Está en la página 1de 26

Rock-Fabric/Petrophysical Classification of Carbonate

Pore Space for Reservoir Characterization1

F. Jerry Lucia2

ABSTRACT should be subtracted from total porosity to obtain


interparticle porosity for permeability estimation.
This paper defines the important geologic param- Separate-vug pore space is generally considered to
eters that can be described and mapped to allow be hydrocarbon filled in reservoirs; however, intra-
accurate petrophysical quantification of carbonate granular microporosity is composed of small pore
geologic models. All pore space is divided into inter- sizes and may contain capillary-held connate water
particle (intergrain and intercrystal) and vuggy within the reservoir. Touching vugs are nonfabric
pores. In nonvuggy carbonate rocks, permeability selective and form an interconnected pore system
and capillary properties can be described in terms independent of the interparticle system.
of particle size, sorting, and interparticle porosity
(total porosity minus vuggy porosity). Particle size
and sorting in limestones can be described using a INTRODUCTION
modified Dunham approach, classifying packstone
as grain dominated or mud dominated, depending The goal of reservoir characterization is to
on the presence or absence of intergrain pore describe the spatial distribution of petrophysical
space. To describe particle size and sorting in dolo- parameters, such as porosity, permeability, and sat-
stones, dolomite crystal size must be added to the uration. Wireline logs, core analyses, production
modified Dunham terminology. Larger dolomite data, pressure-buildup data, and tracer tests pro-
crystal size improves petrophysical properties in vide quantitative measurements of petrophysical
mud-dominated fabrics, whereas variations in parameters in the vicinity of the well bore. These
dolomite crystal size have little effect on the petro- well bore data must be integrated with a geologic
physical properties of grain-dominated fabrics. model to display the petrophysical properties in
A description of vuggy pore space that relates to three-dimensional space. Studies that relate rock
petrophysical properties must be added to the fabric to pore-size distribution, and thus to petro-
description of interparticle pore space to complete physical properties, are key to quantifying geologic
the petrophysical characterization. Vuggy pore models in numerical terms for input into computer
space is divided into separate vugs and touching simulators (Figure 1).
vugs on the basis of vug interconnection. Separate Geologic models are generally based on observa-
vugs are fabric selective and are connected only tions interpreted in terms of depositional environ-
through the interparticle pore network. Separate- ments and sequences. In the subsurface, cores and
vug porosity contributes little to permeability and wireline logs are the main source of data for these
interpretations. Engineering models are based on
wireline log calculations and average rock proper-
©Copyright 1995. The American Association of Petroleum Geologists. All ties from core analyses. Numerical engineering data
rights reserved. and interpretive geologic data are joined at the
1Manuscript received June 27, 1994; revised manuscript received March
21, 1995; final acceptance May 1, 1995.
rock-fabric level because the pore structure is fun-
2Bureau of Economic Geology, University of Texas at Austin, Austin, damental to petrophysical properties and the pore
Texas 78713-7508. structure is the result of spatially distributed depo-
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
sitional and diagenetic processes.
Economic Geology’s Reservoir Characterization Research Laboratory, which The purpose of this paper is to define the impor-
is funded by industry sponsors Agip, Amoco, ARCO, British Petroleum, tant geologic parameters that can be described and
Chevron, Conoco, Exxon, Fina, JNOC, Marathon, Mobil, Phillips, Shell,
Texaco, Total, and Unocal. I am grateful for intense discussions with mapped to allow accurate petrophysical quantifica-
colleagues Charles Kerans, Gary Vander Stoep, Fred Wang, Susan Hovorka, tion of carbonate geologic models by (1) describ-
Mark Holtz, Steve Ruppel, and Don Bebout. An in-depth review of the ing the relationship between carbonate rock fab-
manuscript by Philip W. Choquette was especially useful. Published with
permission of the director, Bureau of Economic Geology, University of Texas rics and petrophysical properties, (2) presenting a
at Austin. generic petrophysical classification of carbonate

AAPG Bulletin, V. 79, No. 9 (September 1995), P. 1275–1300. 1275


1276 Classification of Carbonate Pore Space

Table 1. Pore-type Terminology and Abbreviations Used in


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

Abbreviations
Sedimentation Wireline Logs Choquette and
Term Lucia (1983) Pray (1970)
Core Analysis
Porosity Interparticle IP BP
Diagenesis Permeability Production Intergrain IG –
Saturation Intercrystal IX BC
Pressure Vug VUG VUG
Separate vug SV –
Tectonics Tracer Tests Moldic MO MO
Intraparticle WP WP
Intragrain WG –
Figure 1—Integration of spatial geologic data with Intracrystal WX –
numerical engineering data through rock-fabric studies. Intrafossil WF –
Intragrain
microporosity µG –
Shelter SH SH
Touching Vug TV –
pore space, and (3) suggesting that rock-fabric Fracture FR FR
units are systematically stacked within high-fre- Solution-enlarged
quency cycles. fracture SF CH*
Cavernous CV CV
Breccia BR BR
Fenestral FE FE
PORE-SPACE TERMINOLOGY AND
CLASSIFICATION *Channel.

Pore space must be defined and classified in


terms of rock fabrics and petrophysical properties fabrics to petrophysical rock properties in carbon-
to integrate geological and engineering information. ate rocks. The Archie classification focuses on esti-
Archie (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
CAVERNOUS matrix porosity. Therefore, pore space is divided
INTERGRAIN MOLDIC
INTERCRYSTAL INTRAFOSSIL FRACTURE into matrix and visible porosity (Figure 2). Chalky
SHELTER SOLUTION-ENLARGED
FRACTURE
texture indicates a matrix porosity of about 15%,
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 accord-
MATRIX ing to pore size, and ranges from no visible pore
VISIBLE (A,B,C, and D)
space to larger than cutting size. Porosity and per-
meability trends and capillary pressure characteris-
tics are also related to these textures.
LUCIA (1983)
Although the Archie method is still useful for esti-
VUGGY
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
FABRIC SELECTIVE NONFABRIC provision is made for distinguishing between visible
SELECTIVE
interparticle pore space and other types of visible
Figure 2—Petrophysical classification of carbonate pore
pore space, such as moldic pores. Research on car-
types used in this paper (Lucia, 1983) compared with bonate pore space (Murray, 1960; Choquette and
Archie’s 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
sizes from pinpoint to larger than cutting size. (intergrain and intercrystal) and other types of pore
Lucia 1277

