itself, give an indication of lithology.
However, we can gather much useful information by using combinations of porosity tool
measurements.

The most useful combinations are

• crossplots such as bulk density versus neutron porosity, bulk density versus
interval transit time, and interval transit time versus neutron porosity

• the "M" and "N" plot

• the "MID" plot

• one or more of the above with Pe

In many cases it is possible to scale porosity logs in such a way that two curves, when overlaid,
immediately give a visual indication of the rock type. The use of these methods is to be
encouraged. A very good picture of a geologic column may be gained by simply color coding an
appropriately scaled combination porosity log suite. In mixed lithology it is essential to identify the
rock type to choose the correct parameters needed to perform other log analysis calculations for
porosity and water saturation. Correct identification of lithology will also assist in tasks of well-to-
well correlation.

Conventional Porosity Logs

The neutron-density crossplot appears in two versions. Figure 1 is to be used in the case of fresh
mud filtrates with f = 1.0, and Figure 2 is to be used in the case of salt mud filtrates with f =
1.1. The differences between the two are slight. The lithology is indicated by the location of
the plotted point. The positions of various nonporous minerals are shown on these charts as
points. However, the porous reservoir rocks appear as lines. Shales typically fall in the southeast
quadrant of the chart in Figure 1 .

A visual reading of a log display on compatible limestone scales yields similar information ( Figure
3 ). Note that the scale for the density log is from 2 to 3 gm/cc across Tracks 2 and 3. The
neutron, in limestone porosity units, is scaled from 42 to -18% across the same tracks. The 0%
limestone point thus coincides for both devices at four divisions from the left of Track 3.

Where the neutron curve lies to the right of the density curve, sandstone is indicated. This
corresponds to the area to the northwest of the limestone line in Figure 1 . When the two
coincide, limestone is indicated. Large separations, such as in anhydrite, are easily recognized.
Another frequently used pair are the neutron and sonic devices. When a density log is missing or
is error prone due to bad hole effects, this plot should be used. Figure 4 illustrates the neutron-
sonic crossplot. It is used in much the same way as the neutron-density crossplot.

Shales typically fall near or to the right of the dolomite line.

The final combination of our trinity of porosity devices is the density-sonic crossplot ( Figure 5 ).
This plot is particularly useful in identifying minerals as well as reservoir rocks.
5.5 M-N PLOT

Pairing of Porosity devices does not make full use of all the data when three devices are
available. If a three-dimensional graph could be built with X, Y, and Z axes corresponding to the
neutron, density, and sonic responses, then identifiable minerals would occupy unique Points in
space. Cases in which a lithologic mixture exists could be more easily interpreted.

For example, a mixture of sand and dolomite appearing as limestone on the neutron-density and
neutron-sonic Plots would be correctly identified by the density-sonic plot. Various attempts have
been made to resolve the Problem of reducing three log readings to a two-dimensional crossplot.
One of the first, the M-N plot, requires that the two parameters M and N be defined as

where f is fluid density. Note that  N needs to be in fractional units and that  N fluid is assumed
to be 1.0. Effectively, these two definitions are algebraic methods of finding the slope of a line that
passes through a Plotted point and the 100% Porosity point. Since any pure reservoir rock will
plot on a line on the crossplots, and the slope of this line is substantially constant, that slope is a
characteristic of the rock type. Having defined M and N and determined them for a given point on
the log, their values may be plotted on an M-N plot ( Figure 1 ).

The M-N plot has a number of shortcomings. For example, it is not truly porosity independent in
the case of dolomite since the neutron response is not linear. Another annoying feature is that the
plotted point depends on f, the fluid density. Finally, the actual values of M and N for common
minerals and reservoir rocks are not easy to remember and have no particular significance in
themselves. Although the M-N plot is still used by some analysts, it has largely been superseded
by another plot that accomplishes the same end result more elegantly – the MID plot.
The Matrix Identification Plot (MID)

This plot requires three porosity tools for input. Its main advantages over the M-N plot include its
independence from porosity and mud type and the fact that it uses meaningful parameters
directly related to rock properties that are easily remembered.

The neutron and density logs are combined to define an apparent matrix density, (ma)a . The
neutron and sonic logs are combined to define an apparent matrix travel time, (tma)a These two
parameters are then crossplotted to define lithology mixtures.

The value of (ma)a a can be computed from the equation

whence
Likewise,

However, in practical log evaluation, these values can be found more simply from charts. For
example, Figure 2 defines (ma)a for any b,  N pair and Figure 3 defines (tma)a for any t,  N
pair. Figure 4 should be used in place of Figure 3 when f = 1.1 (salt mud filtrate). Once
(ma)a and (tma)a have been determined, they are crossplotted on the MID plot chart shown in
Figure 5 .
When using either the MID plot or the M-N plot, it is useful to have available a reference table
which summarizes the values of  ma, t ma, M and N for common minerals and reservoir rocks
5.6 PHOTOELECTRIC FACTOR (Pe)

In most lithologies, minerals exist in combination. Since the overall photoelectric index is not a
linear function of the Pe values of the individual components, a new term, the volumetric
photoelectric absorption index (U), has to be calculated. This index is the product of electron
density and the photoelectric absorption index.

An approximation is usually made by using bulk density  b rather than electron density  e . In a
complex lithology formation, the measured volumetric photoelectric absorption index is the sum of
the individual volumetric photoelectric absorption indexes weighted by their relative proportions in
the formation:
U = U1V1 + U2V2 + ...
U1 = volumetric photoelectric absorption index of mineral 1

V1 = volumetric fraction of mineral 1 in the formation, etc.

For a porous, single-mineral, shaly formation containing hydrocarbons, a general equation can be
written as

matrix water hydrocarbon shale

term term term term

Table 1 shows that the absorption coefficient, Uf, of fresh water is substantially lower than any of
the matrix coefficients, Uma, and can therefore be neglected without introducing a major error.
Only if very salty muds are used would the term need to be included. The hydrocarbon
contribution can also be neglected since Uhy is less than 0.12.
If the shale content of the formation is included in the matrix, the foregoing equations can be
combined to produce the relationship

or more strictly

This equation can be solved using the nomogram shown in Figure 1 . For a quick-look
interpretation, the porosity  can be taken from a density-neutron crossplot.

In complex lithology (a mixture of up to three minerals), a second physical parameter is needed to


define the individual minerals and their volume percentage in the formation. The apparent matrix
density (ma)a derived from the density-neutron crossplot may be used or (ma)a can be calculated
using the standard formula:

where: f = fluid density of the invaded zone

f= apparent porosity from density and neutron data

Once the values of (ma)a and (Uma)a have been calculated, they may be crossplotted against one
another to help in identifying lithology. Figure 2 shows such a crossplot, made over a depth
interval that includes anhydrite, dolomite, and shaly sand.
An overlay can be constructed ( Figure 3 ) which, when placed over the crossplot, indicates the
most probable mineral composition. Of course, geological knowledge and cuttings analysis are a
substantial help in selecting the main contributing components.
Once the main mineral components of the formation have been identified, a plot can be made to
determine their relative proportions. Figure 4 illustrates a valid plot for a quartz-calcite-dolomite
composition, with a grid already established.

Continuous computation of the relative proportions of up to three minerals is possible using this
approach. Wellsite computation and display can be illustrated by an example. A section logged
with neutron, density, and Pe curves is first analyzed by a (Uma)a versus (ma)a crossplot ( Figure 5
), which indicates the three main components of the formation to be magnesite, dolomite, and
anhydrite.

The relative proportions of the three minerals at each depth are computed according to the data
points within the defined triangle. Figure 6 shows the corresponding playback over the same
interval. The average grain (matrix) density (RHGA) is displayed in Track 1; the lithology in the
depth track; and the photoelectric absorption index (PEF), density porosity (DPL), average
porosity (PHIA), and neutron porosity (NPL) in
Tracks 2 and 3. The porosity values are scaled in limestone units. Because of the statistical
nature of the measurements, not all data points will plot inside the selected solution triangle
defined by the three minerals. When a point falls outside the triangle, a flag is raised on the left
edge of Track 1, and only one or two minerals will be displayed.

Drilling muds loaded with barite present a burdensome problem for the lithodensity log. If there is
a high concentration of barite in the mud cake or in the drilling mud, the very high P e value of
barite can severely affect the quality of the Pe curve.
Both gas in the formation pore space and barite in the mud can be detected on a (U ma)a versus
(ma)a crossplot ( Figure 7 ). Because of its high atomic number, barite moves the (Uma)a points
toward the right (higher (Uma)a Gas influences the (ma)a values, but has little effect on (Uma)a, and
so moves the points upward.

Chaveroo Method

A classical method of solving multimineral log problems is historically known as the Chaveroo
method. It was developed when the Chaveroo field was being actively drilled and logged in New
Mexico in the l960s. The method reduces the problem to a simple inversion of a "response"
matrix. It is most easily understood by reference to the Martini Problem.*

* I am indebted to John Doveton of the Kansas Geological Survey for the original Martini
Problem, which I have adapted slightly for the present purpose.

A log analyst, after a hard day’s work, sought comfort and deserved repose in a bar. He ordered
a martini and was immediately struck by the harmonious and mellow proportions of its
ingredients. Wishing to learn the secret of the "perfect martini," he asked the bartender to tell him
how much gin, dry vermouth, and sweet vermouth had been used. The bartender (a geologist
perhaps?) could only reply that it was mostly made of gin "with occasional vermouthian
tendencies." The analyst thereupon set out to back-calculate the relative proportions of the mix
using the alcohol and sugar contents of the ingredients and of the mixture. He "logged" his martini
to find it had 35% alcohol and 3.4% sugar. He consulted a reference book to find the amount of
alcohol and sugar in each ingredient. Here is what he found:

Alcohol Sugar

martini 35% 3.4%

gin 47% 0%

dry vermouth 18% 3%

sweet vermouth 16% 16%

He solved the problem by setting up three simultaneous equations:

alcohol: 35 = 47 VG + 18 VDV + 16 VSV


sugar: 3.4 = 3 VDV + 14 VSV
material balance: 1 = VG + VDV + VSV
where VG is the fraction of gin, VDV the fraction of dry vermouth, and so on.

He then solved the three equations to find VG, VDV, and VSV, the respective fractions of gin, dry
vermouth, and sweet vermouth.

This is not a trivial example. Each logging tool can be used to set up a response equation
describing a lithology mix. If VL, VD, and VS are the bulk volumes of limestone, dolomite, and
sandstone in some lithology mixture, for example, the density of the mixture can be expressed in
terms of the relative quantities of each, and the density tool response to each:
In a similar fashion, the sonic and neutron logs can also define a response equation.
t = 47.5 VL + 43.5 VD + 55.5 VS + 189 

 = 0.07 VD + (-0.03 VS) + 

Finally, the material balance equation reveals that


1 = VL + VD + VS + 
Elaboration of examples such as these should convince even the most mathematically inclined
analyst that level-by-level solutions to matrix inversion problems using only a hand-held calculator
are not recommended. Clearly, this type of processing is best left to a computer that has been fed
log data in digital form.

In the last example, three independent formation measurements were combined to solve for four
unknowns: porosity and the volume fractions of limestone, dolomite, and sandstone. Since it
requires four equations to solve for four unknowns, the material balance equation (stating that the
sum of the component fractions is equal to one) provided the missing equality. Thus, in general,
given "n" independent logging measurements, a solution may be found for "n + 1" components.

For example if, in addition to neutron, density, and sonic measurements, measurements of P e
and gamma ray were also available, then a solution could be sought for two additional
components -- anhydrite and salt, for example.

While the mathematics of the method is straightforward, its weakness lies in the analyst’s choice
of components for which the equation is intended to solve. What happens, for example, if a heavy
mineral (e.g., pyrite with Pma = 4.99 gm/cc) is present but not specified as a component for
solution in the equation? There is a good chance that, in such a case, the point in question will fall
outside the solution polygon and result in an apparent negative amount of one of the other
components meant to be solved for in the equation.
Many computer programs have been written to handle the Chaveroo approach to log analysis in
multimineral environments. Some have elegant logic to handle cases where an unspecified
component appears or a negative amount of a specified component results. Such programs have
their place and their application in situations where either the lithology is reasonably well known
(from experience, core analysis, etc.) and where a knowledgeable analyst is employed to select
both the components and their correct end points. An example of the result obtained from this
kind of processing is shown in Figure 8 . The column logged is Predominantly anhydrite and
dolomite with occasional appearances of sand and gypsum. The well was completed in the
intervals shown, acidized and fractured and put on production at 609 BOPD and 6 BWPD.

Application of the Chaveroo method generates only one answer set for a given input set, since
the system is exactly determined. Once the response equations are written and the end points
chosen for each component, one, and only one, solution can be found for any given set of log
data. What happens then if our log data are subject to statistical variations (neutron, density,
gamma ray) or our response model is inaccurate? Logically, the answers produced will also suffer
from such variations and inaccuracies and there will be no way to assess the magnitude of the
error. In general, it can be stated that the most accurate assessment made by a Chaveroo-type
processing will be that of porosity. However, if the inaccuracies are to be properly gauged,
another type of processing is called for.

Over determined Systems

Naturally, the interpretation model selected by the log analyst depends a great deal on the
logging program and the log responses. Ideally, the system should remain balanced or over
determined; in other words, the number of log inputs should not be less than the number of
unknowns. For example, with five log measurements (b,  N, t, Pe, and GR) a solution is sought
to the relatively simple problem of finding the relative proportions of three components (sand,
lime, and dolomite). Now, instead of only one solution being possible, many are. Any three log
measurements suffice to form an exact solution subset, and in this case there are 10 different
ways to choose three logs out of five. (The reader may choose to write down all possible
combinations of any two log measurements to be omitted.) If the log measurements are perfect
and the model perfect, then all 10 possible solutions will coincide. In practice this never happens,
and the result is a set of 10 possible answers for each component. Which one is correct? For
each possible answer, a value for the original log measurements may be back-calculated. For
example, the density tool response equation may be written

where  1, 2 etc. are the log responses to components 1, 2, etc., and V1, V2, etc. are the
volume fractions of components 1, 2, etc.

If V1,1 is the estimated value of V1 by the first of our 10 possible answer sets, then the back-
calculated value of  b will be

This may then be compared with the actual log reading of b and an error function defined:

If the process is repeated for all the possible solutions, the individual error function may be
summed. The same procedure may then be repeated in turn for each of the logging
measurements.
The problem now transforms itself into one of minimizing the error function; i.e., a solution set in
which the differences between the observed log readings and the back-calculated log values are
minimal is the most probable solution set.

This type of processing, generically referred to as global logic, may equally be applied to any log
response equation, including resistivity log responses to both the invasion process and to water
saturation variations. Figure 9 shows a generalized flow diagram of the logical steps involved in
such a processing chain.

Advantages of global logic processing are that places where the raw log data are unreliable will
become evident and that in zones of special interest the analyst will have a rigidly mathematical
estimate of the probability of the analysis being correct.

A major disadvantage is that the processing is so complex it requires specialists to apply it.
6. INTERPRETATION METHODS

6.1 LAMINATED SHALE MODEL

This model proposes a multilayer sandwich of alternate layers of clean sand and shales (lithified
clay materials). The thickness of each layer is small in relation to the vertical resolution of the
porosity and resistivity devices used to log it. Thus, both porosity and resistivity devices "see" an
"average" that is not a true indicator of the properties of the clean sand laminae.

Figure 1 schematically illustrates that as the proportion of shale laminae (V lam) increases, the
amount of clean sandstone is reduced. The porosity in the clean sand remains unchanged,
but there is progressively less and less of the clean sand present.

In the laminated shale model, porosity measured by the density tool is expressed by

where  Dsh is the apparent density porosity in the shale and  e is the true porosity in the clean
sand. A similar equation can be written for the neutron log:

Combining these two equations results in equations for true porosity and shale content in a
laminated shale:

and
This is represented graphically in Figure 2 .
The water saturation model considers two resistivities in parallel: that of the shale and that of the
sandstone. This may be shown schematically as a measurement of R t being that of the parallel
resistivity of the sand and shale fractions; i.e., total rock conductivity is the weighted sum of the
conductivities of its components.

Ct = Csand (1 - Vlam) + Csh Vlam


If Csand is considered the conventional Archie conductivity, then

and
Hence, Sw will be related to the other parameters by

where Rsh is the shale resistivity. With rearrangement we have

If  e is available from a neutron and density crossplot, it should be used. If only one porosity
device is available, the density porosity is preferred and  e may be approximated as  D (1-
Vlam), where Vlam is estimated from another indicator (GR or SP).

The method runs into difficulties when the resistivity device "sees" the sandwich not as a parallel
system, but as a series or hybrid system. Indeed, depending on the geometry of the laminated
sand, the electrode placement on the resistivity tool, and the mode of measurement, the device
may in fact "see" the sand and shale layers in the sandwich in series rather than in parallel.

Note that the hydrocarbon pore volume should be determined using

HCPV =  e (1 - Sw) (1 - Vlam) h


since, in each foot of formation, only a fraction (1 - Vlam) will actually be sandstone at porosity  e
and saturation Sw.

The laminated model should be used where the individual laminations are small with respect to
the vertical resolution of the logging device.
6.2 DISPERSED SHALE MODEL

This model proposes that clay fines and clay overgrowths on the sand grains progressively
replace pore space. They bring with them very high surface areas on which large quantities of
water are absorbed. Thus, high water saturations are likely to be calculated if Archie's clean
model is employed.

Figure 1 illustrates schematically how the dispersed clays replace porosity. The maximum
value of Vdis is equal to the original porosity, but the volume of matrix material (sandstone)
remains unchanged.

In the dispersed shale model, porosity measured by the density log is expressed by

For the neutron log, porosity is expressed by the equation

By combining these two equations, an expression for true porosity can be developed:

and

This equation is solved graphically using Figure 2 .


The electrical model of the dispersed clay system considers the total porosity to be filled with a
mixture of clay slurry of resistivity Rdis and free water and hydrocarbons, if any. Thus the total
formation conductivity is considered to be the sum of an Archie term referred to as the total
porosity (i.e., both the freely interconnected pores and the clay slurry-filled pores) and a clay
conductivity term that depends both on water saturation and the clay fraction. Figure 3 shows this
electrical model.

For the dispersed case, the saturation equation may be written as

and

For practical purposes,   ,  e, and Vdis can be calculated from the neutron/density crossplot
(some prefer to take   from the sonic). Rdis may be calculated at the shale point as

.
6.3 STRUCTURAL SHALE MODEL

In the structural shale model ( Figure 1 ), clay grains, often aggregates of clay particles or
mudstone clasts, take the place of sand grains. The porosity and permeability of the sand is
affected very little. To the extent that the replacement grains may have a different grain density
and hydrogen index, this condition will have an effect on neutron and density response. The
maximum theoretical fraction of shale in this case is

The response of the neutron and density tools can be written as

which gives

and

This is illustrated graphically in Figure 2 .


The electrical model for the structural case assumes that, in addition to the Archie term, a simple
shale conductivity term can be added so that

Whence

When in doubt as to the actual distribution of shale/clay in the formation, an alternative approach
is to use one of the so-called total shale relationships that is nonspecific to the model selected.
Although none of these relations is based on any sound experimental basis, the most widely used
is probably this modified total shale relationship:
6.4 ELECTROCHEMICAL MODEL

Physical chemists can measure the ability of a crystal surface to absorb water by finding the
number of sites available for ionic exchanges. This is referred to as the cation exchange capacity
(CEC). Different materials have different CEC values. Quartz in the form of sandstone has
practically none, but illite and montmorillonite, because of their high specific (surface) areas, have
high CEC values .

Waxman and Smits derived an equation to express S w in shaly sands as a function of the cation
exchange capacity of the clay disseminated in the sand:

where both F* and Sw refer to total interconnected pore space, B is a constant that is dependent
on the value of Rw, and Qv is a constant determined by multiplying the volume of clay and the
CEC.

This method, while undoubtedly based on sound laboratory measurement and good petrophysical
principles, lacks one vital factor for day-to-day application: the availability of CEC values from
conventional log measurements. Correlations for a given field can be made between core
measurements of CEC and other log measurements (GR, N, ma, etc.). However, no practical
means exists to use this method if only well log data are available. Modifications of the Waxman-
Smits method by Juhasz have made it more directly applicable to measurements from logs
6.5 DUAL WATER MODEL

The dual water model proposes that there are two distinct waters to be found in the pore space.
Close to the surface of the grains is found "bound" water of resistivity RwF. This water is fresher
(more resistive) than the remaining water farther away from the grain surface. This "far" water, of
resistivity RwF, is more saline (less resistive) than the bound water and is free to move in the
pores.

The model suggests that the amount of bound water is directly related to the clay content of the
formation. Thus, as clay volume increases, the portion of the total porosity occupied by bound
water increases. This is illustrated schematically in Figure 1 .

For a water-bearing shaly sand, the model proposes a bulk volume of bound water T SwB and a
bulk volume of free water T Sw , where T is a total porosity. The resistivity model for this mixture
of waters requires that the effective mixed water resistivity be such that

whence

In hydrocarbon-bearing sections, the model proposes a bulk volume of bound water  T SwB, a
bulk volume of hydrocarbon T (1 - SwT) and a bulk volume of free water  e Swe. The total
conductivity of such a system, after taking into account both clay and hydrocarbon, leads to the
relation

where CwB is the conductivity of the bound water and the term VQ Qv is the effective product of the
amount of clay water and its ability to conduct. Fo refers to total interconnected porosity.

In practice,  T is taken as the neutron-density crossplot porosity, RwF is considered equal to a


conventional Rw, and RwB is estimated from porosity and resistivity readings in neighboring shales
(making this method somewhat empirical).
7. COMPUTER APPLICATIONS

7.1 DATA MANAGEMENT

The computer is now an integral part of practically all facets of wireline logging and formation
evaluation. Computers are especially useful in

• design of logging equipment (CAD)


• data acquisition
• data transmission
• data analysis
• presentation of computed results
While the log analyst need not know how to program or use a computer, or indeed, even how it
works, he does need to know its advantages and limitations. This section should serve to orient
the log analyst who must work with computer-generated formation evaluation products.
Methods of Recording, Digitization, and Transmission
Practically all wireline logging data are now recorded in digital form, so that the information
gained from a logging survey may be processed, communicated, stored, and retrieved in an
efficient manner. Modern transmission and processing systems require that the data be in digital
form. where the logging system makes raw measurements in analog form, an analog-to-digital
converter is necessary. The digitization process may occur in the downhole equipment, in the
surface equipment, or at a digitizing service company’s facilities after the logging has been
completed.
When data are expressed in digital form, two items are of importance: the precision and the
sampling rate. The precision depends on how many binary bits are available or used to specify
each number. In early digital logging equipment the precision was not good, but as computers
have advanced, imprecision has all but vanished. Floating-point 32-bit systems are capable of
expressing any item of logging data to a precision far beyond that delivered by traditional logging
sensors.
The sampling rate of digital data is a function of the parameter and the rate of variance. For
"normal" logging data it is conventional to sample data every 6 in. as the logging sensor moves
along the borehole. The 6-in. sampling rate was chosen for two reasons: (1) the vertical
resolution of most logging tools does not justify a shorter sampling interval and (2) the limitations
of most standard tape recording systems preclude sampling more frequently without changing to
higher-density tape drives. Nine-track, 800 or 1600 B.P.I. (bits per inch) tapes are the norm.
Some formation evaluation problems – such as dipmeter recording and laminated shaly sands
evaluation – require more frequent sampling of the data, in which case the sampling rate can go
as high as 60 or 120 samples per foot.
Other recordings require the sampling to be made on a time, rather than a depth, basis, as when
a wireline formation tester is run and pressure is to be recorded as a function of time while the
tool is held stationary in the hole.
Another situation requiring a mixture of both depth and time sampling is the recording of sonic
wave trains as a function of
depth. In this case, for each depth level sampled, several hundred samples (in the time domain)
must be taken of the sonic amplitude at a receiver. Such surveys acquire very large data bases
requiring large amounts of mass storage.
Ever since the well logging industry began to use computer processing, a standard format has
been needed for the exchange of well log information. To fill this need, an information interchange
format called the Log Information Standard (LIS) Format has been defined. This format is
designed to work with various types of computers and information in a real-time data acquisition
environment. This information ranges from sensor response (log) data to interpretation
parameters and binary bit patterns.
The LIS format was designed primarily for use with magnetic tape, although it may also be used
with other types of storage media. A hierarchy is used for storing and retrieving data; this
separates physical representation from logical information interpretation. Each magnetic tape is
kept logically complete by recording on each tape the information required to interpret its data.
The LIS format permits adaptation to new concepts while maintaining compatibility with previously
written tapes so as to have no significant effect on processing software.
The LIS format is made up of two structures: logical and physical. Logical structure refers to the
type of information and its organization; physical structure refers to the physical dimensions of the
information (e.g., bytes on the tape).
Transmission of logging data requires very high baud rates in order to complete a task in a
reasonable length of time with an acceptable level of accuracy. It is now possible to transmit logs
from the wellsite to a computing center or directly to the client’s office. An example of such a
satellite network is shown in Figure 1 .

Referring to the figure, an extensive chain of software and hardware at the hub and field log
interpretation center ensures delivery of log data to the client. When a log transmission from a
wellsite reaches the hub, it enters the log information standard access procedure, which
understands the data structure and allows a log graphics generator to read the data and make a
graphics file. The graphics renderer converts this picture description into rasters and passes them
to the compressor, which prepares rasters for speedier transmission without compromising
information accuracy. The compressor takes a rasterline of, say, 500 white bits followed by 10
black bits, and encodes a message that says, "The line contains 500 white bits followed by 10
black bits." This single message can be transmitted faster than all 510 bits. Finally, Faxdriver
software manages a machine that sends compressed rasters from the hub to the client’s office.
The auto call utility manages the automatic-dialing equipment to establish communication
between the hub and the client’s office equipment.
7.2 COMPUTATION METHODS
Interpretation of well logs can be made on a variety of levels, depending on the requirements of
the user. These levels can be classified as
(A) one point from one zone in one well
(B) a few zones in one well
(C) an entire well
(D) many wells in a field
Level A might be used to pick a top or the location of some marker bed as a guide to further
drilling, choosing a casing point, etc. Level B might be used for completion and testing decisions.
Level C might be used for reserve studies, correlation of the well to nearby wells, and completion
and testing decisions. Level D might be used in some field studies as a starting point for mapping
porosity, oil in place, etc., over an entire field.
According to the requirements, the analyst may use nothing more than a quick-look evaluation or
consume hours of computer time in the analysis of the log data. Consciously or not, all analysts
use some kind of physical model as a basis for analysis. It is worthwhile to reflect on these
models before discussing the mathematical interpretation techniques.

