Está en la página 1de 8

TECHNICAL NOTE

Petrophysical Evaluation of Low


Resistivity Sandstone Reservoirs
G.M. HAMADA, M.N.J. AL-AWAD
King Saud University

Abstract
There are many reasons for low resistivity pay zones. It is of
crucial importance to know the origin of this phenomenon. The
problem with these zones is that the resistivity data interpretation indicates high water saturation, but oil or even dry oil will
be produced.
This paper discusses the different reasons sandstone reservoirs can have low resistivity. Clean bearing sandstone has high
resistivity, but when this rock contains shale, or heavy minerals
such as pyrite, the resistivity can become low. This resistivity
change depends on the nature, the volume of clay and heavy
minerals. Pyrite shows a good electrical conductivity, that is
usually comparable to or even higher than the conductivity of
formation water, and can therefore have a larger effect than
shale. Low resistivity pay zones are not necessarily shaly sand
problems.
In this study, different shaly sand models will be discussed, to
propose a suitable shale sand model. Field examples were
analysed using different shaly sand models.
It has been found that the modified total shale sand model
gives good results. In the case of low shale volumes, the normal
saturation equation might give acceptable water saturation
values.

Introduction
The reasons for low resistivity phenomena are classified mainly
into two groups. The first consists of reservoirs where the actual
water saturation can be high, but water free hydrocarbons are produced. The mechanism responsible for the high water saturation is
usually described as being caused by microporosity. The second
group consists of reservoirs where the calculated water saturation
is higher than the true water saturation. The mechanism responsible for the high water saturation is described as being caused by
the presence of conductive minerals such as clay minerals and
pyrite in a clean reservoir rock. The resistivity data must be corrected for the effect of these conductive minerals to reduce the
calculated water saturation to more reasonable levels associated
with water free hydrocarbon production.
Most formations logged for potential oil or gas production consist of rocks which without fluids would not conduct an electrical
current. There are two types of rock conductivity: a) Electrolytic
conductivity which is a property of water containing dissolved
salts and b) Electronic conductivity which is a property of solids

such as graphite and metal sulfides such as pyrite.


Pyrite is a common heavy mineral associated with marine sedimentary rocks. It has a good electrical conductivity that is usually
comparable to, or even higher than the conductivity of the formation water. The crystals of pyrite may form a continuous network
even at low pyrite concentrations. Measured resistivity on dry
pyrite ranges between 0.03 and 0.8 m. Pyrites conduction is of
a metallic (electronic) nature and consequently any transfer of
current between water and pyrites is based on conversion from
ionic to electronic conduction and vice versa. This leads to polarization at the water-pyrite interfaces with the corresponding frequency dependent electrical properties. Thus the electrical properties of porous rocks with pyrites are strongly dependent on the
amount and distribution of pyrite and the frequency of measuring
the electrical current. The main problem with minerals such as
pyrite is how to estimate their volume and the distribution from
well logs(4, 10).
The conductive properties of clean reservoir rocks are of electrolytic origin. The presence of brine alone or mixed with hydrocarbons in the pore spaces will allow the passage of the electrical
current. The water phase must of course be continuous in order to
contribute to the rock conductivity. Dry minerals, with the exceptions of Pyrite, Hematite, Graphite, and a few others, have
extremely high resistivities. Certain minerals appear to be solid
conductors, clay minerals are an example(18, 15). According to
Waxman and Smits(21), a clayey sediment behaves like a clean formation of the same porosity, pore pattern and fluid saturation,
except that the water appears to be more conductive than expected
from its bulk salinity.
This paper deals with the case of low resistivity pay zone
resulting from the presence of clay minerals and Pyrite. Field
examples will be provided to show how to deal with this problem.

The Effect of Shaliness on Rock


Resistivity
The effect of clays on the conductivity of Mastic sediments is
the subject of several petrophysical models as concisely reviewed
by Worthington(23). It is well known that the dispersive dielectric
behaviour of shaly sands is strongly controlled by the geometry
and surface area of the solid grains as well as by electrochemical
processes acting on the interfaces between clays and aqueous
electrolyte(17, 18, 16).
The resistivity of a shaly formation will depend on the shale
type, the shale volume and the mode of distribution in the rock.

This paper is being published as a technical note and has not been peer reviewed.
July 2000, Volume 39, No. 7

FIGURE 1: Shows the structure of the main clay minerals(18).

