Vous êtes sur la page 1sur 21

GEOHORIZONS

AUTHORS

Quantification of pore structure


and its effect on sonic velocity
and permeability in carbonates
Ralf J. Weger, Gregor P. Eberli, Gregor T. Baechle,
Jose L. Massaferro, and Yue-Feng Sun

ABSTRACT
Carbonate rocks commonly contain a variety of pore types
that can vary in size over several orders of magnitude. Traditional pore-type classifications describe these pore structures
but are inadequate for correlations to the rocks physical properties. We introduce a digital image analysis (DIA) method
that produces quantitative pore-space parameters, which can
be linked to physical properties in carbonates, in particular
sonic velocity and permeability.
The DIA parameters, derived from thin sections, capture
two-dimensional pore size (DomSize), roundness (g), aspect
ratio (AR), and pore network complexity (PoA). Comparing
these DIA parameters to porosity, permeability, and P-wave
velocity shows that, in addition to porosity, the combined effect of microporosity, the pore network complexity, and pore
size of the macropores is most influential for the acoustic behavior. Combining these parameters with porosity improves
the coefficient of determination (R2) velocity estimates from
0.542 to 0.840. The analysis shows that samples with large simple pores and a small amount of microporosity display higher
acoustic velocity at a given porosity than samples with small,
complicated pores. Estimates of permeability from porosity
alone are very ineffective (R2 = 0.143) but can be improved
when pore geometry information PoA (R2 = 0.415) and DomSize (R2 = 0.383) are incorporated.
Furthermore, results from the correlation of DIA parameters
to acoustic data reveal that (1) intergrain and/or intercrystalline

Copyright 2009. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received January 7, 2009; provisional acceptance March 27, 2009; revised manuscript
received May 2, 2009; final acceptance May 27, 2009.
DOI:10.1306/05270909001

AAPG Bulletin, v. 93, no. 10 (October 2009), pp. 12971317

1297

Ralf J. Weger  University of Miami,


Rosenstiel School of Marine and Atmospheric
Science, Division of Marine Geology and Geophysics, 4600 Rickenbacker Causeway, Miami,
Florida 33129; rweger@rsmas.miami.edu
Ralf J. Weger was a postdoctoral researcher with
the Comparative Sedimentology Laboratory at
the University of Miami when the article was
written. He received his B.S. degree in systems
analysis (2000) and his Ph.D. in marine geology
and geophysics (2006) from the University of
Miami. His dissertation focuses on quantitative
pore- and rock-type parameters in carbonates
and their relationship to velocity deviations. His
main interests range from processing and visualization of geophysical data to petrophysical
characterization of carbonate rocks.
Gregor P. Eberli  University of Miami, Rosenstiel School of Marine and Atmospheric
Science, Division of Marine Geology and Geophysics, 4600 Rickenbacker Causeway, Miami, Florida
33129; geberli@rsmas.miami.edu
Gregor P. Eberli is a professor in the Division of
Marine Geology and Geophysics at the University
of Miami and the Director of the Comparative
Sedimentology Laboratory. He received his Ph.D.
from the Swiss Institute of Technology (ETH) in
Zrich, Switzerland. His research integrates the
sedimentology, stratigraphy, and petrophysics
of carbonates. With laboratory experiments and
seismic modeling, his group tries to understand
the physical expression of carbonates on log and in
seismic data. He was a distinguished lecturer for
AAPG (1996/97), Joint Oceanographic Institutions
(1997/1998), and the European Association of
Geoscientists and Engineers (2005/2006).
Gregor T. Baechle  University of Miami,
Rosenstiel School of Marine and Atmospheric
Science, Division of Marine Geology and Geophysics, 4600 Rickenbacker Causeway, Miami,
Florida 33129
Gregor T. Bchle graduated from the University
of Tbingen in 1999 with a Diploma (equivalent to
M.Sc. degree) in geology. In 2001, he joined the
Comparative Sedimentology Laboratory (CSL)
with a Scholarship of the German Academic Exchange Service to obtain a Ph.D. from the University of Tbingen. From 2004 to 2008, he was a
research associate in the CSL, managing the rock
physics laboratory. He is currently working for
ExxonMobil Upstream Research Company, Quantitative Interpretation, Houston, Texas.

Jose L. Massaferro  Gerencia Geologa y


Estudios Integrados, Direccin Exploracin y Desarrollo de Negocio, Macacha Gemes 515,
(C1106BKK), Puerto Madero, Buenos Aires,
Argentina
Jose Luis Massaferro is a geology manager in
Repsol YPFs exploration office in Argentina. He
received his Ph.D. from the University of Miami in
1997. He was a Fulbright Fellow while pursuing
his studies in Miami. Prior to his Ph.D. studies, he
worked for Texaco as a geologist. In 1998, he joined
Shell E&P and was involved in different projects,
including 3-D seismic volume interpretation, highresolution sequence stratigraphy, and kinematic
modeling of compressional structures. In 2005,
he joined Repsol in Madrid.
Yue-Feng Sun  Department of Geology
and Geophysics, Texas A&M University, College
Station, Texas 77843
Yue-Feng Sun is an associate professor at Texas
A&M University. He received his Ph.D. (1994)
from Columbia University. He has 25 years of
experience as a geoscientist in the industry and
academia. His professional interests include
carbonate rock physics, poroelasticity, poroelectrodynamics, reservoir geophysics, and petroleum
geology. He is a member of AAPG, the American
Geophysical Union, American Physical Society,
and the Society of Exploration Geophysicists.

ACKNOWLEDGEMENTS

The methodology presented in this paper was


developed in collaboration with Shells carbonate development team in Rijswijk, Holland, and
the Comparative Sedimentology Laboratory
of the University of Miami. We acknowledge financial support from Shell and the Industrial
Associates of the Comparative Sedimentology
Laboratory. Discussions with Guido Bracco
Gartner, Gene Rankey, and Peter Swart were
essential to the technical development of the
equipment and methodology. Comments and
reviews on several versions of the manuscripts
by Wayne Ahr, Stephen Ehrenberg, Jerry Lucia,
Mark Longman, David Kopaska-Merkel, and
Jeroen Kenter greatly improved the manuscript.
The AAPG Editors thanks the following reviewers
for their work on this paper: Jeroen Kenter,
David C. Kopaska-Merkel, and Mark W.
Longman.

1298

Geohorizons

and separate-vug porosity cannot always be separated using


sonic logs, (2) P-wave velocity is not solely controlled by the
percentage of spherical porosity, and (3) quantitative pore geometry characteristics can be estimated from acoustic data and
used to improve permeability estimates.

INTRODUCTION
Several attempts have been made to find a rock or pore-type
classification that would capture rock texture, pore type, and
petrophysical characteristics (Archie, 1952; Choquette and
Pray, 1970; Lucia, 1983, 1995; Lny, 2006). In this article,
we describe a digital image analysis (DIA) method for measuring quantitative pore-structure parameters derived from
thin sections and introduce four parameters that are most reliable for capturing the geometrical character of pore structure in carbonates.
Many studies have recognized that acoustic velocity in carbonates is dependent upon pore geometry (Anselmetti and
Eberli, 1993, 1997, 1999; Kenter et al., 1995; Wang, 1997; Sun
et al., 2001; Eberli et al., 2003; Baechle et al., 2004; Weger
et al., 2004; Weger, 2006). In many theoretical studies, the
pore aspect ratio is assumed to be the main geometric variable
influencing acoustic velocity (Assefa et al., 2003; Saleh and
Castagna, 2004; Agersborg et al., 2005; Kumar and Han,
2005; Rosseb et al., 2005). The theoretical concept is that
high-aspect-ratio pores, such as molds and vugs, provide more
grain-to-grain contact than interparticle and intercrystalline
pores, thus decreasing the pore compressibility and providing more stiffness to the rock at equal porosity (Mavko and
Mukerji, 1995; Saleh and Castagna, 2004). Consequently, a
sequence of rocks with mostly moldic and/or vuggy porosity
will have a higher acoustic velocity than a formation with predominantly intercrystalline and/or interparticle porosity with
the same amount of total porosity. Many scientists exploited
this fact to quantitatively estimate the amount of secondary
porosity (Schlumberger, 1972, 1974) and separate-vug porosity by modeling porosity from acoustic logs (e.g., Nurmi,
1984; Lucia and Conti, 1987; Wang and Lucia, 1993; Lucia,
1999). This modeling was based on (1) Wyllies time-average
equation (Wyllie et al., 1956) and (2) the assumption that
separate-vug porosity has a minor influence on the acoustic
log (Schlumberger, 1972, 1974; Lucia, 1987; Doveton, 1994).
Lucia and Conti (1987) and Lucia (1991) calibrated the influence of separate-vug porosity on acoustic logs by point counting
separate-vug porosity on thin sections of oomoldic rocks, and

