Anda di halaman 1dari 201

Master of Science in Petroleum Engineering

. . - ......‘sfe#
............................ ...
_ * ......... { „

PETROLEUM
GEOLOGY
D r M . A I a

Imperial College
. ondon

Centre for Petroleum Studies


D epartm ent of Earth Science and E ngineering
Royal School of M ines Building
Prince C onsort Road
London SW 7 2AZ
United Kingdom
CONTENTS
PAGE

1. INTRODUCTION: RESERVOIR FLUID AND ROCK PROPERTIES..................................... 4


FLUID DISTRIBUTION IN A RESEVOIR...................................................... ........... 4
WETNESS.................................................................................................................... 4
RESERVOIR ROCK PROPERTIES........................................................................... 5
POROSITY..................................................................................................... ....... 5
PERMEABILITY.............................................................................. ...................... 7
PORE GEOMETRY..................................................... .........................................8
OBJECTIVES OF WIRELINE LOGGING.,.......................................................................9
Qualitative Interpretation........................................................................................9
Quantitative Interpretation.... ................................................................................ 9
THE BOREHOLE ENVIRONMENT: INVASION EFFECTS..........................................10
MATRIX CONCEPT.........................................................................................................12
DATA ACQUISITION ...................................................................................................... 13
LOG DATA RECORDING FORMAT ..............................................................................19
TYPES OF LOGS ............................................................................................................20
NOMENCLATURE............................................................................................................20

2. ELECTRIC LO G S.....................................................................................................................24
THE SP LOG.....................................................................................................................24
RESISTIVITY LOGS................................................................. .......................................33
INVASION AND RESISTIVITY PROFILES....................................................................35
RESISTIVITY MEASUREMENT............................................................................. ....... 41
RECENT ADVANCES IN RESISTIVITY LOGGING............................................... 50
QUALITATIVE INTERPRETATION OF ELECTRIC LOGS............................................ 57

3. THE BOREHOLE COMPENSATED (BHC) SONIC LOG.....................................................60


LONG SPACING SONIC TOOL (LSS)........................................................................... 69
THE ARRAY SONIC TOOL (AST)...................................................................................69
DETECTION OF ABNORMAL PRESSURES................... ............................................. 73

4. RADIOACTIVE LOGS 76
THE G AM M A RAY (GR) LOG .................................................................................. 77
THE NATURAL GAMMA RAY SPECTROMETRY TOOL (NGS).................................80
THE NEUTRON LOG ................................................................ ................................... 87
THE FORMATION DENSITY COMPENSATED (FDC) LOG ..................................... 93
THE LITHO-DENSITY LOG.................................................................. ..........................98
DETECTION OF ABNORMAL PRESSURES................................................... ........... 103

5. THE ELECTROMAGNETIC PROPAGATION TOOL (EPT LOG) .................................... 104

6. THE NUCLEAR MAGNETIC RESONANCE LOG (NMR)................................................... 113

7. PLATFORM EXPRESS (PEX)............................................................................................... 126


>

8. LOG INTERPRETATION.............................................................. ...... ........... ..................... 132


QUALITATIVE INTERPRETATION ............................................................................ 133
WIRE LINE LOG CHARACTERISTICS OF POTENTIAL SOURCE ROCKS..... 135
QUANTITATIVE INTERPRETATION ......................................................................... 144

9. SHALY FORMATION INTERPRETATION..........................................................................177

10. INTRODUCTION TO DIPMETER AND FORMATION IMAGE LOG S............................ 182


THE DIPMETER LOG.................................................................................................... 182
FORMATION IMAGE LOGS...... ................................................................................... 190
ELECTRICAL IMAGE TOOLS................................................................................... 190
ACOUSTIC (ULTRASONIC) IMAGE TOOLS...........................................................191
IMAGE INTERPRETATION...................................... ...................................... ...192

SELECTED REFERENCES ............................. ........................................................................ 198

EXERCISES 202
1. INTRODUCTION: RESERVOIR FLUID AND ROCK PROPERTIES

FLUID DISTRIBUTION IN A RESERVOIR

The distribution of the fluids in a reservoir rock is dependent on the densities of the fluids and
the capillary properties of the rock. Being the lightest, gas occupies the uppermost zone (the gas
cap), which is underlain, respectively, by oil and then water. In the uppermost zone the pores
are filled mainly by gas while in the middle zone the pores are occupied principally by oil with
gas in solution. In the lowermost zone the pores are filled by water. A certain amount of water
occurs along with the oil in the middle zone, the proportion often being of the order of 10 to 30
per cent. Moving upwards across the water-oil-contact in the reservoir, there is a gradual
increase in oil saturation accompanied by a progressive decrease water saturation, giving rise to
a transition zone from pores occupied entirely by water to pores occupied mainly by oil. The
thickness of this transition zone depends on the densities and interfacial tension between oil and
water, and on the sizes o£ the pores. A similar transition zone occurs between oil and gas when
moving upwards across the oil-gas contact: oil saturation gradually decreases while gas
saturation gradually increases. There is also some water in the pores in the gas zone. It should
therefore be noted that although drawn as sharp boundaries on maps and cross-sections, fluid
contacts are not in fact sharp lines. The so-called gas-oil and oil-water contacts are generally
horizontal. However, in certain circumstances these fluid contacts are inclined, usually only very
gently.

WETNESS

The water found in the oil and gas zones is known generally as interstitial water. This interstitial
water occurs as collars around grain contacts, as a filling of pores with unusually small throats
connecting with adjacent pores and to a much smaller extent as wetting films on the surface of
the mineral grains. This is illustrated Figure 1.1, which shows an enlarged section through a
granular rock.
WATER GAS OR
COLLARS OIL

Fig 1.1 An enlarged section through a granular reservoir illustrating the distribution
of water and hydrocarbons and wettability (After Hobson, 1984)
In this case the reservoir is said to be water wet, which means that the hydrocarbons are not in
direct contact with the grains that make up the reservoir. The mineral grains are coated by a thin
film water which intervenes between them and the hydrocarbons. The water film owes its
existence to a greater force of attraction between the liquid and the grain surfaces than the
cohesive strength of the liquid itself.

Oil can also be a wetting agent, but gas cannot act as a wetting fluid as its physical properties
do not allow it to form a coherent film around the mineral grains. However, gas reservoirs which
were previously oil filled could become oil wet.

Knowledge of the nature of the agent wetting a reservoir is important as it affects the production
behaviour of the reservoir. It must also be considered when designing secondary recovery
programmes. P reservation^ the reservoir wettability in cores is thus important if the subsequent
laboratory tests of electrical properties and fluid flow behaviour are to be truly representative of
the reservoir characteritics.

RESERVOIR ROCK PROPERTIES

The most important properties of a reservoir are its porosity (Ф) and permeability (k). Porosity
determines the storage capacity of the reservoir, while permeability governs its ability to transmit
fluids. These important characteristics are discussed briefly below.

POROSITY

Porosity is expressed as a percentage of the bulk volume of the rock:


Ф = (pore volume)/(bulk volume) X 100

The most common range is 10% - 20% and the highest recorded porosity value is 37%. The
maximum theoretical porosity value is 47%. Fluids occur in the pore spaces within the reservoir
(Fig 1.2).

1-Ф
Fig1.2 Illustration of porosity and
matrix (After Schlumberger)
If ‘pore volume’ represents the total void space within the rock regardless of whether or not the
pores are interconnected, the figure obtained is referred to as the total or absolute porosity, ФА,
of the rock in question. Complete pore interconnection is rare in nature and most reservoirs
contain at least some isolated pores. What is more important in practice is the effective porosity,
ФЕ, which is the ratio between the interconnected pore spaces and bulk volume. ФЕ is usually
lower than ФAand the permeability of the rock depends on its effective porosity (Fig 1.3).

Rock with high ФАand negligible ФЕ and к Rock with high ФЕand к

Fig 1.3 Influence of interconnection on ФЕ and к (After Marshak, 2005)

However, this distinction does not arise in practice, as laboratory methods measure effective
porosity. Log-derived porosity values, on the other hand, approach ФА as most logging tools
respond to total rather than interconnected porosity.

Figs 1.4 and 1.5 provide illustrations of porosity in a sandstone and a carbonate reservoir
respectively.

Pore paces (in blue)

Fig 1.4 Photomicrograph of a Fig 1.5 Photomicrograph of


sandstone reservoir a carbonate
(dolomite) reservoir
PERMEABILITY

Permeability is a measure of the ability of the rock to transmit fluids and depends on the degree
of connection between the pore spaces, i.e. ол Ф е - Permeability is a complex quantity and is
>o, —- ----------- - ' '
influenced by several factors including flurd saturation. Fig 1.6 provides an illustration of
permeability in a granular reservoir. The ‘high permeability’ and ‘low permeability’ channels are
controlled by the diameter of the throats or passages connecting the pores: the smaller the pore
throats, the lower the permeability.

Sand grains
Connate w ater

----- -- --------------- - . liyi< perm eability


pore channel------------ pore channel

Fig 1.6 Illustration of permeability

When only one fluid is present and it fully saturates the rock, the permeability of the rock to that
fluid is a maximum and is called the absolute permeability, abbreviated to kA. When more than
one fluid is present, as is the case in most reservoirs, permeability to any one fluid is reduced.
The ability of a reservoir to conduct of one fluid in the presence of others depends on the
saturation of that fluid and is called its effective permeability, abbreviated to kE. kE changes as
the saturation of the fluid in question varies. In reservoir studies, a quantity known as relative
permeability, abbreviated to kr, is used. It is defined as the kE/ kA ratio, its value ranges from zero
to one (depending on fluid saturation) and is expressed as k0/kA for oil, kG/kA for gas and kw/kA
for water.

Permeability is expressed in darcy units. A darcy is the permeability that allows a fluid of one
centipoise viscosity to flow at the rate of one cubic centimetre per second under a pressure
gradient of one atmosphere per centimetre. In practice, however, the darcy is too large, as the
permeability of most reservoirs is considerably less than one darcy. Permeability is therefore
usually expressed in millidarcys, abbreviated to md.
PORE GEOMETRY

In addition to the diameter of the throats or passages, permeability depends also on the way in
which the pores are connected, or on pore geometry. This is illustrated in Fig 1.7, which shows
three types of flow path through a reservoir. ‘Easy’, ‘intermediate’ and ‘tortuous’ flow paths result
respectively from high, intermediate and low permeabilities which are caused by variations in
pore geometry. Ultimately, the flow path complexities affect the production rate.

(a) ‘Easy’ flow path resulting from


high permeability

(c) ‘Tortuous’ flow path resulting


from low permeability

Fig 1.7 Types of flow path through a reservoir (After Schlumberger)


OBJECTIVES OF WIRELINE LOGGING

Logging of oil wells was pioneered by the Schlumberger brothers in the 1920s and quickly
became established as an indispensable source of information in the petroleum industry. Great
advances have been made, particularly in the last 25 years, in logging techniques and the
acquisition and interpretation of wireline log data are now a sophisticated science.

Log interpretation has two aspects: qualitative and quantitative.

Qualitative Interpretation

(a) Identification of porous and permeable beds and their boundaries.


}
(b) Identification of the pore fluids.

(c) Correlation of subsurface strata.

(d) Facies analysis: determination of grain size profiles, diagnosis of depositional


environments and the prediction of the trend of the porous and permeable beds in the
subsurface. However, facies analysis should always be undertaken in conjunction with
independent geological information (e.g. sedimentological observations and core
descriptions) and not on the basis of log responses alone.

Quantitative Interpretation

(a) Quantification of porosity (Ф) and permeability (k).

(b) Calculation of water saturation, Sw, in the uninvaded (by mud filtrate) part of a
hydrocarbon bearing zone, from which oil or gas saturation (Sh) may be deduced: Sh - 1*

Sw-

(c) Calculation of water saturation, Sxo, in the flushed (by mud filtrate) part of a hydrocarbon
bearing zone, from which residual oil or gas saturation, Sor, may be deduced: Sor = 1-Sxo-
A comparison of Sw and Sx0 will provide an indication of the moveable oil saturation
(MOS) in a hydrocarbon bearing zone.

(d) Estimation of the fractional volume of shale (VSh) in a given zone. This is necessary for
making corrections to log readings for the effects of shale.
THE BOREHOLE ENVIRONMENT: INVASION EFFECTS

Invasion is the result of the rotary drilling process which involves the pumping of a fluid (usually
a water- or an oil-based mud) down the inside of the drill pipe and returns to the surface through
the annular space between the drill pipe and the sides of the borehole (Figure 1.8). Invasion
affects only the porous and permeable zones; tight formations permit little or no invasion.

m ud

11

During drilling the mud pressure in the annulus, Pm, must be kept greater than the hydrostatic
pressure of fluid in the formation pores, Pr to prevent a blowout. The differential pressures, Pm-
Pr, which is typically a few hundred psi, forces drilling fluid into the formation. As the mud filters
into the porous layers, it displaces some of their content, replaces them with mud filtrate, and
creates a cylindrical fluid distribution pattern. At the same time, the filtration effect of the process
causes the deposition of some of the material suspended in the mud on the porous rock faces
surrounding the borehole wall. As the mud cake thickens, its low permeability causes it to form a
barrier, and eventually the flow of filtrate into the porous layers virtually ceases. The thickness of

the mud cake is generally between 1/8 in and V* in.

In the immediate vicinity от tne borehole, almost all the formation water and some of the
hydrocarbons, if present, are displaced. This is referred to as the flushed zone,, the width of
which is usually between 3 and 4in. Away from the borehole the effect of the flushing becomes
progressively less marked. The flushed zone is therefore surrounded by a transition zone
beyond which lies the uninvaded part of the porous layer (Fig 1.9). As shown in Fig 1.10,
invasion brings about a cylindrical distribution of the fluids with respect to the axis of the

borehole.

Form ation w ater

U ninvaded zone

M ixture o f m ud filtrate
and form ation water

T ransition zone
Ш Oil

W ater
~1

M ud filtrate

Flushed Zone

Fig 1.9 Invasion effects in a permeable zone (Schlumberger)


IN V A D E D Z O N E

o rig in a l d r illin g mud


fo rm a tio n flu id s filtr a te

d e p th o f in v a s io n

d ia m e te r o f in v a s io n

Fig 1.10 Invasion produces a cylindrical distribution of the


displaced fluids around the well (After Rider, 1996)

Factors that determine the depth of invasion include the type of mud, the differential pressure
between the mud in the borehole and the formation, and the porosity and permeability of the
formation. The most important of these factors, however, are porosity and permeability. Once
the mud cake builds up, due to its low permeability relative to that of the average formation,
almost all of the pressure differential ( P m- P r ) is across the mud cake and little is applied to the
formation. Consequently, in a given time the same volume of fluid will invade different
formations, regardless of their porosities or permeabilities (unless permeability is below about
1.0 md). This means the depth of invasion will be minimum at high porosity where large storage
space is available to accommodate the invading fluid and maximum at low porosity where little
room is available. It is approximately proportional to Other factors being constant, invasion
depth will double as porosity decreases from 36% to 9%, for example.

MATRIX CONCEPT

To a geologist the term 'matrix' refers to the fine-grained material that occurs between sediment
grains and tends to inhibit porosity and permeability. In wireline log interpretation, by contrast,
the term 'matrix' has an entirely different connotation; it refers to the actual mineral grains that
comprise the bulk of a sedimentary rock (Fig 1.11). Certain matrix properties of rocks such as
grain density (pma) and grain acoustic interval transit time (tma) must be considered in the
interpretation of some types of logs. In non-porous rocks bulk density (pb) and interval transit
time (t) measurements approach the values associated with pure minerals (i.e. all matrix, no
porosity).

Matrix
Petrophysical sense

Matrix
Geological sense

Fig 1.10 Geological and Petrophysical definitions of matrix

rt^cCO ^ п ч е Н olo. ViMsO^,


DATA ACQUISITION

A variety of methods are used in the acquisition of log data.

Conventional wireline (WL) logging involves lowering a special instrument down the well. The
instrument is attached to a calibrated cable which also carries the power supply to the tool. It is
lowered into the well and then pulled up, providing a continuous record of the rock
characteristics that the device is designed to detect. To minimise costs, a number of logs are
recorded simultaneously. A logging string is typically 3 5/8in in diameter and 25 to 60ft long,
consisting of several different tools as shown in Fig 1.12. Logging speed range is between 1,800
and 5,400ft/hr and is kept constant during individual surveys. The most commonly used is
18,00ft/hr, the maximum speed for the acquisition of radioactive log data.

1
- >4J а}
03 сз
1= cs
« *
А
Е >, <3 v_
CO
£
С V3
сэ
Oi __
e а>

_p_

я Е а>
о : 04 о Е
о со
| о а>
о 5 _ _J
о ll-
<SJ J2 со со
1 1
с CJ с
! со _о со о
! со
с 1 3
;
с с
3 "3
ГУ э
с •О “
У У
IE S -G R D IL-G R IS F -G R

D ll-S O N IC -G R IS F -S O N IC -G R D I L - R x o - C N l- G R F D C -C N L -G R

Fig 1.12 Logging tool combinations (After Schlumberger)

The advent of horizontal drilling has led to the development of techniques that involve pushing
logging tools down boreholes since it is generally not possible to transport conventional wireline
devices into horizontal or highly deviated wells.

Horizontal drilling has been found to be beneficial in marginal fields, in thin reservoirs, in
reservoirs with thin hydrocarbon columns or in fractured pay zones. In fields with thin reservoirs
or limited oil columns, horizontal wells expose the drainhole to a much larger reservoir area and
minimise water or gas coning since they induce lower drawdown pressures than conventional
wells. Since most fissures are near vertical in fractured reservoirs, a horizontal well will intersect
many more of them than a conventional borehole.

The overall result is to speed up production and reduce hydrocarbon recovery costs since fewer
wells are required to sweep a field. It has also been suggested that horizontal wells increase
recoverable reserves. Horizontal drilling has increased progressively worldwide since the 1990s
and it is estimated that the proportion of horizontally drilled wells in the USA now exceeds 50%.

Since wireline tools will not "fall" in highly deviated and horizontal wells, they must be pushed
down the borehole to reach the target. This may be achieved by using drillpipe conveyed or
coiled tubing techniques.

In drillpipe conveyed logging, conventional WL devices are attached to the end of the drillpipe
and pushed to the intervals of interest. A swivel head is usually used to join the device to the
drillpipe to allow preferential tool orientation. Power is supplied by a wireline cable which is
pumped through a side entry sub and down the drill string where it wet connects to the logging
device. Fig 1.13 shows an illustration of the drillpipe conveyed logging arrangement.

Fig 1.13 Drillpipe conveyed tools for logging in horizontal wells


Coiled tubing (CT) may also be used to convey conventional WL tools to the zones of interest.
The tubing is l.5in in diameter and is coiled around a reel, as shown in Fig 1.14. The logging
device is attached to the end of the tubing which is fed into the well by rotating the reel. A
wireline cable passing through the tubing supplies the power to the logging tool.

Fig 1.14 Coiled tubing conveyed gun for perforating in


horizontal wells (After Schtumberger)

Incorporating sensors into drill collars also provides a means of obtaining logging data during
drilling. The techniques are referred to as Measurement While Drilling (MWD) and Logging
While Drilling (LWD). Nowadays, MWD tools primarily provide control on well depth, inclination
and direction (azimuth). They also record some formation parameters such as resistivity and
gamma radiation. Fig 1.15 shows a diagrammatic representation of an MWD tool.

Oownhcie Steerabie rotary


Density and weight on bit, drilling ted
porosrty Dua: resistivity: ctowrfioie torque,
gamma ray and multi-axis shocks and
annular pressure directional information

Fig 1.15 MWD logging tool (After Schlumberger)

LWD devices include more advanced tools which record resistivity, gamma ray, formation
density, neutron porosity and sonic logs.

MWD/LWD data may be transmitted directly to the surface (real-time data) or stored in memory
chips in the tool. In real time transmission the measurements are converted into mud pulses
which are decoded by the surface data processing system. Data stored in memory chips are
down-loaded onto the surface computer system when the tool is recovered from the borehole.
Fig 1.16 is a diagrammatic illustration of the data transmission modes used in MWD/LWD
operations.

REAL TIME WITH MWD DOWNHOLE


MEMORY

Fig 1.16 MWD/LWD data transmission system


(After Schlumberger)
/ Advantages of MWD/LWD include:

V (a) Savings in rig time.

(b) Current LWD measurements provide resistivity, gamma ray, neutron, density, sonic and
formation image logs as well as pore pressure.

(c) Provision of real time data helping in optimising drilling operations—early detection of
pore pressure changes that require mud weight adjustment, selection of casing and
coring points and continuous directional information; and

(d) "Insurance” logs in case of loss of hole.

Future developments include the introduction of tools measuring microresistivity.

Figure 1.17 provides a comparison between LWD and WL logs. The recording shows that MWD
logging results compare reasonably well with those obtained by WL measurements.

NEUTRON RE5ISTJVFTT

UIRSLJME U fR E lIW - ILO


-0 .1 5
NPH1/f t ; 3
0 .*4 F -0
lL u /Я»3
200. С

ffclO WIRELINE - iL f!
CN0/R;.' j
-1S.0 « 5 .С 200. С

t ГГнО
i EWR/fi;i
?.o c o o .a

,J i
<

<

Z5*

1i
j

I
| i:
1 l! i
! i 1
j | j
г

£ j

Щ
— T r ^ tlt
i;

% i
;

| |
!

i|
.
Fig 1.17 Comparison of WL and LWD logs (After Schlumberger)
Nowadays, LWD is used in the drilling of most production and infill wells.

LOG DATA RECORDING FORMAT

Traditionally, the various measurements are presented graphically alongside a depth-scale. In


the petroleum industry, the API (American Petroleum Institute) grid is the standard log data
recording format. The total width of the log grid is 8.25in, and it is divided into three curve tracks

and a narrow column for recording the depth.

Track one is to the left of the depth column, and tracks two and three are to its right. Each track
is 2.5in wide, while the width of the depth column is 0.75in. The tracks are divided or scaled, the

divisions being referred to as the grid scale.

Three types of grids are in use: linear, logarithmic and split (Fig 1.18). Track one is always linear
while tracks two and three may be linear, logarithmic or split, depending on the data recorded.
The linear grid begins at zero while the logarithmic scale starts at 0.2 and covers a much greater
range of values. It is used for recording parameters that show large variations such as resistivity,
a common scale range being 0.2 ohm-m to 2000 ohm-m.

Fig 1.18 The three common log grids


(After We Iex)
DIGITAL LOGS

Nowadays log data are recorded digitally. Digital recording is not a new phenomenon and digital
logs have been available since the 1960s. Their use, however, dates from the 1980s with the
application of computers to the evaluation of log data through the development of interpretation
programmes and their rapid proliferation in the industry. A large variety of interpretation software
is now available commercially from the major service companies (Schlumberger, Baker Atlas
and Haliburton) and specialist consultancies and many operators have developed their own in
house interpretation programmes.

Log data recordings on CD-Roms have been available since 1985 but did not gain favour in the
industry due to security problems.

The advent of workstation based interpretation further expanded the use of digital logs.
Worksations allow the integration of geophysical and log data in the interpretation of seismic
sections.

TYPES OF LOGS

A large variety of wireline logs is currently in use. Acquisition of some requires an open hole
(uncased) and the presence of a conductive mud, while others may be run in wells containing
non-conductive drilling fluids (such as an oil-based mud, gas or air), or even in cased wells.

As mentioned above, this course is concerned with the acquisition and interpretation of open
hole logs, and these may be conveniently classified as follows:

1. Electric logs.
2. Acoustic or sonic log.
3. Radioactive or nuclear logs.
4. Electromagnetic Propagation Tool (EPT).
5. Nuclear Magnetic Resonance log (NMR).
6. Dipmeter and formation image logs

NOMENCLATURE

The subject lends itself well to the use of abbreviations and symbols. A large number of these,
referring to the various properties of the formations penetrated by the well, borehole parameters
and the measurements made in logging, is in use. The most commonly used abbreviations and
symbols are listed in Tables 1.1-1.7 (Schlumberger).

TABLE 1.1 FORMATION CHARACTERISTICS, DRILLING MUD AND BOREHOLE


PARAMETERS

a Tortuosity factor Ф Porosity

BHT Bottom Hole Temperature Фа Absolute porosity

Фе
BS Bit size Effective porosity
(PHIE)

BVW Bulk Volume Water SPI Secondary Porosity Index

Moveable Oil Saturation


CALI Caliper MOS
(Sxo ■ Sw)

dh Diameter of borehole m Cementation factor

d. Diameter of flushed zone n Saturation exponent

Residual Oil Saturation


dj Diameter of invaded zone ROS
(1.0 -Sxo)
Hydrocarbon saturation
EFT Estimated Formation Temperature sh
(1.0 -Sw)

F Formation Factor Sw Water saturation of uninvaded zone

hmc Thickness of mudcake Swi Irreducible water saturation

Water saturation of
к Permeability Sxo
flushed zone

kA Absolute permeability Sw/Sxo Moveable hydrocarbon index

kE Effective permeability T Formation temperature

Fractional volume of shale in


kr Relative permeability Vsh formation
TABLE 1.2 ELECTRIC LOGS

AIT Array Induction Tool PL Proximity Log


Cased Hole Formation Resistivity
CHFR R
Resistivity Tool
Rilm Resistivity Induction Log
DIL Dual Induction Laterolog Medium
DLL Dual Laterolog R|_Ld Resistivity of Laterolog Deep

HRLA High Resolution Laterolog Array Rlls Resistivity of Laterolog Shallow

IDPH Induction Deep Phasor RLL8 Resistivity of Laterolog 8

IL Induction Log Rm Resistivity of drilling mud

ILD Deep Induction Log Rmc Resistivity of mudcake

ILM Medium Induction Log Rmf Resistivity of mud filtrate

IMPH Induction Medium Phasor Rmll Resistivity of Microlaterolog


Rmsfl Resistivity of MicroSpherically
LL Laterolog Focused Log
Ro Resistivity of 100% water
LLD Deep Laterolog saturated formation (‘wet resistivity’)
LLS Shallow Laterolog Rs Resistivity of adjacent beds

LL8 Laterolog 8 R, Resistivity of uninvaded zone

LN Long (64” ) Normal Rw Resistivity of formation water

MINV Micro Inverse curve Rxo Resistivity of flushed zone

ML Microlog SN Short (16") Normal

MLL Microlaterolog SP Spontaneous Potential

MNOR Micro Normal curve PSP Pseudostatic Spontaneous Potential

MSFL MicroSpherically Focused Log SSP Static Spontaneous Potential

TABLE 1.3 ACOUSTIC (SONIC) LOG


t (At;
AST Array Sonic Tool Interval transit time of formation
Dt)
Bcp Acoustic porosity compaction Interval transit time of fluid in formation
tf
factor
вне Borehole Compensated Sonic Interval transit time of formation matrix
tma
Log
LSS Long Spacing Sonic Log
TABLE 1.4 RADIOACTIVE LOGS
CGR Computed Gamma Ray Pe (P E F ) Photoelectric absorption factor
CNL/CNT Compensated Neutron Log/Tool pb (RHOB) Bulk density of the formation
FDC Formation Density Compensated
Pf Density of fluid in formation
Log
GR Gamma Ray Log Ph Hydrocarbon density
G Rclaen Gamma Ray reading from clean
Pma Density of the formation matrix
zone
G R shale Gamma Ray reading from shale SGR Standard Gamma Ray
G Rzone Gamma Ray reading from
SNP Sidewall Neutron Porosity
formation
GST Gamma Ray Spectrometry Tool TDT Thermal Decay Time Log
LDT Litho-Density Tool TNPH Total Neutron Porosity
Natural Gamma Volumetric photoelectric
NGS U
Ray Spectrometry Log absorption index
NPHI Neutron porosity

ATBLE 1.5 ELCTROMAGNETIC PROPAGATION TOOL (EPT)


EATT Electromagnetice signal attenuation rate
tpl EPT travel time of formation
tp m a EPT travel time of matrix
tp o EPT travel time of a low attenuation (lossless) medium
tp w EPT travel time of water

TABLE 1.6 NUCLEAR MAGNETIC RESONACE LOG (NMR)

ADEPT Adaptable Electromagnetic Propagation Tool


CMR Combinable Magnetic Resonance Tool
EATT Electromagnetic Wave Attenuation
FFI Free Fluid Index
MRIL Magnetic Resonance Imaging Log

TABLE 1.7 PLATFORM EXPRESS


AIT Array Induction Imager Tool
HALS High Resolution Azimuthal Laterolog Sonde
HGNS Highly Integrated Gamma Ray Neutron Sonde
HLLD High Resolution Deep Laterolog
HRMS High Resolution Mechanical Sonde
MCFL Micro-Cylindrically Focused Log
TLD Three Detector Lithology Density Tool
2. ELECTRIC LOGS

INTRODUCTION

Methods of measuring the electrical properties of rocks penetrated by boreholes were the first to
be developed and used in the petroleum industry. The instrument used consists of a system of
electrodes attached to a cable which carries also the electric current. The acquisition of most of
the electric logs requires an open or uncased well containing a water-based, conductive drilling
fluid. Only the electric induction log can be run in the presence of a non-conductive drilling fluid
such as an oil-based mud (see below). Drilling with non-conductive fluids, therefore, limits the
choice of the electric logs that can be run.