Figure 3—Geological and petrophysical classification of carbonate interparticle pore space based on size and sort-
ing 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 Choquette and Pray (1970) discussed the geolog-
prompted modification of Archie’s classification. ic concepts surrounding carbonate pore space and
The petrophysical classification of carbonate presented a classification that is widely used. They
porosity presented by Lucia (1983) emphasized emphasized the importance of pore-space genesis
petrophysical aspects of carbonate pore space, as and used genetic, not petrophysical, divisions in
does the Archie classification. However, by com- their classification. They divided all carbonate pore
paring rock-fabric descriptions with laboratory space into two classes: fabric selective and nonfab-
measurements of porosity, permeability, capillarity, ric selective (Figure 2). Moldic and intraparticle
and Archie m values, Lucia (1983) showed that the pore types were classified as fabric-selective porosi-
most useful division of pore types for petrophysical ty by Choquette and Pray (1970) and grouped with
purposes was of pore space between grains or crys- interparticle and intercrystal porosity. However,
tals, called interparticle porosity, and all other pore Lucia (1983) demonstrated that moldic and intra-
space, called vuggy porosity (Figure 2). Vuggy pore particle pores have a different effect on petrophysi-
space was further subdivided by Lucia (1983) into cal properties than do interparticle and intercrystal
two groups depending on how the vugs are inter- pores and thus should be grouped separately. Pore-
connected: (1) vugs interconnected only through type terms used in Lucia’s (1983) classification are
the interparticle pore network are separate vugs, listed in Table 1, which compares his terms with
and (2) vugs that form an interconnected pore sys- those suggested by Choquette and Pray (1970).
tem are touching vugs. Although most of the terms defined by Choquette
1278 Classification of Carbonate Pore Space

180
22
Mercury/air extrapolated displacement pressure (psia)

21 Percent porosity
160

140 19

10
120
14
15
100
27
27
80 26
22
22
12
60
21
16 18
16
40

12
18 18
20 18
16 21
Figure 5—Porosity-air permeability relationship for var-
0 ious particle-size groups in nonvuggy carbonate rocks
(Lucia, 1983).
0 20 40 60 80 100 120 140
Average particle size (µm)

Figure 4—Relationship between mercury displacement pores, but it is consistent with the Archie termi-
pressure and average particle size for nonvuggy carbon- nology and with the widespread, less restrictive
ate rocks with greater than 0.1 md permeability (Lucia, use in the oil industry of the term “vuggy porosi-
1983). The displacement pressure is determined by
extrapolating the data to a mercury saturation of zero.
ty” in referring to visible pore space in carbonate
rocks.

and Pray are also used here, interparticle and vug ROCK-FABRIC/PETROPHYSICAL
porosity have different definitions. Lucia (1983) CLASSIFICATION
demonstrated that pore space located both
between grains (intergrain porosity) and between The foundation of the Lucia (1983) and the
crystals (intercrystal porosity) are petrophysically Archie classifications is the concept that pore-size
similar. A term identifies these petrophysically simi- distribution controls permeability and saturation
lar pore types: interparticle. “Interparticle” was and that pore-size distribution is related to rock fab-
selected because of its broad connotation. The clas- ric. To relate carbonate rock fabrics to pore-size dis-
sification of Choquette and Pray (1970) does not tribution, one must determine if the pore space
have a term that encompasses these two petro- belongs to one of the three major pore types: inter-
physically similar pore types. In their classification, particle, separate vug, or touching vug (Figure 2).
the term “interparticle” is used instead of “inter- Each class has a different type of pore distribution
grain.” [See Choquette and Pray (1970, page 247) and interconnection. One must also determine the
for their justification for using “interparticle” vs. volume of pore space in these various classes
“intergrain.”] because pore volume relates to reservoir volume
Vuggy porosity, as defined by Lucia (1983), is and, in the case of interparticle and separate-vug
pore space that is within grains or crystals or that porosity, to pore-size distribution.
is significantly larger than grains or crystals; that is,
pore space that is not interparticle. Vugs are com-
monly present as leached grains, fossil chambers, Petrophysics of Interparticle Pore Space
fractures, and large, irregular cavities. Fracture
porosity is categorized as a type of vuggy porosity In the absence of vuggy porosity, pore-size distri-
to be inclusive. This definition deviates from the bution in carbonate rocks can be described by parti-
restrictive definition of vugs used by Choquette cle size, sorting, and interparticle porosity (Figure
and Pray (1970) as nondescript, nonfabric-selective 3). Lucia (1983) showed that particle size can be
Lucia 1279

A B
0 0.5 1.0 0 0.2 0.4
mm mm

C D
0 1.0 2.0 0 0.5 1.0
mm mm
Figure 6—Photomicrographs showing examples of nonvuggy limestone rock fabrics. (A) Grainstone, φ = 25%, k =
1500 md. (B) Grain-dominated packstone, φ = 16%, k = 5.2 md. Note intergranular cement and pore space. (C) Mud-
dominated packstone, φ = 18%, k = 4 md. Note microporosity. (D) Wackestone.

related to mercury capillary displacement pressure porosity-permeability fields can be defined using
in nonvuggy carbonates having more than 0.1 md particle-size boundaries of 100 and 20 µm, a rela-
permeability. Because displacement pressure is a tionship that appears to be limited to particle sizes
function of the larger, well-connected pores, the par- less than 500 µm (Figure 5). These three fields will
ticle size describes the size of the larger pores be referred to as the greater than 100-µm permeabil-
(Figure 4). Whereas the displacement pressure char- ity field, 100–20-µm permeability field, and less
acterizes the larger pores sizes and is largely inde- than 20-µm permeability field.
pendent of porosity, the shape of the capillary pres- Recent work has shown that permeability fields
sure curve characterizes the smaller pore sizes and is can be better described in geologic terms if sorting
dependent on interparticle porosity (Lucia, 1983). and particle size are considered. The approach to
The relationship between displacement pressure size and sorting used herein is similar to the grain-
and particle size (Figure 4) is hyperbolic and sug- support/mud-support principle upon which
gests important particle-size boundaries at 100 Dunham’s (1962) classification is built. Dunham’s
and at 20 µm. Lucia (1983) demonstrated that three classification, however, focused on depositional
1280 Classification of Carbonate Pore Space

(A) (B)
1000 1000
Permeability (md)

Permeability (md)
100 100

10 10

1 1

0.1 0.1
10 20 30 40 10 20 30 40
Interparticle porosity (%) Interparticle porosity (%)
(C) (D)
1000 1000
Permeability (md)

Permeability (md)
100 100

10 10

1 1

0.1 0.1
10 20 30 40 10 20 30 40
Interparticle porosity (%) Interparticle porosity (%)

Figure 7—Porosity-air permeability crossplots for nonvuggy limestone rock fabrics compared with the three per-
meability fields illustrated in Figure 5 (A) 400-µm ooid grainstone, Ste. Genevieve, Illinois (Choquette and Steiner,
1985). Low-permeability, high-porosity data are deleted because they are from oomoldic and wackestone rock fab-
rics (P. W. Choquette, 1993, personal communication). (B) Grain-dominated packstone data, Wolfcamp, west Texas
(Lucia and Conti, 1987), a poorly sorted mixture of 80–300-µm grains and micrite. (C) Wackestones with micropo-
rosity between 5-µm crystals, Shuaiba, United Arab Emirates (Moshier et al., 1988). Data are associated with stylo-
lites 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-µm field.