proposed empirical equations to calculate separatevug porosity from acoustic transit time. Anselmetti
and Eberli (1993, 1997, 1999), however, showed
how in carbonates a variety of pore types produce
variable velocities in rocks with similar porosity.
Other experiments documented that oomoldic carbonate samples with near-spherical pores show large
scatter in velocities with up to 2500 m/s (8202 ft/s)
difference at a given porosity (Baechle et al., 2007,
2008a; Knackstedt et al., 2008).
In an attempt to quantify the influence of pore
structure on permeability, Anselmetti et al. (1998)
defined the DIA parameter g that describes the
roundness of pores and compared it with measured permeability values of plugs with characteristic pore types. The parameter showed a strong
correlation to permeability. Anselmetti and Eberli
(1999) also quantified the pore-structure-induced
scatter of velocities at any given porosity with the
velocity deviation, which is defined as the difference between measured velocities and velocities
estimated using Wyllies time-average equation.
Intervals from MiocenePliocene cores from the
Great Bahama Bank with high positive velocity
deviation and oomoldic porosity show low permeability. This finding corroborated the general notion that rocks with a high amount of separate-vug
porosity have a high velocity and low permeability. The application of the deviation log proved
less successful in Cretaceous carbonates where
the separation between the medium and deep induction curves better detected the high flow zones
(Smith et al., 2003), indicating that the separation
between interparticle or intercrystalline and intragrain or vuggy porosity is insufficient to capture
all pore type-velocity-permeability relationships.
These complications were the motivation behind
the study presented in this article. The goal was to
find a repeatable, independent measure of the pore
structure that is needed to quantitatively evaluate
the influence of pore geometry on acoustic velocity
and other petrophysical properties.
The here-described methodology of DIA produces parameters that quantify the relationship
between pore geometry, acoustic velocities, and permeability. The high correlation between the DIA
parameters and the petrophysical values illustrates

the advantages of quantitative geometrical parameters over qualitative pore-type classifications.

DATA SET
One hundred twenty carbonate core-plug samples
(1-in. [25.4-mm] diameter by 1- to 2 in. [25.4
50.8 mm] long) were selected from cored wells at
several locations in the Middle East, Southeast Asia,
and Australia (Baechle et al., 2004). The Middle
East samples are from the Shuaiba Formation and
are Aptian in age, the Southeast Asian samples are
from an isolated platform of Miocene age, and the
Australian samples are from two drowned coolsubtropical platforms on the Marion Plateau and
are also Miocene in age (Ehrenberg et al., 2006).
Vertical plugs were drilled from reservoir and nonreservoir intervals to capture a wide range of total
porosity, rock types, and pore types. The set of selected samples includes textures ranging from
coarse-grained packstones with interparticle to
vuggy porosity to fine-grained wackestone dominated by interparticle to micromoldic porosity
(mG). All samples are either limestone or dolomite
with less than 2% noncarbonate minerals.
The samples have high-quality measurements
of velocity, porosity, and permeability. Thin sections are impregnated with blue epoxy and cut
from the end of the plug sample on which these
measurements were performed. Petrophysical measurements, geological description, and DIA parameter values are listed in the Appendix.

METHODS
Petrophysical Measurements
Sonic velocity was measured using an ultrasonic
transmitter-receiver pair with piezoelectric transducers forming the core of the equipment. The
transducer arrangement measures one compressional and two independent orthogonally polarized
shear waves simultaneously using a pulse transmission technique developed by Birch (1960). Both
transducers (compressional and shear) generate
Weger et al.

1299

wave signals at frequencies centered on 1 MHz. All


of the samples were saturated with distilled water
and placed between two piezoelectric transducers
and sealed with a rubber sleeve from the surrounding oil in the pressure vessel. The pressure conditions of the pore fluid inside the sleeve were kept
constant at 2 MPa. The confining pressure of the
surrounding oil was varied to produce a range of
effective pressure conditions. For this study, only
measurements with a fluid pressure of 2 MPa and
an effective pressure of 20 MPa were used.
Porosity for all samples was measured by determining the difference between the measured
volume of the core plug (e.g., 1-in. [25.4-mm] diameter  plug length) and the real volume determined by helium injection in a Boyles law porosimeter. The samples real volume is determined by
pumping a known amount of helium gas into the
sample chamber and measuring the pressure. Helium is able to fill all but the smallest connected
micropores of the sample. Gas permeability was
measured at a confining pressure of 20 bar. All values are reported as Klinkenberg-corrected permeabilities in units of millidarcies.

Descriptive Thin-Section Analysis


Thin sections were qualitatively described and classified using traditional carbonate rock and poretype classifications according to Dunham (1962),
the extended Dunham terminology (Embry and
Klovan, 1971), Choquette and Pray (1970), and
Lucia (1995, 1999). Rocks altered by recrystallization that obliterated the original texture are referred
to as recrystallized rocks.
Pore space was described using a limited Choquette and Pray (1970) terminology. In our samples, we determined interparticle, intercrystalline,
moldic, vuggy, and intraparticle porosity, and included intraframe porosity to describe the pore
space within boundstone and rudstone. In addition, we use the term micromoldic (mG) to describe
microscopic molds (<20 mm) that can be detected
with strong magnification. As is characteristic for
carbonates, most samples contain more than one
pore type. The quantification of their respective
1300

Geohorizons

proportions is observer based, resulting in great difficulty to quantitatively assess their respective influence on the petrophysical properties. The pore
types of each sample are listed in the Appendix.
The term dominant pore type is used for the pore
type containing more than 50% of the visible porosity, and minor pore types contribute with at least
5% to the visible pores. In the figures, only the dominant pore type is used for color coding.

Digital Image Analysis Using


Cross-Polarized Light
The DIA method uses three basic steps: (1) image
acquisition, (2) image segmentation, and (3) calculation of pore geometry parameters. A comparison of
images with different orientation in cross-polarized
light (XPL) is used during image segmentation to
differentiate between pore space and matrix (Weger,
2006).
Photomicrographs obtained using conventional digital cameras in combination with an optical light microscope (OLM) have been used in investigations of sandstones and carbonates. The
DIA relies heavily on accurate image segmentation, i.e., the separation of a specific feature such as
pore space from its background. Selection procedures vary widely ranging from simple manual selection to more sophisticated automated thresholding procedures (Crabtree et al., 1984; Ehrlich et al.,
1991; Anselmetti et al., 1998; Fens, 2000; Van den
Berg et al., 2002; Keehm, 2003) or probabilitytheory-based maximum likelihood classification
(Lillesand and Kiefer, 1994; Van den Berg et al.,
2002). In our method, we use variations in extinction under XPL to image and segment pore and
rock space.
We use these differences in extinction behavior
to distinguish white air bubbles and unfilled pore
space from white crystals (Figure 1). To obtain the
differences, we photographed the thin section first
without the polarizing filter and subsequently
with a filter at three different angles (0, 20, 40).
Using a Matlab program, the differences in pixel
intensity during rotation are calculated at the imaging resolution of approximately 30 mm2/pixel.

Figure 1. (a) Image acquired using plane-polarized light shows a thin-section photomicrograph of a carbonate impregnated with blue
epoxy resin. Minerals and grains are beige, whereas pore space is blue except for an air bubble with color identical to the matrix. (b) The
intensity image of absolute cross-polarized-light (XPL) variation covers the same area and is derived using XPL images at different angles.
(c, d) The close ups and the distributions illustrated in panel (e) show that the red-green-blue (RGB) color bands of the subsection are not
capable of separating air bubbles from the matrix mineral, but the XPL variation of intensity is clearly different in those regions.

This difference is then combined with color values


for image segmentation into pore space and rock
(Weger, 2006).
Pore-Shape Parameters from Digital
Image Analysis
Two different types of parameters exist for pore
shape calculation (Russ, 1998): global parameters
that describe the entire pore system on a photograph
or thin section and local parameters that are obtained from individual pores. All shape parameters used here are derived from two-dimensional
(2-D) images. We are aware of the limitation of
2-D-derived geometrical properties for correlation
to the physical property of the three-dimensional
sample volume. However, any kind of thin-section
analysis, quantitative or not, suffers from this limitation. In addition, we performed a variety of tests

on computed tomography (CT) scans of core plugs


at a resolution comparable to that of our OLM
images that suggested that directionality has little
influence on geometrical parameter values.
In our DIA, 37 parameters are measured on
each thin section. A principal component analysis
was performed to identify the most important and
distinguishable parameters (Weger, 2006). Four
DIA parameters proved to best describe several
aspects of the pore system. Definitions and short
descriptions of the parameters characteristics are
given below. More specific explanations on the derivation and characteristics of these parameters are
given by Weger (2006).
Perimeter Over Area
Perimeter over area (PoA) is the ratio between the
total pore-space area on a thin section and the total
perimeter that encloses the pore space. The PoA
Weger et al.

