Anda di halaman 1dari 26

Rock-Fabric/Petrophysical Classification of Carbonate

Pore Space for Reservoir Characterization1


F. Jerry Lucia2

ABSTRACT
This paper defines the important geologic parameters that can be described and mapped to allow
accurate petrophysical quantification of carbonate
geologic models. All pore space is divided into interparticle (intergrain and intercrystal) and vuggy
pores. In nonvuggy carbonate rocks, permeability
and capillary properties can be described in terms
of particle size, sorting, and interparticle porosity
(total porosity minus vuggy porosity). Particle size
and sorting in limestones can be described using a
modified Dunham approach, classifying packstone
as grain dominated or mud dominated, depending
on the presence or absence of intergrain pore
space. To describe particle size and sorting in dolostones, dolomite crystal size must be added to the
modified Dunham terminology. Larger dolomite
crystal size improves petrophysical properties in
mud-dominated fabrics, whereas variations in
dolomite crystal size have little effect on the petrophysical properties of grain-dominated fabrics.
A description of vuggy pore space that relates to
petrophysical properties must be added to the
description of interparticle pore space to complete
the petrophysical characterization. Vuggy pore
space is divided into separate vugs and touching
vugs on the basis of vug interconnection. Separate
vugs are fabric selective and are connected only
through the interparticle pore network. Separatevug porosity contributes little to permeability and

Copyright 1995. The American Association of Petroleum Geologists. All


rights reserved.
1Manuscript received June 27, 1994; revised manuscript received March
21, 1995; final acceptance May 1, 1995.
2Bureau of Economic Geology, University of Texas at Austin, Austin,
Texas 78713-7508.
This classification of carbonate pore space is the result of research
initiated while I was with Shell Oil Company and completed at the Bureau of
Economic Geologys Reservoir Characterization Research Laboratory, which
is funded by industry sponsors Agip, Amoco, ARCO, British Petroleum,
Chevron, Conoco, Exxon, Fina, JNOC, Marathon, Mobil, Phillips, Shell,
Texaco, Total, and Unocal. I am grateful for intense discussions with
colleagues Charles Kerans, Gary Vander Stoep, Fred Wang, Susan Hovorka,
Mark Holtz, Steve Ruppel, and Don Bebout. An in-depth review of the
manuscript by Philip W. Choquette was especially useful. Published with
permission of the director, Bureau of Economic Geology, University of Texas
at Austin.

AAPG Bulletin, V. 79, No. 9 (September 1995), P. 12751300.

should be subtracted from total porosity to obtain


interparticle porosity for permeability estimation.
Separate-vug pore space is generally considered to
be hydrocarbon filled in reservoirs; however, intragranular microporosity is composed of small pore
sizes and may contain capillary-held connate water
within the reservoir. Touching vugs are nonfabric
selective and form an interconnected pore system
independent of the interparticle system.
INTRODUCTION
The goal of reservoir characterization is to
describe the spatial distribution of petrophysical
parameters, such as porosity, permeability, and saturation. Wireline logs, core analyses, production
data, pressure-buildup data, and tracer tests provide quantitative measurements of petrophysical
parameters in the vicinity of the well bore. These
well bore data must be integrated with a geologic
model to display the petrophysical properties in
three-dimensional space. Studies that relate rock
fabric to pore-size distribution, and thus to petrophysical properties, are key to quantifying geologic
models in numerical terms for input into computer
simulators (Figure 1).
Geologic models are generally based on observations interpreted in terms of depositional environments and sequences. In the subsurface, cores and
wireline logs are the main source of data for these
interpretations. Engineering models are based on
wireline log calculations and average rock properties from core analyses. Numerical engineering data
and interpretive geologic data are joined at the
rock-fabric level because the pore structure is fundamental to petrophysical properties and the pore
structure is the result of spatially distributed depositional and diagenetic processes.
The purpose of this paper is to define the important geologic parameters that can be described and
mapped to allow accurate petrophysical quantification of carbonate geologic models by (1) describing the relationship between carbonate rock fabrics and petrophysical properties, (2) presenting a
generic petrophysical classification of carbonate
1275

1276

Classification of Carbonate Pore Space

Table 1. Pore-type Terminology and Abbreviations Used in


This Paper Compared to Abbreviations Used in Lucia
(1983) and Choquette and Pray (1970)

K FABRIC
ROC

Sedimentation

Wireline Logs

Term

Abbreviations
Choquette and
Lucia (1983)
Pray (1970)

Core Analysis
Diagenesis

Porosity
Permeability
Saturation

Production
Pressure

Tectonics

Tracer Tests

Figure 1Integration of spatial geologic data with


numerical engineering data through rock-fabric studies.

pore space, and (3) suggesting that rock-fabric


units are systematically stacked within high-frequency cycles.
PORE-SPACE TERMINOLOGY AND
CLASSIFICATION

Interparticle
Intergrain
Intercrystal
Vug
Separate vug
Moldic
Intraparticle
Intragrain
Intracrystal
Intrafossil
Intragrain
microporosity
Shelter
Touching Vug
Fracture
Solution-enlarged
fracture
Cavernous
Breccia
Fenestral

IP
IG
IX
VUG
SV
MO
WP
WG
WX
WF

BP

BC
VUG

MO
WP

G
SH
TV
FR

SH

FR

SF
CV
BR
FE

CH*
CV
BR
FE

*Channel.

Pore space must be defined and classified in


terms of rock fabrics and petrophysical properties fabrics to petrophysical rock properties in carbonto integrate geological and engineering information. ate rocks. The Archie classification focuses on estiArchie (1952) made the first attempt at relating rock mating porosity, but is also useful for approximating
permeability and capillary properties. Archie (1952)
recognized that not all pore space can be observed
using a 10-power microscope and that the surface
PORE TYPES
texture of the broken rock reflects the amount of
matrix porosity. Therefore, pore space is divided
CAVERNOUS
INTERGRAIN
MOLDIC
FRACTURE
into matrix and visible porosity (Figure 2). Chalky
INTERCRYSTAL
INTRAFOSSIL
SHELTER
SOLUTION-ENLARGED
texture indicates a matrix porosity of about 15%,
FRACTURE
sucrosic texture indicates a matrix porosity of about
7%, and compact texture indicates matrix porosity
ARCHIE (1952)
of about 2%. Visible pore space is described according to pore size, and ranges from no visible pore
MATRIX
space to larger than cutting size. Porosity and perVISIBLE (A,B,C, and D)
meability trends and capillary pressure characteristics are also related to these textures.
LUCIA (1983)
Although the Archie method is still useful for estiVUGGY
INTERPARTICLE
mating petrophysical properties, relating these
SEPARATE
TOUCHING
descriptions to geologic models is difficult because
the descriptions cannot be defined in depositional
CHOQUETTE and PRAY (1970)
or diagenetic terms. A principal difficulty is that no
provision is made for distinguishing between visible
NONFABRIC
FABRIC SELECTIVE
SELECTIVE
interparticle pore space and other types of visible
pore space, such as moldic pores. Research on carFigure 2Petrophysical classification of carbonate pore
types used in this paper (Lucia, 1983) compared with bonate pore space (Murray, 1960; Choquette and
Archies original classification (1952) and the fabric- Pray, 1970; Lucia, 1983) has shown the importance
selectivity concept of Choquette and Pray (1970). A = no of relating pore space to depositional and diagenetic
visible pore space, and B, C, and D = increasing pore fabrics and to distinguishing between interparticle
(intergrain and intercrystal) and other types of pore
sizes from pinpoint to larger than cutting size.