Any petrophysical measurement is a function of the way it is measured ( Figure 1 and Figure 2 ).
Rather like Heisenberg's uncertainty principle, a very accurate value of porosity can be known
(small device), but with no guarantee that the measurement is representative. Likewise a
macro-reading of the value of porosity can be made, but with no certainty as to what volume of
rock it represents.
With any problem of measurement the answer you get is a function of how you make the
measurement. Never was this more true than in well logging. Each practical logging tool, in the
end analysis, is a tradeoff between local precision, which can only be extrapolated fieldwide with
great danger, and global averaging, which is blind to local inhomogeneities.
Where different tools investigate similar rock volumes, their measurements may be combined
with impunity. Where one tool reads a local property and another a global one, their combination
can produce interpretive results that may be either useful or flat wrong. The analyst should bear
this in mind before launching into the purely mathematical aspect of log analysis.
Figure 3 attempts to place these concepts in perspective by crossplotting vertical resolution
against radial investigation for a number of commonly used logging tools.
Single-Level Analysis
Whatever model is chosen, the method used to apply it to the logs at hand is subject to a number
of variables, such as

• the time available to make computations


• the detail required
• the application of the results
Single-level analysis refers to the computation of logs from a single depth interval in a well. Its
purpose is to establish
the presence and extent of porous and permeable zones so that further analysis, testing, and
completion can be planned.
Typically, this kind of analysis is carried out using a programmable calculator or portable personal
computer. A high degree of sophistication is possible with such devices, but the number of levels
in the well that can be analyzed in a reasonable period of time is limited.
Complete Well Analysis
In contrast to the single-level analysis method, the complete well analysis requires that the data
be available in digital form.
Long sections of the formation are evaluated on a continuous foot-by-foot basis, and typically the
output is in the form of a computer-generated "answer" log or tabular listing. Such analysis allows
for corrections to be made to the raw data for a variety of perturbations but does require the
expert input of a log analyst, geological engineer, or petrophysicist.
Field Studies
When a field is partially developed, it may be economically feasible to recompute on a common
base all the logs from all the wells in the field using a common model and a consistent set of
computation parameters.
Results from such studies, which may take weeks or months to complete, lead to more accurate
mapping of reserves and of other potential pay zones in the drilled wells or in undrilled regions of
the field.
7.3 COMPUTATION MODELS

In order to apply quantitative analysis to well logs, some physical model is required to relate log
response to the mineral and fluid content of the formation and its distribution therein.

Common models include:

• clean formations
• shaly formations
• complex lithologies
• fractured formations

Clean Formations

Clean formations conform to Archie’s model, where porosity may be deduced from a single
porosity device (e.g., the density tool) and all formation conductivity is due to the connate water in
the pore space. This water saturation can be determined from the standard relationship

Shaly

Formations Sand-shale sequences present no particular problem, provided that the thickness of
each sand or shale layer is large compared with the vertical resolution of the logging tool. The
shale sections can then be distinguished from the clean sand sections, which are handled in the
conventional manner. The problem arises when either the sand and shale laminae are very thin
or when the shale particles are distributed randomly between the sand grains. In the case of very
thin strata, the logging tools provide "averages" representative of neither sand nor the shale
strata.

Treatment of shaly sand models requires considerable care so that the log readings are matched
to the type, amount, and distribution of the clays present in the formation.

Complex Lithologies

Where the effects of clay materials pose no serious modeling problems, another problem with
mixed minerals (carbonate and evaporite sequences) may exist. This case is more easily handled
by the logical application of the response equations of the individual tools.

Fractured Formations

For the most part, fractured formations are treated as a special subset of complex lithology
models. Particular attention must be paid to the use of an appropriate formation-factor-to-porosity
relationship in a fractured matrix, and to the correct modeling of the resistivity measurements with
respect to water saturation.
7.4 PROCEDURES

Practical Applications for Petrophysical Data


A frequently asked question is "Where do I begin?" This author suggests a structured approach to
log analysis. The approach is divided into logical phases that make the whole process clearer.
These phases are

• data gathering
• quality checks
• reconnoitering
• picking the model
• determining computation parameters
• calculations
• reporting conclusions
Figure 1 offers a flow chart for performing step-by-step desktop analysis, which the reader may
find useful.

Raw Data
Editing Data Gathering You may think you will only need an induction-sonic, but get the rest of
the logs too; for example, the dipmeter may pinpoint a fault. The formation water analysis, core
analysis, drillstem test results -- all can contribute pertinent information. The search for data may
reveal that the file is full of previous log analysis results and the job you had planned to do is
unnecessary.
Quality Checks Read the headings on the log. They contain all sorts of useful information on the
mud type, hole size, operating difficulties, bad tools, and so forth. Scrutinize logs closely for such
readings as baseline shifts and noise.
The raw log data inevitably must be edited before a legitimate set of computations may be made.
The purposes of editing include

• depth matching of all logging data from different sensors or from different logging runs
• removing the environmental effects due to hole size, mud weight, mud salinity,
temperature, pressure, etc.

• removing the effects of filtrate invasion on resistivity devices


Editing requires fairly extensive information about the response characteristics of the logging
sensors, which is not always readily available.
Preinterpretation Techniques
Reconnoitering Start with a log print on a scale of 1 or 2 in. per 100 ft (1/500 or 1/1000 for metric).
Visually examine the log from top to bottom, looking for trends and/or formations of interest. Use
a colored marker pen to delineate shales, evaporites, and porous and permeable zones. Make a
few quick calculations for lithology. Determine, if necessary, whether the formation of interest is a
shaly sand or a carbonate.
One normally makes a preinterpretation pass through a set of logging data before the main
computations to establish the correct values of computation parameters, such as R w.
Common preinterpretation techniques include computation and display of

• Rwa and Rmfa


• (ma)a and (tma)a
• M and N
These indicators assist the analyst in picking both the model and the constants to use in such a
model. Playbacks of curves and crossplotting of selected data pairs are also used extensively in
the preinterpretation stage.
Picking the Model Having made the reconnoitering pass to locate formations of interest, choose a
model to show you which analysis techniques to use. For a shaly sand, try to determine if it is a
laminated shale or a dispersed shale by reference to cores or analysis of cuttings. For a
carbonate reservoir, try to determine which minerals are likely to be present. Geological reports,
sidewall cores, or bit cuttings analysis carry a wealth of information about lithology.
Other, more subtle, pitfalls abound. The user of computed logs is urged to check the raw data
base, the edited data, and the results file for plausibility and conformity with common sense and
independent geologic data.
Determining Parameters You will require a variety of parameters depending on the complexity of
the model used. At a bare minimum, you will need to know the value of R w, the F to 
relationship, and what to use for porosity and R t determination. For more sophisticated models,
you will need tool response end points for various lithologies, shale responses to porosity tools,
shale resistivities, etc. Computer-generated crossplots can help with this parameter selection.
Postinterpretation Checks
Calculations The calculation of saturation and porosity values is the least of your worries. This
can be accomplished by the use of nomograms or by means of a pocket calculator.
Programmable calculators save time. Interactive and batch computer programs are also
available. Service companies offer computer log analysis.
Once a machine computation has been made, it is worthwhile to perform a number of
postinterpretation checks. These serve to qualify the validity of model used and the results
obtained. Such checks should include

• comparison of computed matrix (grain) density with geologic information. For


example, a computed (ma)a value of 2.68 gm/cc in a limestone reef suggests a
raw data error or shift that needs to be corrected before the analysis is rerun.

• monitoring of computed water saturation values. For example, values of Sw =


130% over a long interval strongly suggest that an incorrect value for R w was
used. Unfortunately, many computer programs are preset to cut off high Sw
values at 100%, making them difficult to use

• comparison of computed hydrocarbon densities with PVT data, if available. The


value of  hy is strongly dependent on Sxo, which, in turn, depends on the value
of Rmf used.

• comparison of computed log analysis results with mud logs, wireline formation
tester recoveries, and drillstem tests.
Many more such tests and comparisons have been devised (some of them quite subtle) that
serve to check the validity and accuracy of machine computations. It is up to the end user to
determine if such checks were applied (caveat emptor).
Presentation Methods
The end result of log analysis is a data set describing the variation against depth (or time) of a
number of properties of the formation logged. In order to pass these data to the end user, a
number of presentation methods have been devised. In general, these may be classified as

• graphic displays
• listings
• X-Y plots
Data is most commonly presented in the coded curve plot, which displays the following
information:

• a lithology track using standard geologic codings to display the components of the
matrix

• a porosity analysis track displaying both oil- and water-filled pore space in the form of a
"movable oil plot"

• a hydrocarbon analysis display to distinguish oil and gas, sometimes in the form of a
flag

• a permeability display which, in some presentations, may include estimates of relative


permeability in addition to absolute permeability
Although listings tend to be rather voluminous and to cause "indigestion," they allow the analyst
to extract specific selected sums and averages of computed parameters over limited intervals.
X-Y plots of data pairs are especially useful with the analysis of pressure as a function of time.
Reporting Conclusions Analyses should yield definitive conclusions, whether a simple statement
that sand A is indicated to be productive and should be tested, or a complex computer log
indicating the locations of oil, gas, lithology, net pay, etc. Whatever form the report takes,
document two things: (1) the model, and (2) the parameters (R w, ma’ etc.) used. In this way,
another analyst can make an informed judgment.
Sums and Averages
Of particular use to the analyst are certain sums and averages made over specific depth
intervals. As a minimum, a running total should be available for

h
(1-Sw)·h
Sw h
k h
These may be combined in order to produce values for various averages. For example,

for parallel flow paths,

for series flow paths, and

for random flow paths.


"Pay" may also be summed according to a number of criteria based on tests, at each level, of
porosity, water saturation, permeability, and so on.
Common Pitfalls
A number of common pitfalls lie in wait for the unwary who choose to subject wireline logs to
computer analysis.
It is commonly assumed that the digital log base is without blemish. However, the only
incontrovertible method to assure the accuracy and viability of the digital record is to make a
playback of each raw data curve and visually inspect it for missing data, data busts, mislabeled
curves, and so on.
Once a viable data base has been established, take care with such matters as missing data.
Because of the way logging tools are constructed and strung together, the deepest sensors
always record null data until the lowest sensor on the tool string reaches the depth where the
shallowest sensor was located when the tool string was sitting on bottom. Computations made
over this log interval at the bottom of the well may produce some unexpected results depending
on how the analysis program handles missing data and on what actual number is used in the data
base to indicate missing data. Some service companies use -999 as a number indicating that no
measurement was made.
All log analysis logic is predicated on both porosity and water saturation expressed as fractions. It
is common, however, to list out computed values of porosity and water saturation as percentages.
Problems can occur if the digital data base is not checked prior to computation to see if the input
values for  N or  D are in decimal or percentage form. If the computation logic requires a
decimal form and the data base contains a percentage form, the entire analysis train is derailed.
8. FIELD STUDIES

8.1 INTRODUCTION

One limitation of the single-well approach to log interpretation lies in the fact that the mass of
information gathered during the development of a field cannot be introduced into the processing.
Wells in the same field are never considered to have similarities that could enhance the accuracy
of the results.

In the field study concept, all the logs recorded on the wells in one field are looked at with the
intention of improving the accuracy and consistency of log data and perfecting the interpretation
of each well.

The combination and comparison of logs from well to well provide the log analyst with a powerful
tool for checking the quality of the input data and for defining, at least qualitatively, the errors
associated with each measurement.

Once homogeneity at the input level has been reached, the computation parameters are
selected. Shale characteristics, fluid content, matrix identification, and so forth, are derived both
from logs and from outside data such as lithologic descriptions, core analysis, and test results. In
both cases, field-wide consistency is required.

The field study approach allows a final acceptance test to be applied to the results obtained from
comprehensive log analysis. This test is based on fieldwide continuity controls derived from logs,
cores (when available), and other information (e.g., lithology, R w, salinity). Transforms are
developed for each family of logging programs; these permit the computation of petrophysical
parameters from the logging data.

An uncertainty is also computed for each parameter obtained with each transform. This
uncertainty will be input into the gridding phase, leading to an error estimate at each point of the
grid.

The mapping and gridding phases lead to refined and accurate input to reservoir-simulation
models. Logs used for this purpose are more reliable than cores alone, since logs, unlike cores,
are usually available on all wells within the boundaries of any one field. Numerous checks and
controls introduced into the field study chain ensure the quality of the mappable parameters. Logs
also give more detail than any other measurement, a necessity when planning and controlling
secondary projects with accurate mapping required over small units.

Field study processing itself can be separated into four main parts:

• normalization of input data

• development of field transforms

• individual well processing

• lumping, gridding, and mapping of reservoir parameters

This train is illustrated by the chart shown in Figure 1 .


8.2 LOG DATA INPUT

Data are normalized in two steps: (1) the visual inspection of logs with the help of lithologic
descriptions in order to detect possible markers, anomalies, and zones; and (2) the normalization
itself, to correct the logs for environmental (hole size, mud, invasion) effects and calibration
adjustments.

Quality control must be exercised from the beginning of a logging operation; it is established
through proper shop and well-site calibrations and records. Computers provide the log analyst
with additional methods to check these calibrations and to apply corrections when needed.

Markers such as salt, anhydrite, and shales are recognized, and when present in beds thick
enough to ensure a good statistical analysis, give a first estimate of the calibration accuracy of
the different tools.

Histograms of raw data, preferably normalized, have been used for several years as a calibration
check in homogeneous formations.

Cross plots of raw data often reveal clusters of points that are recognizable throughout a field,
even when changes in lithology, porosity, and the like affect individual histograms beyond the
possibility of use.

Computed values such as ma, and tma develop very stable patterns. In particular, the dispersion
of grain density around its mean values, in a complex lithology formation, remains fairly constant.
A statistical analysis of this histogram in each well has proved to be an effective control both for
log quality and for computation correctness.

8.3 MULTIWELL DATA

The quality and quantity of available data vary widely from well to well. Lack of proper techniques
in the early phase of field delineation, or a reduction in logging programs in the subsequent
development stage, do not always ensure the proper controls necessary to derive high-quality
parameters. On the other hand, a complete body of data is available from some wells. Such
sources are well suited for use as controls.

These key wells are entered into the normal computation chain. Thus it becomes possible to
qualify a computed log, not only for consistency within itself, but also for consistency within the
field. Matrix grain density, travel time, and secondary porosity index often show remarkably
consistent patterns over large sections of a field.

Figure 1 shows  b, t and  N histograms on three wells from a field. The similarity, both in
shape and position, of the three t and  N histograms is outlined by the average curve drawn
over them. This suggests that there is no major lithology and porosity change between the three
wells. The  b shift on Well 3 with respect to Wells 1 and 2 is thus interpreted to be the result of
miscalibration of the density tool used in the third well.

By eliminating some of the variables, computed data often reveal certain formation characteristics
with great spatial consistency. Histograms and crossplots of such parameters have already
proved able to provide excellent information, for both calibrations and computation validity.
Histograms of  ma for four wells are shown on Figure 2 . The grain density peak of 2.84 to
2.85 gm/cc remains a consistent marker throughout the field, even though local changes have
resulted in the development of low density peaks.

Petrophysical parameters within a field are known from sources other than logs, such as core
analysis, test results, and production data. In a field study, these data are compared to log
analysis results, with the following goals in mind:

• to check the quality of the processing by looping back to the input of the
computation chain when discrepancies are not iced

• to define the general petrophysical transforms

• to derive parameters not directly measured by logs -- for example, permeability


Once the general field transform for complete sets of data is known, it becomes possible to set
limits on the quality of available data and to compute the transforms to be applied for reduced
logging programs. For instance, if the neutron were not run in some of the wells, it might become
necessary to derive the best possible porosity from a combination of density and sonic
measurements.

It is then possible, by artificially reducing the number of logs, to develop all the transforms
covering the different logging combinations present in a given study. Obviously, the smaller the
quantity of available data the larger the uncertainty associated with each transform. At the limit, it
becomes impossible to derive some of the parameters for a well that has insufficient information.

In the case of old wells where the only porosity logs are from GNT-type neutron tools, an
intermediate step may have to be considered. In this case,  b or t y may be used as a link
between GNT and porosity. A comparison of GNT and  acoustic on wells where this combination
exists will provide the GNT-to- transform needed when the GNT is the only porosity log
available.

At this point, an ensemble of transforms with associated error estimations has been developed,
covering all log combinations encountered in the field under study. It is a routine matter to
process each well through the proper transform developed for its particular quantity of available
data. Each petrophysical parameter so derived is associated with an error estimation.

8.4 THREE-DIMENSIONAL GRIDDING

Individual reservoirs are described by single-parameter values representative of their


characteristics. The most common parameters required in a reservoir evaluation are listed below:

E - total thickness of reservoir rock

h - net effective thickness

H - net pay thickness

ave - average interval porosity

h - net effective void volume

H - net pay void volume

kave - average permeability

kh - sum of permeability across net effective

thickness

kH - sum of permeability across net pay thickness

Swave - average interval water saturation

(1 - Sw) H - hydrocarbon volume


This list is not exhaustive and other parameters may be computed for special projects.
Productivity may, for instance, be linked to a sand/carbonate ratio; a sand thickness map would
then become interesting.

Whatever the data required, a summation over given intervals inside a zone is necessary.
Besides proper zoning, this requires a definition of the averaging laws (arithmetic, logarithmic . . .)
and cutoff values.

In the zoning itself, the detailed information provided by the log allows the breakdown of a
reservoir into very small, spatially continuous units. This fine zoning becomes necessary to
understand and predict the behavior of a complex reservoir under different conditions of
production and stimulation.

A lumped value, ave for instance, can only be called representative for a given reservoir if the
porosity is not too heterogeneous. A mean value and standard deviation describe only normal
distributions. Bimodal or multimodal distributions of porosity are more common than Gaussian
ones. It is then natural to break large zones into smaller units in which the data are normally
distributed, in order to provide the log analyst with truly representative lumped parameters.

Figure 1 illustrates the gridding of a reservoir into a uniform series of squares. Superimposed on
this grid is a formation thickness map.

Flow profiles derived from production logs are useful in deciding whether a large section can be
considered as a continuous unit.
8.5 TREND ANALYSIS AND KRIGING

The gridding, mapping, and summation may be done by Krigepack, a sophisticated interpolation
technique that takes into account the uncertainty in the data values and computes the overall
error attached to each of its estimations.

The Kriging technique and its capabilities are described in Matheron (1969). Because of its
flexibility, the presentation of results can be adapted to the requirements of the reservoir-
simulation program.

An important addition to the earlier use of this technique is the inclusion of dipmeter results in the
processing. In particular, dipmeter-based extrapolation of formation tops and thicknesses beyond
the area where wells have been drilled is an important aid in defining reservoir boundaries. (For
more information on Kriging, see the presentation on Geostatistics, found under the Exploration
Geology Discipline.)

Output Mapping

The results of a field study must be communicated to those charged with the economic
development of the field. Modern computer mapping allows an elaborate array of presentations to
allow the viewer to "see" the parameter mapped in three dimensions and from any perspective.
Such display methods are to be encouraged. One such display type is illustrated in Figure 1 ,
which shows a trend surface map of a formation top.
9. ADVANCES IN LOW RESISTIVITY PAY EVALUATION

9.1 INTRODUCTION
It is often claimed that, using today's technology, most of the easy-to-find plays have already been
exploited. However, more and more prospects that were once uneconomical to explore or produce just
a few years ago are now targets of ambitious drilling programs. An article by Fanini, et al.(2000), put
the whole problem into perspective, stating that “about 30% of the world’s estimated hydrocarbon
reserves are contained in laminated, low-resistivity, low-contrast, shaly sand formations that are often
either overlooked or are evaluated as unproductive.”
This leaves no doubt that more and more low resistivity reservoirs will have to be evaluated and
produced. And for good reason -low resistivity reservoirs have proven to be good producers. In their
statistical analysis of low resistivity wells throughout the Gulf of Mexico, Shelton and Hill (in Moore,
1993) observe that such completions produced 43% more reserves than the average oil well in the
Gulf, while gas production from such wells is statistically on par with that of typical gas wells.
Bateman (1985) asserts that, in all probability, many potentially prolific producers have already been
plugged and abandoned, owing to lack of adequate formation evaluation. In the early days of log
analysis, it was common practice to regard water saturations above 50% as non-commercial.
However, with advances in logging and seismic technology came the realization that some of those
supposedly non-commercial zones were, in fact, highly productive pays. As more papers and articles
addressed this phenomenon, the problem came to be recognized in basins around the world, and
more companies set pipe on low resistivity wells.
Low resistivity has come to be regarded more as a problem of perception -a problem shared equally
between the logging tool and the Log Analyst. Log data that was once used to summarily condemn a
prospect now serves as a showcase for the Log Analyst in ascertaining the effects of mineralogy, grain
size, porosity, and irreducible water saturation (and hopefully opening new fields to production). We
now know that low resistivity pays can be found any time resistivity tools come across pay zones of
response 0.5 to 5 ohm-m resistivity in which mineralogy, bed thickness, grain size, or water salinity
lead to anomalously low resistivity responses.

Figure 1: Find the Pay; Log provided courtesy of Baker Atlas. The low resistivity zone between
14017 feet and 14128 feet was completed, and tested at 1564 BOPD, 1.6 MMCFD, and 27BWPD.
9.2 HISTORICAL PERSPECTIVE
Low resistivity and low contrast pays have confounded Geologists and Petrophysicists for years. This
special log evaluation problem has existed long before the phenomenon was identified -much less
named.
Imagine the costs (both in terms of unrealized profits and dreams) involved in walking away from an
otherwise sound prospect, where structure, stratigraphy, and economics had previously pointed to a
sure-fire, significant discovery.

AN EARLY APPROACH
Until fairly recently, the first logging tool run in the hole was usually the SP-GR-Resistivity-Sonic
combination. With no pads or bowsprings, these slick tools were able to provide a quick evaluation of
the well, with minimal risk of sticking. Depending on whether these logs showed pay, the decision
would then be made to run more tools in the hole for further porosity, testing, or sidewall cores. But if
the Resistivity curve failed to deflect to the right when the SP deflected left, then the first impression
would be that the well was just a duster -the formation was full of water, and not worth logging any
further.
If there is a lesson to be learned from such an experience, it is that any well worth drilling is worth
logging -thoroughly, and with a full suite of tools, including (at a minimum) sidewall cores. Fortunately,
however, a number of low resistivity pays have been fully logged, cored, and evaluated. Maybe these
wells were drilled “for science”, or maybe they had the backing of an enlightened Management, but
they were thoroughly logged despite their apparent lack of pay. The resulting data provided contrary
indications about the producibility of such wells, especially when core analysis showed oil in zones
which resistivity logs indicated were wet.

STUDYING THE PROBLEM


During the 1960’s and 1970’s a number of papers were published concerning these apparent
contradictions between resistivity and core or production data. Even before this time, a number of
papers were published that identified specific factors which lowered resistivity in a formation. Papers
by Clavier, Coates, and Dumanoir; Patnode and Wylie; Poupon, et al.; Serra; Simandoux; and
numerous others span a range from the 1950’s through the 1980’s. (See reference section.) One
popular brief, published by Schlumberger, entitled A Wellsite Guide to the Recognition of Low
Resistivity Pay Sands in the Gulf of Mexico was widely read, and probably introduced the Low
Resistivity concept to a number of Geologists (and maybe even a few Petrophysicists). Later, another
fine work was published by the American Association of Petroleum Geologists, entitled Log Evaluation
of Shaly Sandstones: A Practical Guide, by George B. Asquith.
As more wells were drilled offshore, the ability to discern low resistivity pay from wet zones became
more important. In 1993, the local geological societies from New Orleans and Houston put together an
atlas of low resistivity pay logs entitled Productive Low Resistivity Well Logs of the Offshore Gulf of
Mexico (Dwight Moore, Editor-in-Chief). Most notable of contributors to this important work is R.M.
Sneider, along with H.L. Darling, W.A. Hill, J.T. Kulha, and G.L. Shelton.

ADVANCES IN LOGGING
Early on, it was recognized that high-resolution dipmeters and their associated visualization products
were useful developments in characterizing low resistivity pay formations. New logging tools have
since been developed that take a different approach to recognizing low resistivity pays. Important
among these are the Array Induction Tool, the 3D Induction Tool, and the Nuclear Magnetic
Resonance Tool.
This module will provide an overview of the geological setting of low resistivity intervals, as well as
common tool limitations, traditional log evaluation methods, and new tools used to discern such pay
zones.
9.3 GEOLOGICAL PERSPECTIVE
Pay zones that produce low resistivity or low contrast log responses are influenced by a variety of
factors associated with mineralogy, water salinity, and microporosity, as well as bed thickness, dip,
and anisotropy. In this module, we will discuss these factors, and their influence in confounding the
response of the resistivity tool.

DISTINGUISHING BETWEEN LOW RESISTIVITY PAY AND LOW CONTRAST PAY


Low-resistivity pay is generally characterized by pay zones that cause deep resistivity log curves to
read around 0.5 to 5 ohm-m. This is often attributed to a combination of shale content, mineralogy,
microporosity and bed thickness.
Low-contrast pay implies a lack of resistivity contrast between pay sands and adjacent shales or wet
zones. This problem is most commonly seen when the resistivity tool encounters a zone that contains
fresh water (or waters of low salinity). As salinity decreases, the electrical pathway through a body of
water becomes weaker and more dispersed, thus causing the water to become less conductive (or
conversely, more resistive). Therefore, while the resistivity of the pay zone may not be low, the
resistivity of the water leg is high enough to make it difficult to distinguish between pay and wet zones.