1301

can be regarded as a 2-D equivalent to a specific


surface, the ratio between pore volume and pore
surface. Generally, a small number indicates a simple geometry. The PoA values in our data range from
less than 40 mm1 to more than 250 mm1. Because of the almost log-normal distribution of
these parameter values, some figures are plotted
in log10(PoA) instead of PoA.
Dominant Pore Size
Dominant pore size (DomSize) is determined as the
upper boundary of pore sizes of which 50% of the
porosity on a thin section is composed. This parameter provides an indication of the pore-size range
that dominates the sample. In our data, DomSize
ranges from less than 100 mm to more than 1 mm
(0.039 in.) (units given in length as equivalent diameter). As in the case of PoA, some figures show values of log10(DomSize) instead of DomSize.
Gamma
Gamma (g) was defined by Anselmetti et al., (1998)
as the perimeter over an area of an individual pore
normalized to a circle; i.e., a perfect round circle
would have a g of 1. The g describes the roundness
of the pore. In our data, the area-weighted mean
of g for the entire thin section ranges from 1.5
to 4.5.
Aspect Ratio
Aspect ratio (AR) is defined here as the ratio between the major and the minor axis of an ellipse
that encloses the pore. The AR describes the elongation of the pore-bounding ellipsoid. The arithmetic means of AR values for the entire thin section range from 1 to 2.5. In acoustic modeling,
pores are commonly assumed to be ideal ellipsoidal
inclusions with a variable AR (Kuster and Toksz,
1974; Norris, 1985). This ideal ellipsoid does not
consider the edginess, complexity, or surface roughness of the pore.
Amount of Microporosity
In our methodology, macropores are defined by
pores, which are vertically connected through the
thin section, resulting in a minimum pore diameter
of approximately 30 mm (the thickness of a thin
1302

Geohorizons

section). The amount of microporosity is calculated as the difference between the observed porosity in DIA and the measured porosity from the core
plug. The geometry of the micropores is not assessed in this study, but the percentage of microporosity is included in the analysis.
In Figure 2, several digital photographs of different thin sections are placed next to a PoA-DomSize
crossplot to demonstrate that these parameters
vaguely recognize and separate traditional carbonate pore classifications. These parameters are, however, not limited to the grouping of samples, as traditional carbonate pore-type classifications are, but
provide a continuous ordered scale of pore geometry. Coarse grainstones with large pores and relatively simple pore systems tend to show large DomSize and small PoA. In contrast, packstones to
mudstones with large amounts of microporosity
commonly show high PoA and low DomSize.
Mutivariate Regressions
Multivariate linear regression is used to quantify
trends among velocity, porosity, and four different
geometrical parameters. We use the coefficient of
determination (R2) between the measured and the
estimated velocity to quantify how well the model
explains the measured data. In addition to direct
linear regression, a semilinear approach is used,
which combines linear regression and Wyllies timeaverage equation. Some rearrangement of the timeaverage equation leads to an explicit formulation
of Wyllies velocity estimate (VpW).
VpW



1f
f 1

VpS
VpF

where VpS and VpF are the compressional velocity


of the solid and the fluid, respectively, and f is porosity. This formulation was used by Anselmetti
and Eberli (1999) to define the velocity deviation as
DVp Vp  VpW

where Vp is the measured compressional velocity


of the sample.

Figure 2. Crossplot of perimeter over area (PoA) versus dominant pore size (DomSize) where the measured acoustic velocity is superimposed in color. (ad) Thin-section images are shown to illustrate carbonate pore types corresponding to certain combinations of digital
image analysis (DIA) parameters and velocity. The samples shown as images are represented by enlarged dots, and exact parameter
values are listed below each thin-section photograph.
Weger et al.

1303

Figure 3. Velocity-porosity crossplot of water-saturated carbonate samples measured at 20-MPa confining pressure. A first-order
inverse proportional relationship between velocity and porosity
can be observed, but individual samples deviate from this trend
in excess of 2000 m/s (6562 ft/s).

This formulation can be used to incorporate


the velocity deviation into a regression model defined as
Vp VpW c0 c1 x e ^y e

where Vp is the measured compressional velocity,


x represents any measured geometrical parameter
(e.g., PoA or DomSize), c0 and c1 are constants to be
determined during the regression, ^y represents the
new velocity estimate, and e is the error term that
in this case would contain both measurement error
and any other influences on velocity that were not
accounted for.

RELATIONSHIP OF PORE STRUCTURE TO


SONIC VELOCITY
The velocity-porosity data of all core-plug samples
show a characteristic first-order trend of increasing
acoustic velocity with decreasing porosity. At any
given porosity, a spread of velocity in excess of
1500 m/s (4921 ft/s) can be observed (Figure 3).
This large scatter cannot be explained by mineralogy because most samples contain only minor
1304

Geohorizons

amounts of dolomite, and variations in grain velocity (e.g., calcite to dolomite) could not produce
such large velocity variations. In addition, all samples were measured saturated with distilled water
so that fluid velocities are constant. Anselmetti and
Eberli (1993) demonstrated that such variation of
velocity at a given porosity is typical in carbonates
and relates to the pore structure. To test this conclusion and to quantify the effect of pore structure,
we relate the four digital image parameters, PoA,
DomSize, AR, and g, to sonic velocity and porosity. Because each of the parameters captures a different characteristic of the pore system, this correlation also assesses the relative importance of each
geometric characteristic for Vp.
Geometry and Trends in
Velocity-Porosity Space
Crossplots of velocity porosity with the digital image parameters PoA, DomSize, AR, g, percentage
of microporosity (% microporosity), and traditional
pore types using the Choquette and Pray (1970)
classification superimposed in color are shown in
Figure 4. Figure 4a displays the samples color coded
with the dominant pore type, which is visually estimated on the thin section. Most samples, however,
contain more than one pore type, and these additional pore geometries (the Appendix lists the minor pore types) might explain some of the scatter.
Nevertheless, some slight trends are visible. For example, samples with vuggy or moldic porosity tend
to fall into the high-velocity area, but several moldic
samples display a low velocity and overlap with
samples containing interparticle porosity. Samples
containing either micromoldic porosity or porosity within particles occupy the lower part of the
velocity-porosity data cloud. In contrast, samples
with interparticle porosity cover the entire velocityporosity space. Samples with high amounts of microporosity (10070%) tend to cluster around the
Wyllie time-average equation (Figure 4b), and at
any given porosity, a trend of increasing velocity with
decreasing microporosity is observed (Figure 4b).
The digital image parameters of the macropores also define trends with similar orientation
in the velocity-porosity space. The PoA shows a

Figure 4. Comparison between (a) Choquette and Pray (1970) pore types, (b) microporosity fraction, and four digital-image-analysis
parameters: (c) dominant pore size, (d) gamma (g), (e) perimeter over area, and (f) aspect ratio. All parameters are superimposed in
color onto velocity-porosity crossplots. All show a gradient that differentiates samples with high velocity from samples with low velocity at
any given porosity.
Weger et al.