Lucia

1277

Figure 3Geological and petrophysical classification of carbonate interparticle pore space based on size and sorting of grains and crystals. The volume of interparticle pore space is important for characterizing the petrophysical
properties.

space. Recognition of the importance of these factors


prompted modification of Archies classification.
The petrophysical classification of carbonate
porosity presented by Lucia (1983) emphasized
petrophysical aspects of carbonate pore space, as
does the Archie classification. However, by comparing rock-fabric descriptions with laboratory
measurements of porosity, permeability, capillarity,
and Archie m values, Lucia (1983) showed that the
most useful division of pore types for petrophysical
purposes was of pore space between grains or crystals, called interparticle porosity, and all other pore
space, called vuggy porosity (Figure 2). Vuggy pore
space was further subdivided by Lucia (1983) into
two groups depending on how the vugs are interconnected: (1) vugs interconnected only through
the interparticle pore network are separate vugs,
and (2) vugs that form an interconnected pore system are touching vugs.

Choquette and Pray (1970) discussed the geologic concepts surrounding carbonate pore space and
presented a classification that is widely used. They
emphasized the importance of pore-space genesis
and used genetic, not petrophysical, divisions in
their classification. They divided all carbonate pore
space into two classes: fabric selective and nonfabric selective (Figure 2). Moldic and intraparticle
pore types were classified as fabric-selective porosity by Choquette and Pray (1970) and grouped with
interparticle and intercrystal porosity. However,
Lucia (1983) demonstrated that moldic and intraparticle pores have a different effect on petrophysical properties than do interparticle and intercrystal
pores and thus should be grouped separately. Poretype terms used in Lucias (1983) classification are
listed in Table 1, which compares his terms with
those suggested by Choquette and Pray (1970).
Although most of the terms defined by Choquette

1278

Classification of Carbonate Pore Space

Mercury/air extrapolated displacement pressure (psia)

180

22
21 Percent porosity

160
19

140

10

120

14
15

100
27
27
26
22
22

80

12

60
21

16

16

18

40
18

20

12

18

18

16 21

0
0

20

40

60

80

100

120

140

Figure 5Porosity-air permeability relationship for various particle-size groups in nonvuggy carbonate rocks
(Lucia, 1983).

Average particle size (m)


Figure 4Relationship between mercury displacement
pressure and average particle size for nonvuggy carbonate rocks with greater than 0.1 md permeability (Lucia,
1983). The displacement pressure is determined by
extrapolating the data to a mercury saturation of zero.

pores, but it is consistent with the Archie terminology and with the widespread, less restrictive
use in the oil industry of the term vuggy porosity in referring to visible pore space in carbonate
rocks.

and Pray are also used here, interparticle and vug


porosity have different definitions. Lucia (1983)
demonstrated that pore space located both
between grains (intergrain porosity) and between
crystals (intercrystal porosity) are petrophysically
similar. A term identifies these petrophysically similar pore types: interparticle. Interparticle was
selected because of its broad connotation. The classification of Choquette and Pray (1970) does not
have a term that encompasses these two petrophysically similar pore types. In their classification,
the term interparticle is used instead of intergrain. [See Choquette and Pray (1970, page 247)
for their justification for using interparticle vs.
intergrain.]
Vuggy porosity, as defined by Lucia (1983), is
pore space that is within grains or crystals or that
is significantly larger than grains or crystals; that is,
pore space that is not interparticle. Vugs are commonly present as leached grains, fossil chambers,
fractures, and large, irregular cavities. Fracture
porosity is categorized as a type of vuggy porosity
to be inclusive. This definition deviates from the
restrictive definition of vugs used by Choquette
and Pray (1970) as nondescript, nonfabric-selective

ROCK-FABRIC/PETROPHYSICAL
CLASSIFICATION
The foundation of the Lucia (1983) and the
Archie classifications is the concept that pore-size
distribution controls permeability and saturation
and that pore-size distribution is related to rock fabric. To relate carbonate rock fabrics to pore-size distribution, one must determine if the pore space
belongs to one of the three major pore types: interparticle, separate vug, or touching vug (Figure 2).
Each class has a different type of pore distribution
and interconnection. One must also determine the
volume of pore space in these various classes
because pore volume relates to reservoir volume
and, in the case of interparticle and separate-vug
porosity, to pore-size distribution.
Petrophysics of Interparticle Pore Space
In the absence of vuggy porosity, pore-size distribution in carbonate rocks can be described by particle size, sorting, and interparticle porosity (Figure
3). Lucia (1983) showed that particle size can be

Lucia

1279

A
0

0.5
mm

1.0

0.2
mm

0.4

0.5
mm

1.0

C
0

1.0
mm

2.0

Figure 6Photomicrographs showing examples of nonvuggy limestone rock fabrics. (A) Grainstone, = 25%, k =
1500 md. (B) Grain-dominated packstone, = 16%, k = 5.2 md. Note intergranular cement and pore space. (C) Muddominated packstone, = 18%, k = 4 md. Note microporosity. (D) Wackestone.

related to mercury capillary displacement pressure


in nonvuggy carbonates having more than 0.1 md
permeability. Because displacement pressure is a
function of the larger, well-connected pores, the particle size describes the size of the larger pores
(Figure 4). Whereas the displacement pressure characterizes the larger pores sizes and is largely independent of porosity, the shape of the capillary pressure curve characterizes the smaller pore sizes and is
dependent on interparticle porosity (Lucia, 1983).
The relationship between displacement pressure
and particle size (Figure 4) is hyperbolic and suggests important particle-size boundaries at 100
and at 20 m. Lucia (1983) demonstrated that three

porosity-permeability fields can be defined using


particle-size boundaries of 100 and 20 m, a relationship that appears to be limited to particle sizes
less than 500 m (Figure 5). These three fields will
be referred to as the greater than 100-m permeability field, 10020-m permeability field, and less
than 20-m permeability field.
Recent work has shown that permeability fields
can be better described in geologic terms if sorting
and particle size are considered. The approach to
size and sorting used herein is similar to the grainsupport/mud-support principle upon which
Dunhams (1962) classification is built. Dunhams
classification, however, focused on depositional

1280

Classification of Carbonate Pore Space

(B)

(A)

1000

Permeability (md)

Permeability (md)

1000

100

10

0.1

100

10

0.1
10

20

30

10

40

Interparticle porosity (%)

(C)

30

40

30

40

(D)
1000

Permeability (md)

1000

Permeability (md)

20

Interparticle porosity (%)

100

10

100

10

0.1

0.1
10

20

30

40

Interparticle porosity (%)

10

20

Interparticle porosity (%)

Figure 7Porosity-air permeability crossplots for nonvuggy limestone rock fabrics compared with the three permeability fields illustrated in Figure 5 (A) 400-m ooid grainstone, Ste. Genevieve, Illinois (Choquette and Steiner,
1985). Low-permeability, high-porosity data are deleted because they are from oomoldic and wackestone rock fabrics (P. W. Choquette, 1993, personal communication). (B) Grain-dominated packstone data, Wolfcamp, west Texas
(Lucia and Conti, 1987), a poorly sorted mixture of 80300-m grains and micrite. (C) Wackestones with microporosity between 5-m crystals, Shuaiba, United Arab Emirates (Moshier et al., 1988). Data are associated with stylolites not shown. (D) Coccolith chalk, Cretaceous (Scholle, 1977). The presence of intragranular pore space in the
coccoliths causes the data to plot below the less than 20-m field.