Electric logs fall into two main categories: the spontaneous potential (SP) which measures a
naturally occurring phenomenon, and the resistivity devices which record the resistance offered
by the rocks surrounding the borehole to the passage of an electric current.

THE SP LOG

The SP curve is a recording versus depth of the difference in electric potential between a fixed
electrode at the surface and a moving electrode in the borehole. It is measured in millivolts, and
there is no absolute zero; only changes in potential are recorded. It is recorded on track 1, and is
always linear. The SP log has the following applications:

(a) Identification of permeable beds and the location of their boundaries.

(b) Determination of the formation water resistivity in the uninvaded zone (Rw).

(c) Estimation of the degree of shaliness of reservoir rocks.

Two types of potential may contribute to the SP effect. These are the electrochemical (Ec) and
the electrokinetic (Ek) potentials. The latter, also known as the electrofiltration or streaming
potential, is in most cases negligible, and in log analysis the observed SP response is assumed
to be solely due to the electrochemical component. The origins of these potentials are discussed
briefly below.

Ek

In general, an Ek is produced by the flow of an electrolyte through a porous, non-metallic


medium. In the case of the SP response, the Ek results from the movement of filtrate through the
mud cake that builds up on a porous and permeable formation. Its magnitude is influenced by a
number of factors, the most important of which are the differential pressure producing the flow
and the resistivity of the formation water (Rw)- During the initial stages of mud cake formation Ek
is significant, but as the mud cake thickens its permeability diminishes rapidly, causing it to
isolate the porous bed from the borehole. As the flow of filtrate into the porous bed virtually
ceases, all the differential pressure is expended on the mud cake, and this effectively ends the
generation of Ek. However, in cases where unusually high differential pressures prevail, Ek
effects may be substantial. Such cases result from drilling with very heavy muds, or when
low-pressure formations are penetrated. Large Ek effects may also be observed in very low
permeability (less than 5 md) formations. Low permeability results in a low rate of filtrate
invasion, and this in turn means that little or no mud cake will build up. Consequently, the
formation remains in communication with the borehole, and nearly all the differential pressure is
applied to the formation. If the formation is clean (shale-free), contains brackish water and the
drilling mud is resistive, the low permeability Ek effect may cause a large deflection in the SP
curve. Such a deflection cannot be used in quantitative interpretation, nor is it indicative that the
zone involved will produce any fluid.

Although these effects occur infrequently, the conditions that cause them are a possible source
of large Ek values.

Ec

The Ec component is the main source of the deflections in the SP curve and results from the
transfer of ions from a more concentrated electrolyte (usually the uninvaded zone formation
water) to a less concentrated electrolyte (usually the mud in the borehole) through a
semi-permeable membrane (e.g. a sand-shale contact). Sodium chloride is the main source of
ions both in the formation water and the drilling mud.

The transfer of ions constitutes an electric current, and as shown in Fig 2.1, these currents flow
through four different media, namely, the mud in the borehole, the invaded part of the porous
and permeable bed, the uninvaded part of the same and the surrounding shales. Movement of
ions takes place in two ways: (a) through the shales, above and below the porous and
permeable bed, and (b) at the boundary between the invaded and uninvaded parts of the porous
and permeable bed where two solutions of different salinities are in direct contact. The
movement of ions through the surrounding shales gives rise to a membrane potential, while the
direct transfer of ions at the invaded - uninvaded zone boundary produces a liquid

Fig 2.1 Diagrammatic representation of the membrane


potential component of the SP (After Welex)

junction potential. The sum of these two independent potentials makes up the electrochemical
component of the SP phenomenon.

The membrane potential is related to the selective passage of ions through the shales above
and below the porous and permeable bed. Due to their layered structure and the charges on the
layers, shales are permeable to the Na+ cations but impervious to the СГ anions. When a shale
separates sodium chloride bearing solutions of different salinities, the Na+ cations move through
the shale from the more concentrated solution. This movement of charged ions is an electric
current, and the force causing them to move constitutes a potential across the shale. The curved
arrow in the upper half of Fig 2.2 shows the direction of current flow corresponding to passage of
Na+ ions through the adjacent shale from a more saline formation water in the bed to the less
saline mud.

Fig 2.2 Diagrammatic representation of the liquid junction


component of the SP (After Schlumberger)

The liquid junction potential arises from the transfer of ions across the invaded
zone/uninvaded zone interface. Here Na+ and СГ ions can transfer from either solution to the
other. Since СГ ions have a greater mobility than Na+ ions, the net result is a flow of negatively
charged particles from the more concentrated solution to the less concentrated solution. This is
equivalent to a conventional current flow in the opposite direction, indicated by the straight
arrow, A, in the upper half of Fig 2.2. The liquid-junction potential is only about one-fifth the
membrane potential.

The first step in the interpretation of the SP log is the establishment of 'sand' and ’shale’ lines as
shown in Fig 2.3. These are arbitrary limits, with the former normally representing the maximum
deflection to the left, the latter representing the maximum deflection to the right. Deflections to
the left of the shale line are regarded as normal or negative, and correspond to porous and
permeable zones containing a more saline interstitial water than the drilling mud (i.e. Rw< Rmf).
In these cases the SP currents flow in the direction shown in Fig 2.4.
Fig 2.3 SP log presentation in a sand-shale sequence (After Schlumberger)

If the mud is more saline than the formation water (this could happen if the mud is very salty or
the formation water is brackish or fresh in which case Rw> R m f), the SP currents flow in the
opposite direction to that shown in Fig 2.4, and the corresponding deflection will be to the

Fig 2.4 S P response associated with a clean, permeable bed containing


formation water more saline than the drilling mud filtrate (Rw < Rmf)
(After Welex)
right of the shale line. Such a deflection is considered as reversed or positive (Fig 2.5).

Fig 2.5 Example of a reversed SP deflection

If there is no salinity contrast between the mud and the formation water (i.e. Rw= Rmf), no SP
currents are generated, and no deflection will be observed in the SP curve - there will be no
departure from the shale line (i.e. SP = 0).

Several factors affect the amplitude of the SP deflections. These are shown diagrammatically in
Fig 2.6 and include:

(a) Bed thickness - The SP deflections associated with porous and permeable formations less
Rrf»R„
+ + for all sands

Shale Shale
Baseline Baseline

Thick clean
SSP
wet sand

sp < T

Thick shaly
PSP wet sand

SP Thick clean
gas sand

Thick shaly
PSP
gas sand

mm

Fig 2.6 Illustrations of the factors that influence the amplitude and
shape of SP deflections (After Asquith & Krygowski, 2004)

than 10ft thick are narrow, rounded and reduced in amplitude. Responses associated with
beds thicker than 20ft, by contrast, are well developed and flat-ended. In both cases the bed
boundaries should be placed at points of inflection on the SP curve.

(b) Bed resistivity - High resistivities (usually resulting from very high hydrocarbon
saturations)reduce the SP deflection.

(c) Shaliness - The presence of shale in a porous and permeable bed reduces the SP
deflection.

(d) Contrast between Rmf and Rw - This is the most important factor. The larger the salinity
contrast between the formation water and the mud filtrate, the more pronounced the
corresponding SP deflection. It should be emphasised that there is no direct relationship
between the value of permeability and the amplitude of the SP deflection, nor does the
deflection bear any direct relation to porosity; a high porosity and permeability zone is not
necessarily associated with a high amplitude SP response.
I

In summary, the SP log qualitatively distinguishes between permeable and impervious beds,
and provides information on the salinity of the formation water relative to that of the drilling mud.
The quantitative application of the SP curve will be discussed later.

STATIC SP (SSP)

As stated above, the amplitude of the SP deflection associated with thin beds is reduced. The
full SP, or the SSP, is developed only in thick, clean (shale-free) and water bearing zones in
which Rmf - Rw. The minilmum bed thickness required for the development of the SSP is about
20ft. In thin beds a correction must be applied to the SP reading in order to obtain the SSP. This
is done by reference to charts provided by the various service companies. One such chart is
presented in Fig 2.7. All these charts are entered by plotting Ri/Rm against bed thickness, and

Fig 2.7 SP bed thickness correction (After Baker Atlas)

reading the correction factor on the x-axis. It should be noted that Rm must be at formation
temperature. This value of Rm is obtained by converting mud resistivity at surface temperature
(measured and recorded on the log heading by the service company engineer) to its
corresponding value at formation temperature.
Knowledge of the SSP is essential for the derivation of Rw which in turn is required for the
calculation of water saturation in the uninvaded zone. There is an equation which relates the
SSP to the conductivities of the mud filtrate (cmt) and the formation water (cw):

SSP — К log (cw/Cmf)i 2.1


2. v

where К is a constant, the value of which is dependent on the formation temperature. Usually,

К = 61 + 0.133 T(°F) or К = 65 + 0.24T (°C).

In practice, however, cmf and cw are of little value, as they cannot be readily quantified. It would
therefore be more useful to express these in terms of measurable quantities, namely, Rw and
Rmf. For pure sodium chloride solutions that are not too concentrated, resistivities are inversely
proportional to chemical activities (Fig 2.8), and equation 2.1 may therefore be written as:

SSP = - К log (R m f/R w ) 2.2

N a + A ctivity (G r-lon/Liter, Total Na)

Fig 2.8 Na+ NaCI Resistivity relationship (After Schlumberger)

However, as shown in Fig 2.8, resistivity and chemical activity are no longer linearly related in
solutions containing more than 0.7 gm-ions/litre of Na+. Consequently, R mf and Rw must be
converted to values that are linearly related to their respective chemical activities. These values
are referred to as equivalent resistivities, and are denoted by R we and R mfe- Thus the standard
equation that relates the SSP to the mud filtrate and uninvaded formation water resistivities is:

SSP = - К log (R m fe /R w e ) 2.3


Equation 2.3 is used in the quantitative interpretation of the SP log. Equivalent resistivity values
are derived through the use of the chart in Fig 2.9.

0.001

0.002

0.005

0.01

0.02

f
s
3
J 0.05
CC
о
2
CC 0.1

0.2

0.5

1.0

2.0
0.005 0.01 0.0 2 0.03 0.05 0.1 0.2 0.3 0.5 1.0 2 3 4 5

R„ or Я,* (ohm-m)

Fig 2.9 Rw" Rwe" formation temperature relationships


(After Schlumberger)

RESISTIVITY LOGS

Resistivity logs measure and record the resistance offered by the rocks surrounding the
borehole to the passage of an electric current. It is a fundamental property of a material, and is
defined as the electrical resistance of a one metre cube (i.e. a cube 1m x 1m x 1m) of the
material concerned (Fig 2.10).

Fig 2.10 Illustration of resistivity

1m
Electric current
Resistivity (R) is the reciprocal of conductivity (c):

R = 1/c 2.4

Resistivity is related to electrical resistance by the following equation:

R = rA/L 2.5

r = resistance in ohms (Q)


A = cross-sectional area of the conducting medium (m2)
L = length of the conducting medium (m)
Substituting for Q resistance, m2 for A, and m for L in equation 2.5:

R = £2m2/m 2.6

Resistivity is therefore expressed in ohms m2/m or ohm.m.

In sedimentary rocks the ability to conduct is related to the movement of ions present in the
formation water; clean, dry reservoir rocks and hydrocarbons are insulators, characterised by
low conductivity and therefore high resistivity. Consequently, the only part of a formation which
conducts electricity is the interstitial water. The resistivity of the water depends on the quantity of
dissolved salts present (mostly NaCI); the more saline the formation water, the higher its
conductivity and the lower its resistivity. Temperature is another important factor. Ionic activity
increases with increasing temperature, lowering resistivity.

In summary, the factors influencing the resistivity of a clean (shale-free) rock are:

1. Formation water resistivity (Rw)


2. Temperature
3. Presence of hydrocarbons
4. Magnitude of porosity (Ф).

The influence of porosity is due to the fact that conductivity depends on the number of ions
available in a solution to carry the electric charge; other factors being equal, the higher the
number of ions, the greater the conductivity. The number of ions depends on water-filled
connected porosity; therefore, the higher the porosity, the greater the number of available ions
and the higher the conductivity.

INVASION AND RESISTIVITY PROFILES

As already discussed, all porous and permeable beds penetrated by a borehole become
invaded by mud filtrate. The invaded formation consists of a flushed zone, close to the borehole,
surrounded by a transition zone which in turn is surrounded by the uninvaded or undisturbed
part of the porous and permeable bed where the original formation fluids remain uncontaminated
by the mud filtrate. The detailed distribution of these zones and the associated resistivities and
saturations are shown in Fig 2.11, and the various parameters shown were defined in Tables 1.1
and 1.2. Depth of invasion is a function of porosity and permeability, as discussed above.

t
(~~j Resistivity of the zone
О Resistivity of the water in the zone
Д Water saturation in the zone

Adjacent bed

Adjacent bed

(Invasion diameters)

dj/dj = 2 indicates high Ф and к


d/di = 5 indicates intermediate Ф
d /d p Ю indicates low Ф and к

Fig 2.11 Distribution of resistivity and saturation in an invaded formation


(After Schlumberger)

If the porous and permeable zone contains oil, the fraction of the original oil saturation displaced
from the flushed zone by mud filtrate invasion is represented by the difference between water
saturations in the flushed and the uninvaded zones (i.e. Sxo - Sw)- Usually, between 70% and
95% of the oil is flushed out; the remaining fraction is called residual oil, and its saturation, Sor,
equals 1 - SXo- In the uninvaded zone the original hydrocarbon saturation, Sh, remains intact,
and is given by the following equation:

Sh = 1 - Sw 2.7

Determination of Sh is one of the main objects of quantitative log interpretation.

It should be clear from the above discussion that the invasion of a porous and permeable bed
creates zones, radially distributed with respect to the borehole axis, containing different fluids
with different resistivities. This distribution of resistivity gives rise to resistivity profiles which
represent cross-sectional views of the invaded formation. There are three commonly recognized
invasion profiles: (a) step, (b) transition, and (c) annulus. These three invasion profiles are
illustrated in Figure 2.12.

STEP PROFILE TRANSITION PROFILE ANNULUS PROFILE


b o reh o le wall borehole wall
^borehole wall

*Distance from the borehole I D istance fro m the borehole

R0: Resistivity of the zone 100% saturated with formation water of resistivity R w- R o is also called ‘wet resistivity’

Fig 2.12 Resistivity profiles (After Asquith & Krygowski, 2004)

The step profile has a cylindrical geometry with an invasion diameter equal to dj. Shallow
reading, resistivity logging tools read the resistivity of the invaded zone (R), while deeper
reading, resistivity logging tools read true resistivity of the uninvaded zone (Rt).

The transition profile also has a cylindrical geometry with two invasion diameters: d. (flushed
zone) and d. (transition zone). It is probably a more realistic model for true borehole conditions
than the step profile. Three resistivity devices are needed to measure a transitional profile; these
three devices measure resistivities of the flushed, transition, and uninvaded zones, Rxo, Ri, and
Rt respectively (Fig 2.12). By using these three resistivity measurements, the deep reading
resistivity tool can be corrected to a more accurate value of true resistivity, Rt, and the depth of
invasion can be determined.

An annulus profile is only sometimes recorded on a log because it rapidly dissipates with time
and can be detected only by logging soon after a well is drilled. However, it is very important as
the profile can occur only in zones which bear hydrocarbons. As the mud filtrate invades the
hydrocarbon-bearing zone, hydrocarbons move out first. Next, formation water is pushed out in
front of the mud filtrate forming an annular (circular) ring at the edge of the invaded zone (Fig
2.12). The annulus effect is characterised by a higher Rt reading than a simultaneously recorded
Ri measurement.
)
Resistivity profiles are developed when three resistivity curves (Rxo, Ri and Rt) are recorded
/simultaneously. They are useful aids in quick-look qualitative interpretation; together with an SP
curve, the resistivity responses are used to (1) identify porous and permeable formations, and
(2) detect hydrocarbon-bearing zones. Because of their importance, resistivity profiles for both
water-bearing and hydrocarbon-bearing zones are discussed here. These profiles vary,
depending on the relative values of Rwand Rmf.

Water-bearing Zones

Fig 2 .1 3 illustrates the borehole and resistivity profiles for water-bearing zones where the
resistivity of the mud filtrate ( R mf) is much greater than the resistivity of the formation water ( R w)

in fresh water muds, where resistivity of the mud filtrate ( R mf) is approximately equal to the
resistivity of the formation water (R w) in salt water muds and where the mud filtrate resistivity
( R mf) is less than that of the formation water ( R w ). A fresh water mud (i.e. R mf > 3 R W) results in a
'wet' log profile where the shallow ( R xo), medium ( R ,) , and deep (R t) resistivity tools separate and
record high ( R xo), intermediate ( R j) , and low ( R t. ) resistivities. A salt water mud (i.e. Rw - R mf)

results in a wet profile where the shallow ( R xo), medium ( R ,) and deep ( R t) resistivity tools all
read low resistivity. Fig 2 .1 4 illustrates the resistivity curves for wet zones invaded with a fresh
water mud.
H o riz o n ta l .section th rou gh a
p erm ea b le w a te r-b ea rin g bed

(a)

(b)

Fig 2.13 Resistivity profiles in a water bearing zone invaded by freshwater (a)
and saltwater (b) muds (After Asquith & Krygowski, 2004)
1LP
2000
jLM _
F eet 0 .2 "2000
SFLU
MD 7Г т~

Fig 2.14 A Dual Induction-Spherically Focused Log suite (DIL-SFL) in a water


bearing zone where Rmf > 3RW(After Asquith & Krygowski, 2004)
Hydrocarbon-bearing Zones

Fig 2.15 illustrates the borehole and resistivity profiles for hydrocarbon-bearing zones where the
resistivity of the mud filtrate (Rmf) is much greater than the resistivity of the formation water (Rw)
for fresh water muds, and where Rmf is approximately equal to Rw for salt water muds. A
hydrocarbon zone invaded with fresh water mud results in a resistivity profile where the shallow
(Rxo), medium (R), and deep (Rt) resistivity tools all record high values. In some instances, the
deep resistivity will be higher than the medium resistivity. When this happens, an annulus is
present. A hydrocarbon zone invaded with salt water mud results in a resistivity profile where the
shallow (Rx0), medium (RJ, and deep (Rt) resistivity tools separate and record low (Rx0),
intermediate (Ri) and high (Rt) resistivities.

Horizontal section through


a perm eable oil-bearing bed

Fig 2.15 Resistivity profiles in a hydrocarbon bearing zone invaded by freshwater (a)
and saltwater muds (b) {After Asquith & Krygowski, 2004)
Fig 2.16 illustrates the resistivity curves for hydrocarbon zones invaded with fresh water mud.

JLP.
ohm-m
0.2 2000'
JLM _
ohm-m
Feet 0,2 '2000
SP SFLU
MD
-*-
O
-1
<
3
о

4
ohm-m
0.2 2000

!--
P 8700

fc
t=

r
SuL
—1 8
-
r~
t-

V-
'h

=t
- 1
r-
a i
--- -H


' X

“ 1~ ~
8800
1
-

Fig 2.16 A Dual Induction-Spherically Focused Log suite (DIL-SFL) in a hydrocarbon


bearing zone where Rmf> 3RW {After Asquith & Krygowski, 2004)

In conclusion, it is emphasised once more that invasion and the associated resistivity profiles
are unique to porous and permeable zones. Impervious formations remain uninvaded, and do
not therefore exhibit resistivity profiles. All three resistivity curves read approximately the same
value opposite an impervious bed.
RESISTIVITY MEASUREMENT

Resistivity devices are designed to measure Rxo, Ri and Rt, resistivities of the flushed (1-6in),
transition (0.5-3ft) and uninvaded (3+ft) zones respectively.

All deep and shallow/medium logs are obtained with electrodes or coils mounted on cylindrical
tools that are run more or less centralized in the hole. By contrast, the flushed zone
(microresistivity) curves are obtained with pad-mounted electrodes in contact with the borehole
wall. Nowadays the three curves are obtained simultaneously on a single pass in the hole.

Resistivity logging has advanced enormously since its introduction in the late 1920s. The
resistivity logs may be divided into conventional or non-focused devices (also known as
Electrical Survey tools - abbreviated to ES tools), focused tools, and induction systems. The
conventional resistivity devices are now obsolete, but many such logs survive in oil company
archives and it is therefore necessary to describe the way in which the tools functioned.

THE CONVENTIONAL RESISTIVITY LOGS

Until about 1950, all resistivity measurements were made with simple electrode systems shown
in Fig 2.17. These measurements produced a Short Normal (SN), a Long Normal (LN), and a
Lateral curve, depending on the spacing between the current electrode (A in Fig 2.17) and
voltage measuring electrode (M in Fig 2.17) in the borehole. The electrode spacing was 16in,
64in and 18ft 8in in the SN, LN and the Lateral curves respectively. In general, the greater the
spacing between A and M the greater was the depth of investigation. In the case of the Lateral
curve, there were two voltage recording electrodes (M and N in Fig 2.18) in the borehole, and
the resistivity was measured between these and the current electrode A. In practice, the
measurement was made between A and a point O. midway between M and N. All three curves
were recorded simultaneously.

The 16in Normal recorded Ri, while the 64in Normal and the Lateral curve responded primarily
to Rt. Fig 2.19 presents a suite of conventional resistivity logs.

The conventional logs were difficult to interpret. Extensive charts were required to correct for
borehole, bed thickness, and adjacent-bed resistivity effects. In particular, the curves were
relatively inaccurate for bed thicknesses less than about 1.5 times the spacing, i.e., 28ft for the
Lateral and 8ft for the Long Normal. The Short Normal curve was the most usable, but it was

Meter
Meter

Generator
H 0H Generator

Spacing

Spacing J MoJ

Fig 2.17 Normal device - schematic Fig 2.18 Lateral device - schematic
diagram (After Schlumberger) diagram (After Schlumberger)

Fig 2.19 A suite of conventional resistivity logs (After Baker Atlas)


severely affected by invasion. The basic problem with the conventional logs was that the
direction of the survey current was not controlled (Fig 2.20). It took the path of least resistance,
favouring conductive mud and conductive shoulder beds over high resistivity beds at the level of
the tool.

Fig 2.20 Schematic representation of Fig 2.21 Schematic representation of


focused and non-focused focused and non-focused
current flow from a current flow from a
logging tool microresistivity logging tool
(After Rider, 1996) (After Schlumberger)

A non-focused microresistivity device (for measuring Rxo) was also available. Known as the
Microlog, the device consisted of three electrodes, spaced 1 in apart, mounted on a pad and
made its measurement in contact with the borehole wall (Fig 2.21). It recorded a Microinverse
(also called the 1" x 1") and a Micronormal (also referred to as the 2") curve simultaneously.

The micronormal device investigated three to four inches into the formation, measuring Rxo, and
the microinverse investigated approximately one to two inches and measured the resistivity of
the mud cake, Rmc. The detection of mud cake by the Microlog indicated that invasion had
occurred and the formation was permeable. Permeable zones showed up on the Microlog as
positive separation when the micronormal curve read higher resistivity than the microinverse
curve (Fig 2.22). Shale zones were indicated by no separation or negative separation (i.e.
micronormal = microinverse).

The Microlog did not work well in salt water-based muds. These muds cause the formation of
conductive mud cakes, and the non-focused current tended to flow between the A and M
electrodes through the mud cake rather than penetrate the more resistive formation behind the
mud cake.
As a result of these problems, the Long Normal and Lateral curves were replaced in the 1950s
by focused logs in which the path of the survey current was controlled. The focusing minimized
borehole and adjacent bed effects and provided simultaneously both deep penetration and good
bed resolution.
SP

Fig 2.22 Example of a Microlog (After Asquith & Krygowsky, 2004)

FOCUSED RESISTIVITY LOGS

These devices were introduced in the 1950s and include the Laterologs. They are designed to
measure RXo, Ri and Rt in boreholes containing salt water muds, have excellent vertical
resolution (about 2ft) and their readings are little affected by the resistivities of the adjacent
beds.

Laterolog systems contain an array of electrodes to focus the survey current and force it to flow
laterally into the formations surrounding the borehole. Focusing is achieved by two bucking
electrodes that emit a current of the same polarity as the surveying electrode but are located
above and below it (Д and A’i in Fig 2.23). The focusing, or guard electrodes, prevent the
surveying current from flowing up the borehole filled with salt water mud. The effective depth of
Laterolog investigation is controlled by the extent to which the surveying current is focused.

Fig 2.23 Schematic diagram of a focused resistivity


logging tool (After Schlumberger)

Deep reading Laterologs (LLD in Fig 2.24) are therefore more strongly focused than shallow
reading Laterologs (LLS in Fig 2.23).