Berg(3) proposed that decreased resistivity caused by the dispersed


clays can be associated with the matrix instead of the bound
water. Inherent in this assumption is that formation water resistivity Rw, in the sand pore spaces is the same as the Rw of the bulk
of the water in the clay pore spaces. Bergs assumption might
seem unusual to some, but it is justified for several reasons. Clay,
although relatively impermeable to oil and gas, has permeability
with respect to water, and over geologic time the water in clay and
sand pore spaces should equilibrate. In addition the amount of
excess conductivity should be directly dependent on the amount
of clay cations, which, in turn, is directly dependent on the
amount of clay. Therefore anomalous resistivity caused by the
clay counterions can be treated as a surface effect associated with
the clay grains and the bound water can be treated as the far
water(4).

Clay Mineral Types


There are several clay minerals classified according to the
thickness of the platelets or the spacing of the crystals lattice.
Most clay minerals consist of platelets of alumina-octahedron and
silica-tetrahedron lattices as shown in Figure 1. There is usually
an excess of negative electrical charge within clay platelets. The
substitution of Al+++ Si++ and Al++ ions is the most common cause
of this excess.
Clay found in sedimentary series usually include several mineral types and may be mixed with silts and carbonates in varying
proportions. Their pore space will depend on the arrangement of
the particles and on the degree of compaction of the rock. These
pores generally contain water but it is quite possible to find solid
or liquid hydrocarbons or gases in them. Hence it is well known
that the log response to a clay depends on its composition, porosity, and hydrocarbon content.

Resistivity of Clays
Clays are composed of thin sheet like crystal structures, which
have a large surface depending on the clay mineral type as depicted in Figure 1. The observed deficiency of the electrical charge
8

within the clay sheet must be compensated to maintain the electrical neutrality of the clay structure. The compensating agents are
positive ions (cations or counter ions) which cling to the surface
of the clay sheets. The amount of these compensating ions constitutes the Cation Exchange Capacity which is commonly referred
to as the CEC (meq/100 g dry rock) or Qv (meq/cm3 total pore
volume). CEC is related to the specific surface area of a clay mineral. It has its lowest value in kaolinite and its highest values in
montmorillonite and vermiculite as shown in Figure 1. When the
clay particles are immersed in water the coulomb forces holding
these positive ions are reduced by the dielectric properties of
water. Part of the counter ions leave the clay surface and move
relatively free in a layer of water close to the crystal surface and
contribute to the conductivity of the rock. The thickness of this
conductive layer expands as the salinity of the solution decreases
and the conductivity of clay is proportional to the volume of the
counter ions exclusion layer.
The excess of conductivity observed within clays is due to
these additional cations held loosely captive in a diffuse layer surrounding the clay particle from which counterions are effectively
excluded virtually as shown in Figure 2.
The conductivity of clayey sediment is thus the sum of two
terms: a) one associated with free water or the water filled porosity, and b) the other associated with the CEC. This can be
expressed in another way: a clayey sediment has a conductivity
which depends on its effective porosity on one hand and on the
effective conductivity of the water it contains on the other(4, 18, 15).

Shale Models and Water Saturation


In recent years, a considerable number of shaly sand models
relating resistivity to water saturation have been proposed. For
details of algorithms for shale sand models see References(1), (5). (7),
(8), (12), (15), (20), (22), and (24). All these models are composed of two
terms, a clean sand term described generally by Archies equation
and a shale term. The shale term could be fairly simple or very
complicated. All models return to Archies formula when the
shale volume becomes zero. For low shale volume, it is not really
Journal of Canadian Petroleum Technology

a) View of outer
Helmholtz plane.

b) Diffuse layer ionic


concentration.

FIGURE 2: Schematic view of the bound water layer(4).

recommended to use these models unless very accurate water saturation values are required.