texture, whereas petrophysical classifications are are filled with mud, even if the grains appear to
focused on contemporary rock fabrics that include form a supporting framework.
depositional and diagenetic textures. Thus, minor Grainstone is clearly a grain-dominated fabric,
modifications must be made in Dunham’s classifica- but Dunham’s (1962) packstone class bridges a
tion before it can be applied to a petrophysical clas- boundary between large intergrain pores in grain-
sification. stone and small interparticle pores in wackestones
Instead of dividing fabrics into grain supported and mudstones. Some packstones have intergrain
and mud supported as in Dunham’s classification, pore space, and some have intergrain spaces filled
fabrics are divided into grain dominated and mud with mud; therefore, the packstone textural class
dominated (Figure 3). The important attributes of must be divided into two rock-fabric classes: grain-
grain-dominated fabrics are the presence of open dominated packstones that have intergrain pore
or occluded intergrain porosity and a grain-support- space or cement, and mud-dominated packstones
ed texture. The important attribute of mud-domi- that have intergrain spaces filled with mud (Figure
nated fabrics is that the areas between the grains 3). The pore-size distribution is controlled by the
Lucia 1281

1000 (1987) reported on a nonvuggy, grain-dominated


Wolfcampian packstone found in a core taken in
Oldam County, Texas. The grain-dominated pack-
stone is described as a poorly sorted mixture of
Permeability (md)

100
150–300-µm grains in a matrix composed of 80-µm
pellets and 10-µm calcite crystals. A porosity-per-
10 meability crossplot of these data (Figure 7B) shows
that it plots on the boundary between the 100–20-
µm and the less than 20-µm permeability fields.
1
Additional data points have been gleaned from the
literature, and they plot in the 100–20-µm perme-
ability field.
Figure 7D is a crossplot between permeability
0.1 and porosity for North Sea coccolith chalk (Scholle,
10 20 30 40 1977). The average size of coccoliths is about 1
Interparticle porosity (%) µm. The data points plot below the less than 20-µm
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 8—Composite porosity-air permeability crossplot in the high-porosity ranges.
for nonvuggy limestone fabrics compared with the Figure 8 illustrates all the data for limestones
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, grain-
pore space. dominated packstone, and mud-dominated fabrics
are reasonably well constrained to the three perme-
ability 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 pro-
Permeability/Rock-Fabric Relationships duce systematic changes in the pore-size distribu-
tion. 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
grain size; in mud-dominated fabrics, the size of the Examples of nonvuggy dolomite petrophysical
micrite particles controls the pore-size distribution. rock fabrics are illustrated in Figure 9. Dolomitiz-
In 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 Steiner’s (1985) work on the Ste. mud fabric. Grain and mud fabrics in fine crys-
Genevieve oolite (Mississippian). The average grain talline dolostones are easily recognizable; however,
size of the oolite is about 400 µm. The points on as the crystal size increases, the precursor fabrics
this graph are concentrated within the greater than become more difficult to determine.
100-µm permeability field. Dolomite crystals (defined as particles in this
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 µm. Micrite particles are
mud-dominated fabric containing microporosity. usually less than 20 µm 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 µm (Moshier et al., 20 to more than 100 µm (Figure 9). The crossplot
1988). The data are concentrated in the less than of interparticle porosity and permeability (Figure
20-µm permeability field. 10A) illustrates the principle that in mud-dominat-
Because grain-dominated packstone is a new fab- ed fabrics, permeability increases as dolomite crys-
ric class, data are difficult to find. Lucia and Conti tal size increases. Fine crystalline (average 15 µm)
1282 Classification of Carbonate Pore Space

A D
0 0.2 0 0.5 1.0
mm mm

B E
0 0.5 1.0 0 0.5 1.0
mm mm

Figure 9—Examples of nonvuggy dolomite fabrics. (A)


Dolograinstone, 15-µm dolomite crystal size, φ = 16.4%,
k = 343 md, Dune field (Bebout et al., 1987). (B) Dolo-
grainstone, 30-µm dolomite crystal size, φ = 7.1%, k = 7.3
md, Seminole San Andres unit, west Texas. (C) Dolo-
grainstone, crystal size 400 µm, φ = 10.2%, k = 63 md,
Harmatton field, Alberta, Canada. (D) Grain-dominated
dolopackstone, 10-µm dolomite crystal size, φ = 9%, k =
1 md, Farmer field, west Texas. (E) Grain-dominated
dolopackstone, 30-µm dolomite crystal size, φ = 9.5%, k
= 1.9 md, Seminole San Andres unit, west Texas. (F)
Crystalline dolowackestone, 10-µm dolomite crystal
size, φ = 11%, k = 0.12 md, Devonian, North Dakota. (G)
Crystalline dolowackestone, 80-µm dolomite crystal
size, φ =16%, k = 30 md, Devonian, North Dakota. (H)
Crystalline dolowackestone, 150-µm dolomite crystal
size, φ = 20%, k = 4000 md, Andrews South Devonian
C field, west Texas.
0 1.0 2.0
mm
Lucia 1283

Figure 9—Continued.

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-µm permeability
field. Medium crystalline (average 50 µm) mud-
dominated dolostones from the Dune field,
Permian basin (Bebout et al., 1987), plot within the
100–20-µm permeability field. Large crystalline
(average 150 µm), mud-dominated dolostones from
Andrews South Devonian field, Permian basin
F (Lucia, 1962), plot in the greater than 100-µm per-
0 0.5 1.0
meability field.
Grainstones are commonly composed of grains
mm much larger than the dolomite crystal size (Figure
9), so that dolomitization does not have a signifi-
cant effect on the pore-size distribution. This prin-
ciple is illustrated in Figure 10B, a crossplot of
interparticle porosity and permeability measure-
ments from dolomitized grainstones. The grain size
of the dolograinstones is 200 µm. The fine crys-
talline dolograinstone from Taylor-Link field,
Permian basin, the medium crystalline dolograin-
stone 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-µm permeability
field. The large crystalline mud-dominated fab-
rics also plot in this permeability field, indicating
that they are petrophysically similar to grain-
stones (Figure 10A).
G Interparticle porosity and permeability measure-
0 0.5 1.0 ments from fine to medium crystalline, grain-domi-
mm nated 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
100–20-µm permeability field. The medium crys-
talline 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 100-
µm permeability field. Grains are very difficult to
recognize in dolostones with greater than 100-µm
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-domi-
H nated dolostones constitute the 100–20-µm perme-
0 0.5 1.0 ability field. Fine crystalline mud-dominated dolo-
mm stones constitute the less than 20-µm field.
1284 Classification of Carbonate Pore Space

(A) (B)
1000 1000
Fine crystal size
Medium crystal size
Large crystal size
Permeability (md)