Fig 2.24 Schematic representation of deep and shallow


Laterologs (After Schlumberger)

Invasion can influence the Laterolog. However, because resistivity of the mud filtrate is
approximately equal to the resistivity of formation water when a well is drilled with salt
water-based muds, invasion does not strongly affect Rt values derived from a Laterolog. But,
when a well is drilled with fresh water-based muds (where Rmf > 3RW), the Laterolog can be
strongly affected by invasion. Under these conditions, a Laterolog should not be used. The
borehole size and formation thickness affect the Laterolog, but normally the effect is small
enough so that Laterolog resistivity can be taken as Rt.

Modern focused resistivity devices (Fig 2.25) include the deep Laterolog (LLD), the shallow
Laterolog (LLS and LL8) and microresistivity tools that are designed to measure RXo- The LLS is
usually combined with the LLD, the combination being called the Dual Laterolog (DLL); LLD
measures Rt, while LLS responds to R j. The DLL combination is recorded in tracks 2 and 3
together with a focused RXo log (Fig 2.26). A Gamma Ray curve is often displayed in track 1.
The LL8 forms part of the Deep and Medium Induction suite.
DUAL L A T E R O L O G S S P H E R IC A L L Y
s h a llo w deep F O C U S E D TOOL

Fig 2.25 Schematic representation of the Dual Laterologs (DLL)


and the Spherically focused Log (SFL)
) (After Rider, 1996)

Fig 2.26 A DLL-MSFL suite

Another R| measuring device is the Spherically Focused Log (SFL), and Fig 2.25 illustrates the
tool. It is offered as an alternative to the LL8 as part of the Deep and Medium Induction suite.
The SFL device carries nine electrodes, with the survey current emanating from the centre
electrode, A0 (Fig 2.25). Focusing electrodes enforce an approximately spherical shape on the
equipotential surface, and hence the name. The depth of penetration of the SFL is smaller than
that of the LL8. This means that the SFL gives greater weight to R j, which is desired, but in
general it still reads too deep to give an accurate measurement of flushed zone resistivity, Rxo.
The vertical resolution of the SFL and the LL8 is about 1ft.

Focused microresistivity (Rxo) devices include the Microlaterolog (MLL), the Proximity Log
(PL) and the Microspherically Focused Log (MSFL). As mentioned above, these are contact
type devices, mounted on pads and the measurement is made by pressing the tool against the
borehole wall by means of spring-loaded arms (Fig 2.27).

Fig 2.27 Schematic representation of microresistivity tool pads (After Rider, 1996)

In the Microlaterolog the survey current flows from the centre electrode, A0, and is focused by
the outer electrode Av It works well in salt water-based muds which result in the development of
conductive mud cakes. However, the use of the MLL is limited to cases where the mud cake
thickness does not exceed a quarter of an inch, otherwise the mud cake resistivity ( R mc) makes
a significant contribution to the R Mll reading.

The Proximity Log resembles the MLL in principle, except in this case the electrodes are
mounted on a somewhat wider pad. Its depth of investigation is also greater, and consequently
mud cakes up to three-quarters of an inch thick have little effect on RPL. On the other hand,
unless the diameter of invasion (dj) is 40in or more, RpL becomes affected by.

The Microspherically Focused Log operates under a wider range of conditions than either the
MLL or the PL. It gives good Rxo resolutions in the presence of thick mud cakes and does not
require an invasion depth as great as that necessary for the PL. It can be combined with the DLL
tool, and together with the LLD and LLS forms the DLL-MSFL suite (Fig 2.26).

THE INDUCTION LOG

The induction device measures the conductivity of the rocks surrounding the borehole by
inducing an electric current through them. Figure 2.28 illustrates the principle of the tool, which is
shown as consisting of one transmitter coil and one receiver coil. This simple two-coil system
does not, however, represent the tool currently in use, which is a multi-coil device. The response
of a multi-coil system is derived from all possible combinations of two-coil transmitter-receiver
pairs. These responses are added algebraically, providing information on conductivity.

Direct Coupling
(X Signal)
Receiver C oil

Secondary
Magnetic _
Field "■'-- » v
(Created 4 \
by the —*■''
Ground
Loop)

Transm itter Primary


Coil i , Magnetic Flux '
j (Created by 1
Constant ' Transm itter) i
Current ;

Tool movement

Fig 2.28 Diagrammatic representation of the induction tool (After Schlumberger)

The current induction tool is focused. Focusing improves vertical resolution by suppressing the
response of the adjacent beds (also known as the shoulder beds) and increases the depth of
investigation by reducing the influence of the mud and the part of the formation close to the
borehole wall.
Principle of Measurement

A constant, high frequency alternating current is sent through the transmitter coils. This
generates an alternating magnetic field which induces secondary currents (also known as
Foucault or eddy currents) in the rocks surrounding the borehole. These currents flow in circular
paths coaxial with the transmitter coils through the surrounding rocks. The resulting magnetic
field, in turn, induces signals in the receiver coils. These signals are proportional to the
conductivity of the formations from which resistivity is derived and recorded on the log.

Since the device does not require the transmission of the survey current through the mud, it can
be run in boreholes drilled with air, gas or an oil-based mud. It works well also in the presence of
conductive muds, provided the mud is not very saline, the borehole diameter is not very large
and the resistivities of the surrounding formations are less than 20 ohm-m.

Current induction systems include a deep reading device (ILD) which measures R t. and a
medium reading tool (ILM) which measures R j. The ILD-ILM combination is called the Dual
Induction Log (DIL) and, together with a shallow Laterolog (either LL8 or SFL), is recorded in
tracks 2 and 3, forming either a DIL-LL8 or a DIL-SFL suite as shown in Figure 2.29. Normally,
an SP curve is recorded in track 1.

RECENT ADVANCES IN RESISTIVITY LOGGING

Advances have been made in recent years in both focused resitivity and induction logging. The
tools and examples of the logs are discussed below.

HIGH-RESOLUTION LATEROLOG ARRAY TOOL (HRLA)

This tool provides five independent resistivity measurements that improve Rt resolution in thin
and deeply invaded formations. Enhanced focusing ensures that all signals are measured at the
same time and tool position, producing depth- and resolution-matched measurements.
Automatic corrections are carried out for borehole, shoulder or adjacent bed and invasion,
effects, yielding a more robust Rt.

The tool operates in six different ‘modes’ and delivers an array of five resitivity curves (RLA1-
IL D _
0.2 ohm-m 2000
SP _ILM_
-1 6 0 5V 40 Feet 0.2 ohm-m '2000
_____________ GR SFLU
0 API anils....................1 5 0 MD 0 .2 ohm-m 2000

RLA5 in Fig 2.30), each with increasing depth of investigation - RLA1 < RLA2 < RLA3 < RLA4 <
RLA5 - providing a detailed resistivity profile. The resistivity curve produced in Mode 0, not
shown in Fig 2.30, primarily represents the borehole environment and is used to estimate Rm.

Fig 2.31 shows an illustration of the HRLA tool.


Array Resistivity RLA1
2 (ohm-m) 200
MSFl
(Logarithmic scale) Array Resistivity RIA2 LLS
0.2 (ohm-m) 200 2 (ohm-m) 200 2 (ohm-m) 200
Bit Size IBS) Array Resistivity RLA3 LLG
5 (in.) 10 2 (ohm-m) 200 2 (ohm-m) 200
Gamma Ray(GR) Array Resistivity RLA4 LLD
0 SgAPI) 150 2 (ohm-m) 200 2 (ohm-m) 200
Caliper (CALI) Щ
Array Resistivity RLA5 -
MD Gromngert Separation
5 (in.) 10 m 2 (ohm-m) 200
........... C_ : V ... ..... Ц

НЯ L A

Fig 2.30 Example of an HRLA log Fig 2.31 HRLA tool


(After Schlumberger) (After
Schlumberger
ARRAY INDUCTION TOOL (AIT)

A significant new development in induction logging is the Array Induction Tool (AIT). Unlike
conventional induction tools consisting of one transmitter and one receiver, the AIT operates at
multiple frequencies and consists of one emitter coil and four receiver coils. Conductivity of the
formations surrounding the borehole is measured as a function of both depth and distance. The
signals are processed to generate a series of resistivity curves with different depths of
investigation, producing, like the HRLA, a detailed invasion profile. The vertical resolution is
reduced from 4ft to 2ft and to 1ft in smooth borehole walls.

An example of an AIT log is shown in Fig 2.32 and the tool is illustrated in Fig 2.33.

Fig 2.32 Example of an AIT log Fig 2.33 AIT tools


(After Rider, 1996) (After
Schlumberger)
THE PHASOR INDUCTION TOOL

Advances in electronics technology and signal processing have resulted in the development and
introduction of the Phasor Induction tool. The tool employs a new digital transmission and
processing system that reduces the signal and cable noise, improves the thin bed resolution of
the induction measurements to 2ft, increases the depth of investigation of the of the device and
automatically corrects the readings for borehole and shoulder bed effects. The deep and
medium induction measurements are combined with an SFL device to record resistivity data at
three depths of investigation, one curve representing Rt and two reading R i.

IDPH is the deep Phasor Induction log ( R t) and IMPH the medium Phasor Induction log ( R i) . The
Phasor-SFL combination may also include an SP electrode.

Fig 2.34 shows the improvement gained by Phasor processing over the conventional Induction
measurement.

Fig 2.34 Comparison between the responses of the traditional


ILD and IDPH curves to Rt (After Schlumberger)

CASED HOLE FORMATION RESISTIVITY TOOL (CHFR)

Although the need to measure resistivity through casing has long been recognized, only very
recent advances in electronics technology have made this possible. Using a 12-electrode
configuration, the CHFR tool delivers a deep resistivity measurement. The tool operates in
contact with the casing and injects a current into it with a return at surface (Fig 2.35). Since

Fig 2.35 Diagrammatic representation of the


CHFR tool (After Schlumberger)

typical formation resistivities are about a billion times that of steel casing, the current passes
through it easily and flows into the rocks surrounding the borehole. Low resistivity cements do
not degrade the CHFR measurement but data obtained through high resistivity cements require

environmental corrections.

The measurements generally show good correlation with the openhole deep laterolog resistivity
data and examples of CHFR logs are shown in Figs 2.36 and 2.37. CHFR logging makes it
possible to monitor the movements of hydrocarbon/water contacts through the reservoir and
identify bypassed pay zones.

Fig 2.36 Comparison between CHFR and open hole and deep resistivity data
in a largely shaly section (After Schlumberger)

Fig 2.37 Comparison between CHFR and open hole and deep laterolog (HLLD) resistivity
data in a gas bearing zone (After Schlumberger)
QUALITATIVE INTERPRETATION OF ELECTRIC LOGS

As mentioned earlier, the main objective of qualitative interpretation is the identification of


permeable beds, their boundaries and pore fluids. Fig 2.38 presents idealized SP and resistivity
responses for various combinations of lithologies and fluid contents. The following
interpretations of units shown may be made on the basis of their log responses:

1. Units 1 to 6 are interpreted as shale for the following reasons:

(a) SP curve does not depart from the shale line, indicating non-permeable intervals
[Fig 2.38(a)],
(b) Units have relatively low resistivity.
(c) Both resistivity curves have the same value [Fig 2.38(b)], indicating that the beds are
impervious to drilling mud, i.e. there is no invasion.

SPONTANEOUS POTENTIAL RESISTIVITY LOGS


---------- deep
Scale: M illivo lts MV ---------- shallow

Scale: оИгм/т2/т(Ш

I
SHALE
1 I

Permeable bed POROUS


A
R - salt OIL
•SANDSTONE
-V R , - fresh SALT

25m 4
S h a le Permeable bed POROUS
line в : Rw - fresh •SANDSTONE

R , - salt

3i POROUS

c E :x:‘: l_ Impermeable bed •SANDSTONE

4
” ■
|

Shaly sand
TIGHT SANDSTONE
* 'QUARTZITE'

/ d :
R. < R...
Clean sand
FINING UP
•SANDSTONE,
SHALY

POROUS, CLEAN
SALT WATER
51

(a) (b)
Fig 2.38 Idealised SP and resistivity responses for various combinations of rock types and
fluid contents (After Rider, 1996)
2. Unit A is interpreted as a salt water bearing sandstone for the following reasons:
(a) SP has a strong negative departure from the shale line, indicating Rw < Rm f
f Cl
[Fig 2.38(a)], '
(b) Shallow resistivity curve shows a higher value than deep resistivity since the part of
formation investigated by the latter (uninvaded zone) is filled with conductive salty
formation water [Fig 2.38(b)].

3. Unit В is interpreted as a fresh water bearing sandstone for the following reasons:
(a) SP has a positive departure from the shale line, indicating Rw > Rmf [Fig 2.38(a)],
(b) Both resistivity curves show high values as fresh water is a poor conductor [Fig
2.38(b)]. Deep resistivity curve reads higher than shallow resisttvity since the part of
formation investigated by the former (uninvaded zone) is filled with low conductivity
fresh water. )

4. Unit С is interpreted as an impermeable (tight), dry (no pore fluids) unit for the following
reasons:
(a) SP does not depart from the shale line, indicating no movement of ions, i.e. no
formation water [Fig 2.38(a)].
(b) Resistivity curves track each other and show very high values indicating a non
conductive, dry formation [Fig 2.38(b)],

5. The upper part of unit D is interpreted as a shaly sandstone for the following reasons:
(a) SP curve has a reduced amplitude compared to the underlying clean sand section [Fig
2.38(a)], The gradual upward amplitude reduction reflects a progressive decrease in
grain size, indicating an upward fining sequence and a transitional contact with the
overlying shale.
(b) The resistivity curves show some separation compared to shale [Fig 2.38(b)].

6. The upper part of unit E is interpreted as containing hydrocarbons and the lower part as salt
water bearing for the following reasons:
(a) Resistivity curves show high values as hydrocarbons are a poor conductor[Fig 2.38(b)].
(b) Deep resistivity shows a higher reading since the part of formation investigated by the
device (uninvaded zone) is filled with non conductive hydrocarbons.
(c) The underlying salt water bearing section is indicated by low resistivity, with the deep
curve reading lower than the shallow curve. The oil/salt water contact is indicated by the
cross over between the deep and shallow curves (see also Fig 2.39).

SPONTANEOUS
POTENTIAL DUAL LATEROLOG.MICRO SFL
CALIPER
DEEP LATEROLOG
(mVt
20
4 —U SHALLO W LATEROLOG

8 C A L IP E R 16
MICRO SFL
(in !

100 1000
^ S it sue — "r—1
iflm)

Hydrocarbon/salt water contact

Fig 2.39 Hydrocarbon detection (After Schlumberger)


3. THE BOREHOLE COMPENSATED (BHC) SONIC LOG

The conventional Sonic or Acoustic log provides a continuous record of the time taken, in
microseconds per foot (^sec/ft) or microseconds per metre ((isec/m), by a sound wave to travel
through one foot or one metre of formation. This is known as the interval tra nsit time,
abbreviated to t or At. The travel time measured is that of a compressional or ‘P’ wave which
travels the fastest through the formations and represents the first arrival. Shear and Stoneley
waves follow the P waves but are not recorded by the conventional tools. Fig 3.1 shows the full
sonic wave form and an example of the arrival times of the various sonic waves is presented in
Fig 3.2.

Travel time (psec/ft)

Fig 3.2 Arrival times of various sonic


waves (After Schlumberger)

The velocity of sound through a given formation is a function of its lithology and porosity. Dense,
low porosity rocks are characterised by high matrix velocities (Vma), while porous and less dense
formations are characterised by low Vma values. Since t and Vma are inversely related, high
porosities correspond to high t values and low porosities to low t values. The conventional Sonic
log that records the P wave arrivals is therefore a porosity measuring device. Listed in Table 1
are the Vma and tma ranges of some of the most commonly encountered rock types and casing.
Modern sonic tools are of the BHC (borehole compensated) type in which automatic corrections

Table 3.1 Some typical sonic matrix velocities and travel times

tm a ( jA S e c /f t)
Vma (ft/sec) tm a (И-S e C /ft)
(com m only used)

Sandstone (Ф=0) 18,000-19,500 55.5-51.0 55.5 or 51.0

Limestone (Ф=0) 21,000-23,000 47.6-43.5 47.5

Dolomite (Ф=0) 23,000 43.5 43.5

Anhydrite (Ф=0) 20,000 50 50

Salt (Ф=0) 15,000 66.7 67

Casing (Ф=0) 17,500 57 57


t ..

are applied to the log reading for the effects of changes in borehole size as well as for errors
arising from the tilting of the device during the logging operation.

Fig 3.3 is a schematic representation of the BHC Sonic tool, which consists of a pair of

3'

Fig 3.3 Diagrammatic representation BHC Sonic tool


(After Rider, 1996)
transmitters and two pairs of receivers. The transmitters are puised alternately, and t values are
read on alternate pairs of receivers. When one of the transmitters is pulsed, the sound wave
generated passes first into the mud and then enters the formation, travelling through the
formation close to the borehole wall. At a critical (lower) velocity it is refracted back into mud and
reaches the tool again where it is detected. The time elapsed between the detection of the first
arrival at the two corresponding receivers is measured and represents the formation reading.
The ray paths in Fig 3.3 indicate the course followed by the first arrivals of compressional sound
energy. Sound waves also travel directly between the transmitters and the receivers through the
mud. However, since the velocity of sound is greater in the formations surrounding the borehole
than in the mud due to the higher density of the former, the first arrivals are from the borehole
wall. The measured travel times of the upgoing and the downgoing signals are averaged and
this is presented linearly as the t curve.

The Schlumberger BHC tool has a spacing of 3ft between transmitter and near receiver and a
span of 2ft between receivers. The transmitters are pulsed a total of 20 times per second so that
five complete measurements are made each second. Logging speed is 5,000ft/hr, which means
a measurement is made about every 3in of hole. Normally, the Sonic tool is run centred so the
contributions to a receiver signal from different sides of the hole will be in phase (if the hole is
round) and the signal-noise ratio will be maximized. The tool can be run off centre, but
significant degradation in the signal-noise ratio must be tolerated.

LOG PRESENTATION

Typical presentation of the Sonic log, when run by itself, is shown in Fig 3.4. The interval transit
time, in microseconds per foot, is recorded across tracks 2 and 3. Short transit times are to the
right and long transit times are to the left, such that increase in porosity deflects the curve
toward the depth track consistent with Density and Neutron recording.

In the depth track are small pips, representing integrated travel time of 1 msec between each
pip. Larger pips are recorded at 10 msec intervals. These are useful in comparing Sonic logs
with seismic sections and assist in converting seismic travel times into depths. When a Gamma

Ray log is run simultaneously, it is also recorded in track 1. A resistivity tool is also run
simultaneously and the resistivity curves are displayed in track 2 and the Sonic travel time is
restricted to track 3.
Fig 3.4 BHC Sonic log presentation (After Baker Atlas)

A good check on the accuracy of a Sonic log is to observe the reading in casing. It should be 57
S^sec/ft, the travel time of steel. The log may not jump immediately to this value on entering
casing because there can be a drastic change in signal amplitude to which the system (or the
engineer) must adjust. The reading is most reliable in uncemented pipe where the casing-borne
arrival will have good amplitude and will always arrive ahead of formation-borne signals, no
matter how fast. The opposite can be true in cemented pipe.
CYCLE SKIPPING

Sometimes the first arrival, although strong enough to trigger the receiver nearer the transmitter,
may be too weak by the time it reaches the far receiver to trigger it. Instead, the far receiver may
be triggered by a different, later arrival, and the travel time measured on this pulse cycle will
then be too large. When this occurs, the Sonic curve shows a very abrupt and large excursion
toward higher t values (Fig 3.5); this is known as cycle skipping. Such skipping is more likely to
occur when the signal is strongly attenuated by unconsolidated formations, formation fractures,
gas saturation, or rugose (washed out) salt sections.

EVALUATION OF POROSITY FROM THE SONIC LOG (Ф5)

In the mid- to late 1950s Wyllie et al (1956 and 1958), on the basis of many laboratory
observations, developed an empirical relationship between the porosity and the transit time of a
compressional sound wave through the matrix and interstitial fluids of a porous medium. They
found that the readings of the sonic curve (t) represented the sum of two individual responses,
namely, that of the matrix (tma) and that of the fluid filled porosity (tf). The matrix response is

Fig 3.5 Example of cycle skipping (After Rider, 1996)


given by the amount of matrix (1 - Ф) multiplied by tma, and the fluid filled contribution equals the
amount of fluid (Ф) multiplied by tf. Thus:

t=(1-4>)tma+<Mf 3.1

Solving for Ф:

Ф5 = (t - tma)/(tr tma) 3.2

This is known as the time-average equation, and has been used universally to derive sonic
porosities from the log-recorded travel time, t, provided tf and t™a are known. The fluid in the
zone of investigation is typically mud filtrate. Consequently, tf is normally taken as 189 (.isec/ft in
fresh mud. In salt mud a value of 185 jisec/ft may be used. Matrix travel times vary from 40-50

^sec/ft, depending on lithology.

Fig 3.6 presents a graphical solution to equation 3.2. Log-derived transit time is entered on the
horizontal axis, a line is projected vertically to the appropriate matrix velocity, and porosity is
read on the vertical scale opposite the point of intersection.

As an example, the zone at XX869 - XX874 ft in Fig 3.4 reads a travel time of 71 [isec/ft. Using
the straight, continuous blue lines. Fig 3.6 gives a porosity of 12% if the matrix is sandstone or
as high as 22% if the matrix is dolomite. Clearly, the lithology must be known to obtain accurate
porosity values. Even within the given lithology there is a range of possible matrix travel times,
as indicated by straight continuous lines in Fig 3.6. Local knowledge dictates the value to use -
although the deeper the burial of a formation, the lower tma is likely to be.

Effects of Under-compaction on Ф 5

The time-average relation holds quite well in consolidated or well-compacted formations.


Typically, these have transit times less than 100 usec/ft. However, serious errors arise if the
relation is applied without modification in shallow, unconsolidated sands which occur in
geologically younger formations. If the effective pressure on the rock framework
(overburden-hydrostatic) is less than about 4,000 psi, which is the case at depths less than
about 2,000m, the sand has not reached its fully compacted rigidity. Travel times in
uncompacted sands may reach as much as 150 ^sec/ft, which convert to porosities far above

v, =-5300 ft/sec

30 40 50 60 V0 80 90 100 110 120 130


L, interval transit time ((isec/tt)

Fig 3.6 Porosity evaluation from the Sonic log (After Schlumberger)

the known maximum of 40%. In such cases, the porosity computed from equation 3.2 must be
divided by a com paction co rrection factor, Bcp, as indicated in Fig 3.6. The factor varies from
1.0 to as high as 1.8. Bcp is never less than unity. Equation 3.2 thus becomes:

<£>S = (t - tm a )/(tf - La) x 1/Bcp 3.3

Typical sonic compaction factors are listed in Table 3.2.


Table 3.2 Typical sonic compaction factors (After Schlumberger)

Lithology Depth(m) Bcp Vma(ft/sec)

Sandstone 1,100 1.60 18,000

1,200 1.40 18,000

1,500 1.40 18,000

1,750 1.35 18,000

1,800 1.35 18,000


1,900 1.10 18,000
2,000 1.05 18,000
2,100 1.00 18,000
2,100 1.00 18,000-19,500

Limestone 1.0 21,000-23,000

Dolomite 1.0 23,000-23,000

Ф5 Evaluation by the Raymer-Hunt-Gardner Method

In 1980, Raymer, Hunt and Gardner proposed a new relationship between transit time and
porosity. It is entirely empirical, based on comparison of transit times with core porosities and
porosities derived from other logs. The relationship can be approximated with adequate
accuracy in the regions of interest by the equation:

Ф = 0.63(1 -tma/t) 3.4

where

tma = 54 usec/ft for sands, 49 |.isec/ft for limestone and 44 ^sec/ft for dolomite.

Fig 3.7 presents a graphical solution to equation 3.4. It is coming into use now and has the dual
advantage of not requiring selection of different matrix times for a given lithology and of giving
reasonable porosities in uncompacted sands with transit times in the range 100-150 usec/ft.

Effect o f Gas on Sonic-derived Porosity

Due to its low density, the presence of gas decreases the density of the host rock, and this
Fig 3.7 Porosity derivation from the Raymer-Hunt-Gardner relationship
(After Raymer et al, 1980)

causes an increase in transit time. An increase in transit time results in a spuriously high
computed porosity.

The increase in transit time is almost nil in deeper, low-porosity formations where pore volume is
low and compaction pressure is high, which means that pore fluid contributes little to rock
rigidity. However, it can be as high as 40% in shallow, high-porosity formations where pore
volume is large and compaction pressure is minimum in which case pore fluid has a much larger
contribution to formation rigidity.

Whether the Sonic will sense the presence of gas depends on how much residual gas is left
after invasion in the 1in or so of formation being investigated by the tool. In medium- to
high-porosity gas-bearing formations a residual gas saturation of at least 15% would be
expected in the flushed zone so that gas should be sensed by the tool. This is implied in Fig 3.7,
where a separate curve is included for gas-bearing sandstones.

In summary, the Sonic responds well to primary or intergranular porosity. os values may be
considered as reliable if the formation being evaluated is water-bearing, gas-free, clean and
monomineralic.

SECONDARY POROSITY (Ф2)

In general, the Sonic log tends to ignore vuggy or fracture porosity which occur commonly in
carbonate reservoirs. The Density or the Neutron log, by contrast, respond to total porosity (see
below). A secondary porosity index (SPI or a2) may therefore be derived by taking the difference
between Density (Фо) or Neutron (Фм) porosity and Фэ:

ф2 = (Фо, Фц) - Os 3.5

THE LONG SPACING SONIC TOOL (LSS)

This tool was introduced in the early 1980s and is characterized by a longer transmitter-receiver
distances than the BHC Sonic device. The transmitters are 2ft apart and are situated below the
receivers, which are also separated by a distance of 2ft. The greater transmitter-receiver
separation increases the depth of investigation of the LSS tool to up to 50cm as opposed to 5cm
- 10cm in the case of the BHC Sonic. Due to the increased depth of penetration, LSS
measurements are generally free from the effects of near bore formation alteration, damage
from the drilling process and enlarged boreholes. The readings therefore represent more reliable
formation travel time measurements. This is particularly important when using the sonic data in
seismic interpretation.

Readings are taken at 2 different depth positions of the tool: once when the receivers straddle
the measure point depth and once when the transmitters straddle the measure point depth. Use
of the upper transmitter and receiver yields an 8 ft -10ft t measurement and the use of the lower
transmitter and receiver yields a 10 ft - 12ft t measurement.

A diagrammatic illustration of the LSS tool is shown in Fig 3.8 and comparisons between LSS
and BHC Sonic measurements in a large borehole and altered formations are presented in Figs
3.9 and 3.10.

THE ARRAY SONIC TOOL (AST)

By using an array of receivers this new generation tool records the full sonic wave form. In
T —tr a n sm it te r
L— lower U —u p p e r
ill R receiver
h
\s

Fig 3.8 Diagrammatic representation of the Long Spacing


Sonic tool (After Rider, 1996)

addition to recording the arrival times of the P waves, the AST measures the shear and Stoneley
wave arrivals. The tool produces data for specialist applications and is not run as part of the
standard suite of logs.