Mode of Shale Distribution


Log analysts customarily distinguish three modes of shale distribution, Figure 3. Each mode has a different effect on the resistivity, spontaneous potential, porosity logs, radioactivity, permeability, and water saturation.
Laminated Shales
These are thin beds or streaks of shale deposited between layers
of reservoir rock (sand, limestone, etc.), Figure 3(15). Shales do not
alter the effective porosity, saturation, or permeability of each
intermediate reservoir layer, provided they do not form lateral permeability barriers. They do, of course, impede vertical permeability between porous beds. Laminar shales can be considered to have
the same properties as neighboring thick shale beds, since they
have been subjected to the same conditions of compaction.
Electrically, laminar shales produce a system of conductive circuits in parallel with the more or less conductive porous beds.
Using this model of parallel circuits we can write the resistivity Rt
of a formation containing laminar shales as :
I / Rt = Vsh + ( I Vsh ) / Rsd

.................................................................(1)

where Rsh and Vsh are resistivity and volume of shales and Rsd is
resistivity of the clean sand.
Water saturation Sw has the following form for this case.

)]

Swn = (1 / Rt Vsh / Rsh ) Fsd Rw / (1 Vsh )

......................................(2)

where Rw is formation water resistivity and Fsd is the formation


factor for clean sand.
Dispersed Shales
Shales of this category adhere to the rock grains, either coating
them or partially filling the pore spaces, Figure 3. Shaly sands
with dispersed clays exhibit different properties from laminar
shales, being subjected to different constraints(11). Dispersed shale
in the pores markedly reduce the permeability, firstly because the
space available for fluid movement in the pores and channels is
now restricted and secondly because the wettability of the clay
with respect to water is generally higher than that of quartz. The
consequences are an increase in water saturation and a reduction
in fluid mobility.
Electrically, a dispersed clayey formation acts like an assembly
of conductors consisting of the pore fluid and dispersed clay. De
Witte(6) assumed that the water and the dispersed clay conduct an
electrical current like a mixture of electrolytes. Development of
this assumption yields the following saturation equation.
2 0.5

Sw = aRw / 2 Rt + q( Rshd Rw ) / 2 Rshd

FIGURE 3: Different modes of shale distribution.


July 2000, Volume 39, No. 7

FIGURE 4: Field example 1 shows the effect of shale on


radioactivity log.

q( Rshd + Rw ) / 2 Rshd / (1 q )

.........................................................(3)
Usually, t can be obtained from a sonic log since dispersed
clay in the rock pores is seen as water by the sonic measurement.
The value of q can be obtained from the relation q = s d/s,
where s and d are the sonic and density derived porosity.
In fact the value of the dispersed shale resistivity, R shd in
Equation (3) is difficult to evaluate. It is usually taken as equal to
Rsh in the adjacent shale layer. Fortunately, its value is not very
critical if it is several times greater than Rw. Therefore, when Rw
is small compared to Rshd and the sand is not too shaly, Equation
(3) can be simplified and being independent of Rshd to the following form.

Sw = aRw / ftRt + q 2 / 4

0.5

q / 2 / (1 q )

...................................(4)

Structural Shales
Structural shales are shale grains forming part of the solid
matrix along with quartz and other grains. They are considered to
have many characteristics in common with laminar shales and
nearby massive shales since they have been subjected to similar
diagnostic depositional constraints(18). However, their effects on
permeability and resistivity resemble more closely those of dispersed clays.
Laminar and structural shales are essentially of deposional origin, while dispersed clays evolve from alteration or precipitation.
All three shale types may be encountered in the same shaly
formation.

Shaly Sand Models


Total Shale Model
Simandoux(19) and Poupon et al.(14) have shown that in some
cases it is possible to use the following equation to calculate water
10

saturation, independently of the distribution of shale:

2
Sw = aRw / 2 Rt + aRw Vsh / 2 2 Rsh

aRw Vsh / 2 2 Rsh

0.5

) ...........................................................................(5)

where Vsh is for total shale volume and Rsh is taken as the resistivity of the adjacent shale bed.
The above equation has been widely accepted and applied in
many areas including Nigeria, Argentina, Egypt, the USA, and
Libya. One limitation is that the porosity exponent is taken as two
and the value of aRw has to be accurately identified. To overcome
this limitation, we introduce the both cementation exponent, m,
and saturation exponent, n, as variables and rewrite Equation (5).
The modified total shale equations will be found in the analysis of
field example 3.
Patchett and Rausch Model
The Patchett and Rausch(12) model can be written as:

Swn = aRw Rmf / R1 m Rmf Rw Ash

Ash = 1 10 SP / ( 60 + 0.1333T )

] ...................................................(6)

...................................................................(7)

where T is formation temperature in F and SP is spontaneous


potential. Equation (6) can be rewritten as:

Log Rt / Ash = m log + log aRmf Rw / Rmf Rw n log Sw

......(8)