Permeability (md)
100 100
Mud-dominated
dolostones

10 10

1 1

0.1 0.1
10 20 30 40 10 20 30 40
Interparticle porosity (%) Interparticle porosity (%)
(C)
Fine crystalline dolograinstone
1000
Fine to medium crystalline Medium crystalline dolograinstone
grain-dominated Large crystalline dolograinstone
dolopackstone
Permeability (md)

100

10

0.1
10 20 30 40
Interparticle porosity (%)

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

The dolostone permeability fields are defined by than 100-µm permeability field are (1) limestone and
dolomite crystal size, as well as grain size and sort-
dolomitized grainstones, and (2) large crystalline
ing of the precursor limestone. Within the field, grain-dominated dolopackstones and mud-domi-
permeability is defined by interparticle porosity. nated dolostones. These fabrics are called rock-fab-
Systematic changes in intergrain and intercrystal ric/petrophysical class 1. The effect of grain size in
porosity by predolomite calcite cementation, this field can be seen by comparing Figures 7A and
dolomite cementation, and compaction will sys- 10B. Ooid grainstones, which have a grain size of
tematically change the pore-size distribution. 400 µm, are more permeable for a given porosity
Therefore, interparticle porosity defines the perme- than dolograinstones, which have a grain size of
ability, and dolomite crystal size, grain size, and 200 µm.
sorting define the permeability field. Fabrics that make up the 100–20-µm field are
(1) grain-dominated packstones, (2) fine to medi-
um crystalline grain-dominated dolopackstones,
Limestone and Dolomite Comparison and (3) medium crystalline mud-dominated dolo-
Data from limestone and dolomite rock fabrics are stones. These fabrics are called rock-fabric/petro-
combined 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 interpar-
ticle porosity. Collapse of separate-vug fabrics
because of overburden pressure can produce frag-
ments that are properly considered “diagenetic par-
ticles.” 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. Interbreccia-
block 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 11—Composite porosity-air permeability cross- Several methods have been presented for relating
plot for nonvuggy dolostone fabrics compared with the porosity, permeability, water saturation, and reser-
three permeability fields illustrated in Figure 6. voir 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,
packstones tend to be more permeable for a given- porosity, permeability, and reservoir height without
porosity than limestone grain-dominated pack- regard to rock fabric. The data presented demon-
stones. Further data may show this to be an impor- strate that the three nonvuggy rock-fabric fields have
tant difference attributable to dolomite crystal unique porosity-permeability relationships, suggest-
morphology. The less than 20-µm permeability field ing that capillary properties of nonvuggy carbonates
is characterized by mud-dominated limestone and should also be separated into rock-fabric categories.
fine crystalline mud-dominated dolostones. These To characterize the capillary properties of the
fabrics are called rock-fabric/petrophysical class 3. three rock-fabric fields, capillary pressure curves
Reduced–major-axis permeability transforms are with different interparticle porosities from the three
presented for each class (Figure 12). The transform rock-fabric fields are compared in Figure 13. Figure
for class 2 is slightly skewed to the field bound- 13A shows two curves representative of class 1.
aries, and the following transform is more compati- Both curves are from samples of fine to medium
ble with the field boundaries. crystalline dolograinstones. The 9.2% porosity curve
Class 1: represents the average of two sets of capillary pres-
sure data from dolograinstones in the Taylor-Link
k = (45.35 × 108) × φip8.537 r = 0.71 field (Lucia et al., 1992b), and the 17.6% porosity
curve represents the average of two data sets, one
Class 2: from the Taylor-Link field and one from the Dune
field (Bebout et al., 1987). Three curves representa-
k = (1.595 × 105) × φip5.184 r = 0.80 tive of class 3 are presented in Figure 13C. The curves
are from samples of fine crystalline dolowackestones
or this recommended transform for class 2: and from the Farmer field, Permian basin. Three
curves representative of class 2 are presented in
k = (2.040 × 106) × φip6.38 Figure 13B. Class 2 represents a very diverse class of
rock fabrics, and it is difficult to combine all the fab-
Class 3: rics into a few simple curves. The curves presented
here are from medium crystalline dolowackestones
k = (2.884 × 103) × φip4.275 r = 0.81 of the Seminole San Andres unit and may not be rep-
resentative of all grain-dominated packstones.
where k = millidarcys and φip = fractional interpar- Pore-throat-size distribution for each capillary
ticle porosity. pressure curve was calculated using the following
1286 Classification of Carbonate Pore Space

1000

Class 1
Class 2

µm
Class 2
Class 1 µm

0
Class 3

10
20

100 Class 3
Permeability (md)

m

10
50

0.1
0.05 0.10 0.15 0.20 0.30 0.40
Interparticle porosity (fractional)

Figure 12—Composite porosity-air permeability crossplot for nonvuggy limestones and dolostones showing statisti-
cal reduced–major-axis transforms for each class (dashed lines). See text for equations.

formula. The results (Figure 13) show decreasing pressure of about 650 psia) and plotting saturation
pore-throat size with decreasing porosity within a against porosity for each rock-fabric class. The
petrophysical class and a general decrease in pore- results (Figure 14) show that in nonvuggy carbon-
throat size from class 1 to class 3. Microporosity ate reservoirs, a plot of porosity vs. water satura-
[pore-throat size less than 1 µm (Pittman, 1971)] is tion can separate rock fabrics into three classes.
concentrated in class 3 and decreases in impor- Equations relating water saturation to porosity
tance from class 3 to class 1. and reservoir height are developed in two steps
(Lucia et al., 1992b). First, mercury capillary pres-
Ri = (2Τ × cosθ × C)/Pc sure is converted to reservoir height using generic
values (Table 2). Second, wetting-phase saturations
where Ri = pore-throat radius (µm); T = air-mercury from the capillary pressure curves are plotted
interfacial tension = 480 dynes/cm; θ = air-mercury against porosity for several reservoir heights. Third,
contact angle = 140°; C = unit conversion constant lines of equal reservoir height are drawn assuming
= 0.145; and Pc = mercury injection pressure (psia). equal slopes and a relationship between intercepts
Each group of curves is characterized by similar and reservoir height is developed. This relationship is
displacement pressures and a systematic change in substituted for the intercept term in the porosity vs.
curve shape and saturation characteristics with saturation equation, resulting in a relationship among
changes in interparticle porosity. The relationship water saturation, porosity, and reservoir height.
among porosity, saturation, and rock-fabric class The resulting equations follow, and three-
can be demonstrated by selecting a reservoir height dimensional representations for classes 1 and 3
of 150 m (which equates to a mercury capillary are presented in Figure 15. These equations are
(A) (A')

Mercury injection pressure (psia)


2000 100
17.6% porosity
Permeability = 166 md

Mercury saturation (%)