1305

Figure 5. Illustration of the importance


of pore structure as a factor controlling
acoustic velocity using pore-shape characteristics as the third dimension. Threedimensional crossplot between sonic velocity (Vp), porosity (f), and perimeter
over area (top) and dominant pore size
(bottom) with simple linear regression
surfaces.

clear trend in which at any given porosity, samples


with a low value of PoA (simple pore geometry)
have relatively high velocities, whereas samples
with high values of PoA (more complex pore geometry) have low velocities (Figure 4e). In other
words, samples with simple pore geometries are
faster than samples with a complicated pore structure if porosities are the same. The DomSize also
shows a clear trend of increasing velocity with increasing values of DomSize at a given porosity.
This trend indicates that samples with larger pores
are faster than those with smaller pores at equal
porosities (Figure 4c). The roundness of individ1306

Geohorizons

ual pores is captured by g, which shows generally


low values in samples with relatively low velocities
at a given porosity and vice versa (Figure 4d). This
trend is similar as for the DomSize but not as well
developed (Figure 4c, d). The AR only displays a
very weak trend in velocity-porosity space where
samples with low velocity for their given porosity
are generally those with high ARs (Figure 4f ). The
parameters PoA and AR form trends with similar
orientation. Low values of PoA and AR correspond
to high velocities, and high values of PoA and AR
correspond to low velocities for a given porosity
(Figure 4e, f). The parameters DomSize and g form

trends in the opposite direction (Figure 4c, d), where


low values correspond to low velocities and high
parameter values correspond to high velocities at
any given porosity.
The trends formed by PoA and DomSize
(Figure 4c, e) are very strong (Figure 5), indicating
that pore structure is a second independent parameter influencing velocity. In Figure 5, these quantitative DIA parameters are displayed together with
velocity and porosity in three dimensions. Many
samples align closely with a simple linear best-fit
surface that is displayed for reference. This illustrates that what appears as a 2-D scatter (Figure 3)
is mostly caused by the projection of this surface
into a 2-D crossplot.
A crossplot of PoA and DomSize with acoustic
velocity superimposed in color (Figure 2) illustrates
the link between the parameters PoA, DomSize,
and acoustic velocity and rock texture. Four thinsection images are shown to illustrate the difference in pore structure detected by high, medium,
and low parameter values. Low-velocity samples are
characterized by DomSize below 200300 mm and
PoA above 50 mm1. The corresponding thin-section
images are dominated by small pores, a significant
amount of small particles, and/or abundant microporosity (Figure 2c, d). In contrast, high-velocity samples are characterized by DomSize above 300 mm
and PoA below 50 mm1. The corresponding thinsection images show larger pores, larger particles,
and little to no mud (Figure 2a, b). In general, high
velocities correspond to samples with simple and
large pores with smooth pore surfaces, low specific
surface, and a small amount of microporosity.
Quantitative Assessment of Different
Geometric Characteristics
To explore the link between velocity, porosity, and
pore-space geometry quantitatively, velocity is estimated using multivariate linear regression from
combinations of porosity and the DIA parameters.
The geometrical parameters g, PoA, DomSize,
and AR, and the percentage of microporosity were
used for multivariate linear regression. The correlation coefficients between measured and estimated
velocity are listed in Table 1.

Table 1. Coefficients of Determination from the Correlation


between Measured Velocity and Estimated Velocity from
Regressions with the Following Digital Image Analysis Parameters
as Input Variables: Dominant Pore Size, Gamma, Perimeter
over Area, Aspect Ratio, and Percentage of Microporosity*
Estimators Used for Velocity Prediction

R2

Porosity
Porosity and AR
Porosity and g
Porosity and DomSize
Porosity and % microporosity
Porosity and PoA
Porosity and PoA and AR
Porosity and PoA and DomSize
Porosity and PoA and g
Porosity and PoA and % microporosity
Porosity and PoA and % microporosity
Porosity and PoA and % microporosity
Porosity and PoA and % microporosity
Porosity and PoA and % microporosity
DomSize and AR
Porosity and PoA and % microporosity
DomSize and g
Porosity and PoA and % microporosity
DomSize and AR and g

and AR
and g
and DomSize
and

0.542
0.549
0.639
0.768
0.769
0.786
0.788
0.800
0.810
0.820
0.822
0.832
0.840
0.841

and

0.844

and

0.845

*The geometric parameters PoA and DomSize in addition to porosity significantly


improve the correlation, whereas the combination of several geometrical parameters does not produce significant improvement. DomSize = dominant pore size;
g = gamma; PoA = perimeter over area; AR = aspect ratio; % microporosity =
percentage of microporosity.

As a first step, porosity alone is used as an estimator of compressional velocity. The correlation
between the measured and the estimated velocity
resulted in a coefficient of determination (R2) of
0.542. Second, a linear combination of porosity
and a single DIA parameter (g, AR, PoA, DomSize,
% microporosity) is used to estimate compressional velocity. The parameter AR combined with
porosity produces the least effective velocity estimate (R2 = 0.549, Table 1). The highest correlation coefficient of all estimates (R2 = 0.845) is
obtained by combining porosity with all DIA parameters (g, AR, PoA, DomSize, and % microporosity), but this correlation coefficient is only
slightly better than the estimate from a combination of porosity with PoA, DomSize, and % microporosity (R2 = 0.840).
Weger et al.

1307

Figure 6. Crossplots between velocity deviation and digital image parameters. Both parameters, perimeter over area (PoA) and dominant pore size (DomSize), are capable of explaining more than 60% of the variability in velocity deviation.

Crossplots between the velocity deviation and


the log10 of DIA parameters PoA and DomSize
(Figure 6) result in an R2 of 0.65 and an R2 of
0.62, respectively. This means that these two quantitative geometric parameters are able to explain
6265% of the deviation of acoustic velocity (DVp)
from Wyllies time-average equation at a given
porosity.

PERMEABILITY AND PORE SHAPE


Pore size and specific surface influence permeability. In our data, pore size and pore network complexity (PoA), which is the 2-D equivalent of a
specific surface, have a strong influence on permeability (Figure 7). Samples with low permeability
for their given porosity have high values of PoA

Figure 7. Permeability-porosity (K-f) crossplots with perimeter over area (PoA) and dominant pore size (DomSize) superimposed in
gray scale. Both parameters exhibit trends in porosity-permeability space. Samples with low permeability despite relatively high porosity
have high values of PoA and low values of DomSize. Samples with high permeability have low values of PoA and high values of DomSize,
representing samples with a large and simple pore structure.
1308

Geohorizons

and low values of DomSize. In turn, samples with


high permeability for their given porosity show low
values of PoA and high values of DomSize. Low values of PoA represent a simple pore structure and
low specific surface, whereas high values of PoA
stem from a more complex pore structure and high
specific surface (Figures 2, 7).
Quantitative Permeability Estimation
Bear (1972) refined Kozenys (1927) equation to
express permeability as a function of porosity, specific surface, and tortuosity. Here we estimate permeability using pore geometry parameters and incorporate them into Kozenys equation.
k cf3 =S2

correlation between a measured and estimated


permeability of 0.419 (red dots in Figure 8).
Microporosity and pore-throat diameter are important to properly predict flow properties. Thinsection-based pore-structure analyses like the presented method here do not capture pore geometries
below the 30-mm threshold, but the macro- and
mesopore system represents a large part of the flow
capacity. This is reflected in the improvement of
the permeability estimates from R2 = 0.143 to R2 =
0.415 gained by incorporating the macroscale DIA
parameters into the Kozceny equation (Figure 8).
Further improvements of permeability estimates
will be possible using micro-CT scans (Knackstedt
et al., 2008) or by combining DIA analysis and mercury injection capillary pressure.

where k is permeability, f is porosity, c is Kozenys factor, which can be estimated from porosity (Fabricius et al., 2007), and S is the specific
surface with respect to bulk volume. The PoA is
the 2-D equivalent of the specific surface, and thus,
we estimate S from measured 2-D geometrical
parameters (PoA and DomSize).
We compare four different approaches to estimate permeability. First, estimates of permeability are derived from porosity alone. For comparison, Kozenys S is expressed as a function of PoA
and DomSize and used for permeability estimation. Finally, the relationship between acoustic velocity and pore geometry is used to calculate S directly from acoustic data. This estimate of S is then
combined with porosity to estimate permeability
directly from a combination of measured porosity
and acoustic velocity.
Figure 8 shows a comparison of measured and
estimated permeabilities. Estimation of permeability using porosity alone is extremely ineffective (R2 = 0.143, black dots in Figure 8). These estimates can be improved using pore geometry
information from PoA and DomSize (R2 = 0.415
and R2 = 0.383, green and blue dots in Figure 8).
The good relationship between sonic velocity, porosity, and PoA allows for the substitution of PoA
by a geometry estimate derived from sonic velocity. Using this geometry estimate, we obtain an R2

DISCUSSION
Anselmetti and Eberli (1999) demonstrated how
acoustic velocities in carbonates are influenced by
porosity and a variety of pore structures using traditional carbonate pore-type classification (Choquette
and Pray, 1970). In our data, the separation of samples grouped according to Choquette and Prays
pore-type classification is poor in velocity-porosity
space (Figure 4a), indicating that the classification
of Choquette and Pray is not capable of uniquely
defining ranges of specific acoustic properties. In
comparison, quantitative characterization of porespace geometry using DIA parameters such as PoA
(Figure 4e) has the advantage of providing a continuous numerical parameter that can be used directly in a mathematical formulation used to estimate velocity.
The AR is the geometrical parameter most commonly used in theoretical models to explain variations in rock stiffness and acoustic velocities (Assefa
et al., 2003; Saleh and Castagna, 2004; Agersborg
et al., 2005; Kumar and Han, 2005; Rosseb et al.,
2005), although Rafavich et al. (1984) concluded
that AR does not significantly influence velocity.
The weak correlation for velocity estimates using
porosity and the DIA parameter for roundness
(g) and AR, respectively, questions this assumption. Our results indicate that (1) the amount
Weger et al.