texture, whereas petrophysical classifications are


focused on contemporary rock fabrics that include
depositional and diagenetic textures. Thus, minor
modifications must be made in Dunhams classification before it can be applied to a petrophysical classification.
Instead of dividing fabrics into grain supported
and mud supported as in Dunhams classification,
fabrics are divided into grain dominated and mud
dominated (Figure 3). The important attributes of
grain-dominated fabrics are the presence of open
or occluded intergrain porosity and a grain-supported texture. The important attribute of mud-dominated fabrics is that the areas between the grains

are filled with mud, even if the grains appear to


form a supporting framework.
Grainstone is clearly a grain-dominated fabric,
but Dunhams (1962) packstone class bridges a
boundary between large intergrain pores in grainstone and small interparticle pores in wackestones
and mudstones. Some packstones have intergrain
pore space, and some have intergrain spaces filled
with mud; therefore, the packstone textural class
must be divided into two rock-fabric classes: graindominated packstones that have intergrain pore
space or cement, and mud-dominated packstones
that have intergrain spaces filled with mud (Figure
3). The pore-size distribution is controlled by the

Lucia

1000

1281

Permeability (md)

(1987) reported on a nonvuggy, grain-dominated


Wolfcampian packstone found in a core taken in
Oldam County, Texas. The grain-dominated packstone is described as a poorly sorted mixture of
100
150300-m grains in a matrix composed of 80-m
pellets and 10-m calcite crystals. A porosity-permeability crossplot of these data (Figure 7B) shows
10
that it plots on the boundary between the 10020m and the less than 20-m permeability fields.
Additional data points have been gleaned from the
1
literature, and they plot in the 10020-m permeability field.
Figure 7D is a crossplot between permeability
0.1
and porosity for North Sea coccolith chalk (Scholle,
20
30
40
10
1977). The average size of coccoliths is about 1
Interparticle porosity (%)
m. The data points plot below the less than 20-m
Ooid grainstone
permeability field. The presence of intrafossil pore
Grain-dominated packstones
space within the coccolith grains probably
Mud-dominated fabrics
accounts for the lower-than-expected permeability
Figure 8Composite porosity-air permeability crossplot in the high-porosity ranges.
Figure 8 illustrates all the data for limestones
for nonvuggy limestone fabrics compared with the
three permeability fields illustrated in Figure 5. Chalks compared with permeability fields. Although the
are not included because of the presence of intragrain data have considerable scatter, grainstone, grainpore space.
dominated packstone, and mud-dominated fabrics
are reasonably well constrained to the three permeability fields. Whereas grain size and sorting define
grain size in grain-dominated packstones and by the the permeability fields, the interparticle porosity
mud size in mud-dominated packstones.
defines the per meability within the field.
Systematic changes in intergrain porosity by
cementation, compaction, and dissolution will proPermeability/Rock-Fabric Relationships
duce systematic changes in the pore-size distribution. Therefore, the permeability field is defined by
Limestone Rock Fabrics
interparticle porosity, grain size, and sorting.
Examples of nonvuggy limestone petrophysical
rock fabrics are illustrated in Figure 6. In grainstone
fabrics, the pore-size distribution is controlled by Dolomite Rock Fabrics
Examples of nonvuggy dolomite petrophysical
grain size; in mud-dominated fabrics, the size of the
micrite particles controls the pore-size distribution. rock fabrics are illustrated in Figure 9. DolomitizIn grain-dominated packstones, however, the pore- ation can change the rock fabric significantly. In
size distribution is controlled by grain size and by limestones, the grain and mud fabrics usually can
the size of micrite particles between grains. Figure be distinguished with little difficulty. If the rock has
7A is a crossplot between permeability and inter- been dolomitized, however, the overprint of
grain porosity for grainstones. The data are from dolomite crystals commonly obscures the grain and
Choquette and Steiners (1985) work on the Ste. mud fabric. Grain and mud fabrics in fine crysGenevieve oolite (Mississippian). The average grain talline dolostones are easily recognizable; however,
size of the oolite is about 400 m. The points on as the crystal size increases, the precursor fabrics
this graph are concentrated within the greater than become more difficult to determine.
Dolomite crystals (defined as particles in this
100-m permeability field.
Figure 7C is a crossplot between air permeability classification) commonly range in size from several
and interparticle porosity from a Middle Eastern microns to more than 200 m. Micrite particles are
mud-dominated fabric containing microporosity. usually less than 20 m in size. Dolomitization of a
The data from this graph are abstracted from mud-dominated carbonate fabric therefore can
Moshier et al. (1988). The average crystal size of result in an increase in particle size from less than
the mud matrix is about 5 m (Moshier et al., 20 to more than 100 m (Figure 9). The crossplot
1988). The data are concentrated in the less than of interparticle porosity and permeability (Figure
10A) illustrates the principle that in mud-dominat20-m permeability field.
Because grain-dominated packstone is a new fab- ed fabrics, permeability increases as dolomite crysric class, data are difficult to find. Lucia and Conti tal size increases. Fine crystalline (average 15 m)

1282

Classification of Carbonate Pore Space

D
0

0.2

0.5
mm

1.0

0.5
mm

1.0

mm

E
0

0.5
mm

1.0

Figure 9Examples of nonvuggy dolomite fabrics. (A)


Dolograinstone, 15-m dolomite crystal size, = 16.4%,
k = 343 md, Dune field (Bebout et al., 1987). (B) Dolograinstone, 30-m dolomite crystal size, = 7.1%, k = 7.3
md, Seminole San Andres unit, west Texas. (C) Dolograinstone, crystal size 400 m, = 10.2%, k = 63 md,
Harmatton field, Alberta, Canada. (D) Grain-dominated
dolopackstone, 10-m dolomite crystal size, = 9%, k =
1 md, Farmer field, west Texas. (E) Grain-dominated
dolopackstone, 30-m dolomite crystal size, = 9.5%, k
= 1.9 md, Seminole San Andres unit, west Texas. (F)
Crystalline dolowackestone, 10-m dolomite crystal
size, = 11%, k = 0.12 md, Devonian, North Dakota. (G)
Crystalline dolowackestone, 80-m dolomite crystal
size, =16%, k = 30 md, Devonian, North Dakota. (H)
Crystalline dolowackestone, 150-m dolomite crystal
size, = 20%, k = 4000 md, Andrews South Devonian
field, west Texas.

C
0

1.0
mm

2.0

Lucia

1283

Figure 9Continued.

F
0

0.5
mm

1.0

0.5
mm

1.0

H
0

0.5
mm

1.0

mud-dominated dolostones from Farmer and


Taylor-Link (Lucia et al., 1992b) fields in the
Permian basin and from Lawrence (Bridgeport)
field in the Illinois basin (Choquette and Steiner,
1985) plot within the less than 20-m permeability
field. Medium crystalline (average 50 m) muddominated dolostones from the Dune field,
Permian basin (Bebout et al., 1987), plot within the
10020-m permeability field. Large crystalline
(average 150 m), mud-dominated dolostones from
Andrews South Devonian field, Permian basin
(Lucia, 1962), plot in the greater than 100-m permeability field.
Grainstones are commonly composed of grains
much larger than the dolomite crystal size (Figure
9), so that dolomitization does not have a significant effect on the pore-size distribution. This principle is illustrated in Figure 10B, a crossplot of
interparticle porosity and permeability measurements from dolomitized grainstones. The grain size
of the dolograinstones is 200 m. The fine crystalline dolograinstone from Taylor-Link field,
Permian basin, the medium crystalline dolograinstone from Dune field, Permian basin, and the
large crystalline dolograinstone from an outcrop
on the Algerita escarpment, New Mexico, all plot
within the greater than 100-m permeability
field. The large crystalline mud-dominated fabrics also plot in this permeability field, indicating
that they are petrophysically similar to grainstones (Figure 10A).
Interparticle porosity and permeability measurements from fine to medium crystalline, grain-dominated dolopackstones are crossplotted in Figure
10C. The samples are from the Seminole San
Andres unit and Dune (Grayburg) field (Bebout et
al., 1987), Permian basin. The data plot in the
10020-m permeability field. The medium crystalline mud-dominated dolostones also plot in this
field (Figure 10A).
Figure 11 illustrates all dolomite data compared
with permeability fields. Dolograinstones and large
crystal dolostones constitute the greater than 100m permeability field. Grains are very difficult to
recognize in dolostones with greater than 100-m
crystal size. However, because all large crystalline
dolomites and all grainstones are petrophysically
similar, whether the crystal size or the grain size is
described makes little difference petrophysically.
Fine and medium cr ystalline grain-dominated
dolopackstones and medium crystalline mud-dominated dolostones constitute the 10020-m permeability field. Fine crystalline mud-dominated dolostones constitute the less than 20-m field.