1600 80

1200 9.2% porosity 60

9.2% porosity
800 40 Permeability = 72 md
17.60% porosity

400 20

0 0
100 80 60 40 20 0 2.0 1.5 1.0 0.5 0 -0.5 -1.0 -1.5
Mercury saturation (%) Log pore-throat radius (µm)

(B) (B')
Mercury injection pressure (psia)

2000 100
13.7% porosity
8.7% porosity Permeability = 5.1md

Mercury saturation (%)


1600 80
10.3% porosity
12.5% porosity
Permeability = 2.3 md
1200 12.5% porosity 60

13.7% porosity 10.3% porosity


800 40 Permeability = 1.3 md

400 20
8.7% porosity
Permeability = 0.05 md
0 0
100 80 60 40 20 0 2.0 1.5 1.0 0.5 0 -0.5 -1.0 -1.5
Mercury saturation (%) Log pore-throat radius (µm)

(C) (C')
Mercury injection pressure (psia)

2000 100
15.4% porosity
9.0% porosity Permeability = 2 md
Mercury saturation (%)

1600 80

11.8% porosity 11.8% porosity


1200 60
Permeability = 0.4 md
15.4% porosity
800 40
9.0% porosity
Permeability = 0.04 md
400 20

0 0
100 80 60 40 20 0 2.0 1.5 1.0 0.5 0 -0.5 -1.0 -1.5
Mercury saturation (%) Log pore-throat radius (µm)

Figure 13—(A–C) Capillary pressure curves and (A′–C′) 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

100 Table 2. Values Used to Convert Mercury/Air Capillary


Pressure to Reservoir Height
Water saturation (%)

Laboratory Reservoir
Class 3 (mercury/air/solid) (oil/water/solid)

Class 1 Class 2 T (dynes/cm) 480 28


10 θ (°) 140 44
Water density 1.04 0.88

Class 2 comprises grain-dominated packstones,


fine and medium cr ystalline grain-dominated
1 dolopackstones, and medium crystalline mud-domi-
5 10 20 30 40 nated dolostones:
Figure 14—Crossplot of porosity and water saturation
for the three rock-fabric/petrophysical classes at a reser-
k = (2.040 × 106) × φip6.38
voir height of 150 m. Water saturation (1 minus mer-
cury saturation) and porosity values are taken from cap- Sw = 0.1404 × H–0.407 × φ–1.440
illary pressure curves illustrated in Figure 13.
Class 3 contains mud-dominated limestones and
fine crystalline mud-dominated dolostones:
specific to the capillary pressure curves used in
this paper and will not necessarily apply to specific k = (2.884 × 103) × φip4.275
reservoirs. However, the equations will provide rea-
sonable values for original water saturations when Sw = 0.6110 × H–0.505 × φ–1.210
only porosity and rock-fabric data are available.

Class 1: Petrophysics of Separate-Vug Pore Space


Sw = 0.02219 × H–0.316 × φ–1.745 The addition of vuggy pore space to interparti-
cle pore space alters the petrophysical character-
Class 2: istics by altering the manner in which the pore
Sw = 0.1404 × H–0.407 × φ–1.440 space is connected, all pore space being connect-
ed in some fashion. Separate-vug pore space is
Class 3: defined as pore space that is (1) either within par-
Sw = 0.6110 × H–0.505 × φ–1.210 ticles or is significantly larger than the particle
size (generally greater than 2×), and (2) is inter-
where H = height above capillary pressure equal to connected only through the interparticle porosity
zero and φ = fractional 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;
Rock-Fabric/Petrophysical Classes moldic pore space, such as dissolved grains
(oomolds) or dolomite crystals (dolomolds); and
Because the three rock-fabric groups define per- intragranular microporosity are examples of intra-
meability and saturation fields, the three groups, particle, fabric-selective separate vugs. Molds of
together with interparticle porosity and reservoir evaporite crystals and fossils found in mud-domi-
height, can be used to relate petrophysical proper- nated fabrics are examples of fabric-selective sepa-
ties to geologic observations. These rock-fabric rate vugs that are significantly larger than the par-
groups, herein termed rock-fabric/petrophysical ticle size (Figure 18). In mud-dominated fabrics,
classes (Figure 16), are described, with their gener- shelter pore space is typically much larger than
ic transform equations as follows: the particle size and is classified as separate-vug
Class 1 is composed of grainstones, dolograin- porosity.
stones, and large crystalline dolostones: In grain-dominated fabrics, extensive selective
leaching of grains may cause grain boundaries to
k = (45.35 × 108) × φip8.537 dissolve, producing composite molds. These com-
posite molds may have the petrophysical character-
Sw = 0.02219 × H–0.316 × φ–1.745 istics of separate vugs. If, however, dissolution of
Lucia 1289

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

Wa
ter )
sat (%
ura ity
tion ros
(% Po
)

Wa
ter )
sat (%
ura ity
tion ros
(% Po
)

the grain boundaries is extensive, the pore space The addition of separate-vug porosity to interpar-
may be classified as interparticle porosity. ticle porosity increases total porosity, but does not
Grain-dominated fabrics may contain grains with significantly increase permeability (Lucia, 1983).
intragranular microporosity (Pittman, 1971; Keith Figure 19A illustrates this principle in that perme-
and Pittman, 1983; Asquith, 1986). Intragranular ability of a moldic grainstone is less than would be
microporosity is classified as a separate vug expected if all of the porosity were interparticle
because it is within the particles of the rock and is and, at constant porosity, permeability were to
interconnected only through the intergrain pore increase with decreasing separate-vug porosity
network. Mud-dominated fabrics also may contain (Lucia and Conti, 1987). This principle is also true
grains with microporosity, but they present no for intragranular microporosity. Figure 19B is a
unique petrophysical condition because of the sim- crossplot of data from a San Andres dolograinstone
ilar pore sizes between the microporosity in the from the Permian of west Texas. The crossplot
mud matrix and in the grains. shows that the permeability of the grainstone with
1290 Classification of Carbonate Pore Space

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

intragranular microporosity is less than would be capillary pressure curves typically have a bimodal
expected if all the porosity were interparticle. character.
Separate vugs found within the oil column are
usually considered to be saturated with oil because
of their relatively large size; oil migration into sepa- Petrophysics of Touching-Vug Pore Space
rate vugs is controlled by interparticle pore size.
Intragranular microporosity, however, is unique Touching-vug pore systems are defined as pore
because of small pore size. The small pore size pro- space that is (1) significantly larger than the parti-
duces high capillary forces that trap water and cle size, and (2) forms an interconnected pore sys-
lead to anomalously high water saturations within tem of significant extent (Figures 17, 18). Touching
a productive interval (Pittman, 1971; Dixon and vugs are typically nonfabric selective in origin.
Marek, 1990). Figure 20 shows the capillary char- Cavernous, breccia, fracture, and solution-enlarged
acteristics of a Cretaceous grainstone composed of fracture pore types commonly form an intercon-
grains having microporosity between 2-µm crystals nected pore system on a reservoir scale and are typ-
(Keith and Pittman, 1983) and a San Andres dolo- ical touching-vug pore types (Figure 18). Fenestral
grainstone with grains having microporosity pore space is commonly connected on a reservoir
between 10-µm dolomite crystals. This figure illus- scale and is grouped with touching vugs because
trates that within the oil column, water saturations the pores are normally much larger than the grain
in grainstones with intragranular microporosity are size (Major et al., 1990).
significantly higher than would be expected if no Fracture porosity is included as a touching-vug
intragranular microporosity were present. The pore type because fracture porosity is an important
Lucia 1291