Causes of Low Resistivity or Low Contrast Pay Zones


A number of factors have been found to act on the logging tool to produce low resistivity or low
contrast pays. In Moore (1993), Darling and Sneider cite the following causes:

• Bed Thickness: some pay zones are simply too thin to be resolved by the logging tool.
• Grain Size: very fine grain size can lead to high irreducible water saturation.
• Mineralogy: conductive minerals (such as pyrite, glauconite, hematite, or graphite) or
rock fragments can have a pronounced affect on resistivity response.
• Structural Dip: dipping beds produce significant excursions on the resistivity log when
orientation between the tool and the bed deviates from normal.
• Clay Distribution: classified as either dispersed, structural, or laminated - all capable of
holding bound water.
• Water Salinity: high salinity interstitial water causes low resistivity within the pay zone,
while low salinity water can cause low contrast pays.
• Any combination of the above: often a combination of inter-related factors causes the
logging tool to read lower resistivity than normal inside a pay zone.
Of all of the factors listed above, probably the most common cause of low resistivity pay is the simple
combination of thin beds containing highly conductive shales (and their associated bound water), along
with thin pay sands which are below the vertical resolution of the logging tool, as shown in Figure 1:
Core through a laminated interval.
We will describe the above factors in more detail later. For now, let’s set the stage with a discussion of
the geological environment in which low resistivity pays are found.

DEPOSITIONAL ENVIRONMENT
Low resistivity pays (and low contrast pays) have been logged in wells throughout the world. They
have been found in clastic basins ranging from the Gulf of Mexico to Alaska, Offshore Brazil to the
North Sea, West Africa, and Indonesia.

An Overall View: Clays, Shales, Sands and Shaly Sands


Detrital clays are sediments born of rock that has been worn down from ancient mountains containing
quartz, feldspar, and clay minerals. These rock or mineral fragments were later deposited by low-
velocity currents as the finest products of erosion and mechanical transportation.
Down mountain streams to rivers, the clays, silts, and muds were deposited on floodplains after
waning floods, or in oxbow lakes formed from cutoff meanders. Further downstream, muds were left
behind by ebbing tides along tidal flats which have not been subjected to fierce wave action. Beyond
the shoreline, and below the depth of effective wave transport, the mud settled over much of the
deeper parts of the continental shelves, slopes and rises. And even deeper and further out, mud
deposited by spent turbidity currents built up on the abyssal plains. This slow, steady deposition of fine
particles contributes to the extensive blanket of deep-sea sediments that extend across the ridges,
trenches, abyssal hills, and abyssal plains of the deeps.
Shales are the lithified equivalent of these clays and fine-grained muds.
Sands are similarly transported downstream, but must be borne by faster-moving currents. The sand
grains settle out of the stream before the clays do.
In Shaly Sand some portion of the otherwise clean sand or pore space is replaced by shale or clay. In
shaly sands, we see either:

• a change of flow regimes which allows the clay to settle out shortly after the sand does, or

• a single (fast) flow regime in which sands have been deposited, after which authigenic clays were later
precipitated out of solution.
We will focus more on clays, shales, and shaly sands in a later section.

A Regional Setting: Deltas and Their Deep-Water Equivalents


Thin, inter-bedded layers of sands and shales that are below the resolution of traditional induction
logging tools are found in a variety of depositional environments. They are common in moderate to
low-energy fluvial sequences, including upper point-bar, levee and crevasse splay settings (Figure 2:
Delta depositional environment). Laminated shales and thin, flaser-bedded sand-shale sequences
are typical in sub-aqueous prograding and tidal-estuarian deltas and shelf sands.

In “Productive Low Resistivity Well Logs of the Offshore Gulf of Mexico”: Causes and Analysis (Moore,
ed. 1993) Darling and Sneider note that low resistivity, low contrast reservoirs have been found in all
major siliciclastic depositional environments, with the exceptions of alluvial fans and aeolian deposits.
Their study of the Gulf of Mexico found that chronologically, these reservoirs range from Middle
Miocene to Upper Pleistocene in age. The low resistivity pay sands are found on shelf and shelf-slope
transition zones in (Figure 3 ):

• lowstand basin floor fans,


• deep water levee-channel and overbank deposits,
• transgressive-marine sands especially the highly burrowed intervals.
• and the lower parts (toes) of delta fronts deposits and adjacent shingled turbidites, and in
the upper part of alluvial and distributary channels.

A Newer Trend: Turbidite Plays


As drilling trends reach into deeper waters, more companies are targeting their exploration programs
to the prolific production found in turbidite fan and channel sand systems (Figure 4). These
systems vary from massive sands to thinly bedded sand-shale sequences, and are deposited
throughout a range of energy levels.

• The channel deposits at the upper part of a fan lobe, near the canyon mouth, are
dominated by massive conglomeratic and very coarse-grained sands. Other deposits in
the upper fan result from slumping or debris flows adjacent to the main channels.
• In the mid-fan area, the channels branch into a braided distributary pattern as the gradient
flattens. Deposits here are not as coarse-grained and show graded bedding, with a
coarser material at the base and progressively finer sands above.
• Lower fan material consists of a series of thin, laterally extensive sand beds deposited in
broad lobes separated by muds.
It is in the waning stages of these flows that clays settle out among the just-deposited sands, creating
sand/clay bodies with potential for later development as low resistivity pay reservoirs.

Individual beds in a turbidite deposit are recognized by a distinctive series of bed forms known as a
Bouma sequence, illustrated in Figure 5: Typical lithology of a Bouma sequence. This alternation of
thin sands and shales in turbidite sequences generates a series of kicks or spikes on SP or gamma
ray logs. In this example, we see the same variation in energy and grain size described above, with
a distinctive fining-upward sequence presented here.
Fanini, et al. (2000) describe thin bedded turbidite sand sequences that are found in the channel levee
and overbank-levee environment as well as middle-to-distal fan complexes.

• Turbidite overbank and channel levee sands tend to be erratic in both thickness and areal
extent. However, they may be hydraulically well connected, both laterally and vertically,
due to cross-cutting or cut-and-fill sequences.
• In the middle and outer turbidite fan complex, the laterally extensive sheet sands and
inter-bedded shales can be highly productive. When they occur in sufficient thickness, a
single wellbore can drain large areas and sustain extremely high flow rates as a result of
the vast hydraulic continuity with better-developed adjacent sands in these reservoirs.
Thus, we see again that low-resistivity pay intervals have been found which are highly productive.
9.4 ROLE OF CLAYS & SHALES IN LOW RESISTIVITY LOG RESPONSE

Clay has been identified as a leading cause of low-resistivity pays. To a certain extent, shale, by way
of its intimate association with clay, also shares responsibility for low resistivity readings across pay
zones. (Unfortunately, a good deal of this shared blame might be based on a loose and inappropriate
tendency to use the terms clay and shale somewhat interchangeably.) However, while it may be
difficult to distinguish clay from shale with common wireline measurements, it is clay that complicates
the log analysis in low resistivity pay zones. In this section, we will look at clay and shale, and examine
the various chemical, mineralogical, and structural factors which interact to lower the response of
resistivity tools in pay zones.
KEY DEFINITIONS
We will begin this discussion of clays and shales by defining these two key terms.
Shale
According to the American Geological Institute’s Glossary of Geology, Shale is defined as a fine-
grained, indurated detrital sedimentary rock formed by the consolidation
(by compression or cementation) of clay, silt, or mud.
It is characterized by a finely stratified structure of laminae ranging from 0.1 to 0.4 mm thick. Shale
contains an appreciable content of clay minerals or derivatives from clay minerals, with a high content
of detrital quartz; containing at least 50% silt, with 35% clay or mica fraction, and 15% chemical or
authigenic materials (Krynine, 1948).
The term shale was originally applied to laminated clayey rock, but now applies to thinly laminated or
fissile claystone, siltstone, or mudstone. The term is sometimes used without implication to
composition, and has been loosely applied to massive or blocky indurated silts and clays that are not
laminated, to laminated silts and clays that are not indurated, to fine-grained and thinly laminated
sandstones, and to slates.
Clay
The AGI Glossary says that Clay is: a rock or mineral fragment, or a detrital particle of any
composition (often a crystalline fragment of a clay mineral), which has a diameter of less than 1/256
mm (4 microns). This size is approximately the upper limit at which a particle that can show colloidal
properties.
It contains a considerable amount of clay minerals (hydrous aluminum silicates) derived by weathering
(primarily decomposition) or by precipitation from feldspathic rocks, and also contains subordinate
amounts of finely divided quartz, decomposed feldspar, carbonates, ferruginous matter, and other
impurities. Furthermore, this composition should have more than 50% clay-sized particles (Twenhofel,
1937), and clay minerals must form at least 25% of the total (Pettijohn, 1957).
This term also applies to loose, earthy, extremely fine-grained natural sediment or soft rock composed
primarily of clay-size or colloidal particles and characterized by plasticity. It is commonly applied to any
soft, adhesive, fine-grained deposit (such as loam, or silicious silt) and to earthy material.
OVERVIEW
From the above definitions, it is easy to see why shale and clay are so often used interchangeably. We
can also see that shale is related to clay through its composition -shale is basically a rough mixture of
about half clay and half silt. In the following sections we will move beyond the dictionary definitions for
a broader discussion of clays and shales, and the ways in which they affect reservoir properties.
Shales
In shales, the silt (often mostly super-fine quartz grains or other materials) does not unduly affect
formation properties or log response in itself (though silt might contribute to microporosity, which can
lower resistivity because of bound water). The clay component does, however, affect log response.
Any thorough log evaluation must account for the effects of shale and clay minerals. These effects
reach beyond a simple reduction in resistivity response. Other petrophysical properties of shales and
clays must be considered, such as density, amount of hydrogen, radioactivity, permeability, etc.
A number of log evaluation strategies have been devised to account for the affects of shales and clays.
The extent to which log evaluation and the resulting estimation of S w is affected by clay or shale will
depend on the type and volume, as well as the distribution of shales and clays with respect to the pay
sand.
Clays
We have just seen that clays generally consist of hydrous alumino-silicate minerals, mixed in with
various other minerals and different proportions of silts. Any porosity found in clay will be affected by
particle arrangement and rock compaction. These pore spaces most often contain water, but may also
contain hydrocarbons.
As with other lithologies, the log response to clay will depend on its composition and porosity, as well
as water-or hydrocarbon saturation. In a pore system, clay minerals are capable of exerting a
response that is entirely out of line with their proportional volume to sand, and which is especially
dramatic when applied to resistivity log response. Sandstones with large amounts of clay, such as
Illite, often show very high values of log-calculated water saturations -but these sands often produce
water-free. This log response is due to the presence of bound-water within the micropores of clay
minerals, as well as surface conductivity of the clays and clay-water system (we will discuss these
factors later).
In fact, many problems attributed to clays are caused by its interaction with water. Water in the
molecular lattice of clay causes it to expand as water molecules squeeze between lattice layers. This
increases the volume of the clay particle, reduces its density, and often causes a reduction in rock
permeability. Clay expansion in water may also be responsible for shale sloughing and for movement
of fines.
Clay Classification
We can take three approaches to the way we classify clay:

• Mineralogy, • Grain Size, and Origin


We will discuss each of these in turn below.
Classification According to Mineralogy

Common clay minerals can be classified in broad types; each having a similar base structure, while
exhibiting a wide range of compositions. On a molecular level, these minerals are arranged into sheets
of alumina-octahedron and silica-tetrahedron lattices (Figure 1: Crystal structure of common clay
minerals). We shall see that the most important aspect of these minerals is their ability to hold
adsorbed water on their grain surfaces.

The four major types of clay minerals are commonly found within the pore system of sandstones
(Figure 2:SEM photos of four common clay minerals): Smectite, Illite, Kaolinite, and Chlorite.
• Montmorillonite (Smectite): Al2 Si4 O10 (OH)2 • n H2O FIG 3 MISSING
Figure 3: Montmorillonite is a well-known member of the Smectite family of swelling clay minerals.
Smectites often occur as frilly minerals within a pore system. When this clay imbibes fresh water, it
swells to several times its original (dry) volume and retains a good deal of water between layers in its
mineral structure. This change in volume can cause Montmorillonite clays to dislodge and migrate
within the pore system, thus resulting in plugged pore throats.
• Illite: K2 Al4 (Si6 Al2) O20 (OH)4

Figure 4: Illite clays commonly appear as fibrous masses of fine crystals. Illite is often associated with
migration of fines, along with a reduction in permeability. High values of microporosity and immobile
bound water saturation are often associated with Illite clays.
• Kaolinite: Al4 Si4 O10 (OH)8

Figure 5: Kaolinite commonly occurs in the pore system as discrete particles in the form of large flaky
booklets which do not attach securely to sand grain surfaces. When Kaolinite booklets become
dislodged, they are usually too large to fit through pore throat openings and so are often responsible
for clogging pore throats.
• Chlorite: (Mg, Fe)5 (Al, Fe111)2 Si3 O10 (OH)8

Figure 6: Chlorite commonly occurs as a pore lining around individual sand grains or in clusters.
Chlorite often contains significant amounts of iron and magnesium within its structure.
When clays are found to contain a combination of clay minerals, they are called mixed-layer clays.
Table 1 (Properties of various clays) lists a number of petrophysical properties associated with four
common clay types.
Classification According to Grain Size
Clay can also be defined as a sedimentary particle of any composition that is smaller than a very fine
silt grain having a diameter of less than 1/256 mm. Recall that the Wentworth Scale classifies particles
by setting a standard definition of grain size (based on diameter). The following table presents that
portion of the Wentworth Scale which applies most to particles encountered in logging:
Table 2: Wentworth Classification of Sands -Clays

PARTICLE TYPE GRADE LIMITS (mm) GRADE LIMITS (microns) AGGREGATE

SAND Range: from 2 to 0.625 mm Range: from 2000 to 50 microns Sand / Sandstone

Very coarse sand 2 to 1 mm 2000 to 1000 microns Sand / Sandstone

Coarse sand 1 to 1/2 mm 1000to 500 microns Sand / Sandstone

Medium sand 1/2 to 1/4 to mm 500 to 250 microns Sand / Sandstone

Fine sand 1/4 to 1/8 mm 250 to 125 microns Sand / Sandstone

Very fine sand 1/8 to 1/16 mm 125 to 63 microns Sand / Sandstone

SILT Range: from 0.05 to 0.004 mm Range: from 50 to 4 microns Silt / Siltstone

Coarse silt 1/16 to 1/32 mm 63 to 31 microns Silt / Siltstone

Medium silt 1/32 to 1/64 mm 31 to 16 microns Silt / Siltstone

Fine silt 1/64 to 1/128 mm 16 to 8 microns Silt / Siltstone

Very fine silt 1/128 to 1/256 mm 8 to 4 microns Silt / Siltstone

CLAY Range: less than 0.004 mm Range: less than 4 microns Clay / Shale

Coarse clay 1/256 to 1/512 mm 4 to 2 microns Clay / Shale

Medium clay 1/512 to 1/1024 mm 2 to 1 microns Clay / Shale

Fine clay 1/1024 to 1/2048 mm 1 to 0.5 micron Clay / Shale

It is important to recognize that this classification scheme does not distinguish between grains of clay
minerals and other similarly sized grains which are not composed of clay minerals. Therefore, this
classification does not account for the unique petrophysical properties that set clay minerals apart from
other particles of the same size.
Classification According to Origin
Clays can be further classified to reflect their mode of origin.

• Allogenic (detrital) clays: deposited with the sandstone at the time the sediments are laid
down. Examples of such clays are found in the structural and the laminated clays
(described below).
• Authigenic clays: rather than being transported, these clays precipitate from solution at a
later time. Dispersed clays (described below) fit into this class of authigenic clays.

McDonald and Schmidt (1992) point out an important distinction between detrital and authigenic clays
in sandstones. Detrital clays tend to be trapped between grains and are not part of the effective pore
network. Authigenic (dispersed) clays are situated within the pore system and, by virtue of their
immense surface area relative to the detrital framework grains (Table 3), have an impact on the
chemical sensitivity and petrophysical properties of the sandstone which is greatly out of proportion to
their volumetric contribution.
Table 3: Nitrogen absorption measurements comparing specific surface areas of quartz
and various clay minerals (From Asquith, 1989, modified after Almon, 1979).

Mineral Surface Area

Quartz 0.15 cm2/gm*

Smectite 752 m2/gm

Illite 113 m2/gm

Chlorite 42 m2/gm

Kaolinite 23 m2/gm

* Depends on grain size and distribution

Distribution of Clays and Shales in the Reservoir


As stated previously, the manner in which clays are distributed throughout the reservoir plays a key
role in the approach that should be used for evaluating the reservoir. Do the logs show a clean sand,
with massive shales on either side, or more likely, intervals of shaly sands? In this section, we will
describe the modes of distribution for each scenario.
A reservoir sand can display any of four modes of distribution (Figure 7: Modes of distribution, after
Schlumberger)

• Clean sand: essentially no distribution of clay


• Dispersed clay: interspersed throughout the sand; either as a coating on the sand grains
or by filling pore spaces between the sand grains
• Laminar clay: thin layers of clay between layers of sand
• Structural clay: clay grains, shale interclasts, or nodules in the formation matrix.
We will discuss each of these modes below.

Clean Sands
Clean sands are made up of relatively pure, well washed sand. As shown in Figure 7, they contain
essentially no clay minerals or shales, and consist solely of sand grains. These sands were deposited
as a result of a single-energy level flow regime.
The standard Archie equation would be suitable for log analysis of clean sands.

Dispersed Clay
Dispersed clays generally occur as a pore-filling component of the rock, and have a variety of crystal
sizes and shapes. They are able to produce a broad spectrum of adverse effects on fluid flow and fluid
saturation properties without necessarily having much effect on the total pore volume of the rock.

Overgrowths and Distinct Particles


The two key criteria used to define and contrast types of dispersed clay in sandstone are clay crystal
structure and location. Dispersed clays are distributed throughout the sand in either of two distinct
forms:

• As clay overgrowths that adhere to or coat the surface of the sand grains (typically seen
in the case of Chlorite).
• As distinct particles of clay that fill some portion of the interstices between sand grains. In
this case, the minutely smaller size of the clay particles allows them to line or fill the pore
throats between the comparatively larger sand grains.

Three Common Forms of Dispersed Clay


Neasham (1979) has identified three common forms of dispersed clays seen in sandstone reservoirs.

• Pore-filling clays are the most common mode of occurrence for Kaolinite, which typically develops as
pseudohexagonal, platy crystals that attach loosely to pore walls or occupy intergranular pores
(Figure 8). The crystal platelets may be stacked face-to-face, forming long crystal aggregates or
"booklets." Kaolinite crystals of either single or stacked platelet morphology are characteristically
scattered in a "patchy" manner throughout the pore system. These Kaolinite crystals are usually not
well attached to each other or to pore walls, and can be mobilized by fast-flowing pore fluids. Kaolinite
crystals that extensively fill pores have a random arrangement with respect to one another and affect
rock petrophysical properties primarily by reducing intergranular pore volume, and by behaving as
migrating "fines" within the pore system.
• Pore-lining clays attach to pore walls and form a relatively continuous, thin mineral
coating (Figure 9). Crystals attached perpendicularly to the pore wall surface are
usually intergrown to form a continuous clay layer that contains abundant micropore space (pore
diameters of less than 2 ). Illite, Chlorite and Smectite typically occur as pore linings.

• Pore-bridging clays are essentially similar to pore-lining clays, except that they not only
line pore walls, but they also extend far into or completely across a pore or pore throat to
create a bridging effect (Figure 10). Pore-bridging clays exhibit extensive development
of intertwined plates and fibers that produce an intricate network with abundant
microporosity and tortuous fluid flow pathway. Smectite, Chlorite and Illite all display this
morphology, although it is most typical of Illite.
Sand intervals that contain dispersed clays are deposited under a single flow regime, and the
dispersed clays are subsequently formed within the sand as a result of authigenesis or through post-
depositional bioturbation or diagenesis. It is not uncommon for theses clays to develop as a result of
precipitation or by alteration of preexisting silicate minerals.
Dispersed clays are able to increase total water saturation while significantly reducing the resistivity,
porosity and permeability of the sand. Therefore, when dispersed clay content exceeds about 40% of a
sand’s pore space, it can severely impact the producibility of the sand.

• When dispersed clays take the form of a coating on the sand grains, an increase in
irreducible water saturation takes place, along with a substantial reduction in well log
resistivity.
▪ Completions in shaly sands containing dispersed clays often produce water-free because
of their high irreducible water saturation.
• When dispersed clays fill the pores between sand grains, they take up space that would
normally serve as a channel for fluid movement between pores. Furthermore, the
wettability of such clays tends to be greater than that of the surrounding quartz grains.
The sum total of these factors is a reduction in porosity and permeability, along with an
increase in water saturation.
Juhasz (1986) describes the significant impact which dispersed shale can have on producibility: “A
certain amount of dispersed (pore filling) shale has a far more detrimental effect on the permeability of
the sand than the same amount of shale concentrated into shale laminae between clean sand
laminae. The permeability of a 33% porosity clean sand for instance would be reduced to practically
zero if its pore-space is filled with shale (that is: Vsh -33%), but it would retain two-thirds of its
permeability if this shale is present in laminations only.”
Different Perspectives
Asquith (1990) notes that any log analysis in pay sands which contain dispersed clays must consider
the authigenic origin of this clay. Asquith feels that most dispersed clay is diagenetically formed in
place after deposition of the sand. Because they form under different conditions than those which
formed adjacent shale beds, these dispersed clays may differ in composition and more importantly,
may exhibit different resistivities from that of the adjacent shales. This difference between the
resistivity of dispersed clay in the reservoir and the resistivity of the adjacent shale is especially critical
in cases where the resistivity of the adjacent shale is greater than the resistivity of the shaly sands.
Any log analysis through such zones should therefore use an equation that does not require a
reference resistivity (Rsh) taken from adjacent shales.
Thomas and Stieber (1975) built a model (discussed later) which assumes that “within the interval
being investigated, there is no change in shale type and the shale mixed in the sand is mineralogically
the same as the “pure” shale sections above and below the sand.” In their view, this similarity stems
from the fact that both sand and shale facies “are derived from the same source material, carried by
the same river and emptied into the same basin. The differentiation between sands and shales begins
as the particles settle at differing rates according to their size and transport energy and not mineral
type.” Thus, they feel that “the porosity destroying material introduced into a sand stratum will be of the
same composition as the shales above and below the sand stratum. Of course, this will not be true for
the diagenetic alteration of feldspars into clay within the sand stratum.”

Laminar Clays and Shales

Laminar clays are distributed in a reservoir as relatively thin layers of allogenic clay or shale that have
been deposited between otherwise clean layers of sand. (Figure 11).
Each lamination of compacted clay, mudstone, or siltstone is a distinct layer, and each such layer can
vary in thickness as well as proportion of sand, silt and clay. The overall sand and laminated clay
interval reflects multiple cycles of deposition under a dual flow regime characterized by fluctuations in
energy levels. This regime requires higher energies for deposition of sand grains than are required for
the lighter shale constituents of muds, clay minerals, and silts.
Reservoirs containing laminated shales alternate between layers of reservoir-quality sands and thinner
layers of shales and clays having zero effective porosity. These shale laminations do not affect the
resistivity, porosity or permeability of the surrounding sand streaks themselves, and a shale fraction of
up to 60% can therefore be tolerated in the reservoir. However, many logging tools lack the vertical
resolution to differentiate between individual thin beds of sand and shale (Figure: Tool resolution
versus bed thickness). This lack of vertical resolution causes many standard logging tools to average
their readings over such alternating sequences of sand and shale.
Though laminated shale layers are generally thinner than the adjacent sand layers, the clay
constituents contribute a disproportionate change in resistivity and porosity for their thickness.
Petrophysical and reservoir properties between each layer may vary because of changing proportions
of clays within each lamination. However, Asquith (1990) reasons that because of their detrital origin,
shale laminations between sands normally have the same clays and water content as adjacent thick
shale beds. This similarity leads to the assumption that resistivities of laminated shale will be similar to
those of adjacent thick shales. Therefore, it is safe to use log analysis equations that require clay
resistivities, and in such equations, the resistivity of adjacent shales is used to represent that of the
shale in the shaly sand.
The problem of finely layered sands and shales is fairly common. In the Gulf Coast region of the
United States, laminar shales have been found in about half of the low resistivity zones.

Structural Clays and Shales

Structural clays or shales consist of shale nodules or lithified clay fragments that have been intermixed
with grains of sand to form part of the sandstone rock matrix. Unlike dispersed clays, whose grain size
is so small that they occupy the interstices between framework grains, the structural clays have a grain
size that is at least as large as the sand grains, therefore placing the structural clay fragments into the
framework of the matrix. Figure 12 shows clasts of structural shale intermixed with grains of sand.
Structural clays or shales occur by way of three different modes:

• As reworked fragments of lithified shale that have been deposited simultaneously with sand grains of
comparable size.
• As nodules that replace selected grains through diagenesis (as when feldspar is transformed into
clay).
• As nodules introduced through bioturbation.
Sand intervals that contain structural shales are deposited under a single flow regime, with the shale
fragments being deposited simultaneously along with the sand grains. Because of their size, structural
shales act as framework grains, and so do not alter reservoir properties through clogging interstitial
spaces between grains. Structural shale does not commonly occur in quantities to affect reservoir
quality. However, when evaluating reservoirs that contain structural shales, the approach must account
for the way in which the clay grains will affect log response, as opposed to simply trying to evaluate the
response through a homogeneous sand. According to Visser, et al. (1988), structural and laminated
shales produce similar log responses. These detrital clays can be compared to adjacent thick shale
beds for resistivity, thus enabling shaly sand equations that require a shale resistivity value to be used.
9.5 NON-CLAY CONTRIBUTORS TO LOW RESISTIVITY
While clay is often cited as a prime factor of low-resistivity pays, there are other instances where clay
does not enter into the picture at all. Though not a common occurrence, it is important to be able to
identify these factors and recognize the ways in which they might affect log resistivity readings.

CONDUCTIVE MINERALS
In reservoirs where actual water saturations proved to be lower than initially calculated, it was found
that conductive minerals were occasionally to blame for low resistivity readings.
Pyrite, a common heavy mineral sometimes found in marine sedimentary rock, is a good conductor, as
well as other minerals such as hematite and graphite. Sands containing as little as least 5-7% of
disseminated conductive minerals are able to produce low resistivity readings (Figure 1).