1284

Classification of Carbonate Pore Space

(B)

(A)

100

1000

Fine crystal size


Medium crystal size
Large crystal size

Permeability (md)

Permeability (md)

1000

Mud-dominated
dolostones

10

0.1
20

30

40

Interparticle porosity (%)

10

20

30

40

Interparticle porosity (%)

(C)

Permeability (md)

10

0.1
10

1000

100

Fine crystalline dolograinstone


Medium crystalline dolograinstone
Large crystalline dolograinstone

Fine to medium crystalline


grain-dominated
dolopackstone

100

10

0.1
10

20

30

40

Interparticle porosity (%)


Figure 10Porosity-air permeability crossplots of nonvuggy dolomite fabrics compared with the three permeability fields illustrated in Figure 5. (A) Mud-dominated dolostones with dolomite crystal sizes ranging from 10 to 150
m. (B) Dolograinstones (average grain size is 200 m) with dolomite crystal sizes ranging from 15 to 150 m. (C)
Grain-dominated dolopackstones with fine to medium dolomite crystal sizes.

The dolostone permeability fields are defined by


dolomite crystal size, as well as grain size and sorting of the precursor limestone. Within the field,
permeability is defined by interparticle porosity.
Systematic changes in intergrain and intercrystal
porosity by predolomite calcite cementation,
dolomite cementation, and compaction will systematically change the pore-size distribution.
Therefore, interparticle porosity defines the permeability, and dolomite crystal size, grain size, and
sorting define the permeability field.

than 100-m permeability field are (1) limestone and


dolomitized grainstones, and (2) large crystalline
grain-dominated dolopackstones and mud-dominated dolostones. These fabrics are called rock-fabric/petrophysical class 1. The effect of grain size in
this field can be seen by comparing Figures 7A and
10B. Ooid grainstones, which have a grain size of
400 m, are more permeable for a given porosity
than dolograinstones, which have a grain size of
200 m.
Fabrics that make up the 10020-m field are
(1) grain-dominated packstones, (2) fine to medium crystalline grain-dominated dolopackstones,
Limestone and Dolomite Comparison
and (3) medium crystalline mud-dominated doloData from limestone and dolomite rock fabrics are stones. These fabrics are called rock-fabric/petrocombined into one porosity-permeability crossplot physical class 2. A comparison of Figures 7B and
in Figure 12. The fabrics that make up the greater 10C shows that the dolomitized grain-dominated

Lucia

1285

Unusual Types of Interparticle Porosity


Diagenesis can produce unique types of interparticle porosity. Collapse of separate-vug fabrics
because of overburden pressure can produce fragments that are properly considered diagenetic particles. Large dolomite crystals with their centers
dissolved can collapse to form pockets of dolomite
rim crystals. Leached grainstones can collapse to
form intergrain fabrics composed of fragments of
dissolved grains. These unusual pore types typically
are not areally extensive; however, the collapse of
extensive cavern systems can produce areally
extensive bodies of collapse breccia. Interbrecciablock pores produced by cavern collapse are
included in the touching-vug category because of
their association with fractures and other vugs
formed by karsting processes (Kerans, 1989).
Capillary Pressure/Rock-Fabric Relationships
Figure 11Composite porosity-air permeability crossplot for nonvuggy dolostone fabrics compared with the
three permeability fields illustrated in Figure 6.

packstones tend to be more permeable for a givenporosity than limestone grain-dominated packstones. Further data may show this to be an important difference attributable to dolomite crystal
morphology. The less than 20-m permeability field
is characterized by mud-dominated limestone and
fine crystalline mud-dominated dolostones. These
fabrics are called rock-fabric/petrophysical class 3.
Reducedmajor-axis permeability transforms are
presented for each class (Figure 12). The transform
for class 2 is slightly skewed to the field boundaries, and the following transform is more compatible with the field boundaries.
Class 1:
k = (45.35 108) ip8.537

r = 0.71

k = (1.595 105) ip5.184

r = 0.80

Class 2:

or this recommended transform for class 2:


k = (2.040 106) ip6.38
Class 3:
k = (2.884 103) ip4.275

r = 0.81

where k = millidarcys and ip = fractional interparticle porosity.

Several methods have been presented for relating


porosity, permeability, water saturation, and reservoir height (Leverett, 1941; Aufricht and Koepf,
1957; Heseldin, 1974; Alger et al., 1989). These
methods attempt to average the capillary pressure
curves into one relationship among saturation,
porosity, permeability, and reservoir height without
regard to rock fabric. The data presented demonstrate that the three nonvuggy rock-fabric fields have
unique porosity-permeability relationships, suggesting that capillary properties of nonvuggy carbonates
should also be separated into rock-fabric categories.
To characterize the capillary properties of the
three rock-fabric fields, capillary pressure curves
with different interparticle porosities from the three
rock-fabric fields are compared in Figure 13. Figure
13A shows two curves representative of class 1.
Both curves are from samples of fine to medium
crystalline dolograinstones. The 9.2% porosity curve
represents the average of two sets of capillary pressure data from dolograinstones in the Taylor-Link
field (Lucia et al., 1992b), and the 17.6% porosity
curve represents the average of two data sets, one
from the Taylor-Link field and one from the Dune
field (Bebout et al., 1987). Three curves representative of class 3 are presented in Figure 13C. The curves
are from samples of fine crystalline dolowackestones
and from the Farmer field, Permian basin. Three
curves representative of class 2 are presented in
Figure 13B. Class 2 represents a very diverse class of
rock fabrics, and it is difficult to combine all the fabrics into a few simple curves. The curves presented
here are from medium crystalline dolowackestones
of the Seminole San Andres unit and may not be representative of all grain-dominated packstones.
Pore-throat-size distribution for each capillary
pressure curve was calculated using the following

1286

Classification of Carbonate Pore Space

Class 2

Class 1

10

Class 1
Class 2
Class 3

1000

20

0
m

Class 3

10

50

Permeability (md)

100

0.1
0.05

0.10

0.15

0.20

0.30

0.40

Interparticle porosity (fractional)


Figure 12Composite porosity-air permeability crossplot for nonvuggy limestones and dolostones showing statistical reducedmajor-axis transforms for each class (dashed lines). See text for equations.

formula. The results (Figure 13) show decreasing


pore-throat size with decreasing porosity within a
petrophysical class and a general decrease in porethroat size from class 1 to class 3. Microporosity
[pore-throat size less than 1 m (Pittman, 1971)] is
concentrated in class 3 and decreases in importance from class 3 to class 1.
Ri = (2 cos C)/Pc
where Ri = pore-throat radius (m); T = air-mercury
interfacial tension = 480 dynes/cm; = air-mercury
contact angle = 140; C = unit conversion constant
= 0.145; and Pc = mercury injection pressure (psia).
Each group of curves is characterized by similar
displacement pressures and a systematic change in
curve shape and saturation characteristics with
changes in interparticle porosity. The relationship
among porosity, saturation, and rock-fabric class
can be demonstrated by selecting a reservoir height
of 150 m (which equates to a mercury capillary

pressure of about 650 psia) and plotting saturation


against porosity for each rock-fabric class. The
results (Figure 14) show that in nonvuggy carbonate reservoirs, a plot of porosity vs. water saturation can separate rock fabrics into three classes.
Equations relating water saturation to porosity
and reservoir height are developed in two steps
(Lucia et al., 1992b). First, mercury capillary pressure is converted to reservoir height using generic
values (Table 2). Second, wetting-phase saturations
from the capillary pressure curves are plotted
against porosity for several reservoir heights. Third,
lines of equal reservoir height are drawn assuming
equal slopes and a relationship between intercepts
and reservoir height is developed. This relationship is
substituted for the intercept term in the porosity vs.
saturation equation, resulting in a relationship among
water saturation, porosity, and reservoir height.
The resulting equations follow, and threedimensional representations for classes 1 and 3
are presented in Figure 15. These equations are