1309

Figure 8. Comparison between measured and estimated permeability (k). Estimates are derived using four different models with different
input parameters. Green dots are estimates derived from porosity alone. Both blue and black dots are derived using measured porosity
and the measured geometric parameters perimeter over area (PoA) and dominant pore size (DomSize). Red dots represent permeability
estimates derived using measured porosity (f), measured acoustic velocity (Vp), and assumed grain and fluid velocities (VpS and VpF).

of microporosity and (2) the size and complexity


of the macropore system are much more important
factors for determining the stiffness and, thus, the
acoustic behavior of carbonates (Table 1). A recent
study of oomoldic rocks by Baechle et al. (2007)
also documented that the percentage of spherical
pore shape is not the dominant factor in producing
positive deviations from the Wyllie time-average
equation. They attribute the variable acoustic response of up to 2000 m/s (6562 ft/s) at a given porosity to variations in intercrystalline porosity in
the rock frame; a conclusion that is corroborated
by ultra-high-resolution CT tomography and scanning electron microscope (SEM) analysis on the
same samples (Knackstedt et al., 2008).
1310

Geohorizons

Baechle et al. (2008b) proposed that the fraction of stiff macropores versus soft micropores is
responsible for the variation of velocity at any given
porosity and develop a rock physics model that captures the presence of both macro- and microporosity
to better estimate velocity and permeability. The
percentage of microporosity for this dual porosity
DEM model is derived with the DIA methodology
described here (Baechle et al., 2008b).
The assumption that rocks with mostly moldic
and/or vuggy porosity will have a faster acoustic velocity than a formation with predominantly intercrystalline and/or interparticle porosity has been
used for quantitative estimates of separate-vug porosity from acoustic logs (e.g., Nurmi, 1984; Lucia

Figure 9. Velocity-porosity
crossplot of samples measured
at 20 MPa with annotation of
porosity types separated into two
groups. Open circles are samples with vuggy, moldic, intraframe, and intragrain porosity,
black and gray dots represent
samples with interparticle and
intercrystalline porosity. A large
overlap exists between these
two groups, indicating that rocks
with interparticle and/or intercrystalline porosity can in some
cases have a stiff framework
and high velocity.

and Conti, 1987; Wang and Lucia, 1993; Anselmetti


and Eberli,1999). To test this assumption, we distribute the samples into two groups. The vuggy group
consists of samples whose primary pore types are
vuggy, moldic, intraframe, and intraparticle porosity. The interparticle group consists of samples with
interparticle and intercrystalline porosity as the
primary pore type. Plotting the two groups in the
velocity-porosity space reveals a considerable overlap (Figure 9). The samples of the vuggy group generally plot above the Wyllie time-average equation, and a cluster of interparticle samples in the
low velocity area is observed. However, nearly an
equal amount of samples from each group display
an exceptionally high velocity at a given porosity
(Figure 9).
High velocity at a given, sometimes high porosity is possible if pore compressibility is low and
consequently if the stiffness of the rock is not significantly decreased (Mavko and Mukerji, 1995).
Such a stiff frame is well known to occur in rocks
with vugs or molds, but it also occurs in rocks with
interparticle and intercrystalline porosity. A process that can produce frame stiffening in these latter
rocks is contact cementation (Dvorkin and Nur,

1996). In carbonates, cementation at grain contacts


are meniscus cements derived from meteoric waters
(Harris, 1978; Longman, 1980) or micritic bridging
cements in the marine realm (Hillgrtner et al.,
2001). In a Holocene grainstone, small amounts
of bridging cement (15% of the total rock) produce
a Vp of 4500 m/s (14,764 ft/s) at 20 MPa (Eberli
et al., 2003). Some of the samples displayed in
Figure 9 are dolomites; in this case, the extreme
stiffening of the frame is not caused by early cement
but more likely by interlocking crystals. Anselmetti
et al. (1997) documented this process on Neogene
carbonates, in which the velocity of sucrosic dolomite increases dramatically as isolated rhombohedra grow together to form a stiff framework.

CONCLUSIONS AND IMPLICATIONS


The DIA quantifies the influence of pore types on
velocity and permeability. A combination of porosity
and (image-derived) microporosity is capable of
estimating velocity with R2 = 0.77; a combination of porosity and digital image parameters is
able to explain more than 85% of the variation
Weger et al.

1311

of acoustic velocity (R2 = 0.85). The geometrical


characteristics most influential for acoustic velocity
are the complexity of pore space (PoA) and the sizes
of the pores (DomSize). These parameters combined with porosity estimate velocity with R2 =
0.79 and 0.77, respectively. In short, carbonates with
a large amount of microporosity, a complex pore
structure (high specific surface), and small pores
generally show low acoustic velocity at a given porosity. Samples with a simple pore structure (low
specific surface) and large pores show high acoustic
velocity for their porosity.
Knowledge of roundness (g) and the aspect ratio of pores (AR) does not significantly enhance
the ability to estimate sonic velocity in carbonates.
Thus, incorporating parameters that capture both
size and complexity (e.g., DomSize and PoA) potentially improves acoustic velocity models.
The finding that samples with interparticle
and intercrystalline porosity can display high velocity similarly to samples with separate-vug porosity is an unwelcomed finding because its nonunique
acoustic response adds uncertainty to quantitative
estimates of separate-vug porosity from velocity
logs. It is well documented that separate-vug porosity is mostly ineffective with regard to velocity,
and in reservoirs that are dominated by such po-

rosity (oomoldic), these estimates work well (Lucia


and Conti, 1987; Wang and Lucia, 1993; Anselmetti
and Eberli, 1999). The nonunique acoustic response of separate-vug porosity might explain why
estimates based on these models do not always
yield the expected results. Given the relationship
between permeability and the DIA parameters
PoA and DomSize, in theory, it should be possible to discriminate high and low permeability at a
given porosity directly from well-log data. For example, rocks with high acoustic velocity for their
given porosity generally show low specific surface
(PoA) and large pore sizes (DomSize, Figures 4, 5).
Rocks with low specific surface and large pore
sizes also have high permeability for their given porosity (Figure 7, Appendix). These relationships
imply two things: (1) not all fast intervals in carbonates that produce a positive acoustic impedance
are necessarily tight, low-porosity sequences, and
(2) a quantitative assessment of the pore types by
DIA and their acoustic response is beneficial for
an accurate interpretation of log-based pore-type
estimates.

APPENDIX: DATA TABLE

Appendix. Texture, Pore Type, DIM (Digital Image Analysis) Parameter Values, and Petrophysical

Measurements*
Sample
C5-B1
C5-B100
C5-B101
C5-B102
C5-B103
C5-B104
C5-B105
C5-B106
C5-B107
C5-B108
C5-B109
C5-B110
C5-B111
C5-B112
C5-B113

1312

Dunham
Index**
G
G
G
G
G
G
G
G
P-G
G
P-G
G
G
G
G

Dominant
Pore Type
IP
MO
IP
IP
IP
IP
IP
IP
IP
IP
MO
WF
WF
mG
IP

Geohorizons

Minor
Pore Type

WG
MO
MO
FR, MO
MO

Gamma

DomSize
(mm)

PoA
(mm1)

AR

VP (m/s)

Phi (%)

Micro
Phi (%)

2.15
2.37
2.25
2.61
2.61
2.17
1.78
2.55
2.05
1.98
2.03
2.38
2.38
1.85
2.78

39
48
73
188
188
87
106
202
63
78
113
208
208
52
108

167
151
103
69
69
89
83
58
117
99
79
52
52
137
90

0.52
0.59
0.55
0.54
0.54
0.54
0.59
0.54
0.58
0.52
0.57
0.53
0.53
0.61
0.56

3177
3185
3262
3738
3866
3458
3853
4050
3406
4893
3435
4259
4177
3466
3403

28.0
27.6
30.4
25.8
26.3
29.0
23.7
29.9
27.4
12.8
26.2
22.5
22.1
26.7
27.4

26.8
25.3
25.3
21.8
22.3
24.4
21.7
21.4
25.3
9.8
22.9
17.3
16.9
25.7
19.3

K (md)
6.7
11.3
35.6
26.1
26.1
37.8
4.5
184.0
13.8
9.8
7.7
4.2
4.2
3.9
25.8

Appendix. Cont.
Sample

Dunham
Index**

Dominant
Pore Type

C5-B114
C5-B115
C5-B116
C5-B117
C5-B118
C5-B119
C5-B120
C5-B58
C5-B60
C5-B61
C5-B72
C5-B74
C5-B75
C5-B79
C5-B80
C5-B81
C5-B82
C5-B84
C5-B85
C5-B86
C5-B87
C5-B88
C5-B89
C5-B90
C5-B91
C5-B92
C5-B93
C5-B94
C5-B95
C5-B96
C5-B97
C5-B98
C5-B99
C5-L10
C5-L11
C5-L12
C5-L13
C5-L14
C5-L15
C5-L16
C5-L17
C5-L19
C5-L2
C5-L20
C5-L21
C5-L22
C5-L23
C5-L24
C5-L25
C5-L26