Pyrite has been found to form a continuous electrical network at low mineral concentrations, and
exhibits electrical conductivity that can be greater than formation water conductivity. Al-Awad,
Hamada, and Almalik (2001) state that pyrite resistivity can range from 0.03 to 0.8 ohm-m. Pyrite's
conduction is metallic (electronic), and consequently any transfer of current between water and pyrites
is based on a conversion from ionic to electronic conduction and vise versa. This leads to polarization
at the water-pyrite interfaces, with corresponding frequency-dependent electrical properties. Thus, the
electrical properties of porous rocks containing pyrite will depend on the amount of pyrite and its
distribution inside the rock, along with the frequency of the electric current measurement. While this
type of low resistivity pay is rarely encountered, conductive minerals are capable of exerting a greater
influence on log readings than shale.

SANDS DERIVED FROM IGNEOUS OR METAMORPHIC ROCK

In some fields, productive intervals have been developed in “granite washes” and in fine sands
consisting of igneous or metamorphic rock fragments. These intervals should not be classified as
igneous -they are deposited as the erosional products of volcanic, plutonic, and neighboring country
rock (Figure 2).
According to Darling and Sneider (in Moore, 1993), these intervals exhibit highly variable mineralogy;
usually have a very high percentage of feldspars, igneous and metamorphic rock fragments and
quartz, but may be low in clay mineral content. The fine-grained sands, when combined with alteration
of lithics (volcanics) and variable mineralogy produce high gamma ray readings, low resistivities, and
little or no spontaneous potential development. This type of log response can easily be misinterpreted
as shale.
These low-resistivity pay intervals can vary in thickness from millimeters to hundreds of meters.
Natural gamma ray spectroscopy logs should be used to identify such intervals.
9.6 CATION EXCHANGE CAPACITY
We have previously described how the distribution of shale and clay affects S w values by reducing
porosity and permeability within a sand body. We will now take a closer look at clays, their molecular
structure, and associated electrical charges. This section will describe the process by which clay
lowers resistivity readings.

A MOLECULAR VIEW

The clay minerals we investigate consist of thin sheets made by lattices of alumina octahedra that are
combined with silica tetrahedra (Figure 1).
The clay sheets usually display negative electrical surface charges, which help to lower resistivity
readings seen by downhole logging tools. The most common cause of this excess negative surface
charge is an exchange of Al+++ in the clay lattice with ions of lower positive valence (the most common
exchange cations being Ca++, Mg++, H+, K+, NH4+, and Na+). After such substitution, the structure of the
crystal remains the same.
These cations cling to the surfaces and broken edges of clay minerals, but are easily replaced by other
cations in solution when the clay particles are immersed in water. In such wet clays, the dielectric
properties of the water weaken Coulomb forces that hold the clay’s surface cations, thus releasing the
cations. Electrical balance is maintained by keeping these cations in a layer at the clay-water interface,
which contributes to the conductivity of the rock.
So, while dry clays act as insulators, the wet clays seen by logging make great electrical conductors.
As just described above, ion exchange is responsible for the surface conductance found on wet clays
when cations flow between exchange sites on the clay lattice. These cation exchange sites are caused
by a charge imbalance on the external source of the clay’s molecular building blocks, and provide
electrical pathways through the clay.
The positive surface charge of wet clay is a function of its cation exchange capacity (CEC). The cation
exchange capacity is related to the concentration of compensating cations on the clay surface which
can be exchanged for other cations available in solution. The CEC quantifies the ability of a clay to
release cations. CEC is expressed in terms of milliequivalents of exchangeable ions per gram of dry
clay (1 milliequivalent equals 6 x 1020 atoms). CEC can also be expressed as milliequivalents / unit
volume of pore fluid (Q), where Q= CEC x  (1 x ) /  meq/cc (Dewan, 1983). In Table 1, we see the
CEC values and the components of the four common clay types we normally see.
Table 1: Petrophysical properties of common clays
Clay CEC Neutron Average Minor Potassium Uranium Thorium
Type Porosity density Constituents % ppm ppm
meq/g
g/cc
Montmorillonite 0.8 - 0.24 2.45 Ca, Mg, Fe 0.16 2-5 14 -24
1.5

Illite 0.1 - 0.24 2.65 K, Mg, Fe, Ti 4.5 1.5 <2


0.4

Chlorite 0 - 0.51 2.8 Mg, Fe --


0.1

Kaolinite 0.03- 0.36 2.65 -- 0.42 1.5 -3 6-19


0.06

Because clay’s conductivity is a surface conductivity effect, those clays with large specific surface
areas exhibit larger CEC values. The following table, seen previously in an earlier section, compares
the surface areas of common clay minerals to that of quartz.
Table 2: Nitrogen absorption measurements comparing specific surface areas of quartz and various
clay minerals (From Asquith, 1989, modified after Almon, 1979).

Mineral Surface Area

Quartz 0.15 cm2/gm*

Smectite 752 m2/gm

Illite 113 m2/gm

Chlorite 42 m2/gm

Kaolinite 23 m2/gm

* Depends on grain size and distribution

Note that the table shows the surface area of quartz is measured in terms of square centimeters,
whereas that of clays are measured in terms of square meters. This table shows that clay provides a
surface area which is an order of magnitude greater than that of the typical sand grain. It is this large
surface area that allows clay to bind large quantities of immovable water (bound water) to the surface
of the clay. High-CEC clays have a greater impact on lowering resistivity than clays (or other minerals)
with low CEC. With their higher cation exchange capacity, we see that when Montmorillonite or Illite
form major constituents of shaly sands, they cause higher conductivity (i.e. lower resistivity) than shaly
sands containing Chlorite and Kaolinite.
Figure 2 generalizes the relationship between grain size and grain surface area.
As we shall see, conventional logging tools do not provide a direct measurement of either CEC or
surface area, and can provide no general correlation between measurements of specific area or CEC
and any single clay indicator.

The Role of Water


CEC is also influenced by the salinity of the water which interacts with the clay. Hill, Shirley, and
Klein (1979) describe this solution within the pores of shaly rock as having properties which vary
with distance from the charged clay surfaces. Close to the surface, the solution contains only
water and exchange capacity cations. Farther away, the solution contains equal quantities of
anions and cations. The solution between these two zones is assumed to grade between the two
extremes.
The relation between absorbed water per unit exchange capacity and the concentration of
equilibrium solution is described by the following equation:

Ws = weight of anion-free water


CEC = cation exchange capacity
Co = concentration of equilibrium solution

Figure 3 (below) shows how this equation predicts that, at constant salinity, the relation between
anion-free water and cation-exchange capacity should be linear, pass through the origin, and have a
slope equal to 0.084Co-1/2 + 0.22.
If the density of the absorbed water is assumed to be unity, the above equation can be written as:

where
ØE = Effective porosity
ØT = Total porosity
Vs / Vp = the volume of bound water per unit total pore volume, and
Qv = the cation exchange capacity per unit pore volume milliequivalents/unit volume of pore fluid.

Geological Forces
We have seen that clay minerals are hundreds of times smaller than sand grains, but their high
surface-to-volume ratio allows them to bind large amounts of surface water. And since the clay
particles are electrically charged and the water molecules are polar, the surface water is held by
electrostatic forces. After clays are initially deposited, the subsequent deposition of just a few thousand
feet of overburden will compact the clay/shale, thus squeezing out any loosely held water. However,
the remaining electrostatically bound water will not be removed by further burial.
9.8 IRREDUCIBLE WATER
We have already seen how, on a molecular level, the interaction between clay and water results in
lower resistivity values. Now we will step back somewhat, and readjust our sights for a microscopic
examination of the pores in a pay zone. At this level, we will see that water, rather than clay, is a prime
factor contributing to low resistivity pays.

In this section, we will describe a number of inter-related factors, each of which are intimately tied to
the amount of non-producible bound-water that a reservoir can hold (Figure 1: Water at the
intergranular scale). Though not produced, this bound-water is none-the-less detected and measured
by resistivity tools, which do not distinguish between freely produced water and immovable water.
We will start with a brief review of the concepts of porosity and saturation, and will take a closer
look at permeability and capillarity as they relate to bound-water. We will see that structural position
also plays an important role, along with rock-fluid interactions and fluid-fluid interactions, in determining
whether a low-resistivity pay zone will produce water or hydrocarbons.

POROSITY
Porosity is the ratio of pore space in the rock to the bulk volume of the rock. It is expressed as a
fraction or as a percent of the bulk volume. In equation form,

where:
ø = porosity in fraction
Vp = pore volume
Vb = bulk volume
Vp and Vb can be expressed in any consistent units.

Porosity Classification
In terms of production, three types of porosity are recognized:

• total porosity refers to all pore space in a rock.


• effective porosity refers only to that portion of the total porosity consisting of
interconnected pore spaces; more specifically, effective porosity is that portion of the total
porosity which will allow fluid flow under normal recovery processes in the reservoir.
Effective porosity is a dimensionless quantity, defined as the ratio of interconnected pore
volume to the bulk volume.
• non-effective porosity is the remaining portion of total porosity which occurs either as
isolated pore spaces or as microporosity.
The difference between these porosities may be significant in highly vuggy or fractured reservoirs,
where some vugs or fractures may be isolated.
The presence of clay also complicates the definition of rock porosity. Although the layer of closely
bound surface water on the clay particle can represent a very significant amount of porosity, it is not
available as potential reservoir porosity for hydrocarbons. Thus, a shale or shaly formation may exhibit
a high total porosity, while actually having a low effective porosity as a potential hydrocarbon reservoir.
Bound water is held by non-effective porosity. When we calculate water saturation for producibility
estimations we are must be sure to use effective porosity.

Microporosity
Microporosity refers to pore spaces which are so small in diameter (4 µ or less) that they trap and hold
water immobile through capillary action. Microporosity is commonly associated with authigenic clay
minerals whose open structure is able to trap water. Another example is chalk, which commonly
exhibits a large percentage of microporosity, with very high total porosity but low matrix permeability.
Microporosity is considered non-effective porosity as far as the production potential of the reservoir is
concerned. If it is not recognized as such, microporosity can lead to optimistic predictions of potential
reservoir porosity. On the other hand, bound water associated with extensive microporosity can lower
resistivity readings and lead to pessimistic estimations of water saturation.

SATURATION
Saturation is a measure of the relative volume of each fluid in the pores. Thus, oil saturation is defined
as the ratio of the volume of the oil in a porous rock to the pore volume of the same rock. It is
expressed in fraction or in percent, and ranges from 0 to nearly 100%. Water is always present in all
reservoirs, and its saturation is always greater than zero. In contrast, the oil saturation is zero in gas
reservoirs, and the gas saturation is zero in oil reservoirs when the pressure is above the bubble-point.
Oil or gas saturation is calculated by subtracting the water saturation from unity (in two-phase
reservoirs).

Irreducible Water Saturation


Irreducible water saturation (sometimes called critical water saturation) defines the maximum water
saturation that a formation with a given permeability and porosity can retain without producing water.
This water, although present, is held in place by capillary forces and will not flow. Critical water
saturations are usually determined through special core analysis.
The critical water value should be compared to the reservoir’s in-place water saturation calculated from
downhole electric logs. If the in-place water saturation does not exceed the critical value, then the well
will produce only hydrocarbons. These saturation comparisons are particularly important in low
permeability reservoirs, where critical water saturation can exceed 60% while still producing only
hydrocarbons.

Using Magnetic Resonance to Obtain Bound Water Saturations


After describing total, effective, and non-effective porosity (above), we can now define the saturation of
non-producible bound water in terms of effective and total porosity through the following equation:

where
Swb is bound water saturation
T is total porosity
E is effective porosity
In essence, this bound water saturation equation divides non-effective porosity by total porosity.
Total and effective porosity measurements can be obtained through magnetic resonance logging. For
more information on magnetic resonance logging (NMR) in low resistivity pay zones, see the section of
this module entitled Advances in Logging.
Ostroff, Shorey, and Georgi (1999) use a variation on the Hill, Shirley, and Klein formula (1979) to
show how magnetic resonance log measurements can relate bound water saturation to cation
exchange capacity and in-situ flushed zone water salinity:

where
• Qv is the cation exchange capacity per unit volume
• Co is the salinity of equilibrated saturant brine in equivalents/liter

• Swb is the NMR-derived clay bound water saturation, which equals CBW / total where CBW is
NMR-derived clay bound water.

WETTABILITY AND INTERFACIAL TENSION


The principal fluids in a petroleum reservoir are water, oil and gas. When they exist as free phases,
they are generally immiscible. When these immiscible fluids co-exist within the pore spaces of a
reservoir, their interactions with one another and with the enclosing rock will control their spatial
distribution and movement. The two principal properties used to quantify these interactions are
wettability, which pertains to rock-fluid interactions, and interfacial tension, which relates to fluid-fluid
interactions.

Wettability
Wettability is the tendency for a fluid to spread or adhere to a solid (rock) surface in the presence of
other immiscible fluids. For example, in a hydrocarbon reservoir, we are primarily concerned with
water-oil-solid, gas-oil-solid, or gas-water-solid systems. In those systems involving two liquid phases
(oil and water), one of the liquid phases will preferentially wet the surface, and hence is called the
wetting phase, whereas the other is the non-wetting phase. In those systems involving gas, the liquid
is usually wetting, whereas the gas is non-wetting.

Interfacial Tension
The wettability of a surface is determined by the interaction of interfacial energies that act on the fluids
and the surface. For two immiscible co-existing fluids in a porous media, the one with the lower
interfacial tension is the wetting phase, while the other is the non-wetting phase. Interfacial tension is a
measure of the surface energy per unit area of the interface between two immiscible fluids, such as
water and crude oil, or oil and gas. The lower the solid-fluid interfacial tension, the lower the surface
energy and the higher the tendency for the fluid to wet that surface.

PERMEABILITY
Permeability is a measure of the ability of porous rock to transmit fluid.

Permeability Classification
Permeability is further classified as either absolute or effective, depending on whether one or more
fluids occupy the pore spaces of the rock.
Absolute permeability occurs when only one fluid is present in the rock. It is independent of the fluid
used in the measurement. This assumes that the fluid does not interact with the rock.
Effective permeability is the measured permeability of a porous medium to one fluid, when other fluids
are present. Effective permeability depends on the relative proportion of the fluids present (fluid
saturation).
Consider the case of oil and water together in a pore system. Under a given pressure gradient, the oil
and water flow through a pore system together. Based on Darcy’s equation, we find that:
for oil -

for water -

where:
k = permeability
q = flow rate
 = fluid viscosity
p = pressure differential
L = length
A = cross-sectional area
Furthermore, we find that the total flow rate (Qt,) is expressed by the equation:
Qt = (Qo + Qw),
Qt is less than the flow rate that either phase would have if it were at 100% saturation. Thus it appears
as though the two phases interfere with each other’s progress through the pore system. A useful way
to quantify this phenomenon is to define the relative permeability, (kr).

Relative Permeability
Darcy's definition of permeability was for a porous medium which was 100% saturated with the flowing
phase (the phase was water). Hydrocarbon reservoirs normally have two and perhaps three phases
present: both water and oil; or water and gas; or water, oil, and gas sharing the pore space of the rock.
We have seen that having more than one phase present in the pores reduces the ability of the rock to
transmit any one of the fluid phases. For this reason, we define the effective permeability as the
permeability to one phase when there is more than one phase present in the pore space. Its value
decreases as the phases' saturation decreases. There is an effective permeability value for each
phase present.
Usually the effective permeability is expressed as a fraction of the absolute permeability, which is the
permeability at 100% saturation of the flowing fluid. This ratio of effective to absolute permeability is
termed the relative permeability, and can be displayed as a set of curves as shown in Figure 2, for an
oil and water system.
This graph shows that the relative permeability to oil decreases as the oil saturation decreases and the
water saturation increases above its irreducible (or connate) value. Conversely, the relative
permeability to water increases, reaching a maximum when the oil saturation is at its residual
saturation. This same general principle applies to any two-or three-phase system.
The graph shows that relative permeability is also a function of fluid saturation. When multiple,
immiscible fluid phases flow in a rock, the sum of the effective permeabilities of the various fluids will
commonly be significantly less than the absolute permeability measured with only a single fluid in the
rock. A different way of stating this is that the sum of the relative permeabilities for all the fluids in the
rock will commonly be less than one.
Relative permeability is the ratio of the effective permeability of the rock to one phase divided by the
absolute permeability, and it is quoted at some particular saturation value:

kro = kr / k
krw = kw / k

Relative Permeability and Irreducible Water Saturation


Another important relative permeability concept is that of the irreducible or residual saturation.
If two fluid phases, A and B, are flowing in a rock, the relative permeability of fluid phase A will
decrease as the saturation of fluid A decreases. At some non-zero saturation of fluid A (commonly 5%
to 4%), fluid A will cease to flow, and only fluid B will continue to flow in the rock. The saturation at that
point is termed the irreducible or residual saturation of fluid A for the A-B two phase flow system in this
rock.
Relative permeability to oil at irreducible water saturation is 100% or 1, and as water saturation
increases, kro decreases until it effectively reaches zero at some high water saturation corresponding
to Sor, the residual oil saturation.
Relative permeability to water, on the other hand, commences effectively at zero when the rock is at
irreducible water saturation Swi, and thereafter increases as Sw increases. It should also be noted that
in an oil-wet system, kro is always less at a given Sw than in a water-wet system. Conversely, krw is
always greater in an oil-wet system than in a water-wet one.
Many workers in this field have proposed generalized empirical equations to relate kro and krw to Sw,
Swi, and Sor Of particular note are those cited in Honarpour, Koederitz, and Harvey (1982), Molina
(1983), and Pirson, Boatman, and Nettle (1964). A commonly used approximation gives

Structural Position
If a well is completed above the transition zone where the reservoir is at irreducible water saturation
(krw = 0), then water will not be produced.

CAPILLARY PRESSURE
In everyday experience, water levels in two or more connected containers have the same level if
exposed to the same atmospheric conditions. But when it comes to spaces of capillary size (like those
we encounter in porous media), we cannot take this rule so literally. To illustrate, consider what
happens when a tube of capillary size is dipped in a larger container filled with water (Figure 3) The
water in the capillary tube rises above the water level in the container to a height that depends on
capillary size. Although strictly speaking, the water still finds its level, it does so in such a way as to
maintain an overall minimum surface energy.

In this situation, the adhesion force allows water to rise up in the capillary tube while gravity acts in the
opposite direction. The water rises until there is a balance between these two opposing forces. The
differential force between adhesion and gravity is the capillary force. This force per unit area is the
capillary pressure.
Capillary pressure is defined as the pressure difference between two fluid phases (e.g., oil and water)
at the same point in the reservoir. It is a measure of the rock-fluid adhesion and fluid-fluid interfacial
tension forces that act to hold one fluid phase (e.g., water) at a particular location in the reservoir (e.g.,
above the oil-water contact) against the force of gravity. Capillary pressure is a complex function of the
nature of the contained fluids, the saturation of the fluid phases, the wettability of the rock, and the
pore size distribution of the rock.
As we might surmise from observations of the capillary tube illustrated in the Figure above, there is a
relationship between capillary pressure, Pc , and the interfacial tension between the two fluids (in this
example -water and air).

where
Pc = capillary pressure
wn = wetting/non-wetting phase interfacial tension
r = radius of the tube
 = angle of contact between the solid surface and liquid

Capillary Fluid Rise


An alternative way to express capillary pressure is in terms of height above a free water surface.
Capillary pressure is equal to the product of the height above the free water surface and the density
difference between the two fluid phases at reservoir conditions. In a reservoir, the relationship between
water saturation and height above an oil-water or gas-water contact is, of course, strongly dependent
on the rock pore system, as well as on the wettability and interfacial tension properties of the rock-fluid
system.
We will again use Figure 3 of the capillary tube to illustrate fluid height. The capillary tube of radius r
will support a column of water of height h. If the density of the air is a and the density of the water is
w, then the pressure differential at the air-water contact is simply (w -a)h. This pressure differential
acting across the cross-sectional area of the capillary is exactly counterbalanced by the surface
tension, T, of the water film acting around the inner circumference of the capillary tube. If the contact
angle is  at the interface between the water and the glass face of the capillary tube, then at
equilibrium we have:
2rT cos = (w -a)h • r2
Force = Pressure • Area
By simplifying and rearranging this expression we see that height is expressed as:

As the capillary tube radius (r ) decreases, the height (h) of the water column increases; therefore, the
height of fluid rise above a free water surface in a capillary tube is inversely proportional to the radius
of the tube. We can draw an analogy between the capillary tube radius and the radii of pore throats in
the rock matrix. In the above example, we can correlate the air to oil, water with water, and the tube
with pore throats.
Translating this laboratory observation in terms of reservoir conditions, we can see that water can be
drawn up into what would otherwise be a 100% oil column by the capillary effect resulting from small
pores in the rock system. Thus the maximum height, h, to which water can be raised is controlled by
the following factors:

• the surface tension, T, between the two phases (oil and water)
• the contact angle,  , between the wetting fluid (water) and the rock
• the radius of the pore throats (r)
• the density difference between phases (w -o in this case)
In summary, the capillary rise will be greater in a rock with smaller pore throats than in one with
large pore throats.

Length of Transition Zone


Given the above factors, it is easy to characterize the relative length of a transition zone in a reservoir.
Reservoirs with large pore throats and high permeability have short transition zones, and the transition
zone at a gas-oil contact will be shorter than that at an oil-water contact simply because of the inter-
phase density differences involved (Figure 4 ).
Since a pore system is made up of a variety of pore sizes and shapes, no single pore throat radius can
be assigned to a reservoir. Depending on the size and distribution of the pore throats, certain available
pore channels will raise water above the free-water level. The water saturation above the top of the
transition zone will thus be a function of porosity and pore-size distribution.
In a water-wet system, the water will wet the surface of each grain or will line the walls of the capillary
tubes. At the time oil migrates into the reservoir, the capillary pressure effects will be such that the
downward progress of oil in the reservoir is most strongly resisted in the smallest capillaries. A
particular elevation will limit the amount of oil that can be expected to fill the pores. Large-diameter
pores offer little resistance (capillary pressure, Pc, is low because pore radius, r, is large). Small-
diameter pores offer greater resistance (Pc is high because r is small). For a given reservoir, o and w
determine the pressure differential that an oil-water meniscus can support.

Capillary Pressure and Irreducible Water Saturation

Maximum oil saturation is controlled by the relative number of small and large capillaries or pore
throats. This maximum possible oil saturation, if expressed in terms of water saturation, translates into
a minimum possible water saturation, and this is referred to as the irreducible water saturation, Swi.
Shaly, silty, low-permeability rocks with their attendant small pore throats and high capillary pressures,
tend to have very high irreducible water saturations. Just the opposite is true for clean sands of high
permeability, which have low irreducible water saturations. Figure 5 illustrates this important concept
by comparing capillary pressure curves for four rock systems of different porosity and permeability.

STRUCTURAL POSITION WITHIN THE RESERVOIR


We all know that gravity segregation causes a natural stratification of reservoir fluids, with gas on top
of oil, and oil over water. In the absence of rock pores, the gas, oil, and water will form distinct layers,
with sharp contacts between each phase. In a reservoir, however, the contacts between each phase
are less distinct, as illustrated in the Figure 6: Reservoir containing oil and water.
This diagram divides the reservoir into three levels. The upper level is mainly oil; the lower level is all
water, while the middle level shows ever-increasing concentrations of water as depth increases.
Plotted on the right-hand side is a curve of water saturation, together with a plot of fluid pressure in the
pore spaces.
When a formation is above the transition zone, i.e., at irreducible water saturation, the product of  and
Sw is a constant. Variations of porosity are normal on a local scale, caused both by changes in the
depositional environment and by subsequent diagenesis. If porosity is reduced locally, then either a
greater proportion of pore throats will be small or there will be simply fewer pore throats. Either way,
the mean radius of the pore throat r will be smaller; thus Pc will be larger, and more water can be held
in the pore maintaining the constant:

 • Swi
After a zone has been analyzed on a foot-by-foot basis for porosity and water saturation, a plot of 
versus Sw reveals the presence or absence of a transition zone.
Figure 7 shows a log-log plot, where points at irreducible saturation plot along a straight line (the red
line denoting Zero Water Production in this graphic), and the points in the transition zone plot to the
right of the irreducible line.

By plotting the product of  • Swi, it is possible to predict certain production characteristics. For
points below irreducible saturation, some portion of water production is to be expected, depending on
the mobility ratio, (kwµo/koµw), for the particular fluids present. On the other hand, in a low-porosity, low-
permeability formation, we again see that surprisingly high water saturations can be tolerated without
fear of water production. Conversely, in other formations that exhibit good porosity and permeability,
even moderate values of Sw will mean that water production should be expected.

Again, since both capillary pressure and relative permeability data are a strong function of pore size
distribution and geometry (among other factors), they will, in turn, often fall into groups that correlate
with specific reservoir facies (Morgan and Gordon 1970). Two samples representing two different
reservoir facies may show very different relative permeability relationships — even though their other
reservoir properties (porosity or permeability) may be very similar. Figure 8 shows water-oil relative
permeability curves for two samples with almost identical permeabilities, taken from two different
geologic facies. The difference in the measured relative permeabilities illustrates the importance of
pore system configuration in determining fluid flow characteristics.
One important observation to be derived from evaluating capillary pressure and relative permeability
for individual geologic facies is the determination of the producing oil-water (or gas-water) contact
elevation throughout various parts of a reservoir. This elevation can vary across a reservoir, and can
be substantially different from the free water surface elevation. The free water surface should be at a
constant elevation throughout the reservoir, providing the reservoir is in a state of gravity-capillary
equilibrium with no significant flow occurring. The producing oil-water contact is often taken as the
highest elevation where 100% water is produced. This depth will not necessarily be the same at all
points in the reservoir, even under equilibrium conditions. It will strongly depend on local
capillary pressure (water saturation versus height above the free water surface) and the relative
permeability characteristics of the rock.