(A')

2000

100

Mercury saturation (%)

Mercury injection pressure (psia)

(A)

1600
9.2% porosity

1200
800

17.60% porosity
400
0
100

17.6% porosity
Permeability = 166 md

80
60
9.2% porosity
Permeability = 72 md

40
20
0

80

60

40

20

2.0

1.5

Mercury saturation (%)

-0.5

-1.0

-1.5

(B')
100

2000

13.7% porosity
Permeability = 5.1md

8.7% porosity
1600

Mercury saturation (%)

Mercury injection pressure (psia)

0.5

Log pore-throat radius (m)

(B)

10.3% porosity
12.5% porosity

1200

13.7% porosity
800
400
0
100

80
12.5% porosity
Permeability = 2.3 md

60

10.3% porosity
Permeability = 1.3 md

40
20

8.7% porosity
Permeability = 0.05 md

0
80

60

40

20

2.0

1.5

1.0

0.5

-0.5

-1.0

-1.5

-1.0

-1.5

Log pore-throat radius (m)

Mercury saturation (%)

(C')

(C)
2000

100
9.0% porosity

1600

Mercury saturation (%)

Mercury injection pressure (psia)

1.0

11.8% porosity

1200

15.4% porosity
800
400
0
100

15.4% porosity
Permeability = 2 md

80

11.8% porosity
Permeability = 0.4 md

60
40

9.0% porosity
Permeability = 0.04 md
20
0

80

60

40

Mercury saturation (%)

20

2.0

1.5

1.0

0.5

-0.5

Log pore-throat radius (m)

Figure 13(AC) Capillary pressure curves and (AC) pore-size distribution for petrophysical classes. (A) Class 1.
Data are from dolograinstones. (B) Class 2. Data are from medium crystalline dolowackestones. (C) Class 3. Data are
from fine crystalline dolowackestones.

1288

Classification of Carbonate Pore Space

Water saturation (%)

100

Table 2. Values Used to Convert Mercury/Air Capillary


Pressure to Reservoir Height

Laboratory
(mercury/air/solid)

Class 3
Class 1

10

Class 2

T (dynes/cm)
()
Water density

1
5

10

20

30

40

Figure 14Crossplot of porosity and water saturation


for the three rock-fabric/petrophysical classes at a reservoir height of 150 m. Water saturation (1 minus mercury saturation) and porosity values are taken from capillary pressure curves illustrated in Figure 13.

specific to the capillary pressure curves used in


this paper and will not necessarily apply to specific
reservoirs. However, the equations will provide reasonable values for original water saturations when
only porosity and rock-fabric data are available.

Reservoir
(oil/water/solid)

480
140
1.04

28
44
0.88

Class 2 comprises grain-dominated packstones,


fine and medium cr ystalline grain-dominated
dolopackstones, and medium crystalline mud-dominated dolostones:
k = (2.040 106) ip6.38
Sw = 0.1404 H0.407 1.440
Class 3 contains mud-dominated limestones and
fine crystalline mud-dominated dolostones:
k = (2.884 103) ip4.275
Sw = 0.6110 H0.505 1.210
Petrophysics of Separate-Vug Pore Space

Class 1:
Sw = 0.02219 H0.316 1.745
Class 2:
Sw = 0.1404 H0.407 1.440
Class 3:
Sw = 0.6110 H0.505 1.210
where H = height above capillary pressure equal to
zero and = fractional porosity.
Rock-Fabric/Petrophysical Classes
Because the three rock-fabric groups define permeability and saturation fields, the three groups,
together with interparticle porosity and reservoir
height, can be used to relate petrophysical properties to geologic observations. These rock-fabric
groups, herein termed rock-fabric/petrophysical
classes (Figure 16), are described, with their generic transform equations as follows:
Class 1 is composed of grainstones, dolograinstones, and large crystalline dolostones:
k = (45.35 108) ip8.537
Sw = 0.02219 H0.316 1.745

The addition of vuggy pore space to interparticle pore space alters the petrophysical characteristics by altering the manner in which the pore
space is connected, all pore space being connected in some fashion. Separate-vug pore space is
defined as pore space that is (1) either within particles or is significantly larger than the particle
size (generally greater than 2), and (2) is interconnected only through the interparticle porosity
(Figure 17). Separate vugs (Figure 18) are typically
fabric selective in origin. Intrafossil pore space,
such as the living chambers of a gastropod shell;
moldic pore space, such as dissolved grains
(oomolds) or dolomite crystals (dolomolds); and
intragranular microporosity are examples of intraparticle, fabric-selective separate vugs. Molds of
evaporite crystals and fossils found in mud-dominated fabrics are examples of fabric-selective separate vugs that are significantly larger than the particle size (Figure 18). In mud-dominated fabrics,
shelter pore space is typically much larger than
the particle size and is classified as separate-vug
porosity.
In grain-dominated fabrics, extensive selective
leaching of grains may cause grain boundaries to
dissolve, producing composite molds. These composite molds may have the petrophysical characteristics of separate vugs. If, however, dissolution of

Lucia

1289

Figure 15Three-dimensional
displays of (A) class 1 and (B)
class 3 equations relating
water saturation to reservoir
height and porosity.

Wa

ter

Wa

sat

ter

ura

sat

tion

ura

tion

(%

ity

(%

ros

(%

the grain boundaries is extensive, the pore space


may be classified as interparticle porosity.
Grain-dominated fabrics may contain grains with
intragranular microporosity (Pittman, 1971; Keith
and Pittman, 1983; Asquith, 1986). Intragranular
microporosity is classified as a separate vug
because it is within the particles of the rock and is
interconnected only through the intergrain pore
network. Mud-dominated fabrics also may contain
grains with microporosity, but they present no
unique petrophysical condition because of the similar pore sizes between the microporosity in the
mud matrix and in the grains.

Po

ity

ros

Po

(%

The addition of separate-vug porosity to interparticle porosity increases total porosity, but does not
significantly increase permeability (Lucia, 1983).
Figure 19A illustrates this principle in that permeability of a moldic grainstone is less than would be
expected if all of the porosity were interparticle
and, at constant porosity, permeability were to
increase with decreasing separate-vug porosity
(Lucia and Conti, 1987). This principle is also true
for intragranular microporosity. Figure 19B is a
crossplot of data from a San Andres dolograinstone
from the Permian of west Texas. The crossplot
shows that the permeability of the grainstone with

1290

Classification of Carbonate Pore Space

Figure 16Petrophysical and rock-fabric classes based on similar capillary properties and interparticle-porosity/
permeability transforms.