Equation (8) indicates that cross plot of Rt/Ash vs. (on logJournal of Canadian Petroleum Technology

FIGURE 5: Log data for field example 2.

log paper) should result in a straight line with a slope equal to -m,
provided that a, Rw, Rmf, and Sw are constant. The intercept at
100% porosity equals to Rmf Rw/(Rmf - Rw).
Effective-Medium Resistivity Model
Berg(2) proposed that rock matrix and hydrocarbons can be
treated together as a dispersive phase in a continuous phase of
water. The matrix grains are definitely connected whereas the
hydrocarbons can be but are not necessarily connected. The connectiviness, however, probably does not matter because the more
resistive dispersed elements react passively to current flow in the
more conductive water. The relationship between those elements
can be treated as resistors in parallel, which mathematically is a
volumetric weighted average of conductivities. To derive the
water saturation equation it was necessary to associate the matrix
and hydrocarbon. The easiest part of setting up the equation is the
association of water saturation with porosity. Berg(3) has adapted
the Hanai-Bruggeman (HB) equation for rocks. Such adaptation
of HB equation results in the following expression.
Sw = ( Rw / Rt )

1/ m

( Rt Rd / Rw Rd )

........................................(9)

and

Rd = Rr / (1 ) (1 Sw )

..........................................................(10)

Where Sw is the whole rock saturation, Rt is the partially saturated whole rock resistivity, Rd is the dispersed phase resistivity
(matrix and hydrocarbon), and Rr is the matrix resistivity. We now
substitute Rd in Equation (9) by the expression of Equation (10),
to obtain the following water saturation equation
July 2000, Volume 39, No. 7

Sw = B B 2 + 4. H . Rm .( Rt Rm )

0.5

) /(2..R ) .........................(11)
m

where:
H
= (Rw/Rt)1/m
Rm = Rr/(1-) and
B
= Rw Rm (H+1)
There are only five input parameters (Rw, Rt, , m, and Rr)
needed for calculating water saturation Sw . Therefore, the calculations needed for obtaining the saturation will be as good as the
degree of certainty of m and Rr values, since Rw, Rt, and (author
are usually known).
In fact, the effective medium model uses the same input as the
dual water model. As in the dual water model, it may be possible
to refine the model by crossploting techniques or by improving
calculation of input variables.
The desirability of this theoretical model in general is that as
long as the original assumptions apply to the phenomena being
studied, they should accurately describe those phenomena.
Theoretical assumptions are usually stated explicitly, while the
restrictions imposed by empirical models may not readily be
apparent. The accuracy of an empirical model may decline when
conditions differ from the conditions under which the model was
derived.

Field Examples
Field Example 1
Figure 4 shows field example 1 which has previously been discussed by Hilchie(9). This example is presented to illustrate the
impact of shaliness on radioactive log characteristics and interpretation. In Figure 4a, the density and neutron read essentially the
11

TABLE 1: Resistivity, shale volume, porosity and water saturation for example 2.

Resistivity
(Rt)

Zone
A
B
C

6.7
2.9
0.19

Shale volume
(Vsh)

Porosity
()

0.14
0.34
0.05

0.23
0.19
0.26

same in the clean water sand zone A and separate in gas sand
zones B and C. This case is identified by the mirror imaging of the
density and neutron logs in a gas sand zone. When the producing
section suffers from dispersed shales, we may have the case of
Figure 4b. This figure shows gas sand that becomes progressively
more shaly with depth. The gas shows separation in the upper portion of the sand zone D where DST I shows gas production. In
zone E, at a particular shaliness, the density and neutron both indicate the same porosity although gas is still present. Below that
point (8,094 fl), the density and neutron do not exhibit the gas
cross over although the DST 2 shows that this section still contains gas. To determine water saturation in this shaly sand section,
we have to use one of the shale sand models.

Field Example 2
This is an example of a low resistivity pay zone from tertiary
sandstones, Figure 5. The well was drilled with oil base mud and
thus there is no short normal or SP. At the bottom of both neutron
and density a sandstone porosity scale has been put on to ease the
calculation.