G
G
G
G-P
G
G
G
FR
RD
P
RD-FR
G
P-G
G
P
G-RD
W-P
P-G
G
W
RD-FL
FL-RD
P
G
FL
G-RD
G-P
G-P
G
G-P
P-G
G-P
G
P
rDol
G-P
rDol
rDol
rDol
rDol
G
rDol
G
rDol
rDol
rDol
rDol
P
rDol
G

IP
WF
IP
IP
IP
IP
IP
VUG
IP
VUG
IP
WF
mG
IP
mG
IP
MO
IP
IP
mG
MO
WP
IP
IP
MO
IP
MO
IP
IP
IP
IP
IP
MO
WP
VUG
WP
VUG
MO
IX
VUG
IP
VUG
IP
VUG
VUG
IX
IX
IP
VUG
MO

Minor
Pore Type
WG
IP
MO
MO
WG
IP
WP, MO
IP
VUG, WF
IP
MO
IP
MO
MO, WP
VUG
MO
MO
IP
IP, MO, FR
MO
IP, FR
WP
MO
VUG

FR
MO
IX
IP
IX
IX
VUG
IX
IX
MO
IX
IX
VUG
MO
MO
IX

Gamma

DomSize
(mm)

PoA
(mm1)

AR

VP (m/s)

Phi (%)

Micro
Phi (%)

K (md)

2.87
2.96
2.10
2.43
3.74
2.23
2.44
2.88
2.58
1.95
2.89
2.51
1.84
2.61
1.97
2.19
1.82
1.61
2.31
2.09
2.32
2.15
2.14
2.57
2.50
1.86
3.74
4.22
2.89
2.25
2.70
3.48
1.70
2.17
2.85
2.36
3.05
2.15
3.53
3.74
3.64
3.23
3.77
2.57
2.73
2.29
3.15
2.61
2.45
2.09

97
262
90
170
178
151
143
560
680
421
519
1200
50
31
39
224
129
102
106
50
143
294
92
87
154
135
215
43
68
53
27
20
109
157
345
118
368
440
451
790
310
452
447
466
297
205
372
115
370
121

92
55
84
66
73
71
70
34
30
45
42
18
150
196
157
42
63
71
96
167
78
49
115
109
95
62
81
164
109
169
215
244
74
139
48
147
47
36
40
28
71
43
41
36
51
77
49
111
38
112

0.54
0.54
0.53
0.55
0.57
0.55
0.54
0.54
0.53
0.59
0.49
0.54
0.60
0.50
0.56
0.54
0.57
0.57
0.57
0.59
0.57
0.54
0.56
0.51
0.52
0.58
0.54
0.51
0.48
0.46
0.59
0.44
0.64
0.60
0.57
0.52
0.55
0.55
0.55
0.56
0.55
0.55
0.54
0.55
0.55
0.56
0.55
0.56
0.55
0.52

3377
3867
3520
3974
3714
4782
3513
4703
4555
4564
4628
4362
3466
3179
2898
3856
3936
4171
3768
4413
3374
4102
4084
4023
5156
3974
3786
3266
3481
3535
3324
3692
3156
4753
5791
4011
5747
5797
5180
4737
5333
4658
3894
5991
5949
5890
3274
3961
5430
5361

28.0
29.8
26.8
25.4
29.4
17.9
28.3
21.8
25.7
18.7
15.9
23.6
29.4
26.4
30.8
26.7
28.8
20.1
27.1
15.9
30.0
23.9
21.9
21.4
12.1
24.3
29.7
26.2
29.8
28.4
22.3
21.8
28.0
13.4
14.2
26.3
20.0
19.5
26.0
33.6
17.8
31.9
41.6
11.2
13.0
13.3
44.7
25.2
21.0
10.8

22.6
23.0
18.4
15.1
24.4
14.8
24.4
15.5
15.9
17.1
10.5
10.6
28.2
25.1
30.2
14.8
25.8
17.5
26.3
15.5
28.5
21.5
10.9
18.0
10.1
18.5
26.6
24.8
27.5
13.4
22.1
21.5
26.2
8.4
4.3
25.1
12.7
9.1
8.2
12.4
15.3
12.4
18.3
1.2
5.5
5.2
32.0
15.2
10.8
10.1

23.5
63.8
36.7
64.7
55.3
2.9
71.5
2195.0
1321.1
12.7
9.0
646.0
13.2
2.1
4.1
113.5
20.8
4.1
14.0
0.1
19.9
1.5
4.7
221.5
5.0
99.8
24.4
1.6
18.3
2.4
1.7
2.3
4.5
0.1
2.0
0.6
562.0
2.9
2340.0
15,049.0
91.9
5564.0
15,966.0
123.0
28.7
92.2
525.0
1.0
131.0
0.0

Weger et al.

1313

Appendix. Cont.
Sample

Dunham
Index**

Dominant
Pore Type

C5-L27
C5-L28
C5-L29
C5-L3
C5-L30
C5-L31
C5-L32
C5-L33
C5-L34
C5-L35
C5-L36
C5-L37
C5-L38
C5-L39
C5-L4
C5-L40
C5-L41
C5-L42
C5-L43
C5-L44
C5-L45
C5-L46
C5-L47
C5-L48
C5-L49
C5-L5
C5-L50
C5-L51
C5-L52
C5-L53
C5-L54
C5-L55
C5-L6
C5-L7
C5-L8
C5-L9
C5-M18
C5-M56
C5-M57
C5-M59
C5-M62
C5-M63
C5-M64
C5-M65
C5-M66
C5-M67
C5-M68
C5-M69
C5-M70
C5-M71

B
P
G-B
rDol
rDol
rDol
G
FL
rDol
G
rDol
P
P
P
rDol
rDol
G
G
rDol
rDol
rDol
G
P-G
rDol
P-G
rDol
rDol
rDol
P
G
G-B
G-B
G
P
rDol
rDol
G-P
P-G
G-P
G-P
P-G
P
P
G-P
rDol
P
P
G
P-G
rDol

WF
IP
IP
IX
VUG
IX
IP
IP
MO
MO
IX
WP
MO
mG
VUG
MO
IP
MO
IX
MO
VUG
WP
IP
VUG
IP
IX
VUG
IX
IP
MO
IP
WF
IP
WP
VUG
VUG
IP
VUG
IP
IP
IP
IP
IP
MO
VUG
VUG
VUG
IX
MO
IX

1314

Geohorizons

Minor
Pore Type
IX
MO
WF
VUG
IX
VUG
VUG
IX
WP
VUG
IP
IP
IX
IX
WP
VUG
IX
IX
MO
MO
IX
MO
VUG
IX
MO
IX
WF
IP
MO
MO
IX
IX
MO
MO, IP
VUG, WP
MO
MO
IP
mG
IX
IX
IP
VUG
IP
VUG, MO

Gamma

DomSize
(mm)

PoA
(mm1)

AR

VP (m/s)

Phi (%)

Micro
Phi (%)

K (md)

3.27
3.07
2.22
2.30
2.90
2.95
2.20
2.92
2.71
2.46
3.28
3.50
2.55
2.50
3.13
2.66
2.49
2.39
2.43
3.49
2.62
1.97
3.01
2.62
2.78
3.05
2.78
3.27
2.00
2.67
2.86
2.24
2.68
2.61
3.71
2.44
2.90
2.56
2.98
2.27
2.18
3.73
2.04
1.91
2.56
2.56
1.86
2.38
2.18
2.46

357
453
413
254
702
602
355
643
290
488
652
852
440
325
439
412
425
132
362
874
14
134
471
514
352
344
247
318
93
347
595
388
279
353
1031
331
112
393
233
81
76
230
49
521
521
685
113
267
98
341

51
53
38
55
31
29
41
32
55
304
35
41
51
61
42
53
40
87
44
26
35
79
49
38
42
45
67
53
137
54
38
46
54
108
25
48
105
35
65
125
140
56
149
36
36
43
78
51
97
43