intragranular microporosity is less than would be


expected if all the porosity were interparticle.
Separate vugs found within the oil column are
usually considered to be saturated with oil because
of their relatively large size; oil migration into separate vugs is controlled by interparticle pore size.
Intragranular microporosity, however, is unique
because of small pore size. The small pore size produces high capillary forces that trap water and
lead to anomalously high water saturations within
a productive interval (Pittman, 1971; Dixon and
Marek, 1990). Figure 20 shows the capillary characteristics of a Cretaceous grainstone composed of
grains having microporosity between 2-m crystals
(Keith and Pittman, 1983) and a San Andres dolograinstone with grains having microporosity
between 10-m dolomite crystals. This figure illustrates that within the oil column, water saturations
in grainstones with intragranular microporosity are
significantly higher than would be expected if no
intragranular microporosity were present. The

capillary pressure curves typically have a bimodal


character.
Petrophysics of Touching-Vug Pore Space
Touching-vug pore systems are defined as pore
space that is (1) significantly larger than the particle size, and (2) forms an interconnected pore system of significant extent (Figures 17, 18). Touching
vugs are typically nonfabric selective in origin.
Cavernous, breccia, fracture, and solution-enlarged
fracture pore types commonly form an interconnected pore system on a reservoir scale and are typical touching-vug pore types (Figure 18). Fenestral
pore space is commonly connected on a reservoir
scale and is grouped with touching vugs because
the pores are normally much larger than the grain
size (Major et al., 1990).
Fracture porosity is included as a touching-vug
pore type because fracture porosity is an important

Lucia

1291

Figure 17Geological and petrophysical classification of vuggy pore space based on vug interconnection. The volume of separate-vug pore space is important for characterizing the petrophysical properties.

contributor to permeability in many carbonate


reservoirs and, therefore, must be included in any
petrophysical classification of pore space. Although fracturing is often considered to be of
tectonic origin and thus not a part of carbonate
geology, diagenetic processes common to carbonate reser voir s, such as kar sting (Kerans,
1989), can produce extensive fracture porosity.
The focus of this classification is on petrophysical properties rather than genesis, and must

include fracture porosity as a pore type irrespective of its origin.


Touching vugs are thought to be typically filled
with oil in the reservoirs and can increase permeability well above what would be expected from
the interparticle pore system. Lucia (1983) illustrated this fact by comparing a plot of fracture permeability vs. fracture porosity to the three porosity-permeability fields for interparticle porosity
(Figure 21). This graph shows that permeability in

B
0

0.5
mm

1.0

1.0
mm

2.0

1.0
mm

2.0

F
0

0.2
mm

1.0
mm

2.0

20 microns

G
1 in.

1 in.

J
1 in.

3 in.

K
0

1.0
mm

2.0

Figure 18Photomicrographs
and photographs showing
examples of vug pore types.
Separate-vug types: (A)
oomoldic porosity, = 26%,
k = 3 md, Wolfcampian, west
Texas; (B) intrafossil pore
space in a gastropod shell,
Cretaceous, Gulf Coast; (C)
fossil molds in wackestone,
= 5%, k = 0.05 md;
(D) anhydrite molds in
grain-dominated packstone,
= 10%, k = <0.1 md,
Mississippian, Montana;
(E) fine crystalline
dolograinstone with
intergranular and
intragranular microporosity
pore types, = 10%, k = 3 md,
Farmer field, west Texas;
(F) scanning electron
photomicrograph of
dolograins in (E) showing
intragranular microporosity
between 10-m crystals.
Touching-vug types: (G)
cavernous porosity in a
Niagaran reef, northern
Michigan; (H) collapse breccia,
Ellenburger, west Texas;
(I) solution-enlarged fractures,
Ellenburger, west Texas;
(J) cavernous porosity in
Miami oolite, Florida;
(K) fenestral porosity in
pisolitic dolostone. Note that
the fenestral pores are more
than twice the size of the
enclosing grains.

1294

Classification of Carbonate Pore Space

(A)

(B)

1000

1000

Permeability (md)

Permeability (md)

Grainstone
field
100

10
Svug
porosity
average
8%

Svug
porosity
average
20%

0.1

Grainstone
field

100

10

Dolograinstone
with intragrain
microporosity

0.1
5

10

20

30

40

Porosity (%)

10

20

30

40

Porosity (%)

Figure 19Crossplot illustrating the effect of separate-vug porosity on air permeability. (A) Grainstones with separate-vug (Svug) porosity in the form of grain molds plot to the right of the grainstone field in proportion to the volume of separate-vug porosity. (B) Dolograinstones with separate vugs in the form of intragranular microporosity
plot to the right of the grainstone field.

touching-vug pore systems is related principally to


fracture width and is sensitive to extremely small
changes in fracture porosity. There is no effective
method for measuring fracture width using the
rock-fabric approach.
IMPLICATIONS FOR CONSTRUCTING A
GEOLOGIC MODEL
A fundamental problem in constructing a reservoir model for input into numerical fluid-flow simulators is converting geologic observations into petrophysical rock properties, namely porosity, permeability,
and saturation. The problem has been discussed by
Lucia et al. (1992a), Senger et al. (1993), and Kerans
et al. (1994). These workers suggested that the
petrophysical properties within a rock-fabric unit
are nearly randomly distributed and thus form the
basic building blocks for reservoir models. Therefore, geologic models described in terms of rock-fabric units can be most effectively converted into
petrophysical parameters and into reservoir models.
On the basis of data presented herein, the most
important rock-fabric characteristics to describe and
map in nontouching-vug reservoirs are (1) grain size
and sorting using the modified Dunham classification, (2) dolomite crystal size using 20 and 100 m
as size boundaries, (3) separate-vug porosity, (4) separate-vug type with special attention to intergrain
microporosity, and (5) total porosity.
The stacking patterns of rock-fabric units are
systematic within a depositional and diagenetic

environment (Lucia, 1993). The simplest example


is a shoaling-upward bar complex. Mud-dominated
fabrics grade upward and laterally into grain-dominated fabrics with a corresponding increase in permeability, decrease in water saturation, and little
change in porosity (Figure 22). The permeability
and water saturation are a function of total porosity and rock fabric. The introduction of moldic
porosity (separate vugs) by selective dissolution of
grains in the grain-dominated fabric, caused perhaps by shoaling and the introduction of meteoric
water at a subaerial exposure surface, can produce
a diagenetic overprint that results in no significant
change in porosity, little change in water saturation assuming the molds are filled with hydrocarbons, and drastically reduced permeability in the
grainstone (Figure 22). The permeability is a function of interparticle porosity determined by subtracting separate-vug porosity from total porosity.
Conversion of the shoaling-upward bar complex
from limestone to 50-m dolostone will not change
the grain size and sorting characteristics of the
grain-dominated fabrics, but will alter the particle
size characteristics of the mud-dominated fabrics,
changing the petrophysical class from class 3 to
class 2. Dolomitization may reduce the total porosity of the grain-dominated fabrics, but the permeability and water saturation will still be a function
of grain size, sorting, and intergrain porosity. If the
precursor limestone is a moldic grain-dominated
fabric, the resulting grain-dominated dolostone will
be moldic (Figure 22). Dolomitization of the muddominated fabric to a 50-m dolomite may reduce

Lucia

(A')
2000

1500

100
Unimodal pore system
11.2%, 44.9 md

Bimodal pore system


16.4%, 5.4 md

Mercury Saturation (%)

Mercury injection pressure (psia)

(A)

1000

500

80

80

60

40

20

Unimodal pore system


11.2%, 44.9 md

60
40

Bimodal pore system


16.4 %, 5.4 md

20
0
1.5

0
100

1.0

0.5

-0.5

-1.0

-1.5

Log pore-throat radius (m)

Mercury saturation (%)

(B)

(B')
2000

100
Unimodal pore system
17.6%, 166 md

Mercury Saturation (%)

Mercury capillary pressure (psia)

1295

1500
Bimodal pore system
18.2%, 14.0 md
1000

Unimodal pore system


17.6%, 166 md

500

0
100

80
60
40
Bimodal pore system
18.2%, 14.0 md

20
0

80

60

40

20

Mercury saturation (%)

2.0

1.5

1.0

0.5

-0.5

-1.0

-1.5

Log pore-throat radius (m)

Figure 20(A, B) Capillary pressure curves and (A, B) pore-size distribution illustrating the effect of intragranular
microporosity on capillary properties. Notice the bimodal character of the samples with both intergranular and
intragranular microporosity pore types. (A) Ooid grainstone with intergranular porosity compared with ooid grainstone with both intergranular and intragranular microporosity pore types, Rodessa limestone, Cretaceous, east
Texas (after Keith and Pittman, 1983). (B) Dolograinstone with intergranular porosity compared with dolograinstone with both intergranular porosity and intragranular microporosity pore types, San Andres dolomite, Permian
(Farmer San Andres field), west Texas.

the porosity, but will increase the permeability and


decrease the water saturation significantly.
Another simple example is the high-energy tidalflatcapped cycle where intertidal and supratidal
sediments overlie a shoreface grain-dominated
packstone (Figure 23). Carbonate cement and compaction have reduced porosity in the tidal-flat sediments, resulting in a low-permeability, highwatersaturation f low unit overlying a permeable
grain-dominated limestone and a low-permeability
mud-dominated limestone. In a seaward direction,
shoaling-upward bar complexes replace the tidalflatcapped cycles.