Capillary Pressure and Geologic Facies


A relationship between geologic facies and capillary pressure curves is illustrated graphically in
Figure 9. Three geologic facies (A, B, C) are shown, each having a different capillary pressure
versus water saturation relationship. On the right side of the graph, increasing capillary pressure has
been expressed in terms of height above an oil-water contact (the conversion is for a specific set of
assumed rock and fluid properties and is not a general correlation). Note that each of the curves
becomes asymptotic to some minimum water saturation value (the irreducible water saturation), in
spite of large increases in capillary pressure (or height above the oil-water contact).
The three different geologic facies in this graphic also have greatly contrasting permeabilities, although
their porosities are similar. Capillary pressure versus saturation data often correlate very strongly with
permeability because both properties are strongly dependent on pore throat size distribution. As
illustrated in this graphic, at a measured capillary pressure of 10 psi (equivalent to a height of almost
25 ft (7.6 m) above the oil-water contact in this example) the water saturation in Facies A has been
reduced to the irreducible water saturation of less than 10%. At the same height above the oil-water
contact, the water saturation in Facies B could be almost 40%. Note that this is above the irreducible
water saturation of about 30% for Facies B. At points in the reservoir where the shaly limestone Facies
C is present at the same height above the oil-water contact, the water saturation could be over 95%! In
fact, the capillary pressure versus water saturation curve for Facies C indicates that this facies would
probably NOT be considered as pay in this reservoir.

PUTTING IT ALL TOGETHER - GEOLOGY AND FLUIDS


We have just reviewed a number of factors which influence the Irreducible Water Saturation within a
formation. These factors include interactions between fluids and rock, as well as interactions between
different fluids. By examining porosity, relative permeability, and capillary pressure relationships, along
with rock texture and structural position, it is possible to determine whether a well having high S w
calculations will actually produce water, or instead, will produce water-free.
For example, as we move toward the top of a fining-upward sequence, the decrease in sand grain
diameter will produce a corresponding decrease in pore throat radius. This decrease in pore throat
radius is accompanied by an increase in capillary pressure, thus increasing the amount of water that
can be imbibed into the system. If we add clay or silt to this example, we can expect that microporosity
will constitute a substantial percentage of total porosity. Such a setting is bound to produce high values
of irreducible water saturations.
Both irreducible water and residual hydrocarbon saturations are strongly influenced by rock texture,
which is controlled by depositional environment. Fine-grained sediments, usually characteristic of low-
energy depositional environments, tend to have high irreducible water with high residual hydrocarbon
saturations; coarse-grained sediments, characteristic of high-energy environments, tend to have low
irreducible water saturation and low residual hydrocarbon saturation. In addition, fine-grained
sediments tend to have lower permeability than coarse-grained sediments.

These factors must all be considered together when analyzing low resistivity pay zones. Porosity,
capillarity, relative permeability, structural position and grain size will all influence the final
evaluation of irreducible water saturation in a low resistivity pay zone.

9.9 FORMATION ANISOTROPY


It is not uncharacteristic of formations to display changes in various properties as a function of
direction. For instance, geophysicists have long recognized changes in the velocity of seismic
wavefronts, and petroleum engineers see reservoirs which exhibit changes in permeability over a
certain direction. Such changes with direction are called formation anisotropy.
Just as formations display changes in seismic velocity or permeability, so can they exhibit changes in
resistivity with direction. In this section, we will explore ways in which electrical anisotropy affects
overall resistivity in a low resistivity pay zone. In a different section of this module, we will discuss a
unique logging tool that Baker Atlas has developed to address this problem.

DEFINING FORMATION ANISOTROPY


A formation is considered to be anisotropic when the magnitude of a vector measurement of a given
property changes with direction. In many cases of formation anisotropy, the vector measurement has
constant magnitude in all horizontal directions but exhibits a different magnitude when measured in the
vertical direction.
It is easy to confuse the concept of anisotropy with that of heterogeneity. Anderson, et.al (1994), offer
two important distinctions.

• First, anisotropy is defined as variation in a vectorial value with direction at one point,
whereas heterogeneity is defined by variation in vectorial or scalar values between two
or more points within a given direction.
• The second distinction is that anisotropy tends to describe mainly physical properties,
whereas heterogeneity is typically used to describe point-to-point variations in
compositions, geometries, or physical properties.

A matter of scale
In their insightful article entitled “Oilfield Anisotropy: Its Origins and Electrical Characteristics,”
Anderson, et. al (1994) make a profound observation regarding the phenomenon of anisotropy.
“Fundamental to both anisotropy and heterogeneity is the concept of scale. Whether anisotropy and
heterogeneity are perceived depends on sample size and sampling resolution. Anisotropy, for
example, can be detected only when the observing wavelength is larger than the ordering of the
elements creating the anisotropy (Helbig, K., 1994).”
So, while a crystal may be homogeneous above the molecular scale, it might be anisotropic when
measured by the larger wavelengths of electromagnetic or acoustic propagation. An assemblage of
crystals can form a homogeneous rock that is anisotropic if the crystals are aligned, or isotropic if they
are randomly packed. When different layers of homogeneous and isotropic rock are measured as a
single unit, they might be seen as heterogeneous and anisotropic (Anderson, et al, 1994). Consider
how scale might affect a laminated sand-shale formation, where maximum thickness of each layer is 1
cm.
Because anisotropy depends on the perspectives of sample size as well as sample resolution, it is
easy to see how anisotropy on the scale of meters could escape detection in sidewall core
measurements. Or how anisotropy on the scale of centimeters might escape detection by wireline-
conveyed formation testing, which has such a large wavelength that it averages out any observable
anisotropy. In the former case, the wavelength is too small; whereas in the latter case, the resolution is
not capable of fine detail.

Sedimentary Processes
Anisotropy arises from processes that take place either during or after deposition.

Anisotropy in Carbonates
For the most part, anisotropy in carbonates is controlled by fractures or diagenesis, and so tends to
take place after deposition. However, anisotropy in carbonates may also be attributed to be slight
changes in carbonate mineralogy which are contemporaneous with deposition. In turn, these variations
in mineralogy may be caused by changes in the carbonate balance of the atmosphere and water.
These changes in carbonate mineralogy can produce differences in texture, porosity and permeability.
Anisotropy in Clastics
Anderson, et. al (1994) state that the ultimate cause of deposition-related anisotropy in clastics is
gravity. Movement of clastic grains always begins with under the influence of gravity.
At the bedding scale, anisotropy has been attributed to two factors:

• Periodic layering, usually attributed to changes in sediment type, typically produces beds
of varying material or grain size.
• Ordering of grains induced by the directionality of the transporting medium. (Rajan, 1988).
Upon settling, the grains tend to align in the direction of least resistance to the movement
of the air or water.
Such alignments can create preferential rock stiffness in one direction, with weakness orthogonal to
that direction (Lynn, 1991). The azimuthal distribution of grain assemblages may also result from post-
depositional tectonic deformation that causes shortening in one direction (Giesle, in Anderson, et. al,
1994).

Diagenesis
A number of processes which can produce anisotropy occur after deposition. Anderson, et. al (1994)
cite the following examples:

• rotation of grain axes caused by compaction brought on by overburden pressure


(Manrique, 1994)
• alignment of clay platelets caused by compaction and dewatering of muds results in
anisotropy of shales
• pressure solution in carbonates can produce styolites of insoluble residue produced by
dissolved material, which can act as laterally extensive flow barriers
• burrowing in carbonates or sandstones either enhances or destroys depositional
anisotropy Anderson, et. al (1994) point out that the process of diagenesis can
significantly alter anisotropies that were previously established during deposition.

ELECTRCAL ANISOTROPY
Electrical anisotropy is said to occur when the resistivity of a formation varies with direction. This form
of anisotropy is seen as a variation between resistivity measured parallel to the formation bedding
plane (known as horizontal resistivity, Rh) versus resistivity measured perpendicular to the bedding
plane (vertical resistivity Rv).
Another article by Anderson, et. al (1997) links the cause of electrical anisotropy to the very same
factors that cause permeability anisotropy:

• formation bedding geometry, or


• grain size in homogeneous sand beds.
This same diagnosis corresponds to two important types of low-resistivity pay. The first is formed by
thin-bedded or laminated sand-shale sequences. The second is caused by sand layers of varying
grain-size distribution or sorting, which can produce variations in water saturation, morphology of the
conductive fluid, or distribution of fluid phases.
In laminated shaly sands, the thin shale layers do not alter the porosity and permeability characteristics
of the inter-bedded sands. However, in such reservoirs, the horizontal resistivity is governed by the
conductive shaly laminae of the formation. This is because the highly conductive nature of shales
suppresses the response of the conventional induction log, which measures current flow parallel to the
shale beds.
The measurement of electrical anisotropy provides a valuable tool for obtaining better estimates of
formation hydrocarbon saturation. In the sections below, we will see how knowledge of vertical
resistivity (though beyond the scope of conventional resistivity tools) and horizontal resistivity can play
a crucial role in evaluating low resistivity pays.

Vertical and Horizontal Resistivity

To properly characterize shaly sand reservoirs, it is important to understand the relationship between
resistivities and laminar shale volume. Mollison, et. al (2001) illustrate this point by modeling a 50%
net-to-gross (i.e., 50% laminar shale) formation consisting of thinly bedded sands and shales, as
shown in Figure 1 (Sensitivity of horizontal and vertical resistivity to net sand fraction and pore fluid in
a finely laminated sequence). The sands in this model had resistivities of 10 -m, and the shales
had resistivities of 1 -m.
In this figure, the vertical and horizontal resistivities, R v and Rh respectively, are plotted as a function of
the laminar shale volume Vshale. The equations used in this model for the computation of both R h and
Rv, as a function of laminar shale volume Vshale, and the anisotropy ratio are presented below:
where
Rh = horizontal resistivity,
Rv = vertical resistivity,
Rsand = sand resistivity,
Rshale = shale resistivity,
Vsand = laminar sand volume,
Vshale = laminar shale volume, and
ratio = anisotropy ratio

Model Behavior
The maximum separation between vertical and horizontal resistivity occurs at 50% laminar shale
volume, where the electrical anisotropy ratio Rv / Rh reaches its maximum value.
As laminar shale volume increases from 0% to 30%, the horizontal resistivity decreases very rapidly;
but for laminar shale volumes over 30%, the horizontal resistivity exhibits a lower sensitivity to changes
in shale volume. When laminar shale volume exceeds 30%, the horizontal resistivity, Rh, shows less
sensitivity to variations in the laminar sand-shale resistivity, thus increasing the difficulty of detecting
thin bedded, resistive sands that contain hydrocarbons.
On the other hand, vertical resistivity is a volume-weighted average of both sand and shale resistivity
components and shows a much more gradual reduction as the laminar shale volume increases.
Mollison, et. al (2001) note that the vertical resistivity, and most importantly, the resulting anisotropy
ratio are sensitive to changes in both laminar shale content and laminar shale resistivity throughout the
entire range from 0% to 100% shale volume. Therefore, the vertical resistivity measurement is more
sensitive to the presence of hydrocarbon-bearing laminated sands within the sand-shale sequence.
In Figure 2, Mollison et. al, take the model from the first Figure a step further, to illustrate limitations of
horizontal resistivity sensitivity. A shale having 1 -m resistivity is again used in this laminated
sand-shale model, and horizontal resistivity is calculated versus laminar shale volume for two laminar
sands having resistivities of 5 -m and 10 -m.
In this model, we note that a laminar shale volume of 50% is able to suppress the horizontal resistivity
in the
10 -m sand to 1.8 -m. Similarly, the horizontal resistivity in the 5-m sand is suppressed to 1.65 -
m. On the basis of horizontal resistivity alone, these two sand-shale sequences would be practically
indistinguishable at a 50% laminar shale volume, with such a minute difference of approximately 0.15
-m between horizontal resistivity readings (typical of low-contrast pays).
By contrast, the vertical resistivity shows significantly higher values over the horizontal resistivity
readings. At 50% laminar shale volume, the 10 -m sand yields a vertical resistivity reading of 5.5 -
m, while the 5 -m sand at 50% shale volume yields a vertical resistivity of 3.0 -m. In this case, we
see that vertical resistivity changes by 2.5 -m in the same sand-shale mixture.
From this model, we can draw an important distinction between the performance of the horizontal and
vertical resistivities. We see only a small difference in R h between the laminated sand-shale
sequences. Conversely, the Rv is very sensitive to sand pore-fluid content. Sensitivity remains high
even when the sand fraction drops as low as 20%. Therefore, the availability of vertical resistivity data
can greatly improve the evaluation of hydrocarbon saturation in thin-bedded or laminar sand-shale
sequences.

In a different presentation (courtesy of Baker Atlas), Figure 3 illustrates how much vertical and
horizontal resistivities (track 3) affect oil saturation calculations (track 5).

Other Instances of Electrical Anisotropy


Thinly laminated sand-shale sequences are not the only cause of electrical anisotropy. This behavior
can also be caused by other structural or sedimentary factors.

Apparent Dip
Recall from our previous discussion that electrical anisotropy depends on the direction in which the
current flows in the rock. It is based on a comparison between vertical and horizontal resistivities -
basically it is the difference between resistivity measured parallel to bedding versus resistivity
measured perpendicular to bedding.
Because electrical anisotropy depends on the direction in which electrical current flows in the rock, we
must always be cognizant of the angle at which the borehole penetrates the formation, as well as the
dip of that formation. (Unless the well is drilled perfectly vertically when it penetrates a formation, it will
encounter a formation that is tilted at an angle that is usually less than its true dip. In geological terms,
we would call this apparent dip, which is defined as the angle that a plane -or wellbore -makes with the
horizontal, measured in any randomly oriented section rather than perpendicular to strike.)
Anderson, et. al (1994) notes that the extent to which anisotropy affects the resistivity reading will
depend on the angle between the borehole and the formation. When a vertical well drills through a flat
(0 dip) formation or bedding plane, the conventional induction tool will measure horizontal resistivity.
However, measures of resistivity show that the current flowing normal to bedding (vertical resistivity) is
always at least as much as, if not more than, horizontal resistivity. And as wellbore deviation
increases, the contribution from the vertical resistivity becomes stronger (Anderson, et. al, 1994).

Differences Between the Oil Leg and Water Leg of a Reservoir


Boyd, et. al, (1995) cite a study by Allen, Klein, and Martin (1995), which reports that water-wet
formations with large variability in grain size can be highly anisotropic in the oil leg while being isotropic
in the water leg. They attribute the electrical anisotropy to grain-size variations, which affect irreducible
water saturation.

In Figure 4 (Effect of hydrocarbon saturation on electrical anisotropy), we see the results of the above-
cited experiment by Allen, Klein, and Martin (1995), which shows how hydrocarbon saturation
influences electrical anisotropy.
In Figure 5 (Effect of water saturation on electrical anisotropy), the influence of water saturation is
shown.
With these two graphs, we can compare plots of perpendicular resistivity versus parallel resistivity
measured within a reservoir oil column and a water column. This resistivity model shows that resistivity
readings in the oil column arc sharply to the right, while resistivity values in the water column only take
a slight bend by comparison. The position of data along the oil column arc indicates the sand/shale
content of the formation. In the oil column, we see strong anisotropy being developed as saturation of
the more resistive oil increases. In the water leg, however, the same formation might display little or no
anisotropy.

Differences in Grain Size


Electrical anisotropy can also come from changes in grain size that affect saturation distribution, as
well as from shaliness. Significant resistivity anisotropies (10:1 or greater) have been found in what
were previously assumed to be clean, homogeneous sands. In these formations, the anisotropy is
thought to be due to contributions from variations in grain size, irreducible water saturation or
permeability.
Similarly, electrical anisotropy in the invaded zone of the formation can be different from that in the
uninvaded zone. (Anderson, 1994) In light of our previous discussions on irreducible water saturation,
electrical anisotropy within these zones should come as no surprise.
9.10 CONVENTIONAL LOGGING TOOLS
As we uncover more factors that contribute to the phenomenon of Low Resistivity Pay, we must place
some of the responsibility on the logging tool itself. In this section, we shall see that the way in which
each logging tool samples the formation, and the measurement taken by each tool, will influence the
outcome of the well log evaluation. From this discussion, we may conclude that, while conventional
logging tools are able to obtain perfectly valid data, this data nonetheless can cause water saturation
estimations to be too high.
MEASUREMENT AVERAGING
Though it would be nice to read true resistivity of the smallest lamination, most logging tools must
strike a compromise somewhere between the vertical measurement and the horizontal measurement.
In logging parlance, we would say that log measurements are a compromise between vertical
resolution and the horizontal depth of investigation. This can cause problems when we log thin or
laminated sands. We would ultimately like to see deep enough into the formation to get past the mud
invasion and measure of true resistivity. However, we would like that measurement to be no thicker
than the thinnest bed or lamination. This goal may not always be realized with conventional resistivity
tools, as thin beds or laminated zones may be thinner than the vertical resolution of the tools.
A common characteristic of conventional induction tool performance is a decrease in vertical resolution
as the horizontal depth of investigation increases. This means that the thinnest interval that a
conventional induction tool can measure is taken by the shallow reading curve, close to the wellbore.
This thin measurement grows thicker as it extends into the formation. Stated another way, the true
resistivity measurement of the deep curve is not able to read intervals that are as thin as those seen by
the shallow curve; however, the shallow curve does not measure as deep, and does not provide true
resistivity.
This all leads to an important point about the way in which logging tools operate. A typical induction
tool might have a radial depth of investigation that ranges from 5 ft (1.5 m) for the deep induction
curve, down to 1.25 ft (38 cm) for the shallow curve. This tool’s vertical resolution would range from 7-
8 ft (2 m) for the deep induction log, down to 2.5 ft (76 cm) for the shallow log. How do these fine tool
capabilities measure against your own perception of a single clay lamination? The way in which an
induction tool with a 2.5 foot vertical resolution will react to a thinly laminated pay zone should be
obvious. The typical induction tool will simply average the readings -turning all of those fine laminations
of sand and shale into a gross average of the smallest thickness that the tool can measure.
CLAY’S AFFECT ON A CONVENTIONAL LOG SUITE
Earlier, we stated that a leading cause of low resistivity pay sands was its clay content. In this section
we will review, in broad terms, ways in which clay affects the typical suite of conventional logging tools.
(A detailed discussion of these tools can be found elsewhere in the Formation Evaluation Series.) In
summary, we can say that clay causes resistivity logs to read too low, while generally causing porosity
logs to read too high. We briefly describe the reactions of broad groups of tools below.
Spontaneous Potential Log
The SP curve is a good indicator of grain size, from which we can draw shale-and sand-baselines.
Shaliness will cause the SP curve to deflect to the right, away from the clean sand line. This deflection
between the clean sand line and the shale base line provides a way to measure V sh. The magnitude of
SP deflection is a function of cation exchange capacity, formation water salinity, mud filtrate salinity,
and formation water saturation. Clays have a strong influence on CEC. If we apply our earlier
discussion of cation exchange capacity to the SP log response, we can generalize by saying that clays
which exhibit high CEC will produce a decrease in SP.Be aware, however, that a number of other
borehole phenomena may affect SP response, particularly hydrocarbon suppression of the SP.
Gamma Ray Log
The gamma ray tool measures formation radioactivity. If the clay mixture does not vary, and if no other
radioactive minerals are present, then the gamma ray log can be linearly correlated to determine the
amount of clay contained in the formation. We can characterize GR log response by saying that, in
general, the GR will deflect to the right in most clays and shales. It is important to remember that not
all shales will be radioactive, and that some sands will also be radioactive.
Resistivity Log
Our previous discussion on clay mineralogy showed that resistivity is lowered as a result of the surface
charge of clay and by its bound water content. These factors, in turn, depend on the surface area and
CEC of the clay, in addition to formation compaction.
Recall that Montmorillonite and Illite have higher cation exchange capacities than those of Chlorite and
Kaolinite. Therefore, shaly sands with Montmorillonite and Illite exhibit higher conductivity (i.e. lower
resistivity) than shaly sands with Chlorite and Kaolinite.
It is also worth noting that while thinly laminated sequences cause logging tools to measure a sort of
average, for resistivity logging tools this averaging process is not linear, but takes a complex form that
is further complicated by the apparent dip of the formation with respect to the borehole.

Acoustic Log
The acoustic logging tool records interval transit times of a compressional sound wave. Clays are
characterized by low acoustic velocities, and hence cause the acoustic log to record higher interval
transit times. When the transit time of the reservoir’s matrix is subtracted from the high transit time log
reading in the clay-sand, the resulting acoustic porosity will calculate to be higher than actual porosity
through that zone.

Neutron Log
This tool derives its porosity measurements from the concentration of hydrogen ions in the formation.
In a clean, shale-free sand, the hydrogen ions measured by the tool would reflect water or
hydrocarbons occupying the pore spaces of the formation. This count is increased, however, by the
presence of clay in the formation. Because hydrogen is found in the crystal structure of clays, and
within the irreducible water that is bound to the clay, the tool will register an increase in porosity. Thus,
in a sand-clay mixture, a portion of the total porosity recorded by the neutron log will consist of non-
effective porosity occupied by bound water, a portion will reflect the atomic constituents of the clay
mineral, with the remainder being porosity occupied by water or hydrocarbons. So we can summarize
by saying that clay in a sandstone will cause neutron porosity calculations to be higher than they are in
actuality.
We can characterize neutron porosity with relation to specific clay minerals. It has been found that
Chlorite and Kaolinite have much greater neutron porosities than Montmorillonite and Illite. This means
that shaly sands containing Chlorite or Kaolinite will exhibit higher porosities when compared to shaly
sands composed predominantly of Montmorillonite and Illite.

Density Log
Porosity is derived from the density tool’s measurement of formation bulk density. When calculating
density porosity, if the density of clay is less than that of the reservoir matrix, the porosity will be higher
than the actual formation porosity. When clay density is the same as the reservoir matrix density, the
density log will measure true effective porosity. When clay density is greater or less than the
sandstone matrix density of 2.65 g/cc, the porosity measured by the density log can be less than or
greater than the porosity calculated in a clay-free sandstone.
Asquith (1990) notes that calculated density porosity of a shaly sand is affected by Montmorillonite and
Chlorite. The higher density of Chlorite (2.8 gm/cc) versus quartz (2.65 gm/cc) results in a lower
calculated density porosity. Conversely, the lower density of Montmorillonite (2.45 gm/cc) versus
quartz (2.65 gm/cc) results in a higher calculated density porosity.

IMPROVED TOOL DESIGN


One way to resolve the resistivity of thin beds would be to develop resistivity tools with better
vertical resolution and greater depth of investigation. A completely different approach would
depend on some measurement other than resistivity. The leading service companies now offer
advanced logging tools that have improved resolution capabilities. We will discuss some of these
tools in a later section.
9.11 KEY VARIABLES IN BASIC LOG ANALYSIS
In 1942, Gus E. Archie published his water saturation equation to show how formation resistivity was
related to fluid resistivity, saturation and porosity in the rock. This basic equation serves as the
foundation of modern Petrophysics, from which a number of log evaluation models have sprung. For
years, Geologists and Petrophysicists have used the traditional Archie equation to relate hydrocarbon
production to high formation resistivity and low water saturation values. Unfortunately, many failed to
see that the obverse relationship did not necessarily exist, and that low resistivity values might also
point to highly productive intervals.
The Archie equation is still useful for calculating pay in thick, shale-free, sand reservoirs. (The trick is
in finding such reservoirs - not in evaluating them!) Because most fields are not able to offer thick or
shale-free sands, the log analyst is usually faced with evaluating complex reservoirs. So when the
Archie model is applied to logs from thin and shaly sands, the water saturation calculations tend to be
more pessimistic than core data or subsequent production data would indicate.
The only thing worse than walking away from a pay zone would be setting pipe on a truly wet zone.
Therefore, in the following discussion of the Archie equation, we will describe key variables and their
impact on Sw calculations. We will show how the manipulation of these variables can change the
outcome of the Sw prediction, and will show why the well log must be calibrated against the core
analysis.

FORMATION FACTOR, F
Archie combined three measurable observations into one equation. By saturating a rock sample with
solutions of different salinities, he found that the resistivity of the water-saturated "wet" rock (Ro) was
related to the resistivity of the saturating water (Rw) through the relation:
Ro = FRw where: F is the formation factor.
This formation factor is correlated to rock porosity according to relation:

where:
a is a constant known as the tortuosity factor,
( tortuosity relates to the complex path that the fluid must travel through permeable rock matrix), and
m is known as the cementation exponent, whose value varies with grain size, grain size distribution,
and tortuosity of the rock.
(Tortuosity and cementation are related: as tortuosity increases, so does the value of the cementation
exponent.)
Common values for a and m are given in Table 1 (After Asquith, 1982):

ARCHIE EQUATION
Rocks having less than 100% water saturation were found to obey the rule:

where
n is the saturation exponent.
By substituting the definition of Ro into the above equation, we obtain Archie’s equation:

VALUES OF a, m, AND n

The values of a and m can be set to the generally accepted values listed in the table above when they
are unknown. The table shows that values for a range from 0.62 to 2.45, while values for m range from
1.08 to 2.15. According to Asquith (1982), values of n range from 1.8 to 2.5, but most often equals 2.0.
(However, for serious petrophysical analysis, the values of a, m, and n should be determined by core
analysis.)
Generally accepted values for these constants and formation-factor-to-porosity relationships are
For sandstones:
a = 0.81 m = 2 n = 2, or
a = 0.62 m = 2.15 n = 2
For carbonates:
a=1 m=2 n=2
Thus, three commonly used versions of Archie's equation are:
for sandstones -

or

for carbonates -

The analyst should avoid the mind set that locks the conventional values for a, m, and n into any and
all log analysis evaluations. Crossplot methods, such as the Pickett plot, allow one or another of these
parameters to be estimated; however, there is no substitute for rigorous core analysis to pin down the
precise values required for each and every reservoir unit. While the saturation exponent n is usually
assumed to be 2, its exact value may vary depending on rock wettability. Oil-wet systems usually
exhibit an n greater than 2. There are certain water-wet systems that may show an n of less than 2.
For a fine perspective on Archie’s equation and the evolution of a, m, and n values, see the paper by
John H. Doveton (2001), listed in the reference section.

Determining the Value of the Saturation Exponent n


To find the saturation exponent n, a core sample is prepared at a number of different water
saturations, and for each saturation the ratio of Ro to Rt is measured. By taking the logarithm of
Archie's equation, it is transposed into:

Therefore, a plot of log (Sw) versus log (Ro/Rt) should produce a straight line of slope n. Figure 1
illustrates the method.