0.54
0.55
0.56
0.56
0.56
0.56
0.55
0.56
0.56
0.45
0.54
0.55
0.56
0.55
0.56
0.53
0.54
0.54
0.56
0.53
0.56
0.55
0.57
0.55
0.55
0.55
0.54
0.53
0.56
0.56
0.57
0.56
0.53
0.52
0.55
0.56
0.54
0.58
0.55
0.59
0.57
0.53
0.58
0.57
0.54
0.53
0.55
0.54
0.57
0.56

6148
4249
5662
6080
5918
4650
5908
5356
4951
5297
4791
3956
4381
4246
5303
5640
5604
4520
5132
5407
6325
4615
4784
5271
4860
5871
5335
4259
5442
5082
6183
5910
4093
4977
6325
5850
4038
3725
3612
3671
3978
4346
3524
4357
5604
4285
4105
4477
3829
5531

14.7
27.2
17.5
10.1
14.2
32.1
16.4
24.2
24.8
11.8
36.1
29.7
20.8
25.0
20.3
13.0
18.9
21.1
24.1
21.8
9.7
17.2
18.2
26.5
26.2
13.4
21.0
32.7
8.9
25.2
11.6
17.2
29.5
14.2
10.5
12.4
26.6
33.5
28.9
26.0
23.6
32.0
29.4
20.8
13.0
22.5
26.7
20.8
31.9
11.4

7.0
21.3
10.0
2.9
4.2
12.9
10.5
8.9
12.7
11.7
17.5
23.5
15.3
23.7
3.8
7.5
10.4
18.2
14.2
0.0
0.9
5.2
12.8
16.8
13.9
3.1
15.5
16.7
8.5
11.2
6.3
11.7
24.3
13.8
0.5
6.0
20.0
15.5
19.3
22.4
19.3
24.1
25.0
15.5
7.9
16.5
22.8
16.2
27.6
4.2

8.1
895.0
535.0
12.2
240.0
29,369.0
698.0
12.7
209.0
2.0
11,940.0
25.2
1.5
4.3
29.1
0.4
2550.0
0.1
167.0
0.7
0.7
0.1
1575.0
25,775.0
2032.0
122.0
16.1
2423.0
0.0
331.0
271.0
401.0
1410.0
0.7
94.2
54.3
15.0
390.0
22.0
14.0
13.0
300.0
21.0
3.7
150.0
36.0
26.0
120.0
26.0
150.0

Appendix. Cont.
Sample

Dunham
Index**

Dominant
Pore Type

Minor
Pore Type

C5-M73
C5-M76
C5-M77
C5-M78
C5-M83

G-P
W-P
P
P
P

MO
IP
IP
MO
MO

IP
MO
MO
VUG, IP

Gamma

DomSize
(mm)

PoA
(mm1)

AR

VP (m/s)

Phi (%)

Micro
Phi (%)

2.05
2.81
2.05
1.91
2.20

118
96
174
148
46

110
116
67
67
163

0.57
0.57
0.56
0.57
0.59

4092
3696
4450
4118
4117

23.4
27.9
21.7
30.0
24.1

20.1
24.3
20.2
24.0
21.6

K (md)
8.6
11.0
3.9
63.0
5.5

*DomSize = dominant pore size; PoA = perimeter over area; AR = aspect ratio; VP = compressional acoustic velocity (values of water-saturated conditions with a confining
pressure of 20 MPa at a frequency of 1 kHz); Phi = porosity; K = permeability.
**G = grainstone; P = packstone; W = wackestone; M = mudstone; FL = floatstone; FR = framestone; RD = rudstone; B = boundstone; combinations are separated by a
hyphen; rDol = completely recrystallized rocks.

IP = interparticle; IX = intercrystalline; MO = moldic; VUG = vuggy; WPO = intraparticle; WF = intraframe; mG = micromoldic; FR = fracture. The dominant pore type listed
in the table is estimated to contain more than 50% of the visible pores. Minor pore types are listed if they are more than an estimated 5% of the total visible pores.

REFERENCES CITED
Agersborg, R., T. A. Johansen, and M. Jakobsen, 2005, The
T-matrix approach for carbonate rocks: Society of Exploration Geophysicists, Expanded Abstracts, v. 24, p. 1597
1600.
Anselmetti, F. S., and G. P. Eberli, 1993, Controls on sonic
velocity in carbonates: Pure and Applied Geophysics,
v. 141, p. 287323, doi:10.1007/BF00998333.
Anselmetti, F. S., and G. P. Eberli, 1997, Sonic velocity in carbonate sediments and rocks, in I. Palaz and K. J. Marfurt,
eds., Carbonate seismology: Society of Exploration Geophysicists, Geophysical Developments Series 6, p. 5374.
Anselmetti, F. S., and G. P. Eberli, 1999, The velocity-deviation
log: A tool to predict pore type and permeability trends
in carbonate drill holes from sonic and porosity or density
logs: AAPG Bulletin, v. 83, p. 450466.
Anselmetti, F. S., G. A. von Salis, K. J. Cunningham, and
G. P. Eberli, 1997, Controls and distribution of sonic velocity in Neogene carbonates and siliciclastics from the
subsurface of the Florida Keys: Implications for seismic
reflectivity: Marine Geology, v. 144, p. 931, doi:10.1016
/S0025-3227(97)00081-9.
Anselmetti, F. S., S. M. Luthi, and G. P. Eberli, 1998, Quantitative characterization of carbonate pore systems by digital image analysis: AAPG Bulletin, v. 82, p. 18151836.
Archie, G. E., 1952, Classification of carbonate reservoir
rocks and petrophysical considerations: AAPG Bulletin,
v. 36, p. 278298.
Assefa, S., C. McCann, and J. Sothcott, 2003, Velocities of
compressional and shear waves in limestones: Geophysical Prospecting, v. 51, p. 113, doi:10.1046/j.1365-2478
.2003.00349.x.
Baechle, G. T., R. Weger, G. P. Eberli, and J. L. Massaferro,
2004, The role of macroporosity and microporosity in
constraining uncertainties and in relating velocity to permeability in carbonate rocks: Society of Exploration
Geophysicists, Expanded Abstracts, v. 23, p. 1662.
Baechle, G. T., L. Al-Kharusi, G. P. Eberli, A. Boyd, and A.
Byrnes, 2007, Effect of spherical pore shapes on acoustic

properties: AAPG Annual Convention, Abstracts Volume, v. 16, p. 7.


Baechle, G. T., G. P. Eberli, A. Boyd, J.-M. DeGrange, and L.
Al-Kharusi, 2008a, Oomoldic carbonates: Pore structure
and fluid effects on sonic velocity: Society of Exploration
Geophysicists, Expanded Abstracts, v. 27, p. 1660.
Baechle, G. T., A. Colpaert, G. P. Eberli, and R. Weger,
2008b, Effects of microporosity on sonic velocity in carbonate rocks: Leading Edge, v. 27, no. 8, p. 10121018,
doi:10.1190/1.2967554.
Bear, J., 1972, Dynamics of fluids in porous media: Mineola,
New York, Dover Publications, 764 p.
Birch, F., 1960, The velocity of compressional waves in rocks
to 10 kilobars: 1: Journal of Geophysical Research, v. 65,
p. 10831102, doi:10.1029/JZ065i004p01083.
Choquette, P. W., and L. C. Pray, 1970, Geologic nomenclature and classification of porosity in sedimentary carbonates: AAPG Bulletin, v. 54, p. 207244.
Crabtree Jr., S. J., R. Ehrlich, and C. Prince, 1984, Evaluation
of strategies for segmentation of blue-dyed pores in thin
sections of reservoir rocks: Computer Vision, Graphics
and Image Processing, v. 28, p. 118, doi:10.1016/0734
-189X(84)90136-1.
Doveton, J. H., 1994, Geological log interpretationReading
the rocks from wireline logs: SEPM Short Course 29,
169 p.
Dunham, R. J., 1962, Classification of carbonate rocks
according to depositional texture: AAPG Memoir 1,
p. 108121.
Dvorkin, J., and A. Nur, 1996, Elasticity of high-porosity sandstones: Theory for two North Sea data sets: Geophysics,
v. 61, no. 5, p. 13631370, doi:10.1190/1.1444059.
Eberli, G. P., G. T. Baechle, F. S. Anselmetti, and M. L. Incze,
2003, Factors controlling elastic properties in carbonate
sediments and rocks: The Leading Edge, v. 22, p. 654
660, doi:10.1190/1.1599691.
Ehrenberg, S. N., G. P. Eberli, and G. Baechle, 2006, Porositypermeability relationships in Miocene carbonate platforms and slopes seaward of the Great Barrier Reef, Australia (ODP Leg 194, Marion Plateau): Sedimentology,

Weger et al.