Evaporitic tidal-flat environments are a source of


dolomitizing water, and downward f low of the
hypersaline water can dolomitize underlying carbonate sediments (Figure 23). In Permian basin
reservoirs, intertidal and supratidal dolostones are
dense because of compaction and occlusion of
pore space by anhydrite and dolomite, and graindominated dolostones underlying the tidal-flat sediments commonly are dense because of pore-filling
anhydrite. High porosity is confined to subtidal
mud-dominated dolostones, and permeability and
water saturation are related to intercrystal porosity
and dolomite crystal size. In this example (Figure

Classification of Carbonate Pore Space

103

10

1
0 cm

Permeability (d)

102

0
0.

cm

SUMMARY

Interparticle

>100m

Z =
100

10-1

porosity rather than total porosity less separate-vug


porosity, and permeability and saturation can be
seriously underestimated if the changes in particle
size resulting from dolomitization are not included
in the geologic model.

cm cm
5
1
0.
=
=
w
w

w
Z =
=
Z =
w
100
0
1 c
cm
m = 0 .0 Z = 1
0 cm
.0 1 c
05
m
w
cm
=
0.

104

kf= 84.4 x 10 5 W 3 /Z
f= W/Z x 100
z = Fracture spacing
w = Fracture width
Kf= Fracture permeability
f= Fracture porosity

cm

105

1296

100 to 20m
<20m

0.

00

cm

10-2

10-3
10-4
10-5

10-4

10-3

10-2

10 -1

10

100

Porosity (%)
Figure 21Theoretical fracture air permeability/porosity relationship compared to the rock-fabric/petrophysical porosity, permeability fields (Lucia, 1983).

23), dolomite crystal size increases from fine to


medium with depth and the petrophysical properties change correspondingly.
Grain-dominated bodies seaward of the tidal-flat
deposits are dolomitized by hypersaline water from
an overlying cycle and are not cemented by anhydrite. The permeability and saturation characteristics are related to grain size, sorting, and intergrain
porosity (Figure 23). The underlying mud-dominated dolostone is composed of medium dolomite
crystals and has higher permeability than a muddominated limestone with the same interparticle
porosity. The underlying limestone has low porosity and permeability in this example.
These examples illustrate the importance of
mapping rock-fabric units within the geologic
framework. If the shoaling-upward cycle is mapped
as a bar facies or the tidal-f latcapped cycle is
mapped as a peritidal facies, the permeability and
saturation structure is compromised and the resulting model is oversimplified. In addition, if the diagenetic facies are not included as mapping parameters, the permeability and saturation values in the
resulting reservoir model may be in serious error.
For example, permeability can be significantly
overestimated if the estimate is based on total

The goal of reservoir characterization is to


describe the spatial distribution of petrophysical
parameters, such as porosity, permeability, and saturation. The rock-fabric approach presented here is
based on the premise that pore-size distribution
controls the engineering parameters of permeability and saturation, and that pore-size distribution is
related to rock fabric, which is a product of geologic processes. Thus, rock fabric integrates geologic
interpretation with engineering numerical measurements.
To determine the relationships between rock
fabric and petrophysical parameters, one must
define and classify pore space as it exists today in
terms of petrophysical properties. This is best
accomplished by dividing pore space into pore
space located between grains or crystals, called
interparticle porosity, and all other pore space,
called vuggy porosity. Vuggy pore space is further
subdivided into two groups, depending on how the
vugs are interconnected: (1) vugs interconnected
only through the interparticle pore network are
termed separate vugs, and (2) vugs in direct contact are termed touching vugs.
The petrophysical properties of interparticle
porosity are related to particle size, sorting, and
interparticle porosity. Grain size and sorting of grains
and micrite are based on Dunhams (1962) classification and are modified to make the classification compatible with petrophysical considerations. Instead of
dividing fabrics into grain-supported and mud-supported categories, fabrics are divided into grain-dominated and mud-dominated categories. The important attributes of grain-dominated fabrics are the
presence of open or occluded intergrain porosity
and a grain-supported texture. The important
attribute of mud-dominated fabrics is that the areas
between the grains are filled with mud even if the
grains appear to form a supporting framework.
Grainstone is clearly a grain-dominated fabric,
but Dunhams (1962) packstone class bridges an
important petrophysical boundary. Some packstones, as we see them now, have intergrain pore
space, and some have the intergrain spaces filled
with mud. Therefore, the packstone textural class
must be divided into two rock-fabric classes: (1)
grain-dominated packstones having intergrain pore
space or cement, and (2) mud-dominated packstones having intergrain spaces filled with mud.

Lucia

1297

Figure 22Three examples of how the stacking of rock-fabric units affects the distribution of porosity, permeability, and water saturation in a shoaling-upward sequence with selective dissolution and dolomitization overprints.

1298

Classification of Carbonate Pore Space

Figure 23Examples of how the stacking of rock-fabric units affects the distribution of porosity, permeability, and
water saturation in a tidal-flatcapped sequence with hypersaline reflux dolomitization and sulfate emplacement
overprints.

Lucia

The important fabric elements to recognize for


petrophysical classification of dolomites are precursor grain size and sorting, dolomite crystal size, and
interparticle (intergrain and intercrystal) porosity.
Important dolomite crystal size boundaries are 20
and 100 m. Dolomite crystal size has little effect
on the petrophysical properties of grain-dominated
dolostones. The petrophysical properties of muddominated fabrics, however, are significantly
improved when the dolomite crystal size is less
than 20 m.
Permeability and saturation characteristics of
interparticle porosity can be grouped into three
rock-fabric/petrophysical classes. Class 1 is composed of limestone and dolomitized grainstones,
and large crystalline grain-dominated dolopackstones and mud-dominated dolostones. Class 2 is
composed of grain-dominated packstones, fine to
medium crystalline grain-dominated dolopackstones, and medium crystalline mud-dominated
dolostones. Class 3 is composed of mud-dominated
limestone and fine crystalline mud-dominated dolostones.
Generic porosity-permeability transforms and
water saturation, porosity, reservoir-height equations for each rock-fabric/petrophysical class are
presented in the section Rock-Fabric/Petrophysical Classes.
The addition of separate-vug porosity to interparticle porosity increases total porosity, but does
not significantly increase permeability; therefore,
one must determine interparticle porosity by subtracting separate-vug porosity from total porosity
and use interparticle porosity to estimate permeability. Separate-vug porosity is thought to be typically filled with hydrocarbons in the reservoir.
Intergranular microporosity, however, may contain
significant amounts of capillar y-bound water,
resulting in water-free production of hydrocarbons
from intervals with higher than expected water
saturation.
Touching-vug pore systems cannot be related to
porosity, but are related principally to fracture
width. Because there is no effective method for
making this observation in the reservoir, the rockfabric approach cannot be used to characterize
touching-vug reservoirs.
The key to constructing a geologic model that
can be quantified in petrophysical terms is to select
facies or units that have unique petrophysical qualities for mapping. Petrophysical properties in rockfabric facies or units are nearly randomly distributed
and can be legitimately averaged, making these geologic units ideal for petrophysical quantification. In
nontouching-vug reservoirs, the most important
rock-fabric characteristics to describe and map are
(1) grain size and sorting using the modified
Dunham classification, (2) dolomite crystal size