It is worthwhile to explore the sensitivity of Sw to the value of n. Suppose a calculation of Sw is made


using n = 2 and the resulting value of Sw is 0.5, or 50%. Then we can deduce

hence

Having found that F Rw/Rt is 0.25, we can now play a "what if" game and see the results of raising n to,
say, 2.2:

Lowering n to, say, 1.7 gives


Sw = (0.25)1/1.7 = 0.44 or 44%
In summary, other factors remaining the same, raising n raises Sw; likewise, lowering n lowers Sw.

Determination of a and m Values


Each rock type has its own distinct characteristic formation-factor-to-porosity relationship.
Measurements to determine this relationship require a range of core samples over different porosities.
Then at each porosity, a measurement is made of Ro, the resistivity of the rock at 100% water
saturation. If the value of Rw is known, the formation factor (F) can be calculated by using the definition
F = Ro/Rw. Note that the solution used to saturate the core should be of the same NaCl concentration
as that of the connate water found in the formation.
By taking the logarithm of both sides of the equation that relates formation factor to porosity, we obtain
this result:

Thus, a plot of log (F) versus log () should give a line of slope -m and an intercept a that is found at 
= 1 (100% porosity, or 0 log ).
Figure 2 illustrates this method.

Note that the calculation of water saturation is very sensitive to m as porosity decreases, and use of an
incorrect m value will yield substantial water saturation errors.

RELATING a, m, AND n, TO CALCULATIONS OF SW


Table 2: Comparison of F, a, m, n, and , to Sw, which accompanies this discussion, was created to
illustrate the influences that a, m, n, and , exert on calculations of Formation Factor, and ultimately,
on Water Saturation.

This table is basically split into 3 different porosity sections, and shows F and Sw for different
combinations of a, m, and n. By holding certain of the variables constant, we can see how the other
variable affects the calculation. Thus, if we hold n constant at 2.5, and if we hold  at 20 %, with a held
at 0.81, then we can see that by changing m from 2.00 to 2.15, we will cause Sw to change from 0.46
to 0.50.
This table helps us to prove that the Sw calculation can be manipulated not just by the log readings, but
by values of a, m, and n, as well. It is important that you understand the influence these values have
on water saturation calculations. From the table, we may conclude that if all other variables are held
constant, then:

• By raising the value of , the value of F decreases, as well as that of Sw


• By raising the value of a, the value of F increases, as well as that of Sw
• By raising the value of m, the value of F increases, as well as that of Sw
• By raising the value of n, the value of F increases, as well as that of Sw
Thus, while you may be accustomed to plugging in certain values of a, m, and n, according to the local
standard, it is important to consider whether core analysis through your pay zone would cause these
values of a, m, and n to change. If core analysis showed a deviation from standard values of a, m, and
n, you should question how those deviations would affect your log analysis. For this reason, it should
be emphasized that it is important to calibrate your log analysis results by integrating them with
your core analysis.
9.12 LOG CALCULATIONS
In previous sections we have discussed factors which lead to low resistivity in pay zones. From this,
we can conclude that the leading cause of low resistivity pay is WATER, and the CLAYS that attract
and hold the water in place. We also know that when low resistivity is caused by bound water, the low
resistivity should not automatically condemn the well to the fate of a water producer.
In this section, we will describe log evaluation methods that have been devised to compensate for the
presence of clay in order to estimate a realistic value of effective water saturation.

QUICK OVERVIEW OF LOG CORRECTIONS


The primary issue that we seek to quantify through log analysis is simply: “Will this well produce more
water than hydrocarbons?” We answer this question, not by calculating water saturation, but by
calculating effective water saturation. Before discussing specific log evaluation techniques, we will
quickly outline the overall process and describe the general concepts behind those log evaluation
techniques.
Basically, we correct for the effects of clay in both porosity and water saturation calculations through
the following steps:
1. Start by determining the volume of clay in the formation (Vcl).
2. Next, use clay volume to correct total porosity values (t) and thus obtain effective porosity (e).
3. Use clay volume and effective porosity to obtain effective water saturation (S we).
To account for the effects that clay has on resistivity, porosity, and water saturations, we must
determine how much clay is present in the formation. To determine the relative percentage of water
that the well will produce, we must calculate effective water saturation (Swe), to account for clay content
and its associated bound water. The calculation of effective water saturation requires that we
incorporate values of effective porosity into the saturation equation, since it is only the effective
porosity which will be capable of transmitting reservoir fluids. To do so, the porosity values must be
corrected for the presence of clay.
Remember the effect that clay has on logging tools:

• Clay causes ACOUSTIC porosity values to read too high.


• Clay causes NEUTRON porosity values to read too high
• Clay causes DENSITY porosity values to read too high (usually)
• Clay causes RESISTIVITY values to read too low.
Specific log evaluation techniques are described in the sections below.

DETERMINING CLAY VOLUME


Clay volumes are derived from spontaneous potential, gamma-ray, or combined neutron-density logs.

• The gamma ray log detects clay by an increase in the radiation associated with the presence of
clay.
• The SP log detects clay by the decrease in SP deflection caused by the loss of ionic
permeability in the shaly sections.
• The neutron-density log detects the presence of clay by the increase in neutron porosity over
density porosity. This significant increase in neutron porosity is the result of the high hydrogen-ion
concentration in clays.
The clay volume (Vcl) estimations do not account for such variables as clay mineral composition and
distribution.
In this section, we will outline the steps for calculating clay volume based on different types of logs.
Clay Volume from the Gamma Ray Log
The most preferred way to derive Vcl is by using the relative deflection of the gamma ray as a shale
volume indicator.
The simplest procedure is to rescale the gamma ray between its minimum and maximum values from
0% to 100% shale (where the minimum and maximum deflections are obtained only within one
consistent geologic zone containing both sands and shales). The gamma ray index is used to linearly
rescale the GR between GRmin and GRmax
1. Calculate the Gamma Ray Index (IGR )

Where:
IGR = Gamma Ray Index
GRlog = gamma ray reading from shaly sand
GRmin = gamma ray minimum from clean sand
GRlog = gamma ray maximum from shale
2. Next, determine if the sand is consolidated or unconsolidated. (Most Tertiary sands are assumed to be
unconsolidated.)

3. Use the IGR value to calculate clay volume based on whether the sand is consolidated or not.

(for consolidated sands)

(for unconsolidated sands)

Alternatives for Estimating Shale Content from Gamma Ray Logs


A number of studies indicate that the above method does not necessarily offer the best approach, so
alternative relationships have been proposed. The alternative relationships can be stated in terms of
IGR as follows:

Relationship Equation

Linear Vshale = IGR

Clavier Vshale = 1.7 -(3.38 -(IGR + .7)2)1/2

Stieber

Bateman Vshale = IGR ( IGR + GR factor )

In the Bateman equation, the GR factor is a number chosen to force the result to imitate the behavior
of either the Clavier or the Stieber relationship. Figure 1 illustrates the difference between these
alternative relationships.
Bateman (1985) reminds us not to assume that all clay minerals are radioactive. For example, since
Kaolinite contains no potassium, it would appear as sand if only the gamma ray log was used to
determine shale content. On the other hand, when a gamma ray tool measures a sandstone containing
potassium feldspars, the log will read higher, possibly leading to the incorrect conclusion that the log
was running through shale instead of sandstone.

SP Clay Log Volume Calculations


Clay volume can also be calculated from the Spontaneous Potential log:

where:
Vcl = volume of clay
PSP = pseudo-static spontaneous potential (SP reading in shaly sand)
SSP = static spontaneous potential (SP in thick clean sand)

Neutron-Density Clay Volume Equations


The combination of neutron and density curves provides another method for estimating V cl.
where:
Vcl = volume of clay
n = neutron porosity in shaly sand
d = density porosity in shaly sand
nsh = neutron porosity of adjacent shale
dsh = density porosity of adjacent shale

The Lowest Clay Volume Estimate


Volume of clay can be determined by using all three methods described above, after which the lowest
estimate should be applied to shale corrections for porosity and saturation formulas. Doveton (2001)
observes that “Estimates of Vsh calculated from different logs differ from one another to varying
degrees, and the lowest estimate is usually taken to be the most valid. This decision is made on the
grounds that, if anything, a shale indicator will tend to overestimate the shale content. This is a
reasonable thesis, because accessory components other than shale will generally result in an
increased apparent shaliness as recorded by the logging tools.”
However, if the shaly sand is a gas sand, then the clay volume (V cl) should not be determined from the
neutron-density log. Gas may not evenly affect the neutron or density curves, and so, in this situation,
the gamma ray log and SP methods are compared to find the lowest value for volume of clay (V cl).

USING CLAY VOLUME TO CORRECT POROSITY LOGS


The presence of clay complicates the way in which we evaluate porosity within the reservoir. Though
the layer of closely bound surface water on the clay particle can represent a substantial amount of
porosity, this porosity cannot be counted as part of a potential reservoir for hydrocarbons. So, while a
shaly formation may exhibit a high total porosity, it may only have a low effective porosity as a potential
hydrocarbon reservoir.
After a volume of clay (Vcl) value is determined, this value can be used to correct the porosity log for
the effect of clay. Formulas for correcting the acoustic, density, and combination neutron-density logs
for volume of clay (Vcl) are listed below:

Acoustic log (Dresser Atlas, 1979)

where:
e = effective porosity (acoustic log corrected for clay)
tlog = interval transit time of shaly formation
tma = interval transit time of the formation matrix
tfl = interval transit time of fluid (either 189 for fresh mud, or 185 for salt mud)
tsh = interval transit time of adjacent shale
Vcl = volume of clay

Density log (Dresser Atlas, 1979)

where:
Vcl = volume of clay
e = effective porosity (density log corrected for clay)
ma = density of the formation's clay-free matrix
b = bulk density of shaly formation
fl = density of fluid (1.0 for fresh mud and 1.1 for salt mud)
sh = density of adjacent shale

Combination Neutron-Density log (Dewan, 1983)


nc = n -(Vcl x nsh)
dc = d -(Vcl x dsh)
e = (nc + dc) / 2 for oil

e = [(nc2 x dc2) / 2]1/2 for gas


where:
e = effective porosity (neutron and density logs corrected for clay)
n = neutron porosity of shaly formation
d = density porosity of shaly formation
Vcl = volume of clay
nc = neutron porosity corrected for clay
dc = density porosity corrected for clay
nsh = neutron porosity of shale
dsh = density porosity of shale

Thomas-Stieber Method (1975)


Thomas and Stieber developed a model to quantitatively determine shale content and distribution as
an alternative to the linear (IGR) method of correction. This model relates variations in gamma ray
response to the concentration and distribution of shale. When this information is later combined with
porosity data, it is possible to determine shale configuration, sand fraction, and sand porosity.
It should be noted that this model works on the assumption that shaly sands and shales are
mineralogically similar to the shales above and below the sand. Thomas and Stieber (1975) assume
that “both facies are derived from the same source material, carried by the same river, and emptied
into the same basin. Of course, this will not be true for the diagenetic alteration of feldspars into clay
within the sand stratum.”
This model was intended for Tertiary sand-shale intervals, but was not designed for complicated
mineralogies involving carbonates.

GR Behavior versus Distribution


Thomas and Stieber start out by defining the relation between GR count rates and shale fraction. The
shale models are divided into 3 distinct groups (based on distribution), and are defined by equations
listed below.

Dispersed Shale
For the dispersed mode, Thomas and Stieber begin by developing a formula to show that the observed
count rate is proportional to the amount of shale added to the pore system. According to their notation
below, the count rate equals the count rate of sand, plus the count rate of shale times the fraction of
sand occupied by the shale:
R = Ra + Xb Rb
Where
R = Count rate
Ra = Count rate of a solid block of sand material
Xb = Fraction of total rock volume occupied by dispersed shale
Rb = Count rate in shale
Applying this count rate to the gamma ray log, Thomas and Stieber relate to , which correlates to
porosity in terms of Xb, the bulk volume fraction occupied by shale.
We can express the value of  in a sand stratum in terms of gamma ray readings:
or

the measure of sand radioactivity () is defined as:

For dispersed shales, the relation between  and Vsh then becomes

which can be expressed as:

The minimum porosity available to the dispersed model occurs when shale completely fills all the
original pore volume of the sand, or when xb = a
Minimum dispersed porosity equals a • b
Where
a = porosity of sand
b = porosity of dispersed shale
When xb > a, then the relation between  and Vsh then becomes  = 1 -xb = xa

Laminated Model
In this model, shale is added to the sand stratum when a corresponding sand volume is filled with
shale.

Thus, we see from this equation that the count rate in the laminated model is actually the sand fraction.

Structural Model
In this model, the porosity increases with the amount of shale porosity that is added in place of the
solid sand grain.
Here, Thomas and Stieber use the prime ( ‘ ) symbol to denote a change in counting yield due to a
change in geometry in the material measured. In the structural model, the relationship between count
rate and Vshale is identical to that for the laminated model.

Correlating Count Rates to Porosity


After Thomas and Stieber defined the relation between  (count rates in pure shales or clean sand)
and the shale fraction, they next constructed a theoretical -porosity relationship based on the 3 modes
of shale distribution.

Dispersed Shale:
Thomas and Stieber defined 2 relationships for dispersed shale:

• Starting with pure sand, they next added some amount of shale to that volume. (However, the
amount of shale could not be so much as to completely fill all of the pore space in the sand.) Then
Thomas and Stieber defined dispersed shale porosity as:

In this case, where Vshale < sand, the pore space is decreased by the amount of solid shale added
(exclusive of the pore space contained in the shale).
When Vshale = 0, then maximum dis = sand, and
when Vshale = sand , then minimum dis = sand shale
The relationship between count rate and porosity is:

where
dis = dispersed porosity
a = porosity of sand
b = porosity of shale
 = count rate
 = radioactivity of sand
• At the other end of the spectrum, Thomas and Stieber began with pure shale, and added some
amount of individual sand grains (where no single grain has porosity), then defined dispersed shale
porosity:

In this case, total porosity is reduced by the amount of sand grains added:
(Vshale = sand)
and the relationship between count rate and porosity is:

when  = 0, then dis = shale

Laminated Shale
Thomas and Stieber replaced both the porosity and the solid rock fraction of the sand by shale porosity
and the solid rock fraction of the shale.

The relationship between count rate and porosity is:

When  = 1, then lam = sand, and when  = 0, then lam = shale.


Structural Shale
In this model, the sand porosity is not changed, but solid rock is replaced by an equivalent bulk volume
of shale porosity.

The relationship between count rate and porosity is:

The maximum amount of structural shale which can be added equals the grain volume of the sand, or
Vmax = 1-min = 1-sand.
For structural modes, the count rate is related to porosity and radioactivity of the sand by the equation:

WATER SATURATION CALCULATIONS


The Society of Professional Well Log Analysts has published many water saturation calculations over
the years since Archie developed his well-known equation.
A sample of the wide variety of shaly sandstone water saturation equations
-from Doveton, 2001

Authors Formula
Poupon, et al.
(1954)

Hossin
(1960)

Simandoux
(1963)

Waxman and Smits


(1968)

Bardon and Pied -Modified Simandoux


(1969)

Poupon and Leveaux


(1971)

Schlumberger
(1972)

Clavier, et al.
(1977)
Juhasz
(1981)

As Doveton (2001) observes, “the extension of Archie’s empirical model to shaly sandstones requires
the evaluation of more elaborate (and frequently contentious) alternative models.”
The techniques listed below offer just a sampling of the numerous equations that now exist. They
follow the strategy that we previously outlined above, they are relatively simple to use and (except for
the Waxman-Smits and -by way of extension -the Juhasz equations) they require only basic log data -
just the basic suite of logs must be obtained, consisting of Resistivity, SP / Gamma Ray, Acoustic, or
Neutron-Density.
These techniques account for clay in the formation. When other contributors, such as heavy minerals,
volcanics, or other exotics are thought to cause the low resistivity, then more extensive suites of logs
(such as the Nuclear Magnetic Resonance Log, Array Induction Log, or 3D Multicomponent Induction
Log) should be used to investigate the problem, along with additional coring or testing. Furthermore,
certain techniques have been developed for specific areas of the world, and are known to work quite
well in those areas. In such cases, local knowledge should guide the selection of the log evaluation
technique. Go with what works!

Dispersed Clay method (Alger et al.)


Asquith (1990) states that the acoustic log detects the dispersed clay in the pore water as a slurry, and
provides a porosity value that is equal to the sum of the volumetric fraction (total porosity  t). The
density log detects only water-filled porosity (Dewan, 1983). The fraction of the clean-sand
intergranular pore space occupied by clay is called q.
Where:

q = (s - d) / s
s= acoustic porosity uncorrected for clay

d = density porosity uncorrected for clay

Effective water saturation equation:

where:
Swe = effective (clay-corrected) water saturation
s = acoustic porosity uncorrected for clay
Rw = formation water resistivity
q = (s - d) / s
Rt = deep formation resistivity
An important aspect of the Dispersed Clay method is that the equation does not require a value for
shale resistivity (Rsh) or volume of clay (Vcl), because the shaly factor (q) is determined within the shaly
sand.

Effective porosity equation:


where:
e= effective porosity corrected for clay
ma = matrix density
b = bulk density from shaly sand
fl= fluid density
sh = shale density
Vcl = volume of clay
The Dispersed Clay method may not be as reliable in gas sands, because density porosity may be
greater than acoustic porosity (remember that the q equation assumes s > d )

Simandoux Method
Water saturation equations that require a shale resistivity (Rsh) component will provide higher water
saturations than those provided by the standard Archie equation when resistivity of the adjacent shale
is greater than the resistivity of the shaly sand. Patchett and Herrick (1982) state that, in general, the
Vsh/Rsh term is appropriate in laminated shaly sands only. Therefore, whenever it has been established
that the shale distribution is laminated, it is advisable to use either the Simandoux method or the Dual
Water method (described later).
The reason for restricting the value of Rsh to laminated shaly sands (unless Rcl is known) is that clays
in the shaly laminated sand and the adjacent shale are both depositional in origin (allogenic), and thus
may differ significantly in resistivity.

where:
Swe = effective (clay-corrected) water saturation
e= effective porosity corrected for clay
Rsh = resistivity of adjacent shale
Rw = formation water resistivity
Rt = deep formation resistivity
Vcl = volume of clay
C = 0.40 for sandstones and 0.45 for carbonates
The neutron and density porosities are corrected for clay in order to obtain e, using the method
described above by Dewan (1983).

Modified Simandoux Method


The Simandoux equation was later modified in 1969 by Bardon and Pied.

where:
Sw = water saturation
F = Formation factor
Rw = formation water resistivity
Rt = true resistivity
Rsh = resistivity of adjacent shale
Vsh = volume of shale

Fertl Method
This method, named after Walter Fertl of Dresser Atlas, can be for dispersed clays, or when the clay or
shale distribution is unknown, since this equation does not require a value for shale resistivity. This
equation does not call for a shale resistivity value, and thus avoids the risk of plugging in an R sh value
from an adjacent shale bed that might not represent actual shale resistivity within the shaly sand.
(Asquith notes that the difference in resistivities between dispersed clay in the reservoir and the
resistivity of the adjacent shale is especially pronounced when resistivity of the adjacent shale is
greater than the resistivity of the shaly sands.)

where:
Swe = effective (clay-corrected) water saturation
d = density porosity uncorrected for clay
Rw = formation water resistivity
Rt = deep formation resistivity
Vcl = volume of clay
a = 0.25 in the Gulf Coast or 0.35 in the Rocky Mountains
The effective porosity (e ) is derived from neutron and density logs that have been corrected for clay
using the method by Dewan (1983) previously described above.

Waxman-Smits Equation
We have previously discussed the way in which crystal surfaces absorb water using ion exchange
sites. And we recognize that different materials have different CEC values. (Quartz in the form of
sandstone has practically none, but illite and montmorillonite, because of their high specific surface
areas, have high CEC values.) Thus, we see that CEC is a mechanism by which clay can suppress
rock resistivity, and yield lower values of m and n (the cementation exponent and the saturation
exponent, respectively). By ignoring the cation exchange capacity, we run the risk of estimating water
saturations that are erroneously high. This may cause hydrocarbon productive zones to be overlooked,
or at best, result in low oil-in-place estimates.
A water saturation equation was developed by Waxman and Smits (1968) to express S w in shaly
sands as a function of the cation exchange capacity of the clay disseminated in the sand:

Where:

Where:
Sw = formation water saturation
Rw = formation water resistivity
Rt = true formation resistivity
F* = formation resistivity factor independent of clay conductivity = a*/m*
F = formation resistivity factor
n* = saturation exponent independent of clay conductivity (slope of RI versus S w plot)
B = specific counterion activity, 1/ohm-m/equiv/liter
Qv = quantity of cation exchangeable clay present, meq/ml of pore space
CEC = cation exchange capacity, meq/100 gm
ma = grain density of rock matrix, g/cm3
a = tortuosity coefficient (intercept on F versus  plot)
m* = cementation exponent (slope of F* vs  plot)
 = measured porosity, fraction
The Waxman-Smits equation requires an iterative, trial and error solution, because the water
saturation (Sw) appears on both sides of the equation.
This method, though based on sound laboratory measurement and good petrophysical principles,
lacks one vital factor for day-to-day application: the ability to derive CEC values from conventional log
measurements. Correlations for a given field can be made between core measurements of CEC and
other log measurements (GR, neutron, or density). However, the cores must be measured first -there
is no practical means of using this method if only conventional well log data are available.
In a wry perspective on log analysis, Doveton (2001) builds a case which shows that the Waxman-
Smits approach is not without its shortcomings. Because shale conductivity is a function of cation
exchange capacities of the various types of clay minerals and their abundances, shale conductivity is
influenced more by surface-area of clay minerals than by the volume of clay minerals. Experimental
data confirms that fine-grained clays exhibit higher exchange capacities than coarser grained forms of
the same volume. Doveton then notes that all shale indicators estimate the volume of the shale
component, but are not able to provide an explicit assessment of grain size or clay mineralogical
variation. Thus, without a tool that measures cation exchange capacity directly, it is difficult to design
log analysis procedures that accommodate the above factors. “Consequently, the model equations that
use cation-exchange data have been modified to variants that substitute quantities that can be
measured on logs as surrogate variables, such as that of Juhasz (1981) which is suggested to be a
‘normalized’ representation of the Waxman-Smits equation” (Doveton, 2001).
Asquith (1990) notes that another objection to this equation is that it predicts that effective water
resistivities will decrease as water sands of constant Rw become more shaly -a point which Clavier et
al. (1977) dispute with evidence to the contrary.
The Dual water model devised by Clavier, et al. (1977) circumvents flaws found in the Waxman-Smits
model, and is indirectly based on CEC.

Juhasz Method
Juhasz (1986) developed a system for assessing the shale distribution and determining average
shaliness, porosity, and hydrocarbon saturation of shaly sand laminae. This system is derived partially
from the relationship between total porosity and shale volume described by Thomas and Stieber
(1975), as well as the Waxman-Smits (1968) saturation model. Juhasz uses crossplots of porosity
versus shale volume to evaluate the formation (Figure 2: Two triangles make up the  versus Vsh
crossplot) . The overall evaluation process has been split into separate procedures to cover the
total porosity system as well as the effective porosity system.
Shale Distribution within the Total Porosity System
Juhasz uses the pattern and position of data points within the total porosity versus volume shale (T
versus Vsh) crossplot to evaluate the formation (Figure 3:Total porosity versus shale volume crossplot,
after Juhasz, 1986). The positions of the lines used in the crossplot are based on the maximum
clean sand porosity (max), and the total shale porosity (T, from the density log) observed in the zone
being evaluated. Juhasz assumes that only the combination of laminated and dispersed shale occurs
in Triangle A of the clay distribution plots -while in Triangle B, only the combination of laminated and
structural shale occurs.

Crossplot Boundaries and End Points


The basic equations used to describe the different distributions of shale are also used to define key
lines on the triangles:

• Laminated shale only (Vsh = VL): T = max -VL (max -Tsh)………………Triangles A and B
• Dispersed shale only (Vsh = VD): T = max -VD (1 - Tsh)……………………Triangle A
• Structural shale only (Vsh = VS): T = max + VS Tsh………………………Triangle B
Material balance for shale: Vsh = VL + VD + VS
Where:
VL = Shale volume: laminated
VD = Shale volume: dispersed
VS = Shale volume: structural
Vsh = Shale volume
T = Total porosity
max = Maximum porosity: clean sand porosity
Tsh = Total porosity: shale
Triangle A, which is used only to describe laminated or dispersed shales, is defined by 3 points:

• Sand: (max), such that Vsh = 0%


• Shale: (Tsh), such that Vsh = 100%
• Sand: in which all porosity is entirely filled by dispersed shale (Tsh • max), such that Vsh = max
Triangle B, which is used only to describe laminated or structural shales, shares 2 of 3 points in
common with the other triangle:
Sand: (max), such that Vsh = 0%
Shale: (Tsh), such that Vsh = 100%
Sand: in which all grains have been entirely replaced by structural shale: [max + (1 -max) Tsh], such
that
Vsh = 1 -max
The equations of the lines that describe the triangles can be used to calculate the individual shale
fractions (VL, VD, and VS) and the porosity of the sand laminae can be calculated for any T and Vsh
data pair in the plot.

Choosing the Appropriate Triangle


The evaluation begins by selecting the correct shale distribution triangle. To do so, V sh must be
compared with the theoretical amount of laminated shale (VL) which would be required to yield T :
VL = (max -T) / (max -Tsh)
The result will fall into one of three categories:
1. Vsh = VL, in which only laminated shale is present, such that:
VD = 0
VS = 0, and
Ts = max
In this case, all data pairs will plot along the common hypotenuse shared by both triangles.
2. Vsh < VL, in which only laminated or dispersed shales are present. In this case, all data pairs plot
within Triangle A, such that:
VL = [T -max + Vsh (1 -Tsh)] / (1 -max)

The amount of dispersed shale can be determined by relating it to total bulk volume:
VD = Vsh -VL
VS = 0 (No structural shale will be plotted within Triangle A.)