Complete Shaly Sand Interpretation


In this approach we will use the model of Fertl and
Hammack(7).
Sw = ( FRw / Rt )

0.5

Vsh Rw / 0.4 Rsh

0.22
0.4
1.00

0.19
0.32
1.00

The calculated water saturation values are relatively close from


the two approaches for zone A. In fact, in the case of low shale
volume, the use of Archies formula is advisable before expending
time and effort on any shaly sand model. While in zone B which
is shaly sand there is a significant difference between the simplified and more complete shale sand model.

Field Example 3
This example is taken from a shaly sandstone formation with
low resistivity. This well has been discussed and evaluated previously(1) using one of the shale sand modelling techniques developed by Schlumberger. Figure 6 shows SP, resistivity, neutron,
and density logs. In this example the total shale model [Equation
(5)] was used after modification for the evaluation. This model
was modified to include a, n, and m constants. The Humble formula constants were selected which were a = 0.62 and m = 2.15.
Equation (5) was modified to include shale terms Atsh and Btsh. The
modified total shale model is deduced as:
Rt / Atsh = aRw m Sw n

Simplified Shaly Sand Interpretation


In this approach we use the average value of neutron density
log readings. Rt values in Archies formula. Formation water
resistivity Rw is calculated from the clean water sand where Rt =
0.19 .m (zone interval 10,290-10,330 ft.) and it equals to 0.017
.m. From induction log data Rt values are for zone A = 6.7 .m
and for zone B = 2.9 .m. Shaly corrected porosity for zone A
was 0.23 and for zone B was 0.19. Water saturation value was
then derived using Archies formula for the two zones A and B. It
is found that Sw = 22 % for zone A and Sw = 40 % for zone B,
Table 1.

Water saturation
(Sw)
Archie Equation
Equation (12)

......................................................................(13)

The total shale group Atsh is given by

2
2
Atsh = 1 + m Rt / aRw 2 Btsh
2 Btsh aRw / m Rt + Btsh

0.5

.......(14)

and
Btsh = aRw Vtsh / 2 m Rtsh

...................................................................(15)

For clean formation Btsh becomes zero and Atsh becomes unit.
Btsh increases with Vsh while Atsh decreases with Vsh. For shale
layer, Rt in Equation (14) will be equal to Rsh .
Taking the logarithm of both sides of Equation (13), we get the
following Equation.

.............................................(12)

where 0.4 Rsh is clay resistivity corresponds to the adjacent shale


layer and is the porosity corrected for shale volume.
For the same two zones A and B, shale volume Vsh was calculated, for zone A was 14 % and for zone B was 34%. Shale values
for the model were taken from the adjacent shale layer (10180
10200), Rsh = 0.91 .m, dsh = 0.08 and nsh = 0.33. Using Fertl
and Hammack model, the water saturation was calculated for the
two zones and equals to 19% for zone A and 32% for zone B,
Table 1.

Log Rt / Atsh = m Log + Log ( aRw ) nLog Sw

...........................(16)

Equation (16) indicates that a log-log plot of Rt/A tsh vs.


(should be straight line with slope of - m, providing that aRw and
Sw are constant. The resistivity index for a shaly formation is calculated using the following form.
Itsh = ( Rt / Atsh ) h / ( R o / Atsh ) w

.........................................................(17)

TABLE 2: Data for water saturation calculation in shaly sand formation for example 3.

12

Data
points

Resistivity
(Rt)

Porosity
()

Shale volume
(Vsh)

Btsh

Atsh

Rt/Atsh

1
2
3
4
5

8.40
7.50
8.40
1.80
1.75

0.27
0.24
0.26
0.28
0.31

0.13
0.18
0.15
0.00
0.00

0.072
0.123
0.094
0.00
0.00

0.449
0.378
0.412
1.00
1.00

18.67
19.87
20.35
1.80
1.75

Itsh
6.5
5.84
6.25
1.0
1.0

Journal of Canadian Petroleum Technology

TABLE 3: Water saturation values from different shale sand models for example 3.
Data points
1
2
3
4
5

Archie Equation
0.54
0.62
0.55
1.00
1.00

Schlumberger

Equation (18)

0.39
0.42
0.37
1.00
1.00

0.39
0.41
0.40
1.00
1.00

and water saturation Sw is derived from the relation


1/ n
Sw = Itsh
...........................................................................................(18)

From the log-log plot, it is also possible to calculate Sw with


the use of the equation.
Sw = ( w / h )

m/n

..............................................................................(19)