Applications o f A S T Data

The main application of AST data is to the determination of the rock mechanical properties
which serve as a useful aid in the interpretation of seismic data. These properties include the
Poisson’s ratio, Young’s modulus, shear modulus and the bulk compressibility of the formations
Fig 3.9 Comparison of BHC and LSS Fig 3.10 Altered formation 153-162 m BHC log
in an enlarged hole (After reads too high (After Schlumberger)

penetrated by the borehole. V n C ^ is s o n ^ 'a n Q^is particularly useful since it provides a means
of detecting gas bearing zones. Generally, high porosity gas bearing intervals have markedly
lower Poisson’s ratio than the surrounding rocks and produce flat lying anomalies in seismic
sections. Such a feature is referred to as a direct hydrocarbon indicator (DHI).

Due to variations in the mechanical properties of different rocks, AST data can be used to
estimate lithology. Crossplotting P and S wave travel times allows lithology characterization as
shown in Fig 3.11. This defines the gross lithological content of seismic sections which is
Fig 3.11 Lithology indentification by tcompr - tshear crossplot (After Schlumberger)

useful information in seismic interpretation. The crossplot can also provide information on the
pore fluids. It has been observed that the presence of light hydrocarbons (gas) decreases P
wave velocity relative to brine and increases S wave velocity, making it possible to identify gas
bearing zones.

Under certain tool configurations, the Stoneley wave part of the sonic wave spectrum can be
used to identify fractures. An open fracture causes some of the energy to be reflected back into
the borehole and the magnitude of reflection is related to the degree of openness or width of the
fracture.

The velocity and energy level of the Stoneley wave indicate permeable intervals in some
reservoirs. Stoneley wave velocity and energy level decrease in permeable zones, making it
possible to identify such intervals. However, this diagnosis tends to be qualitative since the
borehole size, mudcake thickness and formation and tool characteristics also affect the Stoneley
wave. Quantitative estimation of permeability requires calibration to core data.

DETECTION OF ABNORMAL PRESSURES

Drilling experience has shown that abnormally pressured porous and permeable formations are
overlain by overpressured shales, the vertical separation between the two varying from a few
tens of feet to as much as 1,000ft.

Shale density normally increases with compaction, i.e. with increasing depth of burial. This
progressive increase in density with depth is associated with a decrease in shale transit time (tsn)
values which is quite apparent on Sonic logs. Decreasing t5h values with increasing depth is
therefore regarded as the normal com paction gradient (Fig 3.12, upper part of the Sonic
curve).

Any change in the trend of increasing compaction with depth is an indication of abnormal
pressures (lower part of the Sonic curve, Fig 3.12). Abnormal pressures develop when-a-sbale
body becomes isolated and therefore unable to lose water. This inability to lose water arrests the
compaction process, which means that the shale body remains under-compacted and contains
some trapped water. As the depth of burial increases, the trapped water begins to exert a pore
pressure to balance the weight of the increasing overburden. Overpressured shales thus contain
an excess of water, and the presence of this water lowers their density, causing the Sonic transit
time to increase. Such zones are common in Tertiary deltaic provinces where rapid burial rates,
accompanied by contemporaneous faulting in certain instances, lead to the isolation of some
sand and shale bodies. Fig 3.13 presents an actual example of an overpressured zone in the
Cretaceous section of the central North Sea.

Sonic logs run at intervals during drilling may thus be used to predict the presence of
overpressured sands, before they are reached, so that the necessary precautions may be taken
to eliminate the possible hazards.
CK

ч
\

Fig 3.12 Detecting overpressured zones with the Sonic log (After Schlumberger) ^ ^

л & Л .

\ (W " ',\XV
y /v A Q*
?\ \
\ ' *.■ • J 4'
A c
\ ‘’

Я
и К 211/29-1
Measured DT
depth
(metres) 400.0 (msecs/ft) 40.0

- 600 — 'Normal’
compaction
— trend
- 800

— 1000

- 1200 Overpressure

- 1400

- 1600

- 1800

Ч 2000

2200

- 2400

2600

2800

Fig 3.13 Example of overpressuring in the Cretaceous section, North Sea


(Millennium Atlas, 2004)
4. RADIOACTIVE LOGS

These logs are entirely different from those discussed previously, and as their name suggests,
they utilise the radioactive or nuclear properties of the rocks surrounding the borehole. Basically,
two measurements are recorded on radioactive logs: (1) the natural gamma ray activity of the
beds; and (2) the effects of neutron or gamma ray bombardment. Radioactive logs provide
information on the lithology and porosity of the formations traversed by the borehole.

The main feature of radioactive lagging is that it is not dependent on the presence of drilling mud
in the well; it can even be carried out after the well has been cased, and this makes it possible to
re-examine the stratigraphy and recorrelate the strata in old-established fields. A
re-interpretation of the geology in this manner shows if any productive layers have been missed.

Acquisition of radioactive log data involves the measurement of the count rates of gamma rays
or certain types of neutrons and electrons. Since the count rates of nuclear particles vary in time,
logging speed has a significant effect on data quality. Fig 4.1 illustrates the effect of logging
speed on the bed boundary resolution ability of the Gamma Ray log. The best results are
obtained at lower speeds and since most tool combinations include at least one radioactive
device (usually a gamma ray recorder), all logs are run at a speed of 1,800ft/hr.

/
/ 5,400 Ft/hr

Fig 4.1 Effect of logging speed on bed boundary resolution of


the GR log (Modified from Dewan, 1983)
Radioactive logs fall into three main categories: (1) Gamma Ray log; (2) the Neutron log; and (3)
the Formation Density log.

THE GAMMA RAY (GR) LOG

This log records the natural radioactivity of formations. The radioactivity arises from the
presence of uranium (U), thorium (Th) and potassium (K40) in the rocks. These three elements
continuously emit gamma rays, which are short bursts of high energy radiation similar to x-rays.
Gamma rays are capable of penetrating a few inches of rock, and a fraction of those that
originate close to the borehole traverse the hole and can be detected by a suitable gamma-ray
sensor. The detector gives a discrete electrical pulse for each gamma ray detected, and the
parameter logged is the number of pulses recorded per unit of time by the detector. Fig 4.2
shows the relative radioactivity of various sedimentary rocks.

0 4 8 И 20 40 60 80 100

Caprock and anhydrite


Coal
Salt
Dolomite
Limestone
Sandstone
Sandy limestone and
limy sandstone
Greenish-gray sandstone
Shaly sandstone
Shaly limestone
Sandy shale
Calcareous shale
Shale
Organic marine shale
Lean potash beds
Rich potash beds

Fig 4.2 Relative radioactivity of various sedimentary rocks (AfterBaker Atlas,1982)

GR logs are calibrated in API units (APIU), an arbitrary scale set up by the American Petroleum
Institute. The scale increases from left to right, and the log is recorded linearly on track 1. Typical
GR responses in various lithologies are illustrated in Fig 4.3 GR log presentation is shown in Fig
4.4.

Fig 4.3 Typical GR responses in various lithologies (After Schlumberger)

'

S -r

Fig 4.4 Presentation of the GR log [track 1] (After Schlumberger)


T ool Response

In sedimentary rocks the log normally reflects the shale content of the formations. This is due to
the tendency of the radioactive elements to concentrate in shales and GR log readings increase
as the proportion shale increases in the formations traversed by the borehole. Carbonates and
sandstones, the common reservoir rocks, are usually associated with low levels of gamma ray
activity, unless volcanic ash is present or the sand is arkosit^ i.e. derived from a granitic parent
rock. Granite contains feldspar, a mineral consisting of silicates of calcium, sodium, potassium,
magnesium, iron, etc. The potassium constituent of the feldspar minerals includes the
radioactive K40 variety and consequently sandstones containing feldspathic material are
associated with high levels of gamma radiation. The term arkose is applied to such sandstones.

Uses o f the GR Log

(a) The GR log is useful in detecting shale beds when the SP curve is featureless (i.e.
Rmf ~ r«). or when the SP cannot be recorded due to the presence of a
non-conductive drilling fluid.

(b) Non-radioactive minerals - e.g. coal beds - may be detected by their characteristically
low GR log response.

(c) The GR log is sometimes used for correlating formations in cased holes.

(d) In a shaly porous and permeable zone, the volume of shale (V5h) can be estimated
from the deflections of the GR curve (Fig 4.5). The steps involved are as follows:

(i) Read the gamma ray activity associated with the zone of interest (GRZOne)-

(ii) Select a clean shale-free zone, and read GRciean-

(iii) Select a 100% shale zone and read GRShaie- The fraction of shale in the zone of
interest will be:

4.1

О
Although there are many more ways of calculating Vsh, the above is the most widely used
method. ci
0 2S SO 75 100

Fig 4.5 Determination of Vsh from the GR log (After Dewan, 1983)

THE NATURAL GAMMA RAY SPECTROMETRY TOOL (NGS)

The Natural Gamma Ray Spectrometry Tool (NGS) was introduced in the 1980s and represents
a new development of the conventional GR log. As shown in Fig 4.6, the NGST is a pad contact
device, held against the borehole wall by means of a bow spring.

Whereas the conventional GR log records the total radiation emitted by U238, Th232 and K40, the
NGS examines the gamma ray spectrum in more detail, detecting and recording the individual
contributions of the three radioactive elements.

This is possible since uranium, thorium and potassium emit gamma rays of different energies as
shown in Fig 4.6. Potassium has a single peak at 1.46 mev (million electron volts), while thorium
and uranium emit gamma rays of various energies, the major distinction being a prominent
Fig 4.5 Natural GR Spectrometry Tool
(After Schlumberger)

thorium peak at 2.62 mev and a predominant uranium peak at about 0.6 mev. Since minerals
have characteristic concentrations of uranium, thorium and potassium, the individual responses
can sometimes be used to identify minerals or mineral type.
1.46

P o ta ssiu m

с.
о
cd
L_
CJ)
О)
с T h o riu m S e rie s
ел
О 2.62
ш
CL
С
О
со
in ±1
сл
е
ш
б
U ra n iu m -R a d iu m S e rie s
-Q

_о 1.76
о
I . !
I I
II ' ■ I . 1 I C I I ■ 1 ____________

0 0.5 i 1.5 2 2.5 3

G a m m a Ray E n e rg y (M eV)

Fig 4.6 GR spectra of K, Th and U (After Schlumberger)

LOG PRESENTATION

Fig 4.7 shows a standard presentation of the NGS data. Five curves are displayed: the total or
standard gamma ray (SGR), the values of the potassium, thorium and uranium components
(POTA, THOR, and URAN), and the computed gamma ray (CGR), which represents the Th + К
contribution. The thorium and uranium curves are scaled in parts per million while the potassium
curve is scaled as a proportion. It can readily be seen that throughout the interval there is a
significant uranium component, and in particular over the interval 185-190m. In this section the
uranium content contributes up to 60 API of the total GR log reading.
Fig 4.7 Example of an NGS log from the Lower Carboniferous Barnett Shale, West
Texas. The shale is underlain by ‘clean’ limestone, the contact between them
occurring at 9,606ft (After Asquith & Krygowsky, 2004)

A presentation is available in which three ratio curves are presented in track 2. This is shown in
Fig 4.8 with the total and computed GR curves in track 1 and the separate thorium, potassium
and uranium data in track 3. The three ratio curves are Th/K (TPRA), Th/U (TURA) and U/K
(UPRA). The TPRA curve is particularly useful: cross plotting TPRA values against certain
Density log readings provides a means of clay mineral identification.

APPLICATIONS

Applications of the NGS log data are many. The major uses of the data, however, are in
distinguishing between the various clay minerals and determining improved Vsh values.
Clay mineral identification may be effected by the use of the chart illustrated in Fig 4.9, which
compares the potassium and thorium contents of several minerals. The chart is entered by
taking values directly from the recorded curves. Usually the result is not unambiguous, and
consequently other data need to be introduced. In particular, the photoelectric absorption

Fig 4.8 Ratio presentation of NGS data (After Schlumberger)


Possible 100% kaolimte
montmorillonite,
illite “clay line" 100% illite point

Kaolmite

£
Cl
CL

E
Tз~
О
s:
h-

•30% glauconite
Glauconite
Feldspar Th/K: 0.Э
Potassium evapontes. -30% feldspar
1 2 3 4

Pot assi um (%)

Fig 4.9 Mineral identification from NGS log data (After Schlumberger)

coefficient (see under the Litho-Density log in this section) and the ratios of the radioactive
families are used: Th/K, U/K and Th/U.

The major occurrences of the three radioactive families are as follows:

Potassium: micas, feldspars, micaceous clays (illite), radioactive evaporites.

Thorium : shales, heavy minerals.

Uranium: phosphates, organic matter.

The significance of the type of radiation is dependent on the lithology of the formation with which
it is associated. In carbonates uranium indicates the presence of organic matter, phosphates
and stylolites, while the thorium and potassium levels are representative of the clay content, in
sandstones the thorium level is determined by the heavy minerals and clay content, and the
potassium is usually contained in micas and feldspars. In shales the potassium content is
indicative of the clay mineral type and the presence of mica, and the thorium levels is dependent
on the amount of detrital material, or the degree of shaliness. The presence of uranium would
suggest that the shale is a potential source rock, since uranium seems to be associated with
organic matter in argillaceous sediments.

In addition to lithology, the distribution of the radioactive minerals in a sedimentary formation is


dependent also on the agent and manner of transportation, the degree of reworking and
alteration. For example, due to its low solubility, Th has limited mobility and tends to accumulate
with the heavy minerals. On the other hand, uranium has a greater solubility and mobility, and so
high uranium concentrations are found along fault planes, fractures and in formations where
water flow has occurred. Similarly, high uranium concentrations can build up in producing oil
wells, in the permeable beds and on the tubing and casing. Chemical marine deposits are
characterised by their extremely low radioactive content, with none of the three radiation families
making any significant contribution.

In areas where abnormal amounts of potassium-bearing micas and feldspars are associated
with the sands (the North Sea, granite wash areas of the southwestern US), the thorium.,,
component alone is the best indicator of shale content.

As the proportion of uranium is dependent more on the organic content of shales than the
amount of clay, it is not a particularly good quantitative clay indicator. For this reason, the CGR
curve, which is the sum of the thorium and potassium components only, is a better indicator of
shliness. When a KCI mud has been used to drill the borehole the high potassium content will
affect the measurement, and so under these circumstances the thorium measurement alone will
probably be the best clay indicator.

A clay index computation can be made using any of the recorded data. Fig 4.10 is an example
comparing five clay indicators, derived from the thorium (VSTC), potassium (VSPC), uranium
(VSUC), total gamma ray (VSSG) and computed gamma ray (VSCG) data. It is evident that the
values obtained from the thorium and potassium content correlate well, but the uranium derived
curve has a very different character. The relative uranium concentrations of the two shale zones
suggest the possibility that the lower shale is an organic rich source rock.
\ 4' v 0^

9 '

tj
Л
v>
,< Ц о

Fig 4.10 Comparison of 5 clay indicators computed from NGS data (After Schlumberger)

vv
m^ v
V s :v ЧК'
THE NEUTRON LOG
t_ /

In neutron logging the formations surrounding the borehole are bombarded by high energy
neutrons from a radioactive source in the device. The Neutron tool and its operation are
illustrated diagrammatically in Figs 4.11 and 4.12. The current second generation tools have two
detectors located above the source.
(After Dewan, 1983)

Neutrons are electrically neutral particles with a mass almost identical to that of a hydrogen
atom. On leaving the source, the neutrons enter the formations and collide with nuclei in the
rocks traversed by the borehole. The interactions between the neutrons and the nuclei in the
rocks are considered as elastic billiard-ball type collision and with each collision a neutron
loses some of its energy. The amount of energy lost per collision depends on the relative mass
of the nucleus with which the neutron collides. The greatest energy loss occurs when the
neutron strikes a nucleus of practically equal mass, i.e. a hydrogen nucleus. Collisions with
heavy nuclei do not slow the neutron down very much. Thus, the slowing-down of neutrons
depends largely on the amount of hydrogen in the formation.

Within a few microseconds the neutrons have been slowed down by successive collisions to
therm al velocity. They then diffuse randomly, until they are captured by the nuclei of atoms
such as chlorine, hydrogen, silicon, etc.

The detectors may be one of three types: a thermal neutron detector monitoring the density of
the thermal neutrons, an epithermal detector sensing the density of the neutrons just above
thermal energy, or a gamma ray of capture detector sensitive to the gamma radiation emitted by
nuclei when they capture thermal neutrons. When the hydrogen concentration of the material
surrounding the neutron source is large, most of the neutrons are slowed down and captured
within a short distance from the source. If, on the other hand, the hydrogen concentration is
small, the neutrons travel further from the source before being captured. Therefore, regardless
of the detector type, the count rates increase with decreasing hydrogen content (low porosity in
clean formations) and decrease with increasing hydrogen content (high porosity in clean
formations). Count rate thus varies inversely with porosity, since all the hydrogen in clean
formations occurs in the pore fluids.

Tools measuring the density of the neutrons in the m oderation phase (Fig 4.12) produce an
Epithermal Neutron Log, while those counting the thermal neutrons generate a Compensated
Neutron Log (CNL), which is a porosity indicator. Recording the gamma ray of capture produces
a Thermal Decay Time Log (Schlumberger’s TDT Log), from which hydrocarbon saturation may
be derived. This discussion is concerned with use of the neutron tool as a porosity measuring
device and therefore focuses on the CNL.

Tool Response

As stated above, the tool responds to the presence of hydrogen. Since oil and water contain
practically the same amount of hydrogen per unit volume, the responses reflect primarily the
liquid-filled porosity in clean formations. The tool does not, however, distinguish between the
hydrogen in the pore fluids and that associated with bound water, i.e. water of crystallisation.
The latter, of course, does not always correspond to effective porosity; for example, shales and
gypsum (Ca S 0 4.2H20) containing bound water have a high hydrogen index, and are therefore
characterised by a large 'neutron porosity'. In general, however, dense and non-porous beds
such as anhydrite and tight limestones are indicated by low porosity peaks on the neutron curve,
while porous zones show higher readings.

The device is calibrated by using a standard piece of fresh water-bearing, pure limestone in the
American Petroleum Institute (API) test laboratories. Consequently, the porosities recorded
normally assume that the matrix is limestone. The logs are therefore scaled in 'Limestone
Porosity Units'.

If the matrix happens to be sandstone, the true porosity will be different by about 4 p.u., i.e., 20 p
u sandstone will register as 16 limestone p u. The effects are evident in Fig 4.13, which can be
used also to estimate porosity in clean, water-bearing and gas-free zones consisting of a single
lithology.

e-M iro.-. apparent limestone neutron porosi’ y (p 11.)

Fig 4.13 Thermal neutron porosity equivalence curves (After Schlumberger)


Liquid hydrocarbons have hydrogen indexes close to that of water. Gas, however, usually has a
considerably lower hydrogen concentration, which varies with temperature and pressure.
Therefore, when gas is present near enough the borehole to be within its zone of investigation, a
Neutron Log reads too low a porosity. This characteristic is called the gas effect, and allows the
Neutron Log to be used to detect gas zones; in a formation known to have uniform porosity, the
Neutron Log alone will often indicate gas/liquid contacts. A Neutron and Density log combination
is more effective in identifying gas bearing formations and allows gas detection in a zone with
variable porosity, a more accurate quantification of porosity and eliminating the effect of shale.

The CNL has a radius of investigation of about 10in. If run slowly, the vertical resolution of the
tool is approximately 2ft.

Log Presentation

The CNL is rarely run by itself because of substantial matrix and clay effects. It is normally run in
combination with the Compensated Density and GR logs in the configuration shown in Fig 1.5
(bottom right-hand diagram). The Neutron is positioned above the Density with its backup spring
lined up with that of the Density so that the array is forced against the same side of the hole.

Fig 4.14 shows the standard presentation of the curves obtained with the Neutron-Density
combination. A GR log, caliper and bit size are recorded in track 1, and Neutron and Density
curves in tracks 2 and 3 with the Neutron curve dashed and the Density solid. The CNL curve is
scaled in porosity units, each division corresponding to 3 p u.

Evaluation o f P orosity From the CNL

The CNL reads the total porosity. In a monomineraiic, clean (shale-free), water bearing and
gas-free permeable zone, the log reading can be converted directly into a true porosity value
provided the lithology is known, as shown in the chart in Fig 4.13. The dashed lines are for
porosity determination from the Sidewall Neutron Porosity (SNP) tool, which is now obsolete.
With the increasing use of LWD in field development, interpretation charts have been developed
for the derivation of porosity from the Compensated Density Neutron (CDN) tool. Fig 4.15
presents such a chart for use in a typical 8in hole.
Fig 4.14 GR/Neutron/Density log suite (After Asquith & Krygowski, 2004)

Осэтчэ'. чрряи-nt Hir:cs tone pcrosity (p )

Fig 4.15 LWD Neutron Porosity Equivalence Curves (After Schlumberger)


THE FORMATION DENSITY COMPENSATED (FDC) LOG

The FDC log records the bulk density (pb) of the formations surrounding the borehole. Fig 4.16
illustrates the principle of the tool which consists of a gamma ray source and two detectors
mounted on a pad. The pad is pressed against the borehole wall by a spring-loaded arm and
carries a plough which scrapes some of the mud cake to minimise its contribution to the bulk
density measurement. Since the measurement is made in contact with the borehole wall, any
loss of contact renders the log reading invalid over the interval where this occurs.

Fig 4.16 Diagrammatic illustration of the Formation Density Compensated


(FDC) tool (After Schlumberger)

Gamma rays are beamed at the formations by the source. These enter the formations and
undergo multiple collisions with electrons in the rocks, as a result of which they lose energy and
become scattered in all directions (Fig 4.17). This is known as Com pton scattering. Some of
the scattered gamma rays return to the borehole and are recorded by the detectors on the
device. Count rates from the two detectors are combined to provide two signals for log
presentation. One is the corrected pb curve, and the other is the Ap curve (Fig 4.14),

*«V
•V

Fig 4.17 Scattering of gamma rays emitted by the Density tool


[Compton effect] (After Welex)

representing the correction that has been applied automatically by the compensating
mechanism to the pb curve to eliminate the effects of mud cake and variations in borehole size.
The Др recordings may be regarded as a quality control curve; large corrections (more than
0-15gm/cm3) tend to lower the reliability of the pb measurements.

The intensity of returned radiation is proportional to the number of electrons in the formation,
and provides a measure of the electron density of the material. Electron density is approximately
equal to the bulk density of rocks, and this is recorded in gm/cm3. Table 4.1 lists the actual bulk
densities and those measured by the FDC tool in the case of the common minerals.

Table 4.1 Densities of the common minerals and the densities measured by the FDC tool
(After Schlumberger)

Actual Density, ph pa, as seen by tool


Compound Formula
(gm/cm3) (gm/cm3)
Quartz S i0 2 2.654 2.648

Calcite СаСОз 2.710 2.710

Dolomite СаСОзМдСОз 2.850 2.850

Anhydrite CaS04 2.960 2.977

Gypsum CaS042H20 2.320 2.351

Sylvite KCI 1.984 1.863

Halite NaCI 2.165 2.032

Anthracite Coal 1.400-1.800 1.355-1.796

Bituminous Coal 1.200-1.500 1.173-1.514

Fresh Water H20 1.000 1.000


Salt Water H20 (200,000 ppm) 1.146 1.135

Oil n(CH2) 0.850 0.850


1.335pmem- 0.188
О
X

Methane Pmelh

Gas C i.iH « Pg 1.325Pg- 0.188

Tool Response

Dense, low porosity formations are characterized by high рь values, while higher porosity zones
are less dense and are associated with lower pb readings. The FDC is therefore a porosity
indicator. Like the Neutron log, the primary calibration standard for the FDC is a freshwater-filled
limestone of high purity and accurately known density. Consequently, the tool reads the porosity
only in a limestone matrix.

The depth of investigation of the FDC log is approximately 4in at mid densities, slightly greater at
lower densities and slightly less at high densities. This means the log senses the flushed zone,
which contains mud filtrate and possible residual hydrocarbon in the pores. There is usually
insufficient difference in density between water and oil for the log to sense residual oil in the
flushed zone. On the other hand, it can readily sense residual gas, especially if porosity is high
and gas pressure is low. The effect of gas is a lowering of the рь reading, resulting in a
spuriously large computed porosity.

The vertical resolution of the tool, if run slowly, is approximately 2ft. Formation density is
averaged over that interval.

Log Presentation

As mentioned before, the FDC is normally run simultaneously with the CNL and the curves are
recorded in tracks 2 and 3 (Fig 4.14). The FDC is recorded as a solid curve and calibrated in
gm/cm3, each division representing 0.05 gm/cm3.

А Др correction curve is also recorded in track 3, with zero at the centre and ± 0.25gm/cm3 at the
extremes. The Др curve indicates the correction that has been applied to the рь curve to
compensate the measurements for the effects of the mudcake and variations in the size of the
borehole. It should be considered as a quality control curve and it is not necessary to add its
readings to or subtract them from the pb curve. In a smooth hole the Др curve should be close to
the zero line, a little to the right for normal (non-barite) mud, and to the left for heavily loaded
barite mud. When mudcake or hole rugosity is encountered, the Др correction will increase. As
long as Др is less than 0.15gm/cm3, the correction is adequate and the pb curve is reliable.
Above O.I5gm/cm3 the correction is likely to be inadequate and the pb curve in error.

Evaluation o f P orosity {aD) From The FDC Log

For a formation with a matrix density pma, containing a fluid of density pf, the pb reading given by
the FDC log represents the summation of the matrix and fluid responses. The matrix response is
given by the amount of matrix (1 - Ф) multiplied bypma, and the fluid contribution equals the
amount of fluid (Ф) multiplied by pf. Thus:

Pb = (1 - Ф)рта + $pf 4-2

Solving for Ф:
V'­
'/

V
Ф0 - (pт а ~ рь)/(рта ” Pf)

Typical matrix densities (gm/cm3) are:


2.65 for sands, sandstones and quartzites
2.68 for calcareous sands or sandy limestones
2.71 for limestones
2.87 for dolomites

Porosity (Ф0) may be derived from Fig 4.18, which provides a graphical solution to equation 4.3.
Bulk density is entered on the bottom scale and porosity is read on the vertical scale for
appropriate values of pma and pf. Ф0 represents the total porosity in clean (shale-free), water
bearing and gas-free porous and permeable zones, provided the lithology is known.

Fig 4.18 Porosity derivation from the Density log (After Schlumberger)
THE LITHO-DENSITY LOG

The Litho-Density Tool (LDT) was introduced in the 1980s and has now replaced the FDC log. In
addition to the pb measurement, it provides a photoelectric absorption curve, Pe, which
measures the average atom ic num ber of a given formation and is therefore a good indicator of
the rock matrix. It is particularly useful in complex lithology interpretation.