1315

v. 53, p. 12891318, doi:10.1111/j.1365-3091.2006


.00817.x.
Ehrlich, R., S. J. Crabtree, K. O. Horkowitz, and J. P. Horkowitz,
1991, Petrography and reservoir physics: I. Objective
classification of reservoir porosity: AAPG Bulletin,
v. 75, p. 15471562.
Embry, A. F., and J. E. Klovan, 1971, A Late Devonian reef
tract on northeastern Banks Island, Northwest Territories: Canadian Petroleum Geology Bulletin, v. 19,
p. 730781.
Fabricius, I. L., G. Baechle, G. P. Eberli, and R. Weger, 2007,
Estimating permeability of carbonate rocks from porosity and vP/vS: Geophysics, v. 72, no. 5, p. E185E191,
doi:10.1190/1.2756081.
Fens, T. W., 2000, Petrophysical properties from small rock
samples using image analysis techniques: Doctoral dissertation thesis, Technical University of Delft, Netherlands, 199 p.
Harris, P. M., 1978, Holocene marine cemented sands, Joulters
Ooid Shoal, Bahamas: Gulf Coast Association of Geological Societies Transactions, v. 28, p. 175185.
Hillgrtner, H., C. Dupraz, and W. Hug, 2001, Microbially induced cementation of carbonate sands: Are micritic cements indicators of vadose diagenesis?: Sedimentology,
v. 48, p. 117131, doi:10.1046/j.1365-3091.2001.00356.x.
Keehm, Y., 2003, Computational rock physics; transport
properties in porous media and applications: Doctoral
dissertation thesis, Stanford University, Stanford, California, 135 p.
Kenter, J. A. M., et al., 1995, Parameters controlling acoustic
properties of carbonate and volcaniclastic sediments
at sites 866 and 869, in E. L. Winterer, ed., Proceedings of the Ocean Drilling Program, Scientific Results,
Leg 143, Northwest Pacific atolls and guyots: College
Station, ODP, Texas A&M University, p. 287303.
Knackstedt, M. A., R. M. Sok, A. P. Sheppard, S. Latham, M.
Madadi, C. H. Arns, G. Baechle, and G. P. Eberli, 2008,
Probing pore systems in carbonates: Correlations to petrophysical properties: Society of Petrophysicists and Well
Log Analysts 49th Annual Logging Symposium, Edinburgh, Scotland, p. 117.
Kozeny, J., 1927, ber kapillare leitung des Wassers im Boden:
Sitzungsberichte der Wiener Akademie des Wissenschaften, v. 136, p. 271306.
Kumar, M., and D.-h. Han, 2005, Pore shape effect on elastic
properties of carbonate rocks: Society of Exploration
Geophysicists, Expanded Abstracts, v. 24, p. 14771480.
Kuster, G. T., and M. N. Toksoz, 1974, Velocity and attenuation of seismic waves in two-phase media: Part I. Theoretical formulations: Geophysics, v. 39, p. 587606,
doi:10.1190/1.1440450.
Lillesand, T. M., and R. W. Kiefer, 1994, Remote sensing and
image interpretation: New York, John Wiley and Sons,
v. 1, 733 p.
Longman, M. W., 1980, Carbonate diagenetic textures from
near surface diagenetic environments: AAPG Bulletin,
v. 64, p. 461487.
Lny, A., 2006, Making sense of carbonate pore system:
AAPG Bulletin, v. 90, p. 13811405, doi:10.1306
/03130605104.

1316

Geohorizons

Lucia, F. J., 1983, Petrophysical parameters estimated from


visual descriptions of carbonate rocks, a field classification of carbonate pore space: Journal of Petroleum Technology, v. 35, p. 629637.
Lucia, F. J., 1987, Rock fabric, permeability, and log relationships in a upward-shoaling vuggy carbonate sequence:
Bureau of Economic Geology, University of Texas at
Austin, Geological Circular 87.5, 22 p.
Lucia, F. J., 1991, Geological engineering aspects of the San
Andres reservoirs in the Lawyer Canyon, Algerita Escarpment outcrop and Seminole, subsurface field, in C.
Kerans, F. J. Lucia, R. K. Senger, G. E. Fogg, H. S.
Nance, E. Kasap, and S. D. Hovorka, eds., Characterization of reservoir heterogeneity in carbonate-ram systems, San Andres/Grayburg, Permian Basin: Bureau of
Economic Geology, University of Texas at Austin, Final
Report, p. 117174.
Lucia, F. J., 1995, Rock-fabric/petrophysical classification of
carbonate pore space for reservoir characterization:
AAPG Bulletin, v. 79, p. 12751300.
Lucia, F. J., 1999, Carbonate reservoir characterization: Berlin, Springer-Verlag, 226 p.
Lucia, F. J., and R. D. Conti, 1987, Rock fabric, permeability,
and log relationships in an upward-shoaling, vuggy carbonate sequence: Bureau of Economic Geology, University of Texas at Austin, Geological Circular 87-5,
22 p.
Mavko, G., and T. Mukerji, 1995, Seismic pore space compressibility and Gassmanns relation: Geophysics, v. 60,
p. 17431749, doi:10.1190/1.1443907.
Norris, A. N., 1985, A differential scheme for the effective moduli of composites: Mechanics of Materials, v. 4, p. 116.
Nurmi, R. D., 1984, Carbonate pore systems: Porosity/permeability relationships and geological analysis: AAPG
Bulletin, v. 68, no. 4, p. 513514.
Rafavich, F., C. H. S. C. Kendall, and T. P. Todd, 1984, The
relationship between acoustic properties and the petrographic character of carbonate rocks: Geophysics, v. 49,
p. 16221636, doi:10.1190/1.1441570.
Rosseb, . H., I. Brevik, G. R. Ahmadi, and L. Adam, 2005,
Modeling of acoustic properties in carbonate rocks: Society of Exploration Geophysicists, Expanded Abstracts,
v. 24, p. 15051508.
Russ, J. C., 1998, The image processing handbook: Boca
Raton, Florida, CRC Press, 771 p.
Saleh, A. A., and J. P. Castagna, 2004, Revisiting the Wyllie
time average equation in the case of near-spherical pores:
Geophysics, v. 69, p. 4555, doi:10.1190/1.1649374.
Schlumberger, 1972, Log interpretation, Vol. I: Principles:
New York, Schlumberger Limited, 113 p.
Schlumberger, 1974, Log interpretation, Vol. II: Applications: New York, Schlumberger Limited, 116 p.
Smith, L. B., G. P. Eberli, J. L. Masaferro, and S. Al-Dhahab,
2003, Discrimination of effective from ineffective porosity in heterogeneous Cretaceous carbonates, Al Ghubar,
Oman: AAPG Bulletin, no. 9, v. 87, p. 15091529, doi:10
.1306/041703200180.
Sun, Y. F., J. L. Massaferro, G. P. Eberli, and Y. C. Teng,
2001, Quantifying the effects of pore structure and fluid
saturation on acoustic wave velocity in carbonates, in E. C.

Shang, Q. Li, and T. F. Gao, eds., Theoretical and computational acoustics: Singapore, World Scientific, p. 335
347.
Van den Berg, E. H., A. G. C. A. Meesters, J. A. M. Kenter,
and W. Schlager, 2002, Automated separation of touching grains in digital images of thin sections: Computers
and Geosciences, v. 28, p. 179190, doi:10.1016
/S0098-3004(01)00038-3.
Wang, R. F. P., and F. J. Lucia, 1993, Comparison of empirical models for calculating the vuggy porosity and cementation exponent of carbonates from log responses: Bureau of
Economic Geology, University of Texas at Austin, Geological Circular 93-4, 27 p.

Wang, Z., 1997, Seismic properties of carbonate rocks: Geophysical Development Series, v. 6, p. 2952.
Weger, R. J., 2006, Quantitative pore/rock type parameters
in carbonates and their relationship to velocity deviations: Ph.D. dissertation thesis, University of Miami,
Coral Gables, 232 p.
Weger, R. J., G. T. Baechle, J. L. Masaferro, and G. P. Eberli,
2004, Effects of pore structure on sonic velocity in carbonates: Society of Exploration Geophysicists, Expanded Abstracts, v. 23, p. 1774.
Wyllie, M. R. J., A. R. Gregory, and L. W. Gardner, 1956, Elastic wave velocities in heterogeneous and porous media:
Geophysics, v. 21, p. 4170, doi:10.1190/1.1438217.

Weger et al.

1317

Vous aimerez peut-être aussi