The shale content of the sand laminae (VDs) are derived from VD corrected for the reduced bulk volume
of the sand laminae:
VDs = VD / (1-VL)

Juhasz then expresses the porosity of the sand laminae (Ts) by the relation:
Ts = (T -VL Tsh) / (1-VL)
3. Vsh > VL, in which only laminated or structural shales are present. In this case, all data pairs plot
within Triangle B, such that:
VL = [max -T + Vsh Tsh)] / max
The amount of structural shale can be determined by relating it to total bulk volume:
VS = Vsh -VL
VD = 0 (No dispersed shale will be plotted within Triangle B.)

The structural shale content of the sand laminae (VSs)are derived from Vs corrected for the reduced
bulk volume of the sand laminae:
VSs = VS / (1-VL)

The porosity of the sand laminae are again expressed by the equation :
Ts = (T -VL Tsh) / (1-VL)

Hydrocarbon Saturation in the Total Porosity System


Assuming that all hydrocarbons are contained within the sand portion of a shaly sand, Juhasz
expresses hydrocarbon saturation in the sand as:
ShTs = T ShT / (1 -VL ) Ts
Where:
ShTs = Total hydrocarbon saturation in the sand
T = Total porosity
ShT = Total hydrocarbon Saturation:
VL = Shale volume: laminated
Ts = Total porosity: sand

Shale Distribution in the Effective Porosity System


Juhasz uses an approach for evaluating shale distribution in the effective porosity system that is
structured similarly to the technique just described above. This approach, however, uses a crossplot of
effective porosity versus shale volume, as shown in Figure 4.

Using the equation for effective porosity as: E= T -Vsh Tsh , Juhasz assumes that shale’s bound-water
does not contribute to effective porosity, and therefore:
Esh = Tsh -Vsh Tsh = 0
Where:
Esh = Effective porosity in shale
Tsh = Total porosity in shale
Vsh = Shale volume

Crossplot Boundaries and End Points


Juhasz uses the equations that describe different distributions of shale to define key lines on the E
triangles :

• Laminated shale only (Vsh = VL): E = max -VL max ……………….Triangles A and B
• Dispersed shale only (Vsh = VD): E = max -VD ……………………Triangle A
• Structural shale only (Vsh = VS): E = max …..…………………….Triangle B
Where:
VL = Shale volume: laminar
VD = Shale volume: dispersed
VS = Shale volume: structural
Vsh = Shale volume
E = Effective porosity
max = Maximum porosity: clean sand porosity

Triangle A, which is used only to describe laminated or dispersed shales, is defined by 3 points:

• Sand: (max), such that Vsh = 0%


• Shale: (Esh), such that Vsh = 100%
• Sand: in which all porosity is entirely filled by dispersed shale (Esh • max), such that Vsh = max

Triangle B, which is used only to describe laminated or structural shales, shares 2 of 3 points in
common with the other triangle:

• Sand: (max), such that Vsh = 0%


• Shale: (Esh), such that Vsh = 100%
• Sand: in which all grains have been entirely replaced by structural shale: [max + (1 -max) Esh],
such that Vsh = 1 -max

Choosing the Appropriate Triangle


The evaluation begins by selecting the correct shale distribution triangle.
Similar to the previous discussion on the total porosity system, we use V sh and E data pairs to
calculate the theoretical value of VL which should be required to satisfy E in the case of laminated
shale distribution:
VL = (max -E) / max
A comparison of the above VL to actual Vsh will lead to one of three results:
1. Vsh = VL, in which only laminated shale is present, such that:
V D = VS = 0

Porosity of the sand laminae is expressed by the equation:


Es = max

In this case, all data pairs of E and Vsh will plot along the common hypotenuse shared by both
triangles.

2. Vsh < VL, in which only laminated or dispersed shales are present. In this case, all data pairs plot
within Triangle A, such that:
VL = (E -max + Vsh) / (1 -max)
The volume of dispersed shale within the sand laminae is expressed as:
VDs = (Vsh -VL) / (1-VL)

Juhasz then expresses the effective porosity of the shaly sand laminae by the relation:
Es = E / (1-VL)
3. Vsh > VL, in which only laminated or structural shales are present. In this case, all data pairs plot
within Triangle B, such that:

VL = (max -E) / max

The amount of structural shale can be determined by relating it to total bulk volume:
VS = Vsh -VL
VD = 0 (No dispersed shale will be plotted within Triangle B.)

The shale content of the sand laminae are derived from Vs corrected for the reduced bulk volume of
the sand laminae:
VSs = VS / (1-VL)

The porosity of the sand laminae are again expressed by the equation:
Es = max or
Es = E / (1-VL)

Hydrocarbon Saturation in the Total Porosity System


Assuming that all hydrocarbons are contained within the sand portion (h(1-VL)) of a shaly sand, Juhasz
expresses hydrocarbon saturation in the sand as:
ShEs = EShE / (1 -VL ) Es
Where
ShEs = Effective hydrocarbon saturation in the sand
Es = Effective porosity in sand
ShE = Effective hydrocarbon saturation
VL = Shale volume: laminated
Es = Effective porosity in sand

Dual Water Model


The Dual Water model (Clavier, et. al, 1977) works on the basis of partitioning pore water into bound
water (Sb) and free water (Sw), each of which contributes to the resistivity (conductivity) of the shaly
sand. Also, the bound water (Sb) and the free water (Sw) have their own associated resistivities (Rb and
Rw). In this model, Rb is fresher (more resistive) than the Rw, which can move freely within the pores.
The following is a summary of the Dual Water formulation:
1. Calculate volume of clay (Vcl): see previous examples.
2. Correct neutron and density porosities for clay
dc = d -(Vcl x dsh)
nc = n -(Vcl x nsh)
where:
Vcl = volume of clay
d = density porosity
n = neutron porosity
nc = neutron porosity corrected for clay
dc = density porosity corrected for clay
dsh = density porosity of shale
nsh = neutron porosity of shale
3. Calculate effective porosity (e)
e = (nc + dc) /2 oil zone
e = [(nc2 + dc2) / 2]1/2 gas zone
where:
e = effective porosity corrected for clay
nc = neutron porosity corrected for clay
dc = density porosity corrected for clay
4. Calculate total porosity of adjacent shale (tsh).
tsh = dsh + nsh
where:
dsh = density porosity of shale
nsh = neutron porosity of shale
5. Calculate total porosity (t) and bound water saturation (Sb).
t = e + Vcl x tsh
Sb = Vcl x tsh/t
where:
Sb = clay-bound water saturation
t = total porosity
e = effective porosity
Vcl = volume of clay
tsh = total porosity of adjacent shale
6. Calculate bound water resistivity (Rb) from adjacent shale.
Rb = Rsh x tsh2
where:
Rb = bound water resistivity
Rsh = resistivity of adjacent shale
7. Calculate apparent water resistivity (Rwa) in the shaly sand.
Rwa = Rt x t2
where
Rwa = apparent water resistivity
Rt = deep formation resistivity
t = total porosity
8. Calculate total water saturation (Swt) corrected for clay.

where:
Swt = total water saturation corrected for clay
b = [Sb (1-Rw /Rb)] / 2
Rw = formation water resistivity
Rwa = apparent formation water resistivity
9. Calculate effective water saturation (Swe) of the shaly sand.
Swe = (Swt-Sb) / (1-Sb)
where:
Swe = effective water saturation
Swt = total water saturation
Sb = clay bound water saturation
Asquith warns that shale resistivity (Rsh) is only equal to the resistivity of the clay in the shaly sand
when the clay in the sandstone is allogenic. When effective water saturations calculated from
equations that require Rsh are much higher than the effective water saturations calculated from
equations that do not require a value for Rsh, then you should suspect that dispersed authigenic clay is
probably present.
Fertl and Hammack (1971) propose that the resistivity of the dispersed authigenic clay can be
approximated by:
Rcl = 0.4 x Rsh
where:
Rcl = dispersed authigenic clay resistivity
Rsh = adjacent allogenic shale resistivity

GUIDELINES
• When the clay is dispersed or when you do not know the clay or shale distribution, use either
the Dispersed Clay or Fertl methods.
• If the shale is laminated, use the Simandoux or Dual Water methods.
• When the clay is dispersed and the resistivity of the dispersed clay is known to be the same as
the adjacent shale (bioturbated or turbidite shaly sands), use Simandoux or Dual Water methods.
• If the resistivity of the dispersed clay in the shaly sand is known, or you feel comfortable in
assuming that the resistivity of the dispersed clay (Rcl) is equal to 0.4 times (Fertl and Hammack,
1971) the resistivity of the adjacent shale (Rsh), then use Simandoux or Dual Water methods.
• In areas where formation water resistivities are high (Rw > 0.1M), water saturations calculated
by the Fertl (1975) and Dispersed Clay methods should be used with caution. The reason for caution is
that it is assumed in both equations that Rsh >> Rw; however, in shaly sands with high formation water
resistivities, this assumption may not be true. If Rsh  Rw use the Dual Water method.

THE ART OF FORMATION EVALUATION


Asquith (1990) states that “Shaly sand analysis is still part science -part art. The most critical
parameter Vcl is often difficult to determine accurately, and the determination of total shale porosity
(tsh) is still questionable (Dewan, 1983). Also, the problem of determining the resistivity of the
dispersed clay (Rcl) is still not solved. There is no substitute for experience in an area.”
We leave the last word on the challenges of log analysis to John H. Doveton (2001). “Laboratory
research and physical and mathematical derivations have done much to improve our understanding of
the complexities that govern the resistivities of shaly sandstones in the search for water saturation
equations with predictive power. However, the failure of any single equation to triumph demonstrates
the variability of shaly sandstone reservoirs and the difficult transition from the laboratory and
theoretical physics to the subsurface reservoir.”
9.13 ADVANCES IN LOGGING
Previous sections showed that while conventional logging tools obtain valid measurements through
low resistivity pay zones, these valid measurements nonetheless lead to higher than actual estimations
of Sw. In this section, we reiterate that conventional resistivity tools are not fooled by low resistivity
pays -but the log analyst might be. In this regard, low resistivity pays may only be a problem for
conventional resistivity tools.
But what if we were to log a low resistivity pay zone with an advanced, high-resolution resistivity tool,
or one that is not based on resistivity at all? Maybe the right technology would deter us from walking
away from an otherwise productive zone. In this section, we describe tools which have been
developed to address the problems caused by low resistivity pay zones. We will describe various
logging approaches that use advanced resistivity measurements, and another which utilizes nuclear
magnetic resonance.
For additional information on the technology described below, consult your local service company
representative, or check out their web pages. Information from leaders in the field of formation
evaluation technology can be found at these websites:

• Baker Atlas: http://www.bakeratlas.com/


• Halliburton Energy Services: http://www.halliburton.com/
• Schlumberger Oilfield Services: http://www.connect.slb.com/

NEW APPLICATIONS FOR RESISTIVITY TOOLS


With advances in logging technology, the leading oilfield technology companies have developed an
impressive array of tools for a broad range applications. In this section, we will discuss a variety of high
resolution resistivity tools.
Resistivity Imaging Tools
Resistivity imaging tools were introduced during the mid-1980s, as an outgrowth of dipmeter
technology. These tools utilize four to six independent arms, each with articulating pads containing
multiple electrodes. This combination of multiple pads and numerous electrodes results in vastly-
improved vertical resolution -to the tune of mere fractions of an inch.
A typical tool emits an electrical "survey" current into the formation, while another current focuses and
maintains a high-resolution measurement. The currents measured by each electrode vary according to
formation conductivity, which reflects changes in fluid properties, permeability, porosity, rock
composition, and grain texture. These variations are processed and converted into synthetic color or
gray-scale images, which are interpreted according to the following convention:

• Light Colors -reflect low micro-conductivity zones, (low porosity, low permeability and
high resistivity)
• Dark Colors -reflect high micro-conductivity zones, (high porosity, high permeability and
low resistivity)
Resistivity Imaging Applications
Borehole imagers use a fixed-contrast presentation for gross correlations, and a dynamic averaging
display to enhance local features.

• The fixed, or absolute contrast allows the viewer to correlate color values between different
zones of interest within the well, or between images from different wells.
• The dynamic averaging display is applied to local events, to allow the viewer to distinguish
features on a smaller scale, such as oil-filled pores, or tight sands.
When integrated with a traditional suite of logs, the images produced by a resistivity imaging tool
enable the analyst to differentiate laminated reservoirs from low-permeability shaly sands. The tool
produces quantitative, high-resolution micro-resistivity measurements that aid in estimating
hydrocarbon saturation and reserves in thin-bedded reservoirs, thus improving the net pay estimation
of laminated reservoirs.
Resistivity Imaging Services
Each of the three leading oilfield technology companies offers their own unique version of the
resistivity imaging tool. And because each company has its own impressive design, we will feature a
photo of each. Examples include:

Halliburton Electrical Micro Imaging (EMI TM)Tool

This tool has six independent arms, with an articulating pad on each arm ( Figure 1: EMITM tool,
courtesy of Halliburton Energy Services). Each pad contains 25 sensors, with a resolution of 0.2
inches. The central button on each pad produces high-definition quantitative resistivity measurements
with a depth of investigation comparable to a short guard or digital focused log (Murphy, 1996). This
tool is rated to 350 F, and 20,000 psi.
The EMI tool maps formation micro-conductivity with its pad-mounted button electrodes. The current of
each button is recorded as a curve, sampled at 0.1 inch (0.25 centimeters), or 120 samples per foot.
The curves reflect the relative micro-conductivity variations within the formation. These current
variations are converted to synthetic color or gray-scaled images. Light colors represent low micro-
conductivity, while dark colors reflect high micro-conductivity zones.
Evaluating thin-bed formations with a high-resolution resistivity curve will help to reduce the risk of
miscalculating hydrocarbon reserves. The log in Figure 2 shows a comparison between a quantitative,
high-resolution resistivity curve obtained from one EMITM button (Track 3) and DFL and HDRS curves
(Track 4) obtained with the High Resolution Induction tool. The neutron/density curves are shown in
Track 5. Using the high-resolution resistivity measurement results in improved saturation
calculations and more realistic net pay estimations.
Schlumberger Formation MicroImager (FMI TM) Tool
In addition to a 24-button microelectrical array pad on each of four arms (192 buttons total), the FMI TM
mounts an extendable pad below each arm, to increase pad coverage to about 80% of an 8-inch
borehole. (Figure 3: FMITM Tool; courtesy of Schlumberger Oilfield Services) Resolution is 0.2 inch
(5mm), and the tool is rated to 350 F, and 20,000 psi.

Hybrid Resistivity Imaging Devices


In this section, we discuss two rather unique imaging devices, each of which features specialized
capabilities and operating modes.

Baker Atlas Simultaneous Acoustic/Resistivity (STARTM) Tool


Rather than taking only resistivity measurements, this tool simultaneously acquires high-resolution
images of borehole features that have resistivity contrast or acoustic impedance. This combination of
acoustic and resistivity measurements partially compensates for any shortcomings inherent in either of
the individual measurements. The six-arms on the tool use a powered standoff to improve pad contact
with the borehole, thus providing resistivity coverage of 60% of an 8-inch hole, and 100% acoustic
coverage. (Figure 4: STARTM Tool; courtesy of Baker Atlas.) The tool is rated to 350 F.

Schlumberger Azimuthal Resistivity Imager (ARITM) Tool


Instead of relying on pad contact, this tool uses an array of 12 azimuthal electrodes, spaced 30
degrees apart. (Figure 5: Conceptual drawing of ARITM tool; courtesy of Schlumberger Oilfield
Services). This dual laterolog array is able to measure deep resistivity, but with a vertical resolution of
only eight inches. This makes for a laterolog reading that is similar to the laterolog deep curve, but with
a vertical resolution that approaches that of the MSFL curve. As an imaging tool, the ARI TM is less
sensitive to borehole rugosity than the FMI electrical imaging tool, and can also provide coarse
structural dip measurements. The tool was developed for evaluation of heterogeneous reservoirs, thin-
bed analysis, and fracture identification. It is rated to 350F, and 20,000 psi.
Consult your logging representative for more information on the resistivity imaging services that their
company can provide.

Digital Array Induction Logs -


Digital array induction tools use multiple receivers and multiple logging frequencies which provide
capabilities that are not available with conventional induction tools.
• Longer receiver coil spacings enable the determination of accurate R t estimates, even in the
presence of deep invasion.
• Short coil spacings provide information that is used to correct for borehole and near-borehole
effects.
• Better vertical resolution.

Array Induction Applications


The array induction device provides improved vertical resolution capabilities in thin beds. These
measurements are used to evaluate complicated reservoirs having thin beds (down to 1 foot thick) or
having deep or unusual invasion profiles.

Array Induction Services


Each of the three leading oilfield technology companies offers their own unique version of the
resistivity imaging tool. Examples of such resistivity imaging devices include:

Baker Atlas High-Definition Induction Log (HDILTM)


The HDIL provides an order of magnitude more information than conventional induction tools. The
HDIL tool investigates the formation at median depths of 10, 20, 30, 60, 90, and 120 inches, and is
thus able to present a precisely defined invasion profile. And identical readings by the 90-inch and
120-inch depth measurements are used to provide a direct indication of Rt.
HDIL data can be processed in either of two modes to best suit reservoir conditions:

• HDIL true resolution data processing provides very accurate formation resistivity
values by minimizing effects that near-borehole features can cast on deep-reading
curves. Using this format, the vertical resolution of the curves varies with the depth of
investigation.
• HDIL resolution-matched processing is used in thin-bed reservoirs, where bed
boundary effects can limit the accuracy of deeper investigating measurements. To
improve interpretation in this situation, high-resolution data near the borehole are
added to the deeper measurements so that all curves are presented with the same
matched vertical resolution, of 1, 2, or 4 feet.

Schlumberger Array Induction Imager Tool (AITTM)


This tool uses 8 induction-coil arrays operating at multiple frequencies to generate five resistivity
curves. The log curves have median depths of investigation of 10, 20, 30, 60, and 90 inches, and
vertical resolution options of 1 foot, 2 feet, and 4 feet. When the logs are radially deconvolved to
produce a detailed radial description of formation conductivity, the conductivity description can be
presented as a color-coded image or as discrete log curves.

Halliburton High Resolution Induction (HRITM) Tool


tool features five radii of investigation (90, 60, 50, 40, 30, and 24 inches). Their log also displays a
resistivity map to indicate formation resistivity as a function of depth and radial distance from the HRI
tool.
Consult your logging representative for more information on the array induction services that their
company can provide.

3D Multicomponent Resistivity Tool


Conventional induction logging tools use transmitter and receiver coils that are aligned with the long
axis of the tool. In wells drilled perpendicular to bedding, these tools measure formation conductivity
parallel to bedding. When a reservoir is composed of thinly bedded, highly conductive shales and
hydrocarbon-bearing sands that are below the vertical resolution of the tool, the result is
measurements experience the problematic “low-contrast, low-resistivity pay” effect (Mollison, 2001).

Baker Atlas 3D Explorer Induction Logging Service (3DEXTM)


Baker Atlas has developed a resistivity tool unique to the industry, which is designed to overcome the
limitations of conventional induction tools in thin bedded, low-resistivity shaly-sand formations. The
Baker Atlas 3D Explorer Induction Logging Service (3DEX TM) provides both vertical and horizontal
resistivity measurements independently of borehole deviation or formation dip.
The 3DEX features three transmitter-receiver coil arrays, which are mounted orthogonally in the X, Y,
and Z planes relative to the tool axis. These coil arrays provide 3-D coverage in their resistivity
measurements:

• two coils (XX and YY) measure resistivity in transverse directions (parallel to the tool body),
• a third coil (ZZ) measures resistivity in the direction of conventional resistivity tools
(perpendicular to the tool body)
• in addition, there are cross component measurements (XY and XZ).
These arrays induce currents that flow, for the most part, across laminated sand-shale sequences, and
are far more sensitive to hydrocarbon-bearing sand resistivity, as shown in Figure 6: Basic principle of
operation: Laminated sand/shale intervals are surveyed by three orthogonal coil arrays.
Inversion processing of XX-YY-ZZ measurements obtained through the tool’s orthogonal coil
configuration are used to determine vertical and horizontal resistivity R v and Rh. The 3DEX horizontal
resistivity is always determined parallel to the bedding plane, consequently, the vertical resistivity is
always measured perpendicular to the bedding plane. Thus, regardless of changes in borehole
deviation or apparent strike and dip, the 3DEX measurements of Rv and Rh remain properly oriented to
the formation bedding.
Where there is a difference in values between Rv and Rh, the formation is said to be electrically
anisotropic.

Electrical Anisotropy Effect


Conventional induction logging tools are limited to measurements in one dimension because their
sensors are aligned along the length of the tool (its Z-axis). Such measurements are satisfactory only
when formations are at least as thick as the tool’s vertical resolution, which is generally several feet.
In the presence of small apparent formation dips, the conventional induction tools induce currents that
mainly flow in the highly conductive beds (typically shales) of hydrocarbon bearing sections. Thus,
when pay zones occur within thinly bedded sand-shale sequences, the conventional horizontal
induction measurement is dominated by the lowest resistivity, usually found in the shale layers. As a
result of this induced current flow pattern, the horizontal resistivity (R h) is relatively insensitive to the
higher resistivity of the hydrocarbon-bearing sands. In this manner, relatively small volumes of
conductive shale can significantly reduce the apparent resistivity, thereby reducing the accuracy of
computed hydrocarbon saturations for the sand layers.
Vertical resistivity, however, is dominated by the highest resistivity component. In a hydrocarbon
reservoir, Rv measurements provide more information on the resistive sand components, thus yielding
more accurate fluid saturations in the sand layers. The 3DEX tool capitalizes on this principal, with coil
arrays aligned to resolve vertical resistivity.
In a thinly laminated sand-shale sequence, effective horizontal and vertical resistivities are derived
through parallel and series resistor models. The corresponding formulae are:

(Equation 1)

where
Rh = horizontal resistivity
Rsh = shale resistivity,
Rsd = sand resistivity
Vsh = shale volume, and
Vsd = sand volume
such that
Vsh + Vsd = 1
and

(Equation 2)

where
Rv = horizontal resistivity
Rsh = shale resistivity,
Rsd = sand resistivity
Vsh = shale volume, and
Vsd = sand volume
These equations are key to understanding the important differences between horizontal and vertical
resistivity. Equation 1 helps to explain how horizontal resistivity is affected by shale or by sand:
• horizontal resistivity (Rh) is strongly dependent on shale resistivity (usually low) and shale volume

• horizontal resistivity exhibits poor sensitivity to sand resistivity.


Conventional induction tools, with their coils aligned along the length of the tool, are only able to
measure perpendicular to formation bedding, and thus are only sensitive to horizontal resistivity.
Equation 2 demonstrates that vertical resistivity averages the contributions from both sand and
shale, and thereby provides a much better indicator of thin hydrocarbon-bearing sands.
The 3DEX tool capitalizes on vertical and horizontal conductivity measurements to determine the
laminar shale volume and laminar sand conductivity. A Thomas-Stieber-Juhasz evaluation technique is
applied to determine the volume of dispersed shale along with the total and effective porosities of the
laminar sand fraction. By removing laminar shale conductivity and porosity effects, the laminated shaly
sand problem is reduced to a single dispersed shaly sand model to which the Waxman-Smits equation
can be applied. (For additional details, see the Petrophysical Evaluation described below.)

Log Example
In this example from Mollison, et al. (2000), the 3DEX tool was used to evaluate a shaly-sand interval
containing three distinct zones, each of which exhibit different ranges of electrical anisotropy and shale
content. The resulting log is shown in Figure 7.

• The upper sand, from x100 to x145 feet, exhibits a fining-upward sequence of moderately
shaly sand. The data show significant electrical anisotropy (Track 1), as demonstrated by
the separation of Rv and Rh (Track 2).
• The middle sand, from x145 to x169 feet, is a gas producing zone with low shale content.
This interval exhibits little anisotropy, as would be expected in a massive, high-porosity
sand.
• The lower sand, from x169 to x220 feet, is characterized by higher shale content and
higher electrical anisotropy than the upper sand. Conventional, deep-induction resistivity
data, HDIL, shown in track 2, would not be able to effectively identify this interval as a
potentially productive sand-shale sequence. However, the Rv and Rh data improve
evaluation accuracy of the lower sand and properly identify this as a finely laminated sand
interval.
Petrophysical Evaluation of the Log
Directional resistivity measurements from the 3D Explorer tool can be used to compute both the
volume of laminar shale and the resistivity of the sand fraction of a laminated formation without
reference to other measurements or shale indicators. The 3DEX petrophysical evaluation model
removes the laminar shale conductivity effects by utilizing electrical anisotropy measurements R v and
Rh.
In sand-shale sequences, Rv and Rh measurements provide a close link to the petrophysical model
through the direct computation of laminar shale. This laminar shale volume may be compared to
Thomas-Stieber style volumetric laminar shale calculations, thus yielding a validation of the both
petrophysical models.
Petrophysical analysis of the above log reveals that the shales are predominantly laminar, with minor
amounts of dispersed shale (Track 4). In the upper sand interval (100’ to 145’), the calculated laminar-
sand resistivity, Rsd, is 3 to 5 -m higher than that indicated by either the deep induction of the HDIL
tool or the horizontal resistivity Rh of the 3D Explorer (Track 2). Water saturation from the laminar sand
analysis is 10% to 15% lower than that obtained by standard saturation analysis (Track 3), indicating
commercial hydrocarbon production rates are probable from this interval.
This comparison of laminar shale volumes may also provide valuable geological information. For
example, the presence of anisotropic resistivity allows important additional interpretation as to the
geometry of the layers, i.e., parallel bedding. The lack of resistivity anisotropy would point to a lack of
parallel bedding, such as disturbed, folded or slumped bedding. Such intervals often tend toward low
producibility.
The lower sand interval, from x169 to x220 feet, is the most interesting in this well. Total shale volume
in this interval is 60% to 70%. The separation between the Rh and Rv curve, together with the resulting
anisotropy ratio, indicate that the formation is almost entirely laminar and thin-bedded, with an average
net-to-gross ratio of 35%. Water saturation through the laminar sand is calculated at 40% to 55%,
which agrees well with water saturation values obtained in the upper sand interval. The net result is a
possible 18 feet of additional pay that might not have been identified by standard resistivity tools and
traditional water saturation analysis methodology.
The 3D Explorer can also provide supplemental measurements for the High Definition Induction Log
(HDIL). It can be run on the same toolstring and and logged simultaneously, at the same logging
speed required by the HDIL tool. Data processing at the wellsite is provided to expedite the decision-
making process (e.g. testing and completion).