The source-detector arrangement of the LDT is basically the same as that of the FDC but the
operation is different. With the LDT, pb and Pe measurements are made by energy selection of
the gamma rays that reach the long-spacing detector. This is shown in Fig 4.19 which is a plot of
the number of gamma rays reaching the detector, as a function of their energy, for three
formations having the same bulk density but different volum etric absorption indices, U,
designated low, medium and high.

A COUNT
RATE

R e g io n o f p h o to e le c tric e ffe c t
a n d C o m p to n s c a tte rin g .

Low U [ jt
M edium U
H ig h U

R e g io n of
C o m p to n s c a t t e r in g . _ Lj

' of

Fig 4.19 LDT photoelectric and bulk density detection windows (After Schlumberger)

The bulk density measurement is made by registering only those gamma rays that fall in the
high-energy region designated H in Fig 4.19. In this range only scattering of the gamma rays
takes place and the number of gamma rays, represented by the area under the curve, depends
on electron density only. Conversion of count rate to bulk density and correction for mudcake
and rugosity are carried out in the same manner as for the FDC log. Statistical fluctuations in
computed density, however, are reduced by a factor of about 2, to the range 0.01 to 0.02g/cm3,
by utilization of more efficient detectors.
The photoelectric absorption curve is produced by the interaction of gamma rays and electrons.
In this context gamma rays are considered as photons and the absorption of a photon by an
electron generates a photoelectron. At high energy levels, represented by window H in Fig 4.19,
the gamma rays undergo Compton scattering and are not absorbed. After several collisions, the
gamma rays lose sufficient energy to be absorbed by electrons. The photoelectric measurement
is made by registering those gamma rays that fall in window L, positioned at very low energy. In
this region gamma rays undergo photoelectric absorption the rate of which depends on the
product of the absorption coefficient per electron. Pe, and the electron density, p0. The count rate
at window L is related to a photoelectric absorption index or capture cross section, U, given by:

U - Pe Pe 4.4

Table 4.2 lists the Peand U values for the compounds that are commonly encountered in the
interpretation of the LDT log.

Table 4.2 Pe and U values for various compounds (After Schlumberger) оЛ г


Actual
Com pound Formula Density, pb p a, as seen by и
pe
(gm/cm3) tool (gm /cm 3)

Quartz S i0 2 1.806 2.654 2.650 4.79

Calcite СаСОз 5.084 2.710 2.708 13.77

Dolomite CaMg (СОз)г 3.142 2.850 2.864 9.00

Anhydrite Ca S 0 4 5.055 2.960 2.956 14.95


Gypsum Ca S 0 4 2H20 3.420 2.320 2.372 8.11

Sylvite KCI 8.510 1.984 1.916 16.30

Halite NaCI 4.650 2.165 2.074 9.65

Siderite FeC03 14.690 3.890 3.810 55.90

Pyrite FeS2 17.000 5.000 4.990 82.10

Barite BaS04 266.800 4.500 4.011 1070.00

Water (fresh) H20 0.358 1.000 1.00 0.40

Water (120,000 ppm) H20 0.807 1.086 1.185 0.96

CH16 0.119 0.850 0.948 0.11


Oil
CH2 0.125 0.850 0.970 0.12

0.095 0.11 9pg


и
X

Gas Pg 1-25Pg
Log Presentation

LDT measurements are recorded across tracks 2 and 3 as shown Fig 4.20. The bit size, GR and
caliper logs are displayed in track 1. Normally, however, the LDT is run simultaneously and
recorded with the CNL and NGS logs. Fig 4.21 shows a modern NGS/LDT/CNL log suite. The
current practice is to present the bit size, NGS and caliper measurements in track 1 and display
the LDT/CNL data in tracks 2 and 3. The greater variations in the Pe values of the common
reservoir rocks makes the LDT log a useful aid in their identification.
PEP
- ДГ “
__C R _______
DIUIO
045
.. . r«t ЮЮ6
- - - --- - Mb ----

— П
-L-
■-I J T T "
'7' - T ; - -
JL

v-

\
“ 7 ------- ------- -
i- > ; _____ ____ _
—t
t-
: - -
---------
------- 4-* .- ------
r i ----------
_V -L
Ш
~r ...—
— -------t -

. 1
tr : : . ■
.Г -

i ___ ;_________

- i—
J .
. ------------------

4.20 LDT log presentation (After Asquith & Krygowsky, 2004)


Fig 4.21 An N G S/LD T/C N L log suite

H eavy Mineral Id entification

The LDT log is particularly useful in heavy mineral identification. Table 4.3 lists the pb and Pe

values of some of the more common heavy minerals.


h

T ab le 4.3 P0 and pb values of some heavy minerals

M ineral Pb Pe
Zircon 4.39 69

Siderite 3.89 14.7

Barite 4.10 267

Haematite 5.15 21

Magnetite 5.08 22

The photoelectric absorption index has a better resolution than the Density log as shown in Fig

4.22. Though the Density-Neutron combination does not show any obvious change in lithology,

the Pe curve indicates clearly the presence of a different mineral (Zone A). The zone is a

dolomite-limestone mixture with a siderite streak. The detection of heavy mineral streaks is

particularly useful in well-to-well correlation.

---1

---- —* H•r —-}—I- m — 4jИ н
■1
—. , .!
-
j J-U-l
. .* L -- i-
Д ^—
—Г- •1-r
= Й; -- 1“1-!-tr:
w - M♦ ■.i..Pp t— —* ’H
--
t
4— ■pr ■'
v ..ML
: ---
—r—
I--: <i

Fig 4 .22 Comparison of the LD T and FD C-C N L response


to a heavy mineral (After Schlumberger)

This sensitivity to high Pe minerals is a drawback when drilling with barite-weighted muds. Even
small concentrations of barite will affect the photoelectric absorption index beyond correction.

D ET E C TIO N O F A B N O R M A L P R E S S U R E S

Like the Sonic interval transit time curve, the pb recordings may be used to identify

overpressured shales. Increasing depth of burial causes a progressive increase in compaction,

which in turn raises shale bulk density. Consequently, increasing pb values with increasing depth

may be regarded as the normal trend. The presence of trapped water in undercompacted shales

lowers their density, and this causes a departure from, or even a reversal in, the normal gradient

which can be seen in the response of the pb values curve.

Fig 2.23 shows an example of a bulk density curve recorded over an interval that contains an

overpressured section below 5,500ft. The normal trend (the dashed line sloping to the right) and

the abnormal gradient (the dashed line sloping to the left) can be observed clearly on the pb

curve.
5. T H E E L E C T R O M A G N E T IC PR O P A G A TIO N T O O L (E P T LOG)

IN TR O D U C TIO N

The introduction of the EP T in the 1980s allows the measurement of a new property of rocks,

namely, their dielectric constant or permittivity. Interpretation of dielectric permittivity

measurements makes it possible to distinguish between water and hydrocarbons in a reservoir

regardless of the salinity of the water. Normally, waters associated with hydrocarbon bearing

formations are saline and deep resistivity log (Rt) readings can be used to distinguish between

them (high in hydrocarbon and low in saline water). In cases of low salinity or fresh formation

water, it becomes difficult to differentiate hydrocarbons from water through resistivity variations

since both are characterized by high resistivity.

Rsistivity devices currently in use operate in the 35H z to 30kHz frequency range. By contrast,

the EPT operates at frequencies in the giga-Hz range and at these frequencies the dielectric

permittivity of water is substantially higher than that of hydrocarbons or the rock forming

minerals. Consequently, the formation response comes almost entirely from its water content,

making it possible to detect the presence of water more or less irrespective of its salinity.

Table 5.1 lists the dielectric constants for common sedimentary rocks and fluids.

Table 5.1 Laboratory measured values of relative dielectric constant (relative to air)
of some common sedimentary rocks and fluids (After Schlumberger)

Relative Dielectric Propagation Time


Mineral Constant
». . • *>. .
tp,(ns/m)
v -w ■

Sandstone 4.65 7.2


Dolomite 6.8 8.7
Limestone 7.5-92 9.1-10.2
Anhydrite 6.35 8.4
Halite 5.6-6.35 7.9-8.4
Gypsum 4.16 6.8
Dry colloids 5.76 8
Shale 5-25 7.45-16.6
Oil 2-2.4 4.7-5.2
Gas 1 3.3
Water 56-80 25-30
Fresh water 78.3 29.5
P R IN C IP LE OF M EA S U R E M E N T

The E P T measures the travel time and attenuation rate of a 1.1 x 109Hz electromagnetic wave

as it passes through the formations surrounding the borehole. The tool is a pad contact device,

pressed against the borehole wall by a backup arm which provides also a caliper measurement.

Fig 5.1 shows an illustration of the E P T which consists of two microwave transmitters (T 1t T2)

and two receivers (R i, R2). Spacing between transmitter and nearest receiver is 8cm and

between the two receivers is 4cm. The two transmitters are alternately pulsed, and upgoing and

downgoing travel times measured between the two receivers are averaged. This eliminates

effects of uneven mud-саке thickness, pad tilt, and instrumentation imbalances. Travel time is

measured by sensing the phase difference in received signals at the two receivers. A complete

measurement of travel time and signal attenuation is made every 1/60 of a second and the

measurements are averaged over 2in or 6in depth intervals.

Vertical resolution of the EPT log is extrem ely good. It is essentially the span between

EP T pad configuration and signal paths Antenna pad of the EPT device
receivers, which is about 2in. Depth of penetration is quite small, varying from about 1in in

low-resistivity formations to about 6in in high resistivity zones. The radius of investigation of the

tool is thus limited to the flushed zone, and the lower limit of formation resistivity for proper tool

operation is approximately 0.3 ohm-m2/m.

Borehole size and mudcake thicknesses of up to 3/8in have no effect on the E P T measurements

as long as the pad makes good contact with the borehole wall. Travel time increases with

increasing mudcake thicknesses and at 3/4in the response comes almost entirely from the

mudcake. Hole rugosity is also a problem, since it reduces the degree of contact between the

pad and the borehole wall.

A later adaptation of the EP T enables the device to provides more reliable measurements of

travel time and signal attenuation rates in rugse sections and in the presence of thick mudcakes.

Known as the A D E P T E lec tro m a g n etic P ropagation Tool, it uses more advanced antennas

that reduce signal scatter and interaction with other electrical phenomena.

LO G P R E SEN TA TIO N

A typical log presentation is shown in Fig 5.2. A GR log and standard caliper (hole diameter, HD)

are recorded in track 1, the electromagnetic wave attenuation (EA TT), measured in decibels/m

and propagation time (TPL), measured in nanosec/m, are presented in tracks 2 and 3. There is

also a small arm caliper measurement (SA) displayed in track 2. This provides a more sensitive

caliper measurement than the standard device and is a better indicator of borehole rugosity.

IN TE R P R E TA TIO N

As mentioned above, at frequencies in the gigaHz range, the dielectric permittivity of water is

substantially higher than that of hydrocarbons or the rock forming minerals. Since substances

with high dielectric permittivity or constant are characterized aslo by high electromagnetic wave

propagation time, EPT measurement in clean formations is affected primarily by the water-filled

porosity. This contrasts with porosity values derived from radioactive logs, which respond to total

porosity, and consequently a combination of LDT, CNL and E P T data makes it possible to
№•к*®
GR (GAPlj . - У
100 200 . r v > - .. ••
•: '
GR(6APi) ______ р ш fc A r r (d 8 /m ) ,
5 ■- ' Too ° 1
' ■ ■. y^tfSCw^.1- r. :• »■/•* .
HO (in.) __r w- (ns/m).' _ _ .
^ ------
^5 ~ i__~ ... ,.

Fig 5.2 EP T log presentation (After Schlumberger)

distinguish between oil, gas and water in reservoirs independent of formation water

characteristics.

It should be emphasised, however, that due to the shallow depth of investigation of the tool

(1-6in), it can usually be assumed that only the flushed zone is influencing the measurement and

the formation water is represented by the mud filtrate.


Q uick look H yd ro carb o n In d icatio n

A combination of EPT, resistivity and radioactive porosity log data may be used in qualitative
evaluation of hydrocarbon bearing zones. Fig 5.3 shows schematically how the combination of
Induction, Density, Neutron, and EPT logs distinguish between fresh water, salt water, oil, and
gas. Resistivity distinguishes fresh water from salt water, whereas the other curves do not. EPT
distinguishes oil from fresh w ater while the other curves show only slight change. Finally, EPT
and Neutron-Density together distinguish gas from oil, while the resistivity is not definitive.

Since the EPT log responds primarily to water-filled porosity in clean formations, a qualitative
comparison of EPT porosity with the total porosity derived from the Density, Neutron or the
Sonic tools allows a quick-look identification of hydrocarbons in the flushed zone.

Fig 5.4 is an example comparing the Sonic porosity (SPHI) with the EPT porosity (EMCP). The
porosity curves are displayed in tracks 2 and 3 and the computed gamma ray (CGR) and total
gamma ray (SGR) from the NGS survey are recorded in track 1. There is a change in lithology at
245m, with a limestone above this depth and a sandstone with calcareous cement below. The
limestone and the lower section o f the sandstone are water bearing, and the hydrocarbon
content of the upper section o f the sand is clearly indicated by the separation of the two porosity
curves. The original oil/water contact is at 267m, while the present contact is at 262m.

Fig 5.3 Comparison of Resistivity, Neutron, Density and EPT log responses
in hydrocarbon and water bearing zones (After Schlumberger)
Fig 5.4 Quick look identification of hydrocarbon bearing zones by
comparing sonic and EPT porosities (After Schlumberger)

Generally, the EPT porosity will read the same as a nuclear or acoustic derived porosity in water
bearing zones, but in hydrocarbon bearing intervals the EPT porosity will be less than the total
porosity. In gas-bearing zones the separation between the Neutron porosity and the EPT
porosity will not be so apparent because of the effect of gas on the Neutron measurement.

An example of the application of the EPT to the identification of the hydrocarbon bearing section
in a reservoir containing fresh formation water is shown in Fig 5.5. The LLD and LLS curves read
high resitivities throughout the section and do not differentiate between hydrocarbon and fresh
water since both are poor conductors and exhibit high resistivity. However, when EPT porosity
(Ф ерт) is compared with that derived from the Neutron-Density log combination (Фда). a
hydrocarbon/water contact is indicated at 6,850ft.
Hydrocarbon/water contact

Fig 5.5 Hydrocarbon detection in a fresh water bearing reservoir


(After Schlumberger)

C o n versio n o f EP T M easu rem en ts to P o ro s ity ( Ф ерт )

Travel time and signal attenuation recordings can be used to calculate Ф ерт and the procedures

are reviewed briefly below.

tpo m ethod

The t ^ method is based on the principle that the travel time (tp0) of electrom agnetic waves in a
clean, lossless (low attenuation), porous, water-bearing medium is the sum of two individual
responses, namely, that of the matrix (tpma) and that of the pore water (tpw), the time being

measured in nanosec/m (ns/m). The matrix response is given by the amount of matrix (1 - Ф)

multiplied by tpma, and the water contribution equals the amount of water (Ф) multiplied by tpw.
Thus:

tpo — {1 " Ф)1рта + 5.1

Solving for Ф:

Ф ё РТ — Opo " tpma)/(tpw" tpma) 5.2

tpo is related to the measured travel tim e (tpi, ns/m) and Ac (db/m), the attenuation of the medium

corrected for spreading loss:

tpo = [tpi2 - (Ac2/3 6 0 4 )]1/2 5.3

For carbonates and clean sandstones, the attenuation is negligible so that tpo is the actual tpi

value read from the log.

Values o f tp™ for various rock matrices and of tp„ for fresh water are given in Table 5.1. Values of
the dielectric constant, e , relative to air = I, are also listed in Table 5.1. Propagation time in ns/m

is related to e by the simple relation:

tpl = (11.1f.r)1/2 5.4

Consequently, the measured travel tim e o f a medium is a strong function of its water content.

Actual porosity will be higher if the formation contains hydrocarbons, since ФЕрт derived from

equation 5.2 represents water-filled porosity only.

tpi m ethod

This is a sim pler approach than the tpo method and is based on the relationship between the

actual measured propagation time, tpi, and Ф Ерт and tpw:

tpl (I “ Ф^р-ПЭ ■*" tpw 5.5


Solving for Ф:

Ф ерт = {tPi - tpma)/(tpw “ tpma) 5 .6

Ac m ethod

Ac, the attenuation of the medium corrected for spreading loss, is related to Aw> the attenuation
of the pore water, by the following relationship:

Ac - Ф ерт Aw

5 .7

Solving for Ф ерт

Ф ерт = 5.8

Ферт values obtained by these methods may differ. No one of them may be claimed to be more
reliable than the other two, and the most consistent results are obtained in homogeneous, high
porosity formations.

D erivation of Sxo From The EP T Log

Ф ерт would be the true porosity if all pores were water filled. It approximates the water-filled

porosity in formations containing hydrocarbons since the tool does not distinguish between
hydrocarbons and the rock matrix; travel times in the two media are similar. Consequently, a

comparison between ФЕрТ and total porosity, generally obtained from Density-Neutron logs,
allows a quick-look determination of Sxo:

Sxo- Ф ерт/Ф 5 .9
6. THE N U CLEAR MAGNETIC RESONANCE LO G (NMR)

NMR logging is a relatively new technology that was introduced in the late 1980s and its use has
since expanded rapidly in the industry. Although it provides information on a wide variety of
physical properties and fluid contents of reservoirs, NMR m easurements become most useful
when combined with other log and core data and should not be considered as a replacement of
the latter.

The unique features of the NMR are that it is a lithology independent tool and while it makes a
nuclear measurement, it does not employ a radioactive source.

PRINCIPLE OF MEASUREMENT

The tool responds to the presence o f hydrogen in the formations traversed by the borehole.
Hydrogen nuclei or protons behave like spinning magnetic dipoles (bar magnets), randomly
oriented in the formations, as shown diagrammatically in Fig 6.1. The tool operates by subjecting
the formation to a strong polarising magnetic field by sending an electrom agnetic wave through
a polarising coil, causing an alignment o f the proton spins approxim ately perpendicular to the
Earth's magnetic field. Figs 2 and 3 illustrate the operating mode of the device which measures
the relaxation time - ‘precession’ - o f hydrogen nuclei or protons in the pore fluids.

Fig 6.1 Spinning, randomly oriented hydrogen magnetic dipoles


Polarizing Coil

M agnetic field

Fig 6.2 Schematic representation of the operating mode of the NMR tool. Proton spin is
orientated perpendicular to the Earth’s magnetic field (HE) following the application
of an electromagnetic wave through the polarsising coil (Hp)
(Modified after Schlumberger)

6.3 Proton spins aligned perpendicular to the Earth's magnetic field following polarisation
The time taken for full polarisation is called 7, and for this to occur the polarising field must be
applied for a period about five times 7i. The electromagnetic wave is then turned off, allowing
the protons to 'relax' or 'precess' back to their original orientation. As the protons relax they emit
a weak signal which is detected by an antenna. The relaxation time T2 is measured and
recorded by the device. The antenna acts as both transmitter and receiver.

TO O L RESPONSE

There are three components to T2, referred to as surface, bulk and diffusion relaxations. The
dominance of any one com ponent as the main relaxation process is governed by the pore fluids
and wettability of the reservoir.

Surface relaxation is the dominant process in totally water saturated pores. In this case collision
with grain surfaces is the most important factor in determining T2. In small pores collision with
grain surfaces is frequent, resulting in rapid relaxation. There are fewer collisions in larger pores
and the protons take longer to relax. Pore size distribution therefore affects relaxation times, as
shown diagrammatically in Fig 6.4. Consequently, there is a direct relationship between the
amplitude of the Тг measurement and porosity and permeability. Short T2 times generally

Large pore ' '? 4 \


C'on

• Hydrogen prc:on rtec, msec


A totally water bearing pore Collisions with grain surfaces are more frequent in smaller
pores, leading to shorter T2 times (After Schlumberger)

Fig 6.4 Effect of pore size on T2 in a totally water saturated reservoir

indicate small pores with low permeability, while longer T2 times indicate larger pores with higher
permeability.

Figs 6.5 and 6.6 show examples of the relationship between porosity and permeability and Гг-
The sandstone depicted in Fig 6.5 has a porosity of 20% and a permeability of 8 md. Fig 6.6
represents a sandstone with a similar porosity but a much higher permeability of 280 md.. The
latter is associated with a longer relaxation time on account of its higher permeability.

Fig 6.5 Effect of low permeability on relaxation time (After Schlumberger)

High permeability, producer


_ c.
ctf О Porosity = 19.5%
(Л — Permeability = 280 md

i
.~ * s s

■ ' f :/ Increasing relaxation time

\ . r $
о
Bulk relaxation is predominant in the non-wetting fluid phase in a reservoir and is controlled by
the physical properties o f the fluid, such as its viscosity and density. The left diagram in Fig 6.7
shows a pore in an oil bearing, water wet reservoir. Being the non-wetting phase, the oil is not in
direct contact with pore walls. In this exam ple the proton relaxation mechanism in the oil phase
is illustrated in the right diagram in Fig 6.7. Surface relaxation proceeds in the wetting water
phase.

{Thickness of the water film) ~ .

A water wet oil bearing pore Bulk relaxation in the non wetting oil phase

Fig 6.7 Illustration of bulk relaxation

Gas, oil and w ater exhibit significant m olecular Diffusion induced relaxation. This is caused by
the molecules moving along gradients resulting from variations in the strength o f the magnetic
field produced by the polarizing coil (Fig 6.8).

/■"' - .-*4V
* >
r *> ..
■. - - ^ 4 ,*

Fig 6.8 Illustration of diffusion relaxation

These three processes act in parallel and the overall T2 time is given by the following
relationship:

1/ 7*2 = 1/7surface + 1/7bulk + l/^diflusjon

IN TE R P R E TA T IO N

In conjunction with other log data, NMR measurements yield a wealth of inform ation on rock and

fluid properties and these are discussed below.


Identification o f clay-bound and capillary-bound water

Fig 6.9 illustrates the various types of fluids in a reservoir with intergranular porosity. These
include oil, clay-bound water (water of crystallization associated with clay minerals), capillary-
bound water (a thin water film coating the mineral grains and represents the irreducible water

Free Sand
water

Capillary- Clay-
bound bound
water water

Fig 6.9 Various types o f fluids in a reservoir with intergranular porosity


(After Schlumberger)

saturation, Sw,) and free water. The shape of the T2 distribution curve allows the clay-bound
water, capillary bound water and the producible fluids to be differentiated, as illustrated in Fig
6.10. The T2 cutoff represents the value of T2 that separates bound and free fluids. It is 33 msec
for sandstones and 100 msec for carbonates.
T2 cutoff

01 l'o 10 0 10^0 1000.0 10000 0

T2cutoff: Value o f T2that separates free and bound fluids


Free water + hydrocarbons represent the producible fluids
Fig 6.10 T2 time distribution relating to various types of fluids in a reservoir

Determ ination of irreducible w ater saturation and movable fluids

A principal measurement of the NMR tool is the free fluid index, FFI. It represents the volum e o f
the fluid that is free to move within the pore system and is not associated with the clay minerals
or bound to the surface o f the rock matrix by capillary forces. This fluid volume includes oil and
water but not irreducible water and residual oil:

FFI = 0 (S ; 6.2

Since S*0 + S0r = 1,


-V

FFI = 0(1 - Swi) 6.3


Hence

Sw,= 1 - (FFI/ 0 ) 6.4

Swi can therefore be determined by comparing FFI with a porosity measurement. This can be a
significant factor in reservoirs with high volumes o f clay or silt. In such cases the water saturation
calculated from standard logs can be very high and yet the reservoir may be capable of
producing dry oil. The reason for this anomaly is that the water is associated with the clay
minerals and is not free to move.

An example of an NMR log over a shaly sand interval with calcareous cement is presented in
Fig 6.11. The topmost section is associated with very high SWj values and very low FFI readings.
The latter indicate that the water is not free to move and therefore bound water.

D e te rm in a tio n o f e ffe c tiv e p o ro s ity

NMR measurements provide effective porosity values independent of lithology. This is useful in
com plex lithology reservoirs and shaly sands where it is difficult to derive porosity from standard
logs. Examples of NMR derived porosity curves are shown in Figs 6.12 and 6.13.

P e rm e a b ility e stim a tio n

As mentioned above, proton relaxation time (Г 2) is related to pore size distribution and can be
used to estimate permeability. In granular reservoirs small pore sizes correspond generally to
lower permeabilities while higher permeabilities are associated with larger pore sizes. A variety
of methods are available to determine NMR permeability and the choice is governed by
operating company preference, reservoir conditions and service company. NMR derived
permeability curves are displayed in track 2 in Figs 6.11 and 6.12 and in many cases NMR
permeabilities are comparable with those from core measurements (Fig 6.11). It should be
noted, however, that the current methods resolve only matrix permeability and underestimate
the permeability of fractured formations. Isolated vuggy pores in carbonates also present a
problem.

The vugs often contain free fluids and are associated with high FFI values but their isolation
prevents them from contributing to permeability. Consequently, the NMR permeability in such

cases is an overestimation.

P ore size d is trib u tio n

Since the proton relaxation time of a fluid within a single pore is proportional to the size of that
pore, as a first approximation, the distribution of T2 measurements within a reservoir reflects the
pore size distribution (Figs 6.12 and 6.13). This may also be related to grain size variations in

the reservoir.
H y d ro c a rb o n d e te c tio n

As shown in Table 6.1, reservoir fluids exhibit different polarization (Trf and relaxation times (T2).
These variations are used in MNR logging to detect hydrocarbons. Viscosity contrasts are

Table 6.1 NMR properties of reservoir properties (After Coates et al, 1999)

R e s e rv o ir flu id T i{ m sec) T2 (m sec) V is c o s ity (cp)

CO
Kj
1 -5 0 0 1 -5 0 0

о
Brine

I
О
Oil 3 ,0 0 0 -4 ,0 0 0 3 0 0 - 1,000 0 . 2 - 1,000
0.011 -0 .0 1 4
Gas 4,000 - 5,000 3 0 -6 0
(methane)

particularly useful in identifying heavy oil and tar deposits. The NMR log responds to these
heavy substances as if they are solids - very little or no FFI is measured. However, for the direct
identification of medium to light oils and gas and determining their saturation without
incorporating data from conventional resistivity and porosity logs, specialised pulse sequences
are required.

Fig 6.12 shows an example of reservoir fluid identification by the NMR log.