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

contributor to permeability in many carbonate include fracture porosity as a pore type irrespec-
reservoirs and, therefore, must be included in any tive of its origin.
petrophysical classification of pore space. Al- Touching vugs are thought to be typically filled
though fracturing is often considered to be of with oil in the reservoirs and can increase perme-
tectonic origin and thus not a part of carbonate ability well above what would be expected from
geology, diagenetic processes common to car- the interparticle pore system. Lucia (1983) illus-
bonate reser voir s, such as kar sting (Kerans, trated this fact by comparing a plot of fracture per-
1989), can produce extensive fracture porosity. meability vs. fracture porosity to the three porosi-
The focus of this classification is on petrophysi- ty-permeability fields for interparticle porosity
cal properties rather than genesis, and must (Figure 21). This graph shows that permeability in
A B
0 0.5 1.0 0 1.0 2.0
mm mm

C D
0 1.0 2.0 0 1.0 2.0
mm mm

E F
20 microns
0 0.2
mm
Figure 18—Photomicrographs
and photographs showing
examples of vug pore types.
Separate-vug types: (A)
oomoldic porosity, φ = 26%,
k = 3 md, Wolfcampian, west
Texas; (B) intrafossil pore
space in a gastropod shell,
Cretaceous, Gulf Coast; (C)
fossil molds in wackestone,
φ = 5%, k = 0.05 md;
(D) anhydrite molds in
grain-dominated packstone,
φ = 10%, k = <0.1 md,
Mississippian, Montana;
(E) fine crystalline
dolograinstone with
intergranular and
intragranular microporosity
pore types, φ = 10%, k = 3 md,
Farmer field, west Texas;
(F) scanning electron
photomicrograph of
dolograins in (E) showing
G H intragranular microporosity
between 10-µm crystals.
1 in. 1 in. 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.

I J
3 in. 1 in.

K
0 1.0 2.0
mm
1294 Classification of Carbonate Pore Space

(A) (B)
1000 1000
Grainstone
field Grainstone
field
Permeability (md)

Permeability (md)
100 100

10 10

Svug
1 Svug 1 Dolograinstone
porosity
porosity with intragrain
average
average
8% microporosity
20%
0.1 0.1
5 10 20 30 40 5 10 20 30 40
Porosity (%) Porosity (%)
Figure 19—Crossplot illustrating the effect of separate-vug porosity on air permeability. (A) Grainstones with sepa-
rate-vug (Svug) porosity in the form of grain molds plot to the right of the grainstone field in proportion to the vol-
ume 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 environment (Lucia, 1993). The simplest example
fracture width and is sensitive to extremely small is a shoaling-upward bar complex. Mud-dominated
changes in fracture porosity. There is no effective fabrics grade upward and laterally into grain-domi-
method for measuring fracture width using the nated fabrics with a corresponding increase in per-
rock-fabric approach. meability, decrease in water saturation, and little
change in porosity (Figure 22). The permeability
and water saturation are a function of total porosi-
IMPLICATIONS FOR CONSTRUCTING A ty and rock fabric. The introduction of moldic
GEOLOGIC MODEL porosity (separate vugs) by selective dissolution of
grains in the grain-dominated fabric, caused per-
A fundamental problem in constructing a reser- haps by shoaling and the introduction of meteoric
voir model for input into numerical fluid-flow simu- water at a subaerial exposure surface, can produce
lators is converting geologic observations into petro- a diagenetic overprint that results in no significant
physical rock properties, namely porosity, permeability, change in porosity, little change in water satura-
and saturation. The problem has been discussed by tion assuming the molds are filled with hydrocar-
Lucia et al. (1992a), Senger et al. (1993), and Kerans bons, and drastically reduced permeability in the
et al. (1994). These workers suggested that the grainstone (Figure 22). The permeability is a func-
petrophysical properties within a rock-fabric unit tion of interparticle porosity determined by sub-
are nearly randomly distributed and thus form the tracting separate-vug porosity from total porosity.
basic building blocks for reservoir models. There- Conversion of the shoaling-upward bar complex
fore, geologic models described in terms of rock-fab- from limestone to 50-µm dolostone will not change
ric units can be most effectively converted into the grain size and sorting characteristics of the
petrophysical parameters and into reservoir models. grain-dominated fabrics, but will alter the particle
On the basis of data presented herein, the most size characteristics of the mud-dominated fabrics,
important rock-fabric characteristics to describe and changing the petrophysical class from class 3 to
map in nontouching-vug reservoirs are (1) grain size class 2. Dolomitization may reduce the total porosi-
and sorting using the modified Dunham classifica- ty of the grain-dominated fabrics, but the perme-
tion, (2) dolomite crystal size using 20 and 100 µm ability and water saturation will still be a function
as size boundaries, (3) separate-vug porosity, (4) sep- of grain size, sorting, and intergrain porosity. If the
arate-vug type with special attention to intergrain precursor limestone is a moldic grain-dominated
microporosity, and (5) total porosity. fabric, the resulting grain-dominated dolostone will
The stacking patterns of rock-fabric units are be moldic (Figure 22). Dolomitization of the mud-
systematic within a depositional and diagenetic dominated fabric to a 50-µm dolomite may reduce
Lucia 1295

(A) (A')
Mercury injection pressure (psia) 2000 100

Unimodal pore system Bimodal pore system


16.4%, 5.4 md

Mercury Saturation (%)


11.2%, 44.9 md 80
1500
Unimodal pore system
60 11.2%, 44.9 md
1000
40

500 Bimodal pore system


20
16.4 %, 5.4 md

0 0
100 80 60 40 20 0 1.5 1.0 0.5 0 -0.5 -1.0 -1.5
Mercury saturation (%) Log pore-throat radius (µm)

(B) (B')
2000
Mercury capillary pressure (psia)

100
Unimodal pore system
17.6%, 166 md

Mercury Saturation (%)


1500 80
Bimodal pore system
18.2%, 14.0 md 60
1000
Unimodal pore system
17.6%, 166 md 40

500 Bimodal pore system


20 18.2%, 14.0 md

0 0
100 80 60 40 20 0 2.0 1.5 1.0 0.5 0 -0.5 -1.0 -1.5
Mercury saturation (%) Log pore-throat radius (µm)