1299

using 20 and 100 m as size boundaries, (3) separate-vug type with special attention to intergrain
microporosity, (4) total porosity, and (5) separatevug porosity.
In touching-vug reservoirs, characterizing the
pore system is difficult because the pore system is
not related to a precursor depositional fabric, but is
typically wholly diagenetic. Although the pore system may conform to bedding, as in evaporite collapse brecciation, it more commonly cuts across
stratal boundaries. Recognizing the presence of a
touching-vug pore system is paramount, however,
because it may dominate the flow characteristics of
the reservoir.
REFERENCES CITED
Alger, R. P., D. L. Luffel, and R. B. Truman, 1989, New unified
method of integrating core capillary pressure data with well
logs: Society of Petroleum Formation Evaluation, v. 4, no. 2,
p. 145152.
Archie, G. E., 1952, Classification of carbonate reservoir rocks
and petrophysical considerations: AAPG Bulletin, v. 36, no. 2,
p. 278298.
Asquith, G. B., 1986, Microporosity in the OHara oolite zone of
the Mississippian Ste. Genevieve Limestone, Hopkins County,
Kentucky, and its implications for formation evaluation:
Carbonates and Evaporites, v. 1, no. 1, p. 712.
Aufricht, W. R., and E. H. Koepf, 1957, The interpretation of capillary pressure data from carbonate reservoirs: Transactions of
the American Institute of Mining, Metallurgical, and Petroleum
Engineers, v. 210, p. 402405.
Bebout, D. G., F. J. Lucia, C. F. Hocott, G. E. Fogg, and G. W.
Vander Stoep, 1987, Characterization of the Grayburg reservoir, University Lands Dune field, Crane County, Texas:
University of Texas at Austin, Bureau of Economic Geology
Report of Investigations No. 168, 98 p.
Choquette, P. W., and L. C. Pray, 1970, Geologic nomenclature
and classification of porosity in sedimentary carbonates: AAPG
Bulletin, v. 54, no. 2, p. 207250.
Choquette, P. W., and R. P. Steiner, 1985, Mississippian oolite and
non-supratidal dolomite reservoirs in the Ste. Genevieve
Formation, North Bridgeport field, Illinois basin, in P. O. Roehl
and P. W. Choquette, eds., Carbonate petroleum reservoirs:
New York, Springer-Verlag, p. 209238.
Dixon, F. R., and B. F. Marek, 1990, The effect of bimodal pore
size distribution on electrical properties of some Middle
Eastern limestones: Society of Petroleum Engineers Technical
Conference, SPE 20601, p. 743750.
Dunham, R. J., 1962, Classification of carbonate rocks according
to depositional texture, in W. E. Ham, ed., Classifications of
carbonate rocksa symposium: AAPG Memoir 1, p. 108121.
Heseldin, G. M., 1974, A method of averaging capillary pressure
curves: Society of Professional Well Log Analysts Annual
Logging Symposium, paper E, p. 17.
Keith, B. D., and E. D. Pittman, 1983, Bimodal porosity in oolitic
reservoireffect on productivity and log response, Rodessa
limestone (Lower Cretaceous), East Texas basin: AAPG
Bulletin, v. 67, no. 9, p. 13911399.
Kerans, C., 1989, Karst-controlled reservoir heterogeneity in the
Ellenburger Group carbonates of west Texas: AAPG Bulletin,
v. 72, no. 10, p. 11601183.
Kerans, C., F. J. Lucia, and R. K. Senger, 1994, Integrated characterization of carbonate ramp reservoirs using Permian San
Andres Formation outcrop analogs: AAPG Bulletin, v. 78, no. 2,
p. 181216.
Leverett, M. C., 1941, Capillary behavior in porous solids:

1300

Classification of Carbonate Pore Space

Transactions of the American Institute of Mining, Metallurgical,


and Petroleum Engineers, v. 142, p. 151169.
Lucia, F. J., 1962, Diagenesis of a crinoidal sediment: Journal of
Sedimentary Petrology, v. 32, no. 4, p. 848865.
Lucia, F. J., 1983, Petrophysical parameters estimated from visual
description of carbonate rocks: a field classification of carbonate pore space: Journal of Petroleum Technology, March,
v. 35, p. 626637.
Lucia, F. J., 1993, Carbonate reservoir models: facies, diagenesis,
and flow characterization, in D. Morton-Thompson and A. M.
Woods, eds., Development geology reference manual: AAPG
Methods in Exploration 10, p. 269274.
Lucia, F. J., and R. D. Conti, 1987, Rock fabric, permeability, and
log relationships in an upward-shoaling, vuggy carbonate
sequence: University of Texas at Austin, Bureau of Economic
Geology Geological Circular 87-5, 22 p.
Lucia, F. J., C. Kerans, and R. K. Senger, 1992a, Defining flow units
in dolomitized carbonate-ramp reservoirs: Society of Petroleum
Engineers, SPE 24702, p. 399406.
Lucia, F. J., C. Kerans, and G. W. Vander Stoep, 1992b,
Characterization of a karsted, high-energy, ramp-margin carbonate reservoir: Taylor-Link West San Andres unit, Pecos
County, Texas: University of Texas at Austin, Bureau of
Economic Geology Report of Investigations No. 208, 46 p.

ABOUT THE AUTHOR


F. Jerry Lucia
F. Jerry Lucia retired from Shell
Oil Company in 1985, where he
had worked as a consulting geological engineer, after 31 years in
research and operations. Currently,
he is a senior research fellow with
the University of Texas at Austin,
Bureau of Economic Geology,
developing new techniques and
methods of characterizing carbonate reservoirs by applying outcrop
studies to subsurface reservoirs.

Major, R. P., G. W. Vander Stoep, and M. H. Holtz, 1990,


Delineation of unrecovered mobile oil in a mature dolomite
reservoir: East Penwell San Andres unit, University Lands, west
Texas: University of Texas at Austin Bureau of Economic
Geology Report of Investigations No. 194, 52 p.
Moshier, S. O., C. R. Handford, R. W. Scott, and R. D. Boutell,
1988, Giant gas accumulation in chalky-textured micritic limestones, Lower Cretaceous Shuaiba Formation, eastern United Arab Emirates, in A. J. Lomando and P. M.
Harris, eds., Giant oil and gas fields: Society of Economic
Paleontologists and Mineralogists Core Workshop 12, v. 1,
p. 229272.
Murray, R. C., 1960, Origin of porosity in carbonate rocks: Journal
of Sedimentary Petrology, v. 30, no. 1, p. 5984.
Pittman, E. D., 1971, Microporosity in carbonate rocks: AAPG
Bulletin, v. 55, no. 10, p. 18731881.
Scholle, P. A., 1977, Chalk diagenesis and its relation to petroleum
exploration: oil from chalks, a modern miracle?: AAPG Bulletin,
v. 61, no. 7, p. 9821009.
Senger, R. K., F. J. Lucia, C. Kerans, and M. A. Ferris, 1993,
Dominant control of reservoir-flow behavior in carbonate
reservoirs as determined from outcrop studies, in B. Linville,
R. E. Burchfield, and T. C. Wesson, eds., Reservoir characterization III: Tulsa, Oklahoma, PennWell, p. 107150.

Anda mungkin juga menyukai