CO M M ERCIALLY A V A ILA B LE TOOLS

The NMR is a highly versatile log with a wide range of applications. It works in the presence of
most drilling fluids with the exception of high salinity water based muds. Currently, there are two
commercial tools: the Combinable Magnetic Resonance (CMR) tool offered by Schlumberger
and the Magnetic Resonance Imaging Log (MRIL) available from Haliburton and Baker Hughes.
(Figs 6.13 and 6.14).
12
Distribution

Neutron
0.3 3000
Porosity Irreducible water
CMR msec

Borehole Permeability
Density Free water
0.1 m d 1000
Porosity

Gamma
Water
Ray
Depth,
Zone ft 0 API 150

XX700

xxsoo

XX900

Fig 6.12 Formation fluid identification by the NMR log


(After Schlumberger)
Bound Water

T2 Distribution

3ound Fiwd Vc'ume

Density Porosity
Perm eability (md) 2000
0 API 2СС Resistivity (ohm -m) 2000 Pcrosity (%)

Fig 6.14 An MRIL presentation (After Baker Atlas)


7. PLATFORM EXPRESS (PEX)

PEX is a new log data acquisition technology, developed and introduced in thejm id-1990sjand
represents a major departure from conventional wireline logging tools. The system integrates
multiple functions into a single package or p la tfo rm with the various sensors incorporated in the
same device rather than as a series of separate tools connected together in a string. The PEX
device is less than half the length and of a conventional trip le c o m b o which has been in use
since the mid-1980s.

Fig 7.1 provides a comparison between the lengths of a PEX device and the conventional triple
combo: the former is less than half the length and weight of the latter. A summary of the

'

■i

90 ft
{27 rpj

Fig 7.1 Comparison between the lengths of PEX and conventional triple combo
(After Schlumberger)
specifications the two systems is presented in Table 7.1.

T able 7.1 Specifications of the triple combo and PEX (After Schlumberger)

S pecificatio n Triple com bo PEX

Length Typically 90ft (27m) 38ft (12m)

Weight 1,5001b (675kg) 690lb (311kg)


Outside diam eter (OD) 3 3/8 to 4 1/2 in 3 3/8 to 4 5/8 in
Temperature rating 350° F (175° C) 2 5 0 ° F (1 2 0 ° C)

Pressure rating (psi) 20,000 10,000


Maximum logging speed 800ft/hr (540m/hr) 3,600ft/hr (1,080m/hr)

The above PEX specifications relate to the first generation tools. The main disadvantage of
these was their lower temperature and pressure ratings, compared to the triple combo, which
reduced the quality of the measurements in high pressure high temperature (HPHT) regimes
that prevail in several parts of the word, e.g. the Central North Sea basin. These shortcomings
were addressed in the second generation PEX tools which are capable of making reliable
measurements under HPHT conditions.

PEX M E A S U R E M E N T S

PEX records seven petrophysical parameters that include GR, neutron porosity, bulk density

(pb), photoelectric effect (Pe), flushed zone and mudcake resistivities (Rxo and Rmc by the Micro-
Cylindrically Focused Log, MCFL) and deep and shallow resistivity. Fig 7.2 illustrates the two
PEX tool configurations offered by Schlumberger, together with the vertical resolutions of the
various measurements, which vary from 2in to 24in. Improved sensor and tool designs result in
greater logging speed, reliability, efficiency and better quality (higher vertical resolution) data.

The uppermost part of the section delivers GR and neutron measurements with standard vertical
resolutions of 24in and is referred to as the Highly Integrated Gamma Ray Neutron Sonde
(HGNS). Below this is an electronics cartridge, which is the source of the gamma rays and fast
neutrons used in density and neutron logging. Bulk density is recorded by the Three Detector
Lithology Density Tool (TLD), which provides an improvement over the standard dual detector
measurements. Other features of the TLD device include higher precision in denser formations
and less sensitivity to barite which results in better Pe measurements. The TLD tool and MCFL
are housed in the High Resolution Mechanical Sonde (HRMS) which also incorporates a caliper.
Hinge joints above and below the HRMS enable the tool to better negotiate borehole wall
irregularities (Fig 7.3). This improves pad contact and maintains density and Rxo log data quality
in washed out and rugose sections. Real time log quality control allows corrections to be made
to off-depth log readings caused by tool sticking. Fig 7.4a presents a standard GR-DLL-MSFL
showing off-depth readings resulting from tool sticking. In Fig 7.4b the ‘speed correcting’ of the
high resolution PEX data has removed the mis-match between the GR and resistivity curves.

Tool
acceleration

GR
Highly Integrated 24 m
Gamma Ray ~ 0N
Neutron Sonde 12 t o 24 r\.
(HGN3)

Bectronics
cartridge
t > b .P e
H inge 2. 8, 18 in
joint R*o, Rrnc

High-Resofuuon
Mechanical
Scnde

Caliper

Hinge
joint

High-Resoiution AIT Array induction


Azimuthal Laterolog —ti Я». Rf
Imager Tool
Sonde iHALS)

Rt
12

*
HALS AIT

Fig 7.2 PEX tool configurations Vertical resolutions of each measurement


are indicated (After Schlumberger)
The lowermost section of the tool delivers deep and shallow resitivity measurements derived
either from a laterolog (High Resolution Azimuthal Laterolog Sonde, HALS) or an Array
Induction Tool (AIT), depending on the resistivity of the mud in the borehole. Other
measurements include mud resistivity (Rm) and temperature.

Fig 7.3 Hinge joints improve pad contact in washed out and rugose sections
resulting in more accurate density and microresistivity measurements
(After Schlumberger)

P l a t f o r m Ex p r e s s

Speea-CofTftC!fc<lK»gh-Resn*u?ron lLS

SPC W ) COn*2Ct«4j H*jh-RsscM -on IL D

5pe®> l -C txfectod H tgh-PesoM wo RXO

—4-~fj—
±

7ТГ—

Fig 7.4a Standard GR-DLL-MSFL suite Fig 7.4b Speed corrected high resolution
showing off-depth log readings PEX data
resulting from tool sticking
(After Schlumberger)
ADVANTAG ES

The high vertical resolution data produced by PEX logging are useful in the identification of thin
beds. In Fig 7.5 three 2-in tight streaks (seen also in the microresistivity log not shown here) are
detected by their high density log readings. These can act as vertical permeability barriers.

Fig 7.5 High resolution density measurements identify three 2-in tight streaks (seen also in
the microresistivity log not shown here) which can act as vertical permeability
barriers (After Schlumberger)

Due to higher speeds, PEX logging operation takes less time and the reduced weight and length
make the tool easier to handle. The savings in time are reflected in reduced rig and average
logging times, lowering the cost of data acquisition. Fig 7.6 shows rig time comparisons between
the conventional triple combo and PEX in land and offshore operations in Venezuela while Fig
7.7 provides the same illustrations in Saudi Arabia and Argentina.
T r ip le C o m b o v s . P l a t f o r m E x p r e s s R ig T im e

Average lost time


Repeat section
■ Calibration
■ Logging time
■ Rig up/down
О Drill rathole

Land O ffs h o re

Fig 7.6 Rig time comparisons between the conventional triple


combo and PEX Venezuela (After Schlum berger)

Saudi Arabia Argentina


7000-ft w ell 7000-ft w ell
2500-ft openhole 2500-ft openhole

Run 1: AIT-LDT-CNL-MSFL-GR
Run 2: DSI
AIT-LDT-CNL-MSFL-GR

Platform Express

7 h r4 0 m in 4 h r2 0 m in 7 hr 3 h r2 0 m in

Time
Drilling rathole ■■ Calibrations Ш Logging
'ЗЕЭ Rig up, rig down Шк Run in, pull out

Fig 7.7 Rig time com parisons between the conventional triple combo and
PEX in Saudi Arabia and Argentina (After Schlumberger)
, v < './Л . . ЙС w*h©ir
h ;<]V
и

8. LOG INTERPRETATION

As discussed in Part 1, log interpretation has qualitative and quantitative aspects. The form er is
concerned with the detection of zones of interest (permeable formations containing
hydrocarbons) and the determination of lithology, while the latter involves further evaluation of
these zones through the quantification of Ф, Sxo, Sw and VSh and, ultimately, the magnitude of the

recoverable reserves.

Q U ALITATIVE INTERPRETATION

Characteristics associated with permeable hydrocarbon bearing zones may be summarized as


follows:

(a) Low GR read ing - potential reservoir rocks are normally characterised by a low GR reading
relative to shales. Typical GR values associated with clean reservoirs range from 10 to 25
API units..Certain types of reservoirs may, however, produce high GR log responses, in
which case the measurements cannot be used as a shale indicator. Examples of these
include micaceous and arkosic sandstones (containing granules of feldspar). The actual
number of gamma ray units will depend on many factors such as:

(i) The mud - KCI mud gives a higher GR reading:

(ii) Hole size - a larger hole gives a lower GR reading for non-KCI muds;

(iii) Any casing present reduces the reading since it absorbs some of the gamma radiation .

(b) Some SP feature - positive or negative (Rmf * Rw).

(c) A Sonic log reading greater than 52 (.isec/ft.

(d) High R, (high LLd or ILd reading) and R, (high R Sfl , R ilm or RLLS) readings. Hydrocarbon/salt
water contacts are marked by a drop in resistivity.

(e) Indications of hydrocarbons in cuttings and cores.


Lithology Determination and Gas Detection
v
$ Fig 8.1 is an illustrated presentation of the log values associated with the common lithologies
and fluids and Fig 8.2 shows the idealized GR-LDT-CNL log combination responses to the rock
§ types frequently encountered in oil and gas wells.

'I

X'

Fig 8.1 Log values to common lithologies and fluids (After Baker Hughes)

4 G The GR-LDT-CNL combination gives a good lithological indication except in the presence of
"4_
X ^ gas. A tight, clean sandstone is associated with a pb reading of about 2.65gm/cm3 and a Фм of

- 4. The Фм curve is to the right of the pb curve, the separation between them being 2 -4 scale
v / \ \i
' a divisions. Tight limestones are characterised by the pb overlying the Фм curve with the former
L- ч
4 reading 2.7gm/cm3 and the latter zero porosity. In a tight dolomite Фм reads about + 4 and pb is

54 about 2.87gm/cm3. In this case, pb is deflected to the right of the Фм curve. Shales or mudstones

are associated with very high Фм values (up to 45 p.u. in some cases). Consequently, the Фм

curve moves to the left and crosses the pb curve. This cross-over is matched by a deflection to
Сл-> ’/ М

the right (in the direction of higher API readings) in the GR curve.

In permeable formations the curves mtive to the left, pb in the direction of lower densities and <t>N

towards higher Neutron porosity values. Fig 8.3 shows GR-FDC-CNL responses to various


lithologies and Figs 8.4 and 8.5 demonstrate their responses to carbonate-evaporite sequences.
If gas is present, the separation between the curves increases considerably; the Фм curve

moves to the right, in the direction of lower porosity (due to the low hydrogen index of gas), and
the pb curve slightly to the left, towards a somewhat lower density value. This is known as the

gas effect. Figs 8.6 is a general illustration of this phenomenon and Fig 8.7 shows a gas
bearing sandstone reservoir. A gas bearing limestone reservoir is shown in Fig 8.8.

A gas-bearing shaly sandstone may cause confusion. Since gas and shale have opposite effects
on the Neutron log, the influence on one may cancel that of the other, resulting in the elimination
of the gas effect on the Density-Neutron combination. However, a reference to the GR curve
should resolve such a situation.

Im p ro ve d C lay M ineral Id e n tific a tio n

In the discussion of the Natural Gamma Ray Spectrometry Tool (NGS), it was pointed out that
clay mineral identification on the basis of NGS data alone is not entirely free from ambiguity, and
that better results could be obtained by cross-plotting these data with photoelectric absorption
(Pe) m easurements made by the LDT log.

Figs 8.9 and 8.10 present two charts established for the express purpose of combining Pe and
NGS data. The chart in Fig 8.9 is entered by plotting the potassium concentration against the
corresponding Pe value, while in the chart in Fig 8.10 the thorium/potassium ratio is computed
and crossplotted versus the Pe measurement. Agreement between the results of these two
crossplots improves confidence in the clay mineral determination. General areas rather than
unique clay mineral poles are indicated on the charts. This is due to the wide variations in clay
mineral composition which cause some scattering of the points at which the minerals plot.

W IRE LINE LOG CHARACTERISTICS OF POTENTIAL SOURCE ROCKS

Source rocks are organic rich clay or carbonate muds that in their natural setting have generated
and released sufficient hydrocarbons to form a commercial accumulation of oil and gas. They
can be of marine or lacustrine (fresh water) origin. To function as a source rock, the sediment
must contain a minimum of 1.5 - 2% by weight of organic matter.

They can be identified qualitatively by their GR, neutron, d e n sity, s o n ic and re s is tiv ity log
responses which are briefly described below.

Л\
Top reservoir

GOC

ow e

Fig 8.6 Identification of gas, oil and water by a Neutron-Density-resistivity combination


(Shell publication, 1994)
Fig 8.8 Limestone interval showing gas effect
(After Schlumberger)
10
с
О
о
0) G la u c o n ite
00
с B io tite
C h lo rite
о 2
О g

о
CD llite
О.' оз □
о со
о M u s c o v ite
.с: M o n tm o rillo n ite t
CL
K ao lim te <£>
а> CO
■£}
О. c*J

О
2 4 6 8 10
K, P o ta ssiu m C o n c e n tra tio n (% )

Fig 8.9 Pe - К crossplot for mineral identification (After Schlumberger)

10 i

c
О
о
a> G la u c o n ite
CO
(Л с B iotite C h lo rite
O 2
о й CD
о uu
о 1 M ixed Layer
_<D Illite
О) пз
О СП M u sco vite
о ’ и -------p
0L
M o n tm o rillo n ite
K a olinite
CL

0
0 .1 0 .2 0 .3 0 .6 1 2 3 6 10 20 30 60 100

T h /K , T h o riu m -P o ta s s iu m Ratio
Marine source rocks are characterised by higher GR log readings than non-organic shales. This
is related to the association of uranium with marine organic matter (Fig 8.11). By contrast,
lacustrine source rocks are not associated with GR anomalies due to the scarcity or absence of
uranium in fresh water (Fig 8.13).

Organic matter is rich in hydrogen. Consequently, organic rich shales show higher neutron log
values than their organic lean counterparts (Fig 8.11).

Fig 8.11 GR and neutron log responses associated with


the Upper Jurassic Kimmeridge shale, North Sea
(Meyer & Nederlof, 1984)

Organic rich shales are charaterised by lower densities than normally compacted shales. This is
due to the low density of organic matter, which is approximately equal to pw, i.e. about 1gm/cm3.

Shales containing organic matter are therefore associated with relatively low pb readings (Figs
Fig 8.12 Density and resistivity log responses associated with the
Lower Jurassic Posidonia shale, southern Germany
(Meyer & Nederiof, 1984)

There is an increase in t in shales containing organic matter compared to those with little or no
organic content. Low density shales are generally charaterised by higher t values than normally
compacted shales. Hence the presence of relatively low density organic matter in source rocks
results in an increase in t in these sediments (Fig 8.13).

The organic matter itself and the hydrocarbons generated as source rocks mature are poor
conductors and are therefore characterised by high resistivity. As the result, organic rich shales
have higher resistivities (Fig 8.13) than those with little or no organic content. Since source rocks
tend to be laminated due to the occurrence of the organic matter thin layers, a resistivity tool
with a high vertical resolution is the most effective indicator of organic rich shales. Such a tool is
a microresistivity device which has a vertical resolution of 2in or 5cm.

Fig 8.14 GR, density, sonic and resistivity log responses associated with an Oligocene
lacustrine source rock sequence in Indonesia _ ,
(Meyer & Nederlof, 1934) л \|
\J U o V v V U » ~
. . А X й -
QUANTITATIVE INTERPRETATION v>

^ A О У*
INTRODUCTION
V - ^ Vv
Much care is needed when reading values from the curves for manual interpretation. It should a
r c\

be emphasised that logs are not perfect recordings of formation parameters and suffer from
defects such as overall depth mis-matching, variations in cable stretch, different vertical
resolutions of different tools and the effect of statistical variations on the radioactive
measurements. There are also problems caused by poor borehole wall conditions.

Depth mis-matching can result in one logging run being at a different depth from another, a
common cause of which is cable stretching. The cable stretches under its own weight as well as
that of the logging tool. As the tool is pulled up the borehole, the length of cable is reduced,
resulting in a progressive decrease in the amount of stretch. This causes the depth-mismatch to
change in the course of the logging operation.

Combining several tools on the same run results in the measure points of the various curves not
being at the same level - some measurements may be memorised so that they can all be
recorded at the same depth. Another factor is the vertical resolutions of the various tools which
vary from 2in to 4ft. Consequently, thin beds can appear as only small changes in one log while
on curves produced by a tool with a better vertical resolution the same bed may appear as a
much larger deflection. In order to minimise these problems, it is recommended that when taking
readings from the logs, zones at least 2m (6ft) thick should be chosen, preferably where the
curves show relatively constant values. An eyeball average should be taken over the selected
interval.

Large washouts and borehole wall rugosity can affect the data quality by keeping the tools away
from the formation. The pad contact devices such as the Density and microresistivity logs are
affected the most. In large washouts, the pad loses contact with the borehole wall and its
readings become invalid, since they represent mud rather than formation parameters. The Sonic
log measurements may be affected by cycle skipping and the CNL may also read too high due
to a large contribution from the mud.

Sticking is also a problem and distorts the thickness of the section. It can be detected from the
tension curve and stationary readings on the non-radioactive logs. It should be noted that the
sticking may occur at different depths for the different curves because of the depth difference
between measure points.
It should be noted that logs are not the only source of information from a well. There is also a
litholog, a mud log, perhaps some core data with measured helium porosities and permeabilities
and side wall sample descriptions. The litholog is very useful since resolves the ambiguities that
sometimes arise in log interpretation.
The interpretation procedures presented here constitute what is known as the 'quick-look'
method, which is adequate for rapid well-site evaluation. Provided the corrections are properly
applied and the interpretation is carefully made, porosities can be expected to be accurate to ±2

units and water saturations to ± 0.10 in an uncomplicated reservoir.

OBJECTIVES

As stated above, the object of quantitative interpretation is to determine Ф, Ф, Sxo, Sw and Vsh.
Determination of Sw is particularly important, since it leads to a value for hydrocarbon saturation
(Sh) in the uninvaded zone:

Sh = 1 "Sw 8.1

Since Sw represents the fraction of the pore space filled by water, it follows that the porosity
fraction occupied by hydrocarbons is:

Ф(1 - Sw) 8.2

Expression 8.2 is an important factor in the estimation of the total volume of hydrocarbons
present in a given reservoir

Sw is derived from the Archie equation, a relationship developed in the 1930s by G.E. Archie,
and published formally in 1942 (see below). Porosity is one of the essential parameters in the
Archie equation, and must therefore be determined as accurately as possible. Knowledge of Vsh
is also important, since if it exceeds 15-20%, then the Sw value derived from the Archie equation
is unreliable.

Porosity may be derived from any one of the porosity logs (Sonic, Neutron or Density), as
demonstrated previously. However, by crossplotting information from two porosity logs, more
reliable values of Ф can be obtained. Crossplotting provides also a resolution of lithology for

mixtures of up to two minerals.

The steps involved in quantitative log interpretation are described below and summarised in the
flow chart presented in Fig 8.14.

P O R O SITY A N D LIT H O L O G Y D E T E R M IN A TIO N BY C R O S S P L O T T IN G

The porosity logs afford four possible crossplots:

Neutron-Density
Sonic-Neutron

Sonic-Density

P e"Pb
For lithology determination, two dom inant minerals must be assumed. The crossplots and their

relative merits are discussed below.

' /
The N e u tro n -D e n s ity C ro ssp lo t

This is the most effective combination for l;ithology determination and porosity estimation. Figs
8.15 and 8.16 present the current Neutron-Density charts, on which the On and pb values of a

given zone are crossplotted. For a clean, gas-free and monomineralic formation the point will fall
on one of the lithology lines shown, and the true porosity of the formation is indicated by the
graduations along these lines. A point representing a mixture of any two of the three lithologies
shown will fall between the lines. Sandstone presents a small problem, since its density depends
to a small extent on the type and amount of cement. If the cement is calcareous, the sandstone
lithology can be displaced slightly towards the limestone line.

In the examples shown in Figs 8.17 and 8.18, Ф0 = 15, corresponding to a pb value of

2.55gm /cm 3, and Фм = 21 p.u. 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 is found to correspond to about 40% dolomite, 60% limestone.

An error in choosing the matrix pair will not result in large error in the porosity value found, as
long as the choice is restricted to quartz (e.g. sandstone or chert), limestone, dolomite, and
anhydrite (shaliness and gypsum are excluded). This is due to the equiporosity graduations on
the lithology trends falling approximately on a straight line. For instance, in the above example, if
the lithology were known to be quartz and dolomite instead of limestone and dolomite, the
porosity found would be 18.3% instead of 18%, and the mineral proportions would be about 45%
quartz and 55% dolomite.

The separations between the quartz, limestone and dolomite lines indicate good resolution for
these lithologies. Also, the most common evaporites (rock salt, anhydrite and gypsum) are easily
identified.

A Neutron-Density crossplot and the resulting lithology determination relating to a carbonate


section within the Fusselman Formation of Silurian age in West Texas is presented in Fig 8.19.
The data points fall between the limestone and dolomite matrix lines, indicating that the
formation is a mixed carbonate unit.

The chart includes also a provision for gas correction (top left-hand corner of Fig 8.15). In the
presence of gas, <J>N is lowered, and consequently gas-bearing zones tend to plot above the

Porosity and Lithology Determination Irom


Formation Density Log and CNL' Compensated Neutron Log Porosity and Ltthoiogy Determination from
For CHL iocs felcre !93o 01 taDH?c NPHl Formation Density Log and CNL‘ Compensated Neutron Log
- V CfcL iOijs fcefcrj 1386, or а-ик-d fiPHi

**!• ЗСиСЫЙЧС*lt

Mud filtrate concentration < 100,000 ppm Mud filtrate concentration > 100,000 ppm

Fig 8.15 FDC-CNL cross plot charts (After Schlumberger)

Porosity and L:thj:ijiyy Determination (т,гп Porosity and Litho'ogy Determination trcrr
Liiho-Densily' Leg and CNL" Compensated Neutron Log [jtho-Density’ Log and CNL' Compensated Neutron Log
CfIL vlir.es ih ti -93? :гэе)и) TJJPrt

1КОo'sw.С...оpar ., 1ЛЛ*Чв\«(1.П C..«rsw*«e,


X
/ }* ■ У у
*
- . Л ■
i . ■ KP
у
j \j/ ' >
* *
.. : . \X . _
• У; У 5!* jj
... ; . У
X ■У '' у ' : .
/ У • * i

У У 1 1

Mud filtrate concentration < 100,000 ppm Mud filtrate concentration > 100,000 ppm

Fig 8.16 LDT-CNL cross plot charts (After Schlumberger)


sandstone matrix line. Porosity can be corrected for the effect of gas by projecting the point,
parallel with the gas correction arrow, onto the appropriate matrix line as shown in Fig 8.20. A
Neutron-Density crossplot of several hundred data points from a Palaeocene age sand-shale-

k £
]
1 / £
40
Ъ /, /
,К ' $ / с / и ? / '- / t s
<$>/ ! /

30 / n d e s p c w e /a n '/'
УЖ ь У л /

P -f I I K o t С
$ 20
s > /
^ 4 1 л»
/ .'■•О0
10

О
\o \
а /
У

CJ
■у
/
-10

0 10 20 30 40
pc n l (L im e s to n e ) c>cnl (Limestone)
Fig 8.17 Porosity and lithology Fig 8.18 Porosity and lithology
from determination determination from the
FDC-CNL crossplot the LDT-CNL crossplot
(After Schlumberger) (After Schlumberger)

к
-A :-,.

si

SP-GR-DIL-SFL-LDT-CNL suite
over the interval evaluated

Fig 8.19 A Neutron-Density crossplot and the resulting lithology determination, Fusselman
Formation carbonates, West Texas (After Asquith & Krygowsky, 2004)

YY\V l ( (A .
tuff (lithified volcanic ash) sequence in a Northern North Sea well is presented in Fig 8.21.

<>n----- •­
Fig 8.20 Hydrocarbon correction (After Schlumberger)

Fig 8.21 8Neutron-Density analysis of the Palaeocene section from a well in the Northern North Sea
Key to numbered clusters:
1, 2 & 5 - claystone, shale and minor tuff 3 - gas bearing sand
4 - Massive tuff 6 - oil bearing sand 7 - water bearing sand
(After Hatton et at, 1992)
Variations in lithology and fluid content allow four clusters to be identified. The oil and water
bearing sands (cluster 6 & 7) appear to be relatively clean while the scatter of points falling
between the sandstone and dolomite matrix lines are shaly. The presence of gas in the sands
represented by the points in cluster 3 causes them to plot above the sandstone trend.

The S o n ic-N eu tron cro ssplo t

This crossplot is also a good indicator of porosity and lithology. Fig 8.22 shows the
Sonic-Neutron crossplot charts relating to pre-1986 (NPHI) and post-1986 (TNPH) CNL logs.
The charts are entered by the t and <&N readings of a given zone. As with the Density-Neutron
plots, resolution between quartz, limestone and dolomite lithologies is good and the equiporosity
graduations on the 3 matrix trends fall approximately on a straight line. Porosity determination is
therefore largely independent of lithology.

P orosity and Litho lo gy D e term in atio n from P orosiry a nd Lithology D eterm ination from S onic Log
S on ic Log a nd C N L ' C o m p en sa te d N eutron Log and CNL* C om p en sa te d N eutron Log
Fo: CNL tngs neJore 19Я6. cr iSDffril rjf-Hl fur CM logs alter 1986 !ab"iad >4?H

!. - 1SO ..-икУП; С, - 0
ppm

I* /
V /
//■ ,/
/
I
/ / / /л f л
j 9
/ ^!
4' /

/ jJ,/ ".
,/ r ! ■
У

Fig 8.22 Sonic-CNL cross plot charts (After Schlumberger)

However, there is no provision for gas correction, and this crossplot should not be used in
gas-bearing formations.
carbonate interval as that used in Fig 8.19 is presented in Fig 8.23. It should be noted that this
crossplot indicates a much greater proportion of dolomite than limestone. This is due to the
occurrence of vuggy porosity in the interval 9,088ft - 9,126ft, which is ignored by the Sonic but
not by the Neutron-Density combination. The Sonic tool responds only to primary intergranular
or intercrystalline porosity while the Neutron-Density logs measure total porosity. Consequently,
the porosity read by the Sonic is lower than that measured by the Neutron-Density, causing the
points to shift downward on the chart, i.e. cluster closer to the dolomite line, making the interval

( ) '
■J . ' /
:/ r ■
• .
i f ' :
:
•■ /

Я
GR-Sonic log suite over the
0.00 0II) 0.20 0.30 interval evaluated
Neutron Pnrositv. N'PHI

Fig 8.23 A Sonic-Neutron crossplot and the resulting lithology determination, Fusselman
Formation carbonates, West Texas {After Asquith & Krygowsky, 2004)

appear more dolomitic. This underlines the importance of having access to independent
geological information - core description in this case - while interpreting logs. Such data would
resolve the inconsistencies and ambiguities of this kind.

The S o n ic -D e n sity crossplot

The chart is entered by t and pb values (Fig 8.24). Porosity resolution is poor in this case, as the
equiporosity graduations on the lithology trends do not fall on a straight line. Consequently,
porosity determination is not independent of lithology and an error in the choice of mineral pair
from the quartz-limestone-dolomite group can result in appreciable error in porosity estimation.