Where w is the porosity reading on 100% water saturation


line of the Picket(1973) crossplot and h is the porosity of the
hydrocarbon bearing interval. Table 2 shows the values of Rt, ,
and Vtsh as reported by Schlumberger. Also shown are values of
Atsh, Btsh, and Itsh values as calculated by Equations (14), (15), and
(17) respectively. The last column includes values of Rt/Atsh that
are cross plotted against on log-log coordinates, Figure 7. The
slope of the straight line through points 4 and 5 gives m = 2.15 the
intersection of the porosity axis at 100% porosity gives a = 0.62.
This validates the assumed values. If the straight line did not go
through points 4 and 5, new values of a and m could be assumed
for a new try. Water saturation was calculated using Equations
(18) and (19). In this calculation we have used Rw = 0.17 .m, Ro
= 1.29 .m, Rtsh = 1.19 .m, a = 0.62, m = 2.15, and n = 2. Table
3 shows water saturation values from Archies equation,
Schlumberger program (Saraband, Coriband) and from Equations
(18) and (19).
The difference between water saturation values derived from
Archies formula and other shaly sand equations indicates the
importance of using a shale sand model for this example. Also, it
is obvious that there is really no difference between Schlumberger
values and Picket cross plot values. Pickets plot of modified total
shale Equation (16), does not need previous knowledge of m and

Equation (19)
0.39
0.42
0.39
1.00
1.00

aRw. Based on the points pattern shown on the cross plot, these
parameters can be determined by trial and error. But it needs the
existence of a nearby water section.

Conclusions
The conductivity of clean reservoir rock is of electrolytic origin. The water phase which carries the electric current must of
course be continuous in order to contribute to the rock conductivity. When clean reservoir rock contains clays or heavy minerals
such as pyrite, the resistivity becomes low. Clay minerals increase
rock conductivity by increasing the conductivity of bulk water in
the pore spaces. Extra rock conductivity is caused by pyrite which
creates electronic conductivity that is usually comparable to or
even higher than formation water electrolytic conductivity. These
are the most common reasons for low resistivity pay zones.
To evaluate a low resistivity pay zone, we must identify the origin of the low resistivity phenomenon. The main problem with
pyrite is how to estimate its volume and distribution and correct
the formation resistivity, and then calculate water saturation by
Archies formula or by a shale sand model. In the case of shaly
formation, there are several models to correct the water saturation
value. The choice of the model is controlled by the type, the distribution and the volume of clay minerals in the pay zones.

REFERENCES
1. AGUILERA, R., Extensions of Pickett Plots for the Analysis of
Shaly Formations by Well Logs; The Log Analyst, Vol. 35, No. 5, pp.
304-313, 1990.
2. BERG, C.R., Effective Medium Model for Water Saturation in
Porous Rocks; Geophysics, Vol. 60, No. 4, pp., 1070-1080, 1995.

FIGURE 6: Log data for field example 3.


July 2000, Volume 39, No. 7

13

21. WAXMAN, M.H. and SMITS, L.J.M., Electrical Conductivities in


Oil Bearing Shaly Sands; Joumal of Petroleum Technology, Vol. 8,
pp. 107-132, 1968.
22. WILEY, R. and SNODDY, M.L., Complex Resistivity of Shaly
Sand and Minerals; The Log Analyst, Vol. xxvii, No. 5, pp. 45-59,
1986.
23. WORTHINGTON, P.F., The Evolution of Shaly Sand Concepts in
Reservoir Evaluation; The Log Analyst, Vol. 26, pp. 23-40, 1985.
24. ZEMANEK, J., Low Resistivity Hydrocarbon Bearing Sand
Reservoir; Society of Petroleum Engineers FE, pp. 515-521,
December 1989.

ProvenanceOriginal unsolicited manuscript, Petrophysical


Evaluation of Low Resistivity Sandstone Reservoirs
(98-10-13). Abstract submitted for review August 5, 1998; editorial comments sent to the author(s) October 23, 1998; revised manuscript received December 8, 1998; paper approved for pre-press
April 21, 1999; final approval June 28, 2000.

FIGURE 7: Picket crossplot for field example 3.