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 grain-
stone 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 dolograin-
stone 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 Evaporitic tidal-flat environments are a source of
decrease the water saturation significantly. dolomitizing water, and downward f low of the
Another simple example is the high-energy tidal- hypersaline water can dolomitize underlying car-
flat–capped cycle where intertidal and supratidal bonate sediments (Figure 23). In Permian basin
sediments overlie a shoreface grain-dominated reservoirs, intertidal and supratidal dolostones are
packstone (Figure 23). Carbonate cement and com- dense because of compaction and occlusion of
paction have reduced porosity in the tidal-flat sedi- pore space by anhydrite and dolomite, and grain-
ments, resulting in a low-permeability, high–water- dominated dolostones underlying the tidal-flat sedi-
saturation f low unit overlying a permeable ments commonly are dense because of pore-filling
grain-dominated limestone and a low-permeability anhydrite. High porosity is confined to subtidal
mud-dominated limestone. In a seaward direction, mud-dominated dolostones, and permeability and
shoaling-upward bar complexes replace the tidal- water saturation are related to intercrystal porosity
flat–capped cycles. and dolomite crystal size. In this example (Figure
1296 Classification of Carbonate Pore Space

105 porosity rather than total porosity less separate-vug


kf= 84.4 x 10 5 W 3 /Z cm cm porosity, and permeability and saturation can be
1 5
φf= W/Z x 100 = 0. seriously underestimated if the changes in particle
w =
104 z = Fracture spacing w size resulting from dolomitization are not included

cm
w = Fracture width cm in the geologic model.

1
5

0.
cm
Kf= Fracture permeability 0
0.

=
103 φf= Fracture porosity

w
=

100
w
SUMMARY

Z =
102

0 cm
The goal of reservoir characterization is to

m = 0 .0 Z = 1
describe the spatial distribution of petrophysical
Permeability (d)

10
parameters, such as porosity, permeability, and sat-

m
Inter-

.0 1 c
cm
particle φ uration. The rock-fabric approach presented here is
05
0
1 based on the premise that pore-size distribution
=
w

controls the engineering parameters of permeabili-


0 cm

>100µm
w

ty and saturation, and that pore-size distribution is


10-1
100

1 c

100 to 20µm related to rock fabric, which is a product of geolog-


Z =

ic processes. Thus, rock fabric integrates geologic


Z =

<20µm
10-2 interpretation with engineering numerical mea-
surements.
cm
1

To determine the relationships between rock


00
0.

10-3 fabric and petrophysical parameters, one must


=
w

define and classify pore space as it exists today in


10-4
terms of petrophysical properties. This is best
accomplished by dividing pore space into pore
10-5 10-4 10-3 10-2 10 -1 1 10 100
Porosity (%)
space located between grains or crystals, called
interparticle porosity, and all other pore space,
Figure 21—Theoretical fracture air permeability/porosi- called vuggy porosity. Vuggy pore space is further
ty relationship compared to the rock-fabric/petrophysi- subdivided into two groups, depending on how the
cal porosity, permeability fields (Lucia, 1983). vugs are interconnected: (1) vugs interconnected
only through the interparticle pore network are
termed separate vugs, and (2) vugs in direct con-
23), dolomite crystal size increases from fine to tact are termed touching vugs.
medium with depth and the petrophysical proper- The petrophysical properties of interparticle
ties change correspondingly. porosity are related to particle size, sorting, and
Grain-dominated bodies seaward of the tidal-flat interparticle porosity. Grain size and sorting of grains
deposits are dolomitized by hypersaline water from and micrite are based on Dunham’s (1962) classifica-
an overlying cycle and are not cemented by anhy- tion and are modified to make the classification com-
drite. The permeability and saturation characteris- patible with petrophysical considerations. Instead of
tics are related to grain size, sorting, and intergrain dividing fabrics into grain-supported and mud-sup-
porosity (Figure 23). The underlying mud-dominat- ported categories, fabrics are divided into grain-dom-
ed dolostone is composed of medium dolomite inated and mud-dominated categories. The impor-
crystals and has higher permeability than a mud- tant attributes of grain-dominated fabrics are the
dominated limestone with the same interparticle presence of open or occluded intergrain porosity
porosity. The underlying limestone has low porosi- and a grain-supported texture. The important
ty and permeability in this example. attribute of mud-dominated fabrics is that the areas
These examples illustrate the importance of between the grains are filled with mud even if the
mapping rock-fabric units within the geologic grains appear to form a supporting framework.
framework. If the shoaling-upward cycle is mapped Grainstone is clearly a grain-dominated fabric,
as a bar facies or the tidal-f lat–capped cycle is but Dunham’s (1962) packstone class bridges an
mapped as a peritidal facies, the permeability and important petrophysical boundary. Some pack-
saturation structure is compromised and the result- stones, as we see them now, have intergrain pore
ing model is oversimplified. In addition, if the dia- space, and some have the intergrain spaces filled
genetic facies are not included as mapping parame- with mud. Therefore, the packstone textural class
ters, the permeability and saturation values in the must be divided into two rock-fabric classes: (1)
resulting reservoir model may be in serious error. grain-dominated packstones having intergrain pore
For example, permeability can be significantly space or cement, and (2) mud-dominated pack-
overestimated if the estimate is based on total stones having intergrain spaces filled with mud.
Lucia 1297

Figure 22—Three examples of how the stacking of rock-fabric units affects the distribution of porosity, permeabili-
ty, and water saturation in a shoaling-upward sequence with selective dissolution and dolomitization overprints.
1298 Classification of Carbonate Pore Space

Figure 23—Examples of how the stacking of rock-fabric units affects the distribution of porosity, permeability, and
water saturation in a tidal-flat–capped sequence with hypersaline reflux dolomitization and sulfate emplacement
overprints.
Lucia 1299