Lithology resolution is poor since the matrix lines are closely spaced, making it difficult to
differentiate between the minerals. The crossplot does, however, distinguish effectively between
the common evaporite minerals. As seen on the chart, good resolution is achieved for salt,
gypsum and anhydrite, since these mineral poles are widely separated.

Lithology Identification from


Formation Density Log and Sonic Log (E

If = 169 usee/ft; p, = 1.0

■*0 50 ftO 70 80 ЭД 100 110 120


1. sonic transit time (usec/fi)

Fig 8.24 Sonic- Density cross plot chart (After Schlumberger)

Fig 8.25 illustrates a Sonic-Density crossplot of the Fusselman Formation. The lithology of the
interval 9,088ft - 9,126ft, characterized by vuggy porosity, is now shown as limestone.
Underestimation of the porosity by the Sonic in the presence of vugs causes the points to shift
leftward on the chart and cluster around the limestone line.

The inconsistencies in lithology determination from the three crossplot techniques in relation to
interval 9,088ft - 9,126ft are in themselves significant: they indicate that the section has vuggy
porosity. Had all the porosity been primary intergranular or intercrystalline, the crossplots would
indicate similar lithology.

It should therefore be noted that lithology determinations that involve the Sonic log may be
unreliable if the interval in question has vuggy porosity.

2.20 FEET
MD Lithology

3. 00
40 50 60 70 80

A coustic W ave Travel Time, DT Dokffiiiii

, t U A
Legend
Shale
Ш22ЕВ1

Fig 8.25 A Sonic-Density crossplot and the resulting lithology determination, Fusselman
Formation carbonates, West Texas {After Asquith & Krygowsky, 2004)

The Pe - pb crossplot

Fig 8.26 presents the Ре-рь crossplot charts which are entered by the Pe and pb readings of a
given zone. A clean, gas-free and monomineralic zone will plot on one of the matrix lines
provided it is one of the lithologies shown on the chart and its porosity is indicated by the
graduations on the lines. A clean mixture of any two of the three minerals shown will fall
between the matrix lines and, assuming two dominant minerals, the relative amounts of these
may be estimated by proportioning the distance between the curves. As in the Density-Neutron
crossplot, the wide spacing between the matrix lines allows good lithology resolution.

Porosity may be determined by interpolating between the equiporosity lines joining the stems
representing the dominant lithologies assumed. However, the like Sonic-Density crossplot, the
equiporosity graduations on the lithology trends do not fall on a straight line and consequently
porosity determination is not independent of lithology; an appreciable error in porosity
determination will result from an incorrect choice of mineral pair. Also, there is no provision for
gas correction.

Porosity and Lithology Determination Porosity and Lithology Determination


from Litho-Density* Log from Litho-Density' Log

t'tniv u

у
■it:
I?

1
U

Mud filtrate concentration < 100,000 ppm Mud filtrate concentration > 100,000 ppm

Fig 8.26 LDT crossplot charts (After Schlumberger)

The advantages and limitations of the crossplot techniques outlined above are reviewed in Table
8.1. The Neutron-Density combination is the most effective and the most commonly used
crossplot.

Table 8.1 Relative merits of the various crossplot techniques

C ro ssp lo t A d van tag es Lim itations

Neutron-Density Good resolution of lithology Large holes and borehole wall


rugosity may render the Density log
Porosity determination
data invalid
independent of lithology
Density measurements affected by
Provision for gas correction
heavy drilling muds
Sonic-Neutron Good resolution of lithology The Sonic under-reads porosity when
vugs are present
Porosity determination
independent of lithology
No provision for gas correction
Less sensitive to poor borehole
conditions
Sonic-Density Effective in evaporite mineral Poor resolution of lithology
identification
Porosity determination not
independent of lithology

No provision for gas correction


Pe " P b Good resolution of lithology Porosity determination not
independent of lithology
Both measurements made by
the same logging tool
No provision for gas correction

V sh D eterm ination

There are many different ways of determining Vstl in a shaly formation. The most rapid and
popular method is by the use of the GR curve, as demonstrated earlier. The equation used is:

Vsh —(GRzone " GRclean)/(GRshale " GRclean) 8.3

Graphical methods are also used commonly. These involve cross-plotting readings from two
porosity logs on a chart on which three components or points - the rock matrix, shale and fluid
(usually water) - must be identified (Fig 8.27):
Matrix - denoted by M and defined as all grains or crystals, no shale, no porosity and no
fluid.
Shale - designated S and defined as all shale, no matrix and no fluid.
Fluid - called F and defined as all fluid (100% porosity, no matrix and no shale.
Porosity increases along the line MF from 0 at M to 100% at F. Shale content, Vsh, increases
from 0 at M to 100% at S. The lines MF and MS are divided into 10 equal parts, each division
representing an increase of 10% in porosity and Vsh respectively. Fig 8.27 illustrates a Neutron-
Density crossplot chart constructed for a sandstone reservoir with a matrix density of
2.65gm/cm3.

Fig 8.27 Vsh and porosity determination in a shaly sand (After Schlumberger)

The matrix point coordinates in this case are therefore 2.65 and 0. The coordinates of the fluid
point are 1 (density of fresh water) and 100 neutron porosity units (<I> n )- The shale point has the
coordinates pbshaie and <&Nshaie, obtained from a shale interval associated with the reservoir.

The clean zones in this example will fall on MF (clean trend). Shaly zones plot below MF and
point A in Fig 8.27 represents such a case. Its shale content can be estimated by projecting the
point parallel with MF onto MS and its shale corrected porosity by projecting it parallel with MS
onto MF, as demonstrated in Fig 8.28.

In practice, formations with pb values of less than 2.00 gm/cm3 and <J>N readings of greater than
i/Vater

1 .2 / /

1.4
C o n d itic >ns:

(?m<a =
@sh
2 .6 5
*
2 .4 5
80,
4? /

£?w
1.6
£
\C)\^\ 60,­
j / 7
// ,
1.8
<p 1
/
<?°
40 j ^ 7 У
6b 2
/ - V
/
2.2 4 / / l 7-7

2 .4 Z /
7
/ "'Shak f
/
S
2.6 Matri> & O’ _I
о [Vlo)
\J s ^
M
________

2.8
i

'<5 1969 Sc erger


I
О 10 20 30 40 50 60 70 80 90 100
о N

Fig 8.28 Determining Vsh and shale corrected porosity (After Schlumberger)

50 PUs are rare and only the bottom left quadrant of the crossplot is relevant in practice. Manual
crossplotting can be undertaken by using the charts supplied by the service companies and the
following procedure outlines the Neutron-Density method (Fig 8.29):

(1). Plot the shale point on the Neutron-Density crossplot chart. The values for the shale point,
Pbshaie and 0 NShaie. can be obtained from the log by finding a neighbouring shale bed. If the
shale is within the reservoir, it is advisable to select maxima on the curves, since such beds
Fresh water, liquid-filled holes (p, = 1.0)

2.0

2.1

2.2

2 3

§ 2 .4
>4

с
В
•о
^ 2 .5
оэ
е?
26

WfRix 7

2.8

2 .9

О 10 20 30 40
OeNLcor. neutron porosity index fp .u.j (apparent lim estone porosity)

Fig 8.29 Neutron-Density cross plot constructed for porosity and Vsh determination in a
shaly sand (M odified from Schlumberger)

are likely to contain some silt.

(2) Straighten the curved, low porosity end of the sandstone lithology trend and establish a
matrix point M as shown in Fig 8.29. Shaly zones such that represented by A plot below the
sandstone matrix line.

(3) Draw equiporosity lines parallel to the zero porosity line M-S.

(4) Divide M-S into ten equal parts and draw lines upwards parallel to the matrix line. These are
equi-shale fraction lines.
(5) For each point plotted, the shale fraction Vsh and Ф can now be estimated from this grid. The
shaly sand represented by point A in Fig 8.29 has а Фм reading of 16 PUs and a pb value of
2.38gm/cm3. The true porosity (corrected for the effect of shale) is found by projecting A
parallel to the equiporosity lines onto the matrix line established in (2) above, and amounts
to 13%. The shale fraction in this zone is determined by projecting A parallel to the
equi-shale lines on to the M-S line, giving a Vsh of about 20%.

(6) The case of a calcareous cemented sandstone is illustrated in Fig 8.30. The crossplot points
from such a reservoir fall below the sandstone matrix line due to the presence of the calcite
cement which causes them to shift towards the limestone lithology trend. The matrix trend in
this case is established by drawing a line originating at the zero porosity point and passing
through the plots closest to the sandstone line. The porosity graduations on the matrix trend
are determined by the intersections with lines joining equiporosity points between the
sandstone and limestone lines.

F re s h w a le r, liq u id - fille d h o le s (p f = 1 .0 )

n e u tro n p o r o s ity m ci^x (p .u .) (a p p a re n t lim e s to n e p o ro s ity )

Fig 8.30 Neutron-Density cross plot constructed for porosity and Vsh determination in a
calcareous cemented sand (M odified from Schlumberger)
It is also possible to use the Sonic-Density combination to evaluate Vsh graphically (Fig 8.31). On
this chart the lithology lines are closer together and an error in the choice of matrix lines does
not significantly affect the Vsh determinations. Plotting the tshaie and pbShaie measurements from a
shale bed defines the shale point. Equi-shale fraction lines are established by dividing M-S into
ten equal parts and drawing lines through the divisions parallel with the sandstone matrix trend.

Shaly zones fall below the sandstone matrix trend and their shale content can be determined by
projecting them parallel with the equi-shale line onto MS.

t f = 189 usec/ft, p( = 1.0

t. sonic transit time (jisec/ft}

Fig 8.31 Sonic-Density cross plot constructed for Vsh determination in a shaly
sand (Modified from Schlumberger)
It is likely that the Vsh values determined for a given zone by the above three methods are
different. If this is the case, the convention is to record the lowest of all three values.

FO R M A TIO N R E S IS TIV IT Y FA C T O R (F) - Ф R E LA TIO N S H IP

The development of the modern petrophysical logs began with a series of experiments by
Conrad Schlumberger in 1912, and the first electric log was run by H.G. Doll in 1927 in
Alsace-Lorraine, eastern France). Modern quantitative interpretation methods were pioneered by
G.E. Archie in the 1930s, and the results published in 1942.

The first step in this direction was the demonstration by Archie that the resistivity of a totally
water-saturated, clean formation bears an almost constant relationship to the resistivity of the
water saturating it. This constant is called the form ation resistivity factor, F. If R0 is the
resistivity of a clean formation 100% saturated with water of resistivity Rw, then:

F - Rq/R w 8.4

and

R0 - FRW 8.5

R0 is also known as w et resistivity. For a given porosity, F remains nearly constant for all
values of Rw below lohm.m. F diminishes if Rw exceeds lohm .m and as grain size decreases.

This led to A rc h ie ’s First Law which established a relationship between F and Ф:

F = а/Фт , 8.6

where:

a = the tortuosity factor, the ratio between the length of flow along a straight path and the

actual length of flow through a permeable medium. It is a measure of the complexity of the
passages that connect the pores of a permeable formation and varies from 0.62 for

unconsolidated sands, to 0.81 for consolidated sands to 1.0 for carbonates.

m = cementation factor, the value of which varies from 1.4 to 2.8. It is a function of grain size,
grain size distribution and tortuosity. In quantitative interpretation a value of 2 is usually
assumed for m.

Ф = porosity (fraction).

Table 8.2 lists in further detail variations in the values of a and m, and Fig 8.32 illustrates F-Ф
relationships for various values of a and m.

Table 8.2 Variations in the values of m and n used to calculate F


(After Asquith & Krygowsky, 2004)

F = а/Фт General relationship

a: T o rto u sitv m : C e m e n ta tio n C o m m e n ts


fa c to r e x p o n en t

1.0 2.0 C arbonates1

0.81 2.0 Consolidated sandstones *


On
tJ

2.15 U nconsolidated sands (H um ble form ula)1


P

1.45 1.54 Average sands (alter C arothers, 1968)

1.65 1.33 Shaly sands (after C arothers, 1968)

1.45 1.70 Calcareous sands (after C arothers, 1968)

0.85 2. 14 C arbonates (after C arothers, 1968)

2.45 1.08 Pliocene sands, southern C alifornia (after C arothers and Porter. 1970)

1.97 1.29 M iocene sands. T exas-L ouisiana G ulf Coast (after C arothers and
Porter, 1970)

1.0 g('2.05-6> Clean granular form ations (after Sethi. 1^79)

RESISTIVITY INDEX

Since hydrocarbons are non electrical conductors, their presence results in an increase in the
overall resistivity of the host rock. For a reservoir containing hydrocarbons and formation water,
Archie proposed a second factor called the resistivity index, lR, given by:

lR - R t/ R 0 8.7

Substituting FRWfor R0(equation 8.5) in 8.7,

lR = Rt/FRW 8.8
Formation Resistivity Factor Versus Porosity

Fa, formation resistivity factor

Fig 8.32 F- 0 relationships for various values of a and m (After Schlumberger)

Replacing F with а/Фт (equation 8.6),

lR = OmR,/aR„ 8.9

Archie then showed empirically that

lR = 1/Sw" 8.10

Equation 8.10 is known as Archie’s Second Law

Combining equations 8.9 and 8.10:


where n = the saturation exponent, the value of which usually varies from 1.8 to 2.5, but it is
commonly assumed to be 2. Values of n equalling 1.1 and 3.2 have been measured in
exceptional cases, the value of 1.1 being associated with shaly sand and 3.2 in a sand with
micropores. Solving equation 8.11 for Sw:

Sw" = (aRw)/(OmRt) Q 8.12


t
Assuming m = n = 2, equation 8.12 becomes: '

Sw = [(aRw)l(<t>2R,)]1/2 8.13,

This is the Archie Saturation Equation.

Since F = а/Фт , equation 8.13 can be re-written as

Sw = [(FRW)/Rt]1/2 8.14

The Archie equation is based on following assumptions:

- Clean formation (Vsh < 20%)


- S w > 15%
- Formation water salinity > 20,000 ppm
- Intergranular Ф
- Unimodal pore-throat size distribution
- Water wet system
- No conductive minerals - e.g pyrite

In equation 8.13 a value can be assumed for a once the lithology of the reservoir is known, Ф
can be derived from the Density-Neutron crossplot, and Rt is given by either the R|L(j or RLLd. The
only unknown is Rw, which must be determined before the Archie equation can be solved to
provide a value for Sw.

Determination of Rw is thus essential.

DETERMINATION OF Rw

In mature oil and gas producing regions, detailed formation water analyses are available and Rw
values can be found in water catalogues. In exploration wells, Rw can be determined by
obtaining a sample of the formation water through extended drill stem tests (DST) and
measuring its resistivity. However, this may not be possible in practice, as formation water
samples are invariably contaminated by mud filtrate. In the absence of direct information on Rw,
it needs to be calculated, and there are three methods available for this purpose:

(a) The SP method


(b) The Archie method
(c) The ratio method

(a) The SP Method

As discussed earlier, provided Rmf * Rw, the deflections of the SP curve may be used to derive a
value for Rw:

SSP = -K log (Rmfe/Rwe). 8.15

where the SSP represents the maximum difference in millivolts between the shale baseline and
the deflection associated with the zone of interest, and К = 61+0.133T (°F) or 65+0.24T (°C).

All the information necessary for a quantitative interpretation is recorded on the log heading, an
example of which is shown in Fig 8.33. This information includes bit size, type of drilling fluid,
Rm, Rmf, and Rmc at surface and bottom hole temperature (BHT) and the maximum recorded
temperature.

Presented below is an example of the derivation of Rw from the electric log suite shown in Figure
8.34. The steps involved may be summarised as follows:

(1) Pick a thick, clean, permeable formation and establish the sand and shale lines as
demonstrated in the section describing the SP log. In Fig 8.34 such a formation is the
interval 5,870 - 5,970ft. From the log, read the difference in millivolts between the shale and
the sand lines. In this case the SSP is read as -98mv, the scale being 20mv per division in
track 1.
COMPANY 1
WELL i
Г lELDt
COUNTYl STATES*
MA ТION *
LATITUDEl LONGI7UDE*
LOCATIONI

SEC: RGE*

PERMANENT DATUM* M. S. L
ELEVATION QF PERMANENT DATUM* .0 METE
LOG MEASURED FRCm R, T
DRILLING MEASURED FROMi R. T
EL EVa TIOM-
K.B.i гз.о METE D.F.i 23.0 METE G.L.i -191.0 METE

о т и е я SERVJCES-
l S r - 3 H C- GR FDC-CNL-GR HDT
UST CST RFT

OLD HEPC

D4T£ i
PROGRAM TftPE VERS I DM N0. » 19.2
RUM n u mb e r ; 3
SERVICE ORDER NUMBER*
API SERIAL MUnSERi
LOGGING UMIT NUN BER * 69 16
LOGGING UMIT LOCATION* AB T
LOGGING COMPANY CODE* 440
ENGINEER' S NAME*
UITNE S S * NANE *

TOTAL DEPTH -
DR ILLER* 3232.0 METE LOGGER* 3232.9 METE
BOTTOM LOG INTERVAL * 3233 .,0 METE
TOP LOG INTERVAL* 3oso. METE

CASING SIZE ucIGMT BOTTOM BOTTOM


DR ILLER LOGGER
19.62 IK 9 7 . 30 L 8/F 313.0 METE 9 13.6 METE
13.36 IN 6 9 . 00 L 8/T 1764.0 METE 1766.3 METE
Э. 6 г 5 IN <7.00 LB/ F 30 30 . 0 METE 3 0 So . 0 ' me t e
BIT SIZE DEPTH
17.30 IN 1777.0 ЛЕТЕ
12.23 IN 3062.0 METE
3.300 IN 3232.Q METE

DRILLING ^LUID TYPE» KCL POLYMER


DENSITY! 1 . 620 C. CC vI SCOS I TYi 44.00 SEC
Pи j 10,00 FLU ID LOSSi 4.500 ML

SOURCE MEASURED TEMPERATURE 3QTTQM MOLE


RES I ST IV ITY RESI ST IVTY
MUD SAMPLE FLOU LINE a . i г 9 OHMS 70 .. 0 DF 0.04 OHMS
MUD riLTRATE SAMPLE PRESS 0 ., 12 6 OHMS 64 ,. 0 Df 0.03 OHMS
n u 3 CAKE SAMPLE PRESS 0 . 36 7 OHMS 62. . 0 DF cnrts 0.16
QЧ I ,ч ип RECORDED TEMPER AT URE * г I 3.0 DF
: : ^ E С IR CUL AT I ON STOPPED» 7.13
' I ME LOGGER AT BOTTOM: г г . зо

JEnfiBr,;-
_;n mqlc n:rL pod

C-UiP-CNT MUM3CRS-
0RG2 SGTE ORG4
Э97

Fig 8.33 Typical log header


?■/

ч '

\ 4..
Й

Information from the log header: R mf = 0.71 at 68° F, BHT = 196° F at 9,400ft

Fig 8.34 Rw derivation by the SP method (Modified from Asquith & Krygowski, 2004)

(2) Read the Rmf value from the log header and convert it to the temperature of the zone of
interest. If the zone is at or within 500ft (150m)of the bottom of the hole, then the estimated
formation temperature (EFT) and BHT will be the same for all practical purposes. If the
vertical distance between the zone of interest and the bottom of the hole is large, then EFT
and BHT will not be the same, and the former is derived by interpolation between surface
temperature (ST) and BHT.

(3) The chart presented in Fig 8.35 may be used for this purpose. It has provisions for annual
mean surface temperatures of 60 and 80°F (16 and 27°C). Rmf can be converted to EFT by
using the chart presented in Fig 8.36.

(4) Find Rmfe - follow the recommendations along the top of Chart SP-1 shown in Fig 8.37, and
use Chart SP-2 (Fig 8.38) where necessary.

For the example of Fig 8.31, BHT = 196°F at the total depth (TD) of 9,400ft and Rmf = 0.71
Temperature gradient conversions: 1°F/1QQ ft ~ 1.823°C/100 m
1°C/100 m = 0.5486°F/i00 ft
Annual mean
surface temperature
Temperature (°C)

Fig 8.35 Estimation of formation temperature (EFT), assuming a linear thermal


Gradient (M odified from Schlumberger)

Assuming an annual mean surface temperature of 60°F, EFT at 5,920ft (middle of zone in
Fig 8.34) is estimated at 148°F.
7 L-

Rmf at 148°F = 0.31 ohm-m (Fig 8.36)


Rmfe = 0.31 x 0.85 = 0.26 ohm-m.

(5) Proceed to Figure 8.37. Mark the value of Rmfe on the Rmfe stem. Mark the SSP value on the
appropriate axis and draw a vertical line from that point to the formation temperature, then
project a horizontal line to obtain an intercept on the R mfe/ R We axis. From this point, project a
Conversion approxim ated by Rz = R, [(T,+ 6.77)/(T2+ 6.77)]=For R2 = R, [(T, + 21.5)/(T2 + 2 1.5)]X

Fig 8.36 Resistivity conversion chart (Modified from Schlumberger)

line through the value of R mfe on the R mfe stem and continue to the R we axis and read a
value for R we. Enter the value of R we on the vertical axis of Fig 8.38, project horizontally to
the formation temperature, and then project vertically to the horizontal scale to read Rw.
(Note: if the well has been drilled with KCI mud, use the chart shown in Fig 8.39 to obtain a
value for R mfe/ R w e ) .

From this procedure, R We = 0.013 ohm.m, and Rw = 0.028 ohm.m (Chart SP2).
Kweq Determination trom hssp
SP-1
Clean formations

I b i s chart and nom ograph calculate th e equivalent form a­ Example S S P = 100 mV a t 250°F
tion w ater resistivity. R trom the static spontaneous R n,i - 0 .7 0 oh m -m at 100CF
potential. Esst». m easurem ent in clean form ations. or 0.33 ohm -m at 250°F
E nter the nom ograph \friih Еч.чи in mV. turning through
T herefore. R r,(M = 0 .3 5 x 0 .3 3
the reservoir tem perature in F nr "C to define the
= 0 28 ohm -m at 2501F
R ntn^R ^cq nitio. From this value. pa&s through the
value to define R >v«,. К = 0.025 ohtn-m at 250r F
For predom inantly NTaCI m uds, determ ine R as H >SP = - K c lOgf R nrfe q/R )
*ец (ohm-m)
follow?: Кг = 61 +0.133 T ’F . 0.001
j. If Rmf at 7v’ F i'2 4 'C i is greater than 0.1 ohm -m . К, - - 6 5 + 0.24 T r
correct Rmi to form ation tem perature using C han
G en-l>. and u.se К.™ - = 0 85 R mj-
b. If R,.,r at 7 5 'F ( ’ I T ) is less than 0.1 ohm -m , u>e
C hart S P -2 to derive a value ol R mic4 at formation
tem perature.

Rwe- 0.013

8.37 Rwe determination by the SP method (Modified from Schlumberger)

(b) The Archie Method

Rw may be calculated from the Archie equation by considering a clean, water bearing zone:

Sw2 = (aRw)/(‘I>2R,), therefore Rw = (Sw2 <I»2R,)/a 8.15

In a water bearing zone Sw = 1, and the above equation becomes:

Rw = (<I>‘ R,)/a л . " j I ( r ' ( 1 8.16


versus Rweq and Formation Temperature
( » Iе- w ■'

7 !<Xo

Rw or Rm, (ohm-m)
Rw = 0.028

Fig 8.38 Rw - RWe- formation temperature relationships (Modified from Schlumberger)

Deriving Ф from a Density-Neutron crossplot, reading R, from the ILd or the LLd andknowing a,
Rw may be calculated.

(c) The Ratio Method

In a clean, water bearing zone the following relationships can be obtained from the Archie
equation for the uninvaded and the flushed zones respectively:

Rt = (aRw)/( Ф2 Sw2), and

Rxo = (aRmf)/( Ф ^ о 2) 11.17

•4
SP ( M i l l i v o l t s )

Fig 8.39 Rwe determination by the SP method in the presence of a KCI mud
Note that when Rmf = Rw, SP * 0 (After Hilchie, 1982)

In such a zone Sw - S*0 - 1, and Rw, Ф and a will have the same value. Dividing Rxo by Rt:

Rxo/R, = Rmf/Rw, therefore Rw = Rt Rmf/Rxo 8.18

R, and Rxo are obtained from resistivity logs, and Rmf at formation temperature is known (see
above).

There may be variations in the Rw values determined by the above methods. If this is the case,
then the value obtained by the Archie method should be used, since it will always give a value
of 1.00 for Sw when applied to the water zone. Once determined, Rw is assumed to be constant
throughout the formation being evaluated, unless local knowledge or clear SP log amplitude
changes indicate otherwise.

D E T E R M IN A TIO N OF Sw

In clean formations, Sw may be determined by the direct application of the Archie equation
(equations 8.13 and 8.14). It should be noted that in calculations, the d ecim al value o f Ф must
be used.

D ETE R M IN A TIO N OF S„0

The Archie equation applies also to the flushed zone. However, in this case Rw is replaced by
Rmf (since the flushed zone is assumed to be saturated with mud filtrate), and Rt by Rxo (flushed
zone resistivity):

Sxo = [(aRmf)/(<t>2R*o)]1/2 8.19

Or
7 Ss •
Sxo = [ ( F R m f) / ( R x o ) ]1/2 ' f ~ 8 .2 0

It is important to know S*0 since it gives an indication of hydro carb o n m oveability. Only the
moveable fraction of the hydrocarbons in a reservoir is producible, and can therefore be
regarded as reflecting the recoverable reserves. This fraction, together with all the original
formation water, is displaced from the flushed zone by the invasion of the drilling fluid. Following
invasion, the fraction of the pore volume occupied by moveable hydrocarbons is filled with mud
filtrate. Consequently, water saturation in a flushed zone that originally contained moveable
hydrocarbons increases, i.e. Sxo > S„. The difference between Sx0 and Sw (obtained from the
uninvaded part of the formation) represents the moveable hydrocarbon saturation in a reservoir.
From this it is possible to find the quantity of hydrocarbons displaced by invasion:

Sxo - Sw = moveable oil saturation (MOS), 8.21

and
The ratio S J S m is q u a lita tive index of hydrocarbon moveability. If S J S m = 1, i.e. Sw = Sxo, no
hydrocarbons have been moved by invasion, implying that any hydrocarbons present are largely
residual. Moveable hydrocarbons are indicated by Sw/Sxo < 0.7 in clastic reservoirs and <; 0.6 in
carbonates.
9. S H A L Y FO R M A TIO N IN T E R P R E TA T IO N

IN TR O D U C TIO N

The calculation of Sw in shaly formations is one of the most troublesome aspects of log
interpretation. Due to its conductivity, the presence of shale can lower Rt and sometimes mask
the hydrocarbon effect. This can be a significant factor in marginal cases where Sw tends to be
on the high side.