NUCLEAR MAGNETIC RESONANCE TOOLS


When microporosity, conductive mineralogy, or altered framework grains are the cause of low-
resistivity pay problems, then perhaps an alternative approach to logging would help the formation
evaluation program. In this case, Nuclear Magnetic Resonance logging, which does not depend on
rock conductivity, can be used to accurately determine hydrocarbon saturation and distinguish
between free water and bound water in the reservoir. (In fact, the esimation of bulk volume irreducible
water, or BVI, is one of the earliest and most widely used applications of NMR logging.)
The conventional neutron, bulk-density, and acoustic-travel-time porosity-logging tools are influenced
by components of the reservoir rock. Because reservoir rocks typically have more rock framework than
fluid-filled space, these conventional tools tend to be much more sensitive to matrix materials than to
pore fluids.
Conventional resistivity-logging tools, while being extremely sensitive to the fluid-filled space, are
traditionally used to estimate the amount of water present in reservoir rocks, but cannot be regarded
as true fluid-logging devices. These tools are strongly influenced by the presence of conductive
minerals and, for the responses of these tools to be properly interpreted, a detailed knowledge of the
properties of both the formation and the water in the pore space is required.
NMR logging tools use a permanent magnet to produce a magnetic field that excites formation
materials. An antenna transmits an oscillating magnetic field in precisely timed bursts of radio-
frequency energy into the formation. Between these pulses, the antenna is used to listen for the
decaying “echo” signal from hydrogen protons that are in resonance with the field from the permanent
magnet. Since this magnetic resonant frequency depends on the local strength of the magnetic field,
the measurement zone of the tool is a function of the magnetic field generated, and the radio
frequency used.

NMR measurements respond primarily to hydrogen protons in the pore spaces of the formation, thus
providing a measure of water or hydrocarbons in the rock. Unlike conventional porosity measurements
(such as the compensated neutron tool), this measure of NMR porosity does not include hydrogen
bound in the matrix of the rock, thus providing porosity values that are not influenced by lithology.
(Figure 8: MRIL porosity model, Courtesy of Baker Atlas) With only fluids visible to the NMR tool, it
does not need to be calibrated to formation lithology. This response characteristic makes NMR logging
tools fundamentally different from conventional logging tools.

Unique Formation Measurements


NMR tools can provide three types of information, each of which make these tools unique among
logging devices:

• information about the quantities of fluids in the rock,


• information about the properties of these fluids, and
• information about the sizes of the pores that contain these fluids.
Specifically, NMR tools are used determine total porosity, effective porosity, capillary bound water
volume, free water volume, hydrocarbon volume, and permeability. The basic physics behind NMR
interpretation is common to all such data; however, each of the current NMR logging service
companies -Baker Atlas, Halliburton, and Schlumberger have their own proprietary interpretation
methods. In addition, there are now several companies that specialize in the interpretation of NMR
data, including NuTech and NMR+.

Schlumberger Combinable Magnetic Resonance Tool (CMRTM)


The Schlumberger Combinable Magnetic Resonance tool uses a directional antenna sandwiched
between a pair of bar magnets to focus the CMR measurement on a 6-in. [15-cm] zone inside the
formation—the same rock volume scanned by other essential logging measurements. As shown in
Figure 9 (CMR tool), it is a compact skid-mounted tool that was designed to be combinable with many
other standard logging tools. The CMR tool is run in an eccentered configuration.
The vertical resolution of the CMR measurement makes it sensitive to rapid porosity variations, as
often seen in laminated shale and sand sequences. The sensitive region of the tool is shown in red in
Figure 10 (Cross-section of the CMR tool). This region is approximately 0.5” x 0.5” by 6” long, and is
located about 1.1 inches inside the formation.

Baker Atlas Magnetic Resonance Imaging Log (MRIL™) Service

The Magnetic Resonance Imaging Log run by Baker Atlas provides the capability to run in combination
with other openhole logging instruments (Figure 11: Schematic of combined tool configuration;
Courtesy of Baker Atlas). The tool is run in a centralized configuration to ensure that the sensitive
volume does not include the borehole fluid, and is unaffected by borehole rugosity. The MRIL
measurements can investigate the formation at diameters of up to 18 inches. This tool can be operated
simultaneously at different frequencies to increase the sensed volume, improve the signal-to-noise
ratio, and allow multiple NMR measurements to be obtained at one time.

Halliburton Magnetic Resonance Imaging Log (MRIL™) Tool


The MRIL-Prime tool was introduced in 1998. Like other MRI tools, this MRI probe can be tuned to be
sensitive to a specific frequency, thereby allowing the MRI to image narrow slices of the rock
formation. Figure 12 (Cylinders of investigation: Courtesy of Halliburton Energy Services) illustrates the
measurement concept behind the MRIL-Prime tool. The diameter and thickness of each thin cylindrical
region are selected by simply specifying the central frequency and bandwidth to which the MRIL
transmitter and receiver are tuned. The diameter of the cylinder is temperature dependent, but typically
ranges from approximately 14 to 16 inches.
Consult your local service company representative for more information on NMR logging tools.

Log Example

In the first log Figure 13, we see a classic example of a low resistivity zone, which does not show any
potential for future completion. This example, provided by Halliburton Energy Services, shows MRI and
resistivity data obtained in a Low Resistivity Pay zone. (Figure 14: MRIL Log presentation, Courtesy of
Halliburton Energy Services)

Log Description -
Following is a list of curves presented in each track of the log.

• Track 1: MRIL porosity derived from T2 bins, along with Caliper, Gamma Ray, and
SP measurements from conventional logs.
• Track 2: MRIL permeability, derived from MRIL Porosity, Bound Water, and Free
Fluid measurements, along with Deep and Shallow Resistivity from conventional logs.
• Track 3: T2 distribution from partially polarized activation with TE of 0.6 ms (left side
of track 3), which is typically indicative of clay bound water, and the T2 distribution
from fully polarized activation of a TE of 1.2 ms (right side of track), usually indicative
of capillary bound water and free fluids.
• Track 4: The difference between two T2 distributions, each taken with a TE of 1.2 ms
at different polarization times, yields hydrocarbon signals within the free fluids.
Relative position indicates hydrocarbon type and viscosity value.
• Track 5: Time Domain Analysis of calculated volumes of oil, gas, and free water in
the pore space, which provides a complete description for the fluids in the invaded
zone.
• Track 6: MRIL-Resistivity display of MRIAN (MRI Analysis) model to calculate
volume of hydrocarbon and free water in the pore space, which provides a complete
description of fluids in the uninvaded zone.

MRI Log Evaluation


On this log, we see a low-resistivity sand interval from xx160 feet to xx260 feet. This interval is
characterized by clean sand at the top, grading down to a shaly sand (fining downward) at the base of
this interval. Average shale resistivity is about 0.8 ohm-m; maximum resistivity through the sand is 1.3
ohm-m at xx184 feet; and minimum resistivity through that zone is 0.6 ohm-m at xx238 feet. The Time
Domain Analysis shown in Track 5 indicates the presence of some free water, but the MRIAN model in
Track 6 shows the interval is water-free. In fact, this interval was tested at 2000 BOPD, with no water.
9.14 EVALUATION OF LOW RESISTIVITY PAY USING MULTICOMPONENT
INDUCTION LOGGING
This case study describes a low resistivity pay zone that was logged using the Baker Atlas 3D
ExplorerSM Multicomponent Induction Logging Instrument. A detailed discussion of this well log can be
found in the original paper by Munholm, et.al (2002).

A number of exploration and appraisal wells had been drilled in the West Delta Deep Marine
Concession, offshore from Egypt’s Nile River delta (see Figure 1: Location map of West Delta
Deep Marine Concession). These wells were notorious for their low resistivity pays, found in
Upper Pliocene slope channel complex reservoirs, which comprise a mix of deep marine channel
and overbank facies. The channel fills are highly heterogeneous, and include a mix of turbidites,
slides, slumps, and debris flows. Lithology varies from clean, unconsolidated sands and
conglomerates, to finely laminated sandstones, siltstones and mudstones. The overbank facies
are typical channel levee deposits.
Dual packer formation tester and well test data proved that many of the fine-grained laminated
sections contain economic pay zones. (One DST flowed in excess of 23 mmscf/d from a finely
laminated interval of 17m.) While these laminated sections make up substantial volumes of the
channel fill and overbank sections, they are difficult to quantify using conventional petrophysical
logs, as demonstrated in Figure 2 (Composite log from WDDM Concession).
Figure 2 presents a composite log showing GR, Density/Neutron, Array Laterolog, and Resistivity
Image logs. While pay is quite clearly evident in the zone between XX37 and XX45 meters, we
cannot be quite so sure about the zone from XX29 to XX37m. Though the neutron/density curves
exhibit gas-effect crossover through this upper interval, the resistivity curves of the array laterolog
cannot resolve the resistivities of individual sand laminae. By zooming in on a 0.5m segment of
core, we see individual laminations of only a few centimeters thickness, which are well below the
vertical resolution of conventional resistivity devices. The array laterolog records resistivities of 1-
3 Ohmm through this interval, resulting in uncertain saturation estimates.
The Operator of this well elected to analyze this zone further, using the Baker Atlas 3D Explorer SM
tool. The composite log In Figure 3 (Interpreted results from the 3DEX tool), shows the following
curves:
Track 1 – Gamma Ray (0 – 100 gAPI) and Caliper (4-14 in);
Track 2 – Bulk Density (1.7-2.7 g/cc) and Compensated Neutron (60-0 pu);
Track 3 – Depth (meters);
Track 4 – HDIL Array Resistivity (2ft vertical resolution) matched with 3DEX
derived resistivity of the sand fraction: Rsand (0.2 to 200 Ohmm);
Track 5 – 3DEX Rh and Rv and HDIL 4ft, 90 inch VRM (0.2-200 Ohmm);
Track 6 – Shale volume from Thomas-Stieber and Tensor calculations (0-100%);
Track 7 – Total porosity for the sand fraction PTSD and total porosity for the entire
formation, PORT (50 – 0 pu);
Track 8 – Total water saturation of the formation, SWT, and total water saturation
of the sand fraction SWTSD (100 – 0%)

In the laminated intervals, the log shows that Rsand is significantly higher than Rh. In Track 5, the
agreement between HDIL and Rh reinforces the fact that while the HDIL tool obtains high resolution
measurements, it nonetheless measures horizontal resistivity. The track on the far right shows that
water saturations derived for the sand portions of the formation (SWTSD) hover around the 10% level.
On the basis of the 3DEX data, the operator of this well was able to open a zone to production that had
previously been in question.

INTEGRATION OF NMR DATA WITH CONVENTIONAL LOGGING DATA


By integrating NMR data with conventional log measurements, it is possible to improve the
petrophysical analysis of shaly sands. Integration enables the determination of several important
reservoir properties that had previously been quite elusive in the absence of core data, such as:
permeability,
shale distribution;
clay bound water volume;
irreducible water saturation;
Qv and;
mineralogically independent total and effective porosities.
This case study describes how NMR measurements have been integrated with conventional
logging measurements to evaluate porosity and permeability in a laminated sand-shale interval.
Further discussion of this log can be found in the paper by Ostroff, et.al (1999).
In laminated sand-shale sequences, where the thickness of individual laminations is less than
the vertical resolution of the logging tool, the porosity values for individual laminations will not be
adequately resolved. Consequently, the log porosity represents the partial volume-weighted
summations of the sand and shale porosity values integrated over the vertical measurement
resolution of the logging tool as follows:
In the above equations, the laminar shale volume is computed using the NMR tool’s shale
distribution model. Figure 4 (NMR response in thin beds) illustrates NMR tool response in a
laminated sand-shale sequence.

After the laminar shale volume is quantified, the pore volumetrics of the sand laminations can be
reconstructed as follows:

In thin-bedded, laminar sand-shale sequences, the NMR-derived “bulk” permeability that is


computed from the “bulk” NMR-porosity data might be much too low when compared with the
“true” permeability of actual sand laminations. If the above re-constructed sand lamination pore
volumes are substituted into the Coates-Timur model, a reconstruction of the sand lamination
permeability can also be made, as shown below.

Coates-Timur Model

This model uses NMR-derived porosity and the ratio of free fluid to bound fluid to compute
permeability. Default values for model parameters are C = 10, a = 4, and b = 2
Log Results

The log example in Figure 5 (Highly laminated pay zone) shows computed “bulk” permeabilities
and the re-constructed sand lamination permeabilities for a highly laminated progradational sand-
shale sequence from the Gulf of Mexico. The combined NMR and resistivity-based saturation
analysis indicates that the entire sequence is at irreducible saturation. The reconstructed sand
lamination permeability (derived from the above models) correctly predicts the high productivity
observed from the completion across the very highly laminated lower interval. The original
completion of the upper interval produced water-free at 800 BOPD. The additional completion of
the lower interval added another 733 BOPD of water-free production.
REFERENCES AND FURTHER READING:

WEBSITES
For information on technology or log analysis techniques described in this module, consult your local
service company representative, or check out their web pages. Information from leaders in the field of
formation evaluation technology can be found at these websites:
• Baker Atlas: http://www.bakeratlas.com/
• Halliburton Energy Services: http://www.halliburton.com/
• Schlumberger Oilfield Services: http://www.connect.slb.com/

REFERENCES AND SIGNIFICANT WORKS


Al-Awad, M.N.J., Hamada, G.M., and Almalik, M.S. 2001: “Low-resistivity beds may produce water-
free,” Oil & Gas Journal, January 1, vol 99.1.
Allen, D.F., Klein, J.D., and Martin, P.R., 1995: “The Petrophysics of Electrically Anisotropic
Reservoirs,” Transactions of the SPWLA 36th Annual Logging Symposium, Paris, France, June 26-29.
Almon, W.R., 1979: “A geologic appreciation of shaly sands,” Society of Professional Well Log
Analysts, 20th Annual Symposium Transactions. Paper WW.
Anderson B., Bryant I., Luling M., Spies B., and Helbig K., 1994: “Oilfield Anisotropy: Its Origins and
Electrical Characteristics” in Schlumberger Oilfield Review, October issue.
Anderson, B., Druskin, V., Habashy, T. Lee, P., Luling, M., Barber, T., Grove, G., Lovell, J., Rosthal,
R., Tabanou, J., Kennedy, D., Shen, L., 1997:” New Dimensions in Modeling Resistivity” Schlumberger
Oilfield Review, Spring issue.
Archie, G.E., 1942: “The electrical resistivity log as an aid in determining some reservoir
characteristics,” Petroleum Technology volume 5, p.54-62.
Asquith, G.B., and Gibson, C.R., 1983: Basic Well Log Analysis for Geologists, Second Printing
(revised), AAPG Methods in Exploration Series, Published by the American Association of Petroleum
Geologists, Tulsa, Oklahoma. 216 p.
Asquith, G.B., 1990, Log Analysis of Shaly Sandstone Reservoirs: A Practical Guide, AAPG
Continuing Education Course Notes, Published by the American Association of Petroleum Geologists,
Tulsa, Oklahoma. 59 p.
Bateman, R. M. 1984.” Watercut prediction from logs run in feldspathic sandstone with fresh formation
waters,” Paper EE, SPWLA 25th Annual Logging Symposium (June 10-13).
Bateman, R.M., 1985: Open-Hole Log Analysis and Formation Evaluation, Published by IHRDC Press,
Boston, MA. ISBN: 0-88746-060-7 (647 pages).
Boyd, A., Darling, H., Tabanou, J. Davis, B., Lyon, B., Flaum, C., Klein, J., Sneider, R., Sibbit, A.,
Singer, J., 1995, “The Lowdown of Low-Resistivity Pay”: Schlumberger Oilfield Review, Autumn issue.
Chemali, R, Gianzero, S, and Su, SM, 1987: “The Effect of Shale Anisotropy on Focused Resistivity
Devices,” Transactions of the SPWLA 28th Annual Logging Symposium, London, England, June 29 -
July 2, 1987, Paper H.
Clavier, C., Coates, G., and Dumanoir, J., 1977: “The theoretical and experimental bases for the ‘Dual
Water’ model for the interpretation of shaly sands,” SPE paper 6859, AIME, 52nd Annual Fall
Technical Conference, Denver, Colorado, October 9-12.
Darling. H.L. and Sneider, R.M., 1993: “Productive Low Resistivity Well Logs of the Offshore Gulf of
Mexico: Causes and Analysis” In Productive Low Resistivity Well Logs of the Offshore Gulf of Mexico,
Dwight Moore, Editor in Chief, Published jointly by the New Orleans Geological Society and Houston
Geological Society.
Dewan, J.T., 1983: Essentials of Modern Open-Hole Log Interpretation, PennWell Publishing
Company, Tulsa, Oklahoma, 361 pages.
de Witte, L. 1950. “Relations between resistivities and fluid contents of porous rocks,” Oil and Gas
Journal (August 24).
Doveton, J. H. 1986: Log analysis of subsurface geology. New York: Wiley Interscience. 273 pp.
Doveton, John H., “All Models are Wrong, but Some Models are Useful: “Solving” the Simandoux
Equation.” From Session J of the International Association for Mathematical Geology Conference,
2001, Cancun, Mexico.
Fanini, O.N., Kriegshauser, B.F., Mollison, R.A., van Popta, J., and Yu, L., 2000: “New multicomponent
induction resistivity tool finds thin-bed pay,” World Oil Magazine, June, Vol. 221, no. 6.
Gary, M., McAfee, R., and Wolf, C.L., 1972: Glossary of Geology, American Geological Institute,
Washington, D.C. ISBN 0-913312-00-2.
Giesel. W., 1963: “Elastische Anisotropie in Tectonisch Verformten Sedimentgesteinen,” Geophysical
Prospecting 11 (1963): 423-458.
Helbig, K., 1994: Foundations of Anisotropy for Exploration Seismics, Handbook of Geophysical
Exploration Section I Seismic Exploration, Volume 22. Oxford, England: Elsevier Science Ltd).
Hill, H.J., Shirley, O.J. and Klein, G.E. (Edited by M.H. Waxman and E.C. Thomas), 1979: "Bound
Water in Shaly Sands - Its Relation to Qv and Other Formation Properties," The Log Analyst, Volume
XX, number 3, Society of Professional Well Log Analysts, May-June, 1979, p. 3-19.
Honarpour, M. L. Keedritz and H.H. Harvey. 1986. Relative Permeability of Petroleum Reservoirs.
Boca Raton, FL: CRC Press, Inc.
Juhasz, I. 1981, “Normalised Qv -the key to shaly sand evaluation using the Waxman-Smits equation
in the absence of core data,” SPWLA Logging Symposium Transactions.
Juhasz, I., 1986: “Assessment of the distribution of shale, porosity, and hydrocarbon saturation in
shaly sands,” Society of Professional Well Log Analysts, Tenth European Formation Evaluation
Symposium, Paper AA.
Krynine, P.D.,1948: "The megascopic study and field classification of sedimentary rocks", Journal of
Geology, v.56, p.130-165.
Kulha, J.T., 2001: “Identification and Evaluation of Low Resistivity, Low Contrast Productive Sands,”
Society of Petroleum Engineers Gulf Coast Section Continuing Education Seminar, Houston, Texas,
February 13, 2001.
Lynn, HB, 1991: “Field Measurements of Azimuthal Anisotropy: First 60 Meters, San Francisco Bay
Area, CA, and Estimation of the Horizontal Stresses’ Ratio from Vs1/Vs2,” Geophysics 56 (June): 822-
832.
Maillet, R. and Doll, H.G., 1932: “Sur un Theorem Relatif aux Milieux Electriquement Anisotropes et
ses Applications a la Prospection Electrique en Courant Continu,” Eganzungshefte fur angewandte
Geophysik 3: 109-124.
Manrique, JF, Kasap, E, and Georgi, DT, 1994: “Effect of Heterogeneity and Anisotropy on Probe
Permeameter Measurements,” Transactions of the SPWLA 35th Annual Logging Symposium, Tulsa,
OK, USA, June 19-22, Paper R.
McDonald, D.A., and Schmidt, V., 1992: Porosity Evolution of Sandstone Reservoirs, Published by
IHRDC Press, Boston, MA.
Molina, N. N. 1983. “Systematic approach aids reservoir simulation.” Oil and Gas Jour. (April).
Mollison, R., 2001: “3-D Induction logging improves evaluation of low resistivity pay zones,” The
American Oil & Gas Reporter, September.
Mollison, R.A., Kriegshauser, B.F.,Yu, L., and van Popta, J., 2001: “3-D Induction Resistivity Improves
Reservoir Characterization in Low-Contrast, Low-Resistivity Shaly Sand Formations,” in Baker Atlas
In-Depth Magazine.
Morgan, J.T. and D.T. Gordon, 1970, Influence of Pore Geometry on Water-Oil Relative Permeability,
Soc. Pet. Eng. J. (October 1970), p. 1179.
Munholm, M., Nashaat, M., Samuel, A., Webb, E., Parsons, C., and Hogarty, P.: "Low Resistivity Pay
in Offshore Nile Delta: Quantification using Multicomponent Induction Logging," Presented in
Alexandria, Egypt, at the Mediteranean Offshore Conference, April 9-11, 2002.
Neasham, J.W., 1977: “The morphology of dispersed clay in sandstone reservoirs and its effect on
sandstone shaliness, pore space, and fluid flow properties,” Society of Petroleum Engineers Paper
6858, presented at the 52nd Annual SPE Meeting, Denver, Colorado.
Ostroff, G., Shorey, D. and Georgi, D., 1999: "Integration of NMR and conventional log data for
improved petrophysical evaluation of shaly sands" Society of Professional Well Log Analysts; 40th
Annual Logging Symposium Transactions; paper OOO, 1-14.
Page, G.C., Fanini, O.N., Kriegshauser, B.F., Mollison, R.A. Yu, L., and Colley, N., 2001: “Field
Example Demonstrating a Significant Increase in Calculated Gas-In-Place: An Enhanced Shaly Sand
Reservoir Characterization Model Utilizing 3DEX Multicomponent Induction Data,” SPE Paper 71724,
Presented at the SPE Annual Technical Conference and Exhibition in New Orleans, Louisiana, 30
September through 3 October, 2001.
Patnode, H.W., and Wylie, M.R.J., 1950: “The presence of conductive solids in reservoir rocks as a
factor in log interpretation,” Petroleum Transactions, AIME 189, p.47-52.
Pettijohn, F.J., 1957: Sedimentary Rocks, 2nd Edition, Harper Press, New York, 718 p.
Pettingill, H.S., 1998: “Worldwide Turbidite Exploration and Production: A Globally Immature Play with
Opportunities in Stratigraphic Traps,” SPE Paper 49245, presented at 1998 Society of Petroleum
Engineers Annual Technical Conference in New Orleans, La., September 27-30, 1998.
Pirson, S. J., Boatman, E. M., and Nettle, R. L., 1964: "Prediction of relative permeability
characteristics of intergranular reservoir rocks from electrical resistivity measurements." J. Pet. Tech.
(May) :565-570.
Poupon, A., Loy, M.E., and Tixier, M.P., 1953: "A contribution to Electrical Log Interpretation in Shaly
Sands," AIME Technical Paper 3800, Presented at the Petroleum Branch Fall Meeting in Dallas, TX.
Oct. 19-21, 1953.
Poupon, A., Hoyle, W. R., and Schmidt A. W., 1971: Log analysis in formations with complex
lithologies. J. Pet. Tech. (August).
Rajan, VSV, 1988: “Discussion of the Origins of Anisotropy” paper SPE 18394, Journal of Petroleum
Technology, 40, (July): 905.
Schlumberger Well Services, Inc., 1984(?): A wellsite guide to the recognition of low resistivity pay
sands in the Gulf of Mexico. Published by Schlumberger Offshore Services Eastern Division, New
Orleans, Louisianna.
Serra, O. (1989), Formation MicroScanner Image Interpretation, published by Schlumberger
Educational Services.
Shelton G. and Hill, W.A., 1993: “Productive Low Resistivity Well Logs of the Offshore Gulf of Mexico:
A Summary” In Productive Low Resistivity Well Logs of the Offshore Gulf of Mexico, Dwight Moore,
Editor in Chief, Published jointly by the New Orleans Geological Society and Houston Geological
Society.
Simandoux, P., 1963: “Measures dielectriques en milieu poreux, application a mesure des saturations
en eau: Etude du Comportment des Massifs Argileux,” Revue de L’Institut Francais du Petrole,
Supplementary Issue.
Twenhofel, W.H., 1937: “Terminology of the fine-grained mechanical sediments,” National Research
Council, Division of Geology and Geography, Annual Report for 1936-1937, Appendix I, Exhibit F,
p.81-104.
Thomas, E.C, and Stieber, S.J., 1975: “The distribution of shale in sandstones and its effect upon
porosity,” Society of Professional Well Log Analysts, 16th Annual Logging Symposium Transactions,
June 4-7.
Visser, R., Baur, K.A.T., and von Baaren, J.P., 1988: “Effective porosity estimation in the presence of
dispersed clay,” Society of Professional Well Log Analysts, 29th Annual Logging Symposium
Transactions, Paper BB.
Waxman, M.H. and L.J.M. Smits, 1968, "Electrical Conductivities in Oil-Bearing Shaly Sands," Soc.
Pet. Eng. J. (June 1968) Trans., AIME, Vol 8, no. 2, p. 107-122.
Waxman, M.H. and E.C. Thomas, 1974, Electrical Conductivities in Shaly Sands-I. The Relation
Between Hydrocarbon Saturation and Resistivity Index; II The Temperature Coefficient of Electrical
Conductivity, J. Pet. Tech. (February 1974) p. 213-225, Trans., AIME, v. 257, p. 213.
Wyllie, M. R., A. R. Gregory, and G. H. F. Gardner. 1958. "An experimental investigation of factors
affecting elastic wave velocity in porous media." Geophysics 23 (3).
Wyllie, M. R. J., and W. D. Rose. 1950. "Some theoretical considerations related to the quantitative
evaluation of the physical characteristics of reservoir rock from electrical log data," J. Pet. Tech. 189
(April):105-108.

You might also like