3. BERG, C.R., Effective Medium Resistivity Models for Calculating
Water Saturation in Shaly Sands; The Log Analyst, Vol. 37, No. 3,
pp. 16-28, 1996.
4. CLAVIER, C., COATES, G., and DUMANOIR, J., The Theoretical
and Experimental Basis for the Dual Water Model for the
Interpretation of Shaly Sand; Journal of Petroleum Technology, Vol.
24, pp. 153-168, 1984.
5. DE KUIJPER, A., SANDOR, R.K.J., HOFMAN, J.P., KOELMAN,
J.M.V.A., HOFSTRA, P., and DE WAAL, J.A., Electrical
Conductivities in OilBearing Shaly Sand Accurately Described
with the SATORI Model; The Log Analyst, Vol. 37, No. 5, pp. 22-32,
1996.
6. DE WITTE, L., Relation Between Resistivities and Fluid Contents
of Porous Rocks; Jounal of Oil and Gas, August 24, 1950.
7. FERTL, W.H. and HAMMACK, G.W., A Comparative Look at
Water Saturation Computations in Shaly Pay Sands; SPWLA
Symposium, 1971.
8. HAMADA, G.M., An Integrated Approach to Determine Shale and
Hydrocarbon Potential in Shaly Sands; Transactions Intl. Symposium
of the Society of Core Analysts, France, September 1996.
9. HILCHIE, D.W., Applied Openhole Log Interpretation; Douglas W.
Hilchie Inc., Colorado, 1982.
10. KLIMENTOS, T., Pyrite Volume Estimation by Well Log Analysis
and Petrophysical Studies; The Log Analyst, Vol. 36, No. 6, pp. I117, 1995.
11. MILLOT, M., Geologie Des Argiles; Masson, Paris, 1969.
12. PATCHETT, J.G. and RAUSCH, R.W., An Approach to
Determining Water Saturation in Shaly Sands; Journal of Petroleum
Technology, Vol. 19, pp. 1395-1405, 1967.
13. PICKETT, G.R., Pattern Recognition as a Means of Formation
Evaluation; Transactions of 14 th Annual Logging Symposium
SPWLA, paper A, 1973.
14. POUPON, A., LOY, M.E., and TIXIER, M.P., A Contribution to
Electric Log Interpretation in Shaly Sands; Transactions of the
American Institute of Mechanical Engineers, Vol. 201, pp. 138-145,
1954.
15. Schlumberger, Log Interpretation Principles/Applications;
Schlumberger, Ltd., New York, 1987.
16. SCHWARTZ, L.M., SEN, P.N., and JOHNSON, D.L., Influence of
Rough Surfaces on Electrolytic Conduction; Physical Review
Bulletin, Vol. 40, pp. 2450-2458, 1989.
17. SEN, P.N., SCALA, C., and COHEN, M.H., A Self Similar Model
for Sedimentary Rocks with Application to the Di-Electric Constant
of Fiiseed Glass Leads; Geophysics, Vol. 46, pp. 781-795, 1981.
18. SERRA, O., Fundamentals of Well Log Interpretation; Elsevier, New
York, 1984.
19. SIMANDOUX, P., Mesures Diaelectriques en Milieu Poreux,
Application a Mesure des Saturations en Eau, Etude du
Comportement des Massifs Argileux; Reveu de LIFP, Vol. IX,
pp. 193-215, 1963.
20. TIXIER, M.P., MORRIS, P.L., and CONNELL, J.C., Log
Evaluation of Low Resistivity Pay Sands in the Gulf Coast;
Transactions of 9th Annual Logging Symposium of SPWLA, paper E,
1968.

14

Authors Biographies
Gharib Hamada is a professor of well logging and applied geophysics. He is currently with King Saud University, Saudi
Arabia. Previously, he was with the
Technical University of Denmark, Sultan
Qaboos University, Oman; and Cairo
University, Egypt. He received a B.Sc. in
1975 and an M.Sc. in 1979 from Cairo
University, Egypt; a DEA in 1980 and a
Doc. DIng in 1983 from Bordeaux
University, France. He is a member of SPE, SPWLA, and SCA.
Musaed Al-Awad is an assistant professor
with King Saud University in Saudi Arabia.
He is a professor of oil and gas well
drilling engineering and petroleum related
rock mechanics. Dr. Al-Awad received a
B.Sc. in 1987 and an M.Sc. in 1990 both in
petroleum engineering from King Saud
University, and a Ph.D. in 1994 from
Heriot-Watt University, Edinburgh, UK.

Size matters.
petroleumshow.com

Journal of Canadian Petroleum Technology

También podría gustarte