Shale normally occurs in sandstone reservoirs. C onsequently, sh aly form ation analysis is
invariably associated with the evaluation of sandstone rather than carbonate reservoirs. In
addition to suppressing Rt, the presence of shale has a detrimental effect on the reservoir
properties (porosity and permeability) of sandstones. In general, its presence results in the
reduction of porosity and permeability, thereby decreasing the productive capacity of the
reservoir. Shales containing swelling clays are particularly troublesome in improved oil recovery
(IOR) projects. Their presence requires the treatment of the reservoir with special, and therefore
more expensive, fluids when applying IOR techniques.

From a petrophysical point of view, the occurrence of shaly material in sandstone reservoirs falls
into three categories or in combinations thereof: (a) lam inated; (b) dispersed; and (c)
structural. Fig 9.1 illustrates the forms of shale occurrence in sandstones.
(a) Laminated shale

In this form, thin shale laminations - fractions of an inch to several inches in thickness - are
interspersed with clean sand. The effective porosity and the permeability of the shale are
essentially zero so that the overall porosity and permeability of the formation is reduced in
proportion to the fractional volume of shale. For example, 40% shale will theoretically reduce
effective porosity and permeability to 60% of the clean sand values. Consequently, 30-40%
laminated shale is the maximum amount normally tolerable for production.

(b) Dispersed shale

In this form, shaly material is disseminated in the pore space of the sand. It replaces pore fluid.
This type of distribution is very damaging because a relatively small amount of shale can choke
pores and reduce effective porosity and particularly permeability to non-producible values.
Maximum tolerable clay content is approximately 40% of the sand pore space or about 15% by
volume.

(c) Structural shale

In this form, shale grains, which may be aggregates of shale particles or mudstone clasts, take
the place of sand grains. Porosity and permeability of the sand is affected very little.
Consequently, this type of shale is least objectionable, but it does not occur frequently.

EFFECT OF SHALE ON THE RESPONSE OF LOGGING TOOLS

Measurements of the formation parameters provided by most logging tools are influenced by the
presence of shale in the intervals examined. Table 9.1 summarises the effects of shale on the
responses of the various tools.

The use of uncorrected shale-affected log readings in quantitative interpretation leads to


erroneous results. While the Archie equation gives acceptable Sw values provided Vsh in the
zone of interest is below about 20%, Vsh quantities greater than 20% affect all the parameters in
the equation, particularly Ф and Rt. Uncorrected log data in a shaly reservoir result in Ф values
that are too high and Rt readings that are too low. Consequently, Sw values based on these data
will be too high, and, as mentioned above, may result in the downgrading of the reservoir
concerned. Correcting log readings for the effects of shale requires knowledge of Vsh and some
of the methods of evaluating this parameter were reviewed in Part 8. In quick-look well site
interpretation, however, the GR method, the Neutron-Density or the Sonic-Density crossplots are
preferred. Furthermore, the use of the Neutron-Density crossplot permits Ф to be corrected for
the effect of shale at the same time (Fig 8.26).

T ab le 9.1 Effect of shale on the responses of various logging tools


(After Asquith & Krygowsky, 2004)

M e a su r e m e n t E ffe c t

S p o n ta n e o u s S P is d e c r e a s e d in m a g n itu d e w ith re s p e c t to the s h a le b a se lin e .


p o te n tia l. S P

G a m m a ray In c re a s e d ra d io a c tiv ity is a p p a re n t as less m o v e m e n t a w a y from th e n e a rb y sh a le v a lu e s th a n an


e q u iv a le n t c le a n san d .

S o n ic S o m e p o ro s ity is h ig h e r th a n the a c tu a l fo rm a tio n p o ro s ity d u e to the h ig h e r tra v e l tim e o f the


c la y s /s h a le s .

N e u tro n N e u tro n p o ro s ity is h ig h e r th a n th e a c tu a l fo rm a tio n p o ro s ity d u e to th e w a te r th a t is p a rt o f th e c la y


s tru c tu re a n d is a d s o rb e d o n the c la y su rfa c e s .

D e n sity D e n sity p o ro s ity is h ig h e r th a n the a c tu a l fo rm a tio n p o ro s ity d u e to the g e n e ra lly lo w e r m a trix


d e n s itie s o f m o st cla y m in e ra ls. If the m a trix d e n sity o f th e c la y is c lo se to th a t o f the fo rm a tio n
m a trix , th e re is little o r no e ffe c t o n p o ro sity .

R e s is tiv ity R e s is tiv ity is le ss th a n th a t in an e q u iv a le n t c le a n fo rm a tio n d u e to the c o n d u c tiv ity o f th e clay . T h is


p ro d u c e s a c a lc u la te d w a te r s a tu ra tio n g re a te r th a n th e a c tu a l fo rm a tio n w a te r s a tu ra tio n . ( A r c h ie ’s
e q u a tio n a s s u m e s th at all c o n d u c tiv ity is fro m the fo rm a tio n w a te r an d th a t the fo rm a tio n m a trix is
c o m p le te ly n o n c o n d u c tin g .)

D E T E R M IN A TIO N OF Sw IN SH A L Y R E S E R V O IR S

The literature on the subject of shaly reservoir log analysis is voluminous and over the years
many relationships for the estimation fluid saturations in such formations have been proposed.
They all contain a clean term, described by the Archie equation, and a shale term which in some
relationships is complex and requires knowledge of the various properties of the shale, e.g. its
cation exchange capacity (Qv). The relationships revert to the Archie equation when Vsh = 0.

Broadly, the proposed saturation equations fall into two categories: Sw can be expressed either
in terms of resistivity or conductivity. Although resistivity measurements are used far more
commonly, it is generally accepted that conductivity models based on total porosity, Фт, which
includes fluid filled or effective porosity and that associated with bound water, produce more
consistent results. However, the conductivity relationships have the following inherent problems:

- Фт cannot be quantified without calibration with core analysis data which may take weeks, if
not months, to generate.

- Cores are not usually taken in shale sections and Фт values derived from low volume clays
in the reservoir may not be representative of those in the adjacent shales.

- Different laboratory methods produce different Фт for the same core.

- The laboratory measurements of Qv and the correlation between these and log data is not
always consistent and can induce errors.

In general, conductivity saturation equations are preferred where sufficient core data are
available, while resitivity models are applied in wildcat wells or where there are insufficient core
analysis results to determine Фт accurately.

A detailed review of the various methods of calculating Sw in shaly formations is beyond the
scope of this course. Instead, two of the most commonly used relationships are presented and
discussed. These are the Indonesia and modified Sim andoux equations, the main advantage
of which is that their application does not require Qv data.

Indonesia Equation

This is an empirical formula proposed by Poupon and Leveaux (1971) and considers the
relationship between R, and Sw in terms of the conductivities of the shale in the reservoir,
formation water and any interaction between the two:

(1-Vsh)/2 m/2
1 /л/R t = [(Vsh N R sh + Ф N aRw] Sw 9.1

In terms of conductivity:

n/2 2-Vsh n/2


V C t = V (C w / F e) S we + V(V<=h C sh) S we 9.2

Ct= Total formation conductivity (1/Rt)


Cw = Conductivity of formation water (1/Rw)
Csh = Conductivity of shale
Fe= ‘Clean’ formation factor
RSh= Resistivity of shale
Swe = Effective (moveable) hydrocarbon saturation
The resistivity equation can be adapted to calculate Sw:

(1-Vsh)/2 m/2 2 -1/n


S w = {[(V sh /VRsh + Ф /л/aRw] Rt} 9.3

The chart shown in Fig 9 .2 solves V Sh 1Vsh/2 graphically. Up to V ci values of 0.3, the slope of the
curve is about 45°, which means that V sh = V sh 1Vsh/z. When V sh = 0, the Indonesia relationship
reverts to the Archie equation

V c la y

Fig 9 .2 Graphical derivation of V sh 1Vsh/2 (After Schlumberger)

Modified Sim andoux Equation

This relationship was proposed by Simandoux (1 9 6 3 ) and is based on resistivity, density and
neutron log data. Like the Indonesia model, the Simandoux equation can be expressed in terms
of both resistivity and conductivity:

Sw = 0.4R„/<I> 2['v(5 Ф 2)/(Rw R,) + (Vsh/R sh) - Vsh/R Sh] 9 .4

C t — (C w S w )/F e + V shC 3h 9.5


10. IN TR O D U C TIO N TO D IP M E TE R A ND FO R M A TIO N IM A G E LO G S

TH E D IPM ETE R LOG

Dipmeter technology was developed in the 1930s and the logs have been available
commercially since the 1950s. Although it is considered by some to be obsolete and replaced by
the current image logs, dipmeter data complement the image logs and are often used in
conjunction with them. The processing and interpretation of dipmeter data are computer based
and various softwares are available for this purpose.

The log provides a continuous measurement of the angle of inclination and the dip direction
(azimuth) of the strata penetrated in the well. Modern dipmeter tools consist of four or six
electrodes mounted on extending arms and record microresistivity. Fig 10.1 shows a four arm
dipmeter device which measures microresistivity at four positions 90° apart around the borehole.
The differences in depth between the curves across the borehole are compared, enabling the
computer to determine dip and azimuth. The principle of measurement and the computation of
dip and azimuth are illustrated in Figs 10.2 and 10.3.

Fig 10.2 Resistivity curves recorded by a four arm dipmeter tool (After Rider, 1996)

tan = —
12
пп m it i m гл
PAD N К S \V N
1

Plane orientation: 27/270

Fig 10.3 Dip computation from dipmeter data (After Schlumberger)

Determinations of the angle and direction of dip are usually presented as a plot of ‘tadpoles’
versus depth as shown in Fig 10.4. The head of the tadpole indicates the dip angle and its tail
points to the azimuth. The dip direction can also be displayed as a histogram or fan plot, referred
to as the Azimuth Frequency Diagram (AFD). AFDs display the dominant dip direction for a
selected depth interval. Fig 10.5 presents an example of an AFD.

10°
1800

Well
deviation

Fig 10.4 Dipmeter log presentation. Standard presentation is a recording of tadpoles


versus depth (After Schlumberger)
The various dipmeter logging tools are listed below:

A b breviatio n N um ber of
C om pany N am e of tool
pads
High Resolution Dipmeter Tool HDT 4
Schlumberger Stratigraphic High Resolution Dipmeter SHDT 4
Oil Based Dipmeter Tool OBDT 4
Diplog Diplog 4
Baker Hughes Hexdip Log HDIP 6
High Resolution Dipmeter Tool HEDT 4
Malll DUriOn 6
Six Arm Dipmeter SED

IN TER PR E TA TIO N

Dipmeter data have applications in structural and sedimentary geology. In structural geology the
data are used in the interpretation of unconformities, folds and faults, while in sedimentary
geology the dipmeter can assist in facies analysis and palaeocurrent direction studies. It must be
emphasised, however, that dipmeter data should never be interpreted alone and need to be
integrated with other standard logs.

Four patterns are recognised in dipmeter log motifs: green, red, blue and random. These are
illustrated in Fig 10.6.
Green p a tte rn is characterised by constant dip and azimuth with depth and usually represents
post depositional structural dip. It is typical in shale sequences.

R ed p a tte rn represents uniform azimuth but the dip angle decreases upward. This pattern is
usually associated with folds, faults, unconformities, reefs, channel sands and valley fills (Figs
10.7-10.12).

Fig 10.7 Dipmeter motif associated with a tilted, asymmetrical anticline


(After Western Atlas, 1987)
Fig 10.9 Dipmeter motif associated with a thrust (After Western Atlas, 1987)
B lue p a tte rn exhibits roughly uniform azimuth but an upward increase in dip angle. This pattern
is usually associated with folds or faults, unconformities, palaeocurret directions and submarine
fan deposits (Figs 10.13 and 10.14).
SP TRUE D IP П
0 Ю 20 30 Л0

Fig 10.13 Dipmeter blue patterns showing sediment transport direction (After Western Atlas, 1987)
R a n do m p a tte rn shows scattered dip and azimuth. Massive beds lacking coherent bedding
planes or slumped deposits may be responsible for such a response. It may also result from tool
malfunction or poor borehole conditions causing the pads to lose contact with the borehole wall.

FO R M A T IO N IM A G E LOGS

Image logs were introduced in the mid-1980s and have undergone rapid development in terms
of both data acquisition technology and image production. The tools do not, however, produce
pictures or photographs of the borehole wall; they provide a computer generated image based
on the measurement of the resistivity or the acoustic reflectivity of the formations surrounding
the borehole. The measurements are limited to the borehole wall and do not penetrate the
formations.

Both wireline and LWD imaging tools are available. Wireline dvices produce high resolution
microresistivity, micro-induction and acoustic images, suitable for detailed structural and
sedimentological analyses. LWD tools generate microresistivity and density images that can be
transmitted to the surface via mud pulses in real time or from data stored in memory and
downloaded when the device is recovered from the borehole.

Imaging tools fall into electrical and acoustic categories, which are briefly described below.

E L E C T R IC A L IM A G E TO O LS

These are based on dipmeter technology and have pad mounted microresistivity electrodes
which are pressed against the borehole wall. They provide partial coverage of the borehole wall,
operate in conductive water based mud only and have a vertical sampling interval of 0.25cm.
The various electrical imaging tools are listed below:

Company Name of Tool Abbreviation Number of pads


Schlumberger Formation Micro Scanner FMS 4
Schlumberger Formation Micro Imager FMI 8

Baker Hughes SimulTaneous Acoustic and STAR 6


Resistivity Imager
Halliburton Electrical Micro Imaging EMI 6
Fig 10.15 presents the FMI and STAR imaging tools.

(After Schlumberger) (After Baker Hughes)

Fig 10.15. Electrical imaging tools

A C O U STIC (U L T R A S O N IC ) IM A G E TO O LS

Also known as borehole televiewers (BHTV), these tools provide full coverage of the borehole
wall, operate in any fluid including oil based mud and are run centred in the well. A rotating
transducer emits and records signals and the vertical sampling interval depends on the logging
speed. Generally, it is between 180 and 250 samples per revolution which is between 6 and 12.
The sound waves emitted by the transducer are reflected by the borehole wall and the transit
time of the first echo and the first echo amplitude are recorded and processed into images. The
operating and measurement principles of the device are illustrated in Figs 10.16 and 10 17.

Fig 10.16 Schematic representation of the operating principle of the acoustic imaging tool
(F rom A s q u ith a n d K rygo w ski, 2004)
Focused \
Transducer

W all

Pulse Transit T im e Echo


1
UBI sign, i
Л / First echo amplitude

UBI measurements:
• Transit time of first echo: distance - speed in mud x Transit time / 2
=> Transit Time image (borehole radii)
• First echo amplitude => amplitude image
Fig 10.17 Principle of measurement of the acoustic imaging tool (After Schlumberger)

The various acoustic imaging tools are listed below

Company Name of tool A bbreviation


Schlumberger Ultrasonic Borehole Imager UBI
Baker Hughes Circumferential Borehole Image Log CBIL
Baker Hughes SimulTaneous Acoustic and Resistivity Imager STAR
Halliburton Circumferential Acoustic Scanning tool CAST

IMAGE INTERPRETATION

The standard presentation of image logs is to split or unwrap the borehole along true north and
unroll the cylinder into a flat strip. Real horizontal and vertical planes will appear as horizontal
and vertical features on the log. Dipping surfaces intersecting the well, however, will appear as a
sine wave, as demonstrated in Fig 10.18 The high point of the surface as it intersects the well is
represented by the crest of the sine curve and the steeper the dip, the greater the wave
amplitude.
Iri electrical image logs conductive, low resistivity features such shales and fluid filled fractures

After Rider, 1996

: f s W
N
\1 r\ f\ Л
\ j.! \1 \
I \
\ 1\ 1\
\\ ' \ i\ \\ 1 \ After Asquith and Krygowsky, 2004
i\
V'- -

Fig 10.18 Representation of dipping surfaces on image logs

are displayed as dark colours, while high resistivity features such as sandstones and carbonates
are shown as lighter shades of brown, yellow and white. In the case of the acoustic image logs,
low amplitude or high transit time features such as shales, borehole irregularities (washouts) and
fluid filled fractures are represented as dark colours. Sandstones and carbonates, charaterised
by high amplitude or low transit times, are displayed in lighter shades of brown, yellow and white
(Fig 10.19).

Low amplitude/resistivity High amplitude/resistivity

Fig 10.19 Image log colour codes

Image processing falls into two categories - static and dynamic. Static images are produced by
applying a single colour contrast setting to the entire well. In the case of dynamic images, a
variable colour contrast is applied in a ‘window’ that moves along the borehole. This improves
image quality and enhances views of features such as bed boundaries, fractures and vugs.

Image logs have a wide variety of applications. As with all log data, they should not be
interpreted in isolation and the best results are obtained by combining them with information
from other sources. The most common applications of image logs include the detection of faults
and fractures (Figs 10 20 and 10.21), borehole breakouts (Fig 10.22 and 10.23), stratigraphic
characteristics such as bedding (Fig 10.24) and identification of sedimentary features such as
cross bedding, slumping and texture (Figs 10.25-10.27).

Fig 10.20 Normal fault detection from FMI (After Schlumberger)


In-sttu Stress Directions
глгатиш
B o rer> o te b re a k o u ts
Рюгиотш
FA»xjfTx*n horizontal wua
Stre*»dkbCDcn drecfo n

tnducod
Iracfurr*

naturai

*
4r
Г'Ю ЯТН

Fig 10.22 Illustration of breakout (After Schlumberger)

Fig 10.23 A fractured sand and shale interval


from CBIL. Borehole breakouts (B)
appear as dark patches 180° apart
from each other
(After Asquith and Krygowsky, 2004)
■■■ ,•

3
*r

Fig 10.24 Regular bedding from LWD Startrack (After Baker Hughes)
SELECTED REFERENCES

Alger, R. P., 1980. Geological use of wireline logs: in Developments in Petroleum Geology - 2,
G.D. Hobson ed. Applied Science Publ., London, p. 207-272.

Allen, D., et a!., 1989. Logging While Drilling: Oilfield Review, April, 1989, p 4-17.

Archie, G.E., 1942. The Electrical Resistivity Log as an aid in determining some reservoir
characteristics: Petroleum Technology, v. 5, p. 54-62.

Asquith, G., 1991. Log Evaluation of Shaly Sandstone Reservoirs: A Practical Guide, A.A.P.G.
Course Note Series, No. 31, 59p.

Asquith, G and Gibson, C., 1982. Basic Well Log Analysis for geologists: A.A.P.G Publ., Tulsa,
216p.

Asquith, G. and Krygowski, D, 2004. Basic Well Log Analysis, 2nd ed: A.A.P.G. Publ., Tulsa,
244 p.

Bassiouni, Z , 1994. Theory, Measurement and interpretation of well logs: SPE Textbook Series,
Vol 4

Bateman, R.M., 1985. Log Quality Control: IHRDC Publication, 398p.

Betts, P. et a!., I 990. Acquiring and interpreting logs in Horizontal wells: Oilfield Review, July,
1990, p. 34-51.

Boyle, K,Jing, X.D. and Worthington, P.F., 2000. Petrophysics in Modern Petroleum Technology,
R.A. Dawe ed, John Wiley, p. 131-206.

Brown, E. and Milne, A., 1990. The Challenge of completing and stimulating horizontal wells:
Oilfield Review, July, 1990, p. 52-63.

Bussian, A .E .,1982. A generalised Archie equation: SPWLA 23rd Ann. Symp. Trans., paper E,
12p.

Bussian, A .E .,1983. A comparison of shaly sand models: SPWLA 24th Ann. Symp. Trans.,
paper E, 16p.

Coates, G.R., Xiao, L. and Prammer, M.G., 1999. NMR Logging Principles and Applications:
Houston, Texas, Haliburton Energy Services, 232p.

Desbrandes, R., 1985. Encyclopaedia of Well Logging: Graham & Trotman, London, 584p.

Dewan, J. Т.,1983. Modern Open Hole Log Interpretation: PennWell Publ. Co., Tulsa, Oklahoma,
361p.

Doveton, J.H.. 1986. Log Analysis of Subsurface Geology - concepts and computer methods:
John Wiley and Sons, 273p.
Dresser Atlas, 1982. Well Logging and Interpretation Techniques. The Course for Home Study,
2nd ed.

DresserAtlas, 1985. Log Inteipretation Charts: 157p.

Elliott, H.W., 1983. Some 'Pitfalls' In Log Interpretation: Log Analyst, v. 24, p. 10-24.

Ellis, D.V., 2007. Well Logging for Earth Scientists, 2nd ed: Springer, 692p.

England, R.E., 1975. Well Log Interpretation: Birdwell Div., Seismograph Service Corp., Tulsa,
Oklahoma.

Fertl, W.H. (ed.), 1976. Abnormal Formation Pressures - Implications to Exploration, Drilling and
Production of Oil and Gas Resources: Elsevier, Amsterdam.

Fertl, W.H., 1984. Advances in well logging interpretation: Oil and Gas Journal, April 16th, p.
85-91.

Hansen, R.R. and White, J., 1991. Features of Logging While Drilling: SPE/IADC Conf.,
Amsterdam, Paper 21989.

Hatton, I.R., Reeder, М., Newman, M. St J. and Roberts, D., 1992: Techniques and applications
of petrophysical correlation in submarine fan environments, early Tertiary sequence,
North Sea: in Geological Applications of Wireline Logs II, A. Hurst, C.M. Griffiths & P.F.
Worthington eds, Geol Soc Spec Pub No 65, p. 21-30.

Helander, D.P., 1983. Fundamentals of Formation Evaluation: Oil and Gas Consultants Int.
Publ., 332p.

Heslop, A., 1975. Porosity in shaly-sands: SPWLA, 16th Ann. Symp. Trans., Paper F. 12p.

Hilchie, D.W., 1982a. Applied openhole log Interpretation: D.W. Hilchie Inc., Golden, Colorado,
330p.

Hilchie, D.W., 1982b. Advanced Well Log Interpretation: D.W. Hilchie Inc., Golden, Colorado.

Jenkins, R.E., 1960. Accuracy of porosity determinations: First SPWLA Logging Symp., Tulsa,
Oklahoma.

Merkel, R.H., 1979. Well Log Formation Evaluation: A.A.P.G. Continuing Education Series, No.
14, 82p.

Meyer, B.L. and Nederlof, M.H., 1984. Identification of source rocks on wireline logs by
density/resistivity and sonic transit time/resisivtity crossplots: A.A.P.G. Bull., v. 68, p.
121-29.

Misk, A., Mowat, G., Goetz, J. and Vivet, B., 1977. Effects of hole conditions on log
measurements and formation evaluation: SAID 5th Ann. European Symp. Trans., Paris,
Comm. 22, 16p.

Patchett, J.G. and Coalson, E.B., 1979. The determination of porosity in sandstones and shaly
sandstones: Part 1 Quality control: SPWLA 20th Ann. Symp. Trans. Paper QQ. 17p.

Patchett, J.G. and Coalson, E.B., 1982. The determination of porosity in sandstones and shaly
sandstones: Part 2 Effects of complex mineralogy and hydrocarbons: SPWLA 23rd Ann.
Symp. Trans. Paper T. 50p.

Petroleum Exploration Society o f Great Britain, 2004. The Millenium Atlas: Petroleum geology of
the central and northern North Sea: Geol Soc Lond Pub.

Pirson, S J., 1983. Geologic well log analysis: 3rd ed., Guif Pub. Corp., Houston, 475p.

Poupon, A., Hoyle, W.R. and Schmidt, A.W., 1971. Log Analysis in Formations with Complex
Lithologies: Jour. Petrol. Tech., v. 23, p. 995-1005.

Poupon, A., Clavier, C., Dumanoir, J.L., Gaymard, R. and Misk, A., 1970. Log Analysis of
Sand-Shale Sequences - A Synthetic Approach: Jour. Petrol. Tech., v. 22, p. 867-881.

Quierein, J.A., Garden, J.S. and Watson, J.T., 1982. Combined natural gamma ray
spectral/litho-density measurements applied to complex lithologies: SPE 11143, 14p.

Ransom, R.C., 1977. Methods based on density and neutron well logging responses to
distinguish characteristics of shaly sandstone reservoir rock: Log Analyst, v. 18, p. 47-63.
Raymer, L.L., Hunt, E.R. and Gardner, G.H.F., 1980. Improved Sonic Transit Time-to-Porosity
Transform: SPWLA Logging Symp. Trans., July, 1980.

Rider, M.H., 1996. The Geological Interpretation of Well Logs: 2nd ed, Gulf Pub Corp, Houston,
280p.

Schlumberger, 1974a. Well Evaluation Conference, North Sea: 171 p.

Schlumberger,1974b. Well Evaluation Conference, Nigeria: 2nd. ed.

Schlumberger, 1975. Well Evaluation Conference, Arabia: 152p.

Schlumberger, 1976. Well Evaluation Conference, Iran: 179p.

Schlumberger, 1979. Well Evaluation Conference, Algeria.

Schlumberger, 1980. Evaluacion de Formaciones en Venezuela.

Schlumberger, 1991. Log Interpretation Principles/Applications.

Schlumberger, 1998. Log Interpretation Charts.

Schlumberger/Anadrill, 1991. LWD general information sheet: Rev 6, 26 March, 1991, 21 p.

Sengel, E.W., 1983. Basic Well Logging: The Institute for Energy Development (IED), Oklahoma
City, 91 p.

Serra, O., 1984. Fundamentals of Well Log Interpretation: 1, The Acquisition of Logging Data,
Elsevier, Amsterdam, 423p.
Serra, О., 1986. Fundamentals of Well Log Interpretation; 2. The Interpretation of Logging Data:
Elsevier, 684p.

Simandoux, P., 1963. Mesures dielectriques en milieu poreux, application a mesure saturation
en eau: Etude du Comportement des Massifs Argileux: Revue de I’instutu Francais du
Petrole, Supplementary Issue (Translated text in SPWLA Reprint Volume Shaly Sand,
July 1982).

Society o f Professional Well Log Analysts, 1975. Glossary of terms and expressions used in well
logging: SPWLA publ., Houston, Texas, 74p.

Thomas, D.H., 1977. Seismic applications of sonic logs: SPWLA 5th European Symp., Trans.,
Paris, Paper 7, 24p.

Threadgold, P., 1971. Some problems and uncertainties in log interpretation: SPWLA 12th Ann.
Symp. Trans., Paper W. 19p.

Welex, 1978. An Introduction to Well Log Analysis: Houston, Texas, 46p.

Wyllie, M.R.J., Gregory, A.R. and Gardner, G.H.F., 1956. Elastic Wave Velocities in
Heterogeneous and Porous Media: Geophysics, v. 21, No. 1, p. 41-70.

Wyllie, M.R.J., Gregory, A.R. and Gardner, G.H.F., 1958. An Experimental Investigation of
Factors Affecting Elastic Wave Velocities in Porous Media: Geophysics, v. 23, No.3, p.
459-93.

Worthington, P.E., 1985. The Evolution of Shaly-Sand Concepts in Reservoir Evaluation: The
Log Analyst, v. 26, No. 1, p. 23-40.

Anda mungkin juga menyukai