The important fabric elements to recognize for using 20 and 100 µm as size boundaries, (3) sepa-
petrophysical classification of dolomites are precur- rate-vug type with special attention to intergrain
sor grain size and sorting, dolomite crystal size, and microporosity, (4) total porosity, and (5) separate-
interparticle (intergrain and intercrystal) porosity. vug porosity.
Important dolomite crystal size boundaries are 20 In touching-vug reservoirs, characterizing the
and 100 µm. Dolomite crystal size has little effect pore system is difficult because the pore system is
on the petrophysical properties of grain-dominated not related to a precursor depositional fabric, but is
dolostones. The petrophysical properties of mud- typically wholly diagenetic. Although the pore sys-
dominated fabrics, however, are significantly tem may conform to bedding, as in evaporite col-
improved when the dolomite crystal size is less lapse brecciation, it more commonly cuts across
than 20 µm. stratal boundaries. Recognizing the presence of a
Permeability and saturation characteristics of touching-vug pore system is paramount, however,
interparticle porosity can be grouped into three because it may dominate the flow characteristics of
rock-fabric/petrophysical classes. Class 1 is com- the reservoir.
posed of limestone and dolomitized grainstones,
and large crystalline grain-dominated dolopack-
stones and mud-dominated dolostones. Class 2 is REFERENCES CITED
composed of grain-dominated packstones, fine to Alger, R. P., D. L. Luffel, and R. B. Truman, 1989, New unified
medium crystalline grain-dominated dolopack- method of integrating core capillary pressure data with well
logs: Society of Petroleum Formation Evaluation, v. 4, no. 2,
stones, and medium crystalline mud-dominated p. 145–152.
dolostones. Class 3 is composed of mud-dominated Archie, G. E., 1952, Classification of carbonate reservoir rocks
limestone and fine crystalline mud-dominated dolo- and petrophysical considerations: AAPG Bulletin, v. 36, no. 2,
stones. p. 278–298.
Generic porosity-permeability transforms and Asquith, G. B., 1986, Microporosity in the O’Hara oolite zone of
the Mississippian Ste. Genevieve Limestone, Hopkins County,
water saturation, porosity, reservoir-height equa- Kentucky, and its implications for formation evaluation:
tions for each rock-fabric/petrophysical class are Carbonates and Evaporites, v. 1, no. 1, p. 7–12.
presented in the section “Rock-Fabric/Petro- Aufricht, W. R., and E. H. Koepf, 1957, The interpretation of capil-
physical Classes.” lary pressure data from carbonate reservoirs: Transactions of
the American Institute of Mining, Metallurgical, and Petroleum
The addition of separate-vug porosity to inter- Engineers, v. 210, p. 402–405.
particle porosity increases total porosity, but does Bebout, D. G., F. J. Lucia, C. F. Hocott, G. E. Fogg, and G. W.
not significantly increase permeability; therefore, Vander Stoep, 1987, Characterization of the Grayburg reser-
one must determine interparticle porosity by sub- voir, University Lands Dune field, Crane County, Texas:
tracting separate-vug porosity from total porosity University of Texas at Austin, Bureau of Economic Geology
Report of Investigations No. 168, 98 p.
and use interparticle porosity to estimate perme- Choquette, P. W., and L. C. Pray, 1970, Geologic nomenclature
ability. Separate-vug porosity is thought to be typi- and classification of porosity in sedimentary carbonates: AAPG
cally filled with hydrocarbons in the reservoir. Bulletin, v. 54, no. 2, p. 207–250.
Intergranular microporosity, however, may contain Choquette, P. W., and R. P. Steiner, 1985, Mississippian oolite and
non-supratidal dolomite reservoirs in the Ste. Genevieve
significant amounts of capillar y-bound water, Formation, North Bridgeport field, Illinois basin, in P. O. Roehl
resulting in water-free production of hydrocarbons and P. W. Choquette, eds., Carbonate petroleum reservoirs:
from intervals with higher than expected water New York, Springer-Verlag, p. 209–238.
saturation. Dixon, F. R., and B. F. Marek, 1990, The effect of bimodal pore
Touching-vug pore systems cannot be related to size distribution on electrical properties of some Middle
Eastern limestones: Society of Petroleum Engineers Technical
porosity, but are related principally to fracture Conference, SPE 20601, p. 743–750.
width. Because there is no effective method for Dunham, R. J., 1962, Classification of carbonate rocks according
making this observation in the reservoir, the rock- to depositional texture, in W. E. Ham, ed., Classifications of
fabric approach cannot be used to characterize carbonate rocks—a symposium: AAPG Memoir 1, p. 108–121.
Heseldin, G. M., 1974, A method of averaging capillary pressure
touching-vug reservoirs. curves: Society of Professional Well Log Analysts Annual
The key to constructing a geologic model that Logging Symposium, paper E, p. 1–7.
can be quantified in petrophysical terms is to select Keith, B. D., and E. D. Pittman, 1983, Bimodal porosity in oolitic
facies or units that have unique petrophysical quali- reservoir—effect on productivity and log response, Rodessa
ties for mapping. Petrophysical properties in rock- limestone (Lower Cretaceous), East Texas basin: AAPG
Bulletin, v. 67, no. 9, p. 1391–1399.
fabric facies or units are nearly randomly distributed Kerans, C., 1989, Karst-controlled reservoir heterogeneity in the
and can be legitimately averaged, making these geo- Ellenburger Group carbonates of west Texas: AAPG Bulletin,
logic units ideal for petrophysical quantification. In v. 72, no. 10, p. 1160–1183.
nontouching-vug reservoirs, the most important Kerans, C., F. J. Lucia, and R. K. Senger, 1994, Integrated charac-
terization of carbonate ramp reservoirs using Permian San
rock-fabric characteristics to describe and map are Andres Formation outcrop analogs: AAPG Bulletin, v. 78, no. 2,
(1) grain size and sorting using the modified p. 181–216.
Dunham classification, (2) dolomite crystal size Leverett, M. C., 1941, Capillary behavior in porous solids:
1300 Classification of Carbonate Pore Space

Transactions of the American Institute of Mining, Metallurgical, Major, R. P., G. W. Vander Stoep, and M. H. Holtz, 1990,
and Petroleum Engineers, v. 142, p. 151–169. Delineation of unrecovered mobile oil in a mature dolomite
Lucia, F. J., 1962, Diagenesis of a crinoidal sediment: Journal of reservoir: East Penwell San Andres unit, University Lands, west
Sedimentary Petrology, v. 32, no. 4, p. 848–865. Texas: University of Texas at Austin Bureau of Economic
Lucia, F. J., 1983, Petrophysical parameters estimated from visual Geology Report of Investigations No. 194, 52 p.
description of carbonate rocks: a field classification of carbon- Moshier, S. O., C. R. Handford, R. W. Scott, and R. D. Boutell,
ate pore space: Journal of Petroleum Technology, March, 1988, Giant gas accumulation in “chalky”-textured micrit-
v. 35, p. 626–637. ic limestones, Lower Cretaceous Shuaiba Formation, east-
Lucia, F. J., 1993, Carbonate reservoir models: facies, diagenesis, ern United Arab Emirates, in A. J. Lomando and P. M.
and flow characterization, in D. Morton-Thompson and A. M. Harris, eds., Giant oil and gas fields: Society of Economic
Woods, eds., Development geology reference manual: AAPG Paleontologists and Mineralogists Core Workshop 12, v. 1,
Methods in Exploration 10, p. 269–274. p. 229–272.
Lucia, F. J., and R. D. Conti, 1987, Rock fabric, permeability, and Murray, R. C., 1960, Origin of porosity in carbonate rocks: Journal
log relationships in an upward-shoaling, vuggy carbonate of Sedimentary Petrology, v. 30, no. 1, p. 59–84.
sequence: University of Texas at Austin, Bureau of Economic Pittman, E. D., 1971, Microporosity in carbonate rocks: AAPG
Geology Geological Circular 87-5, 22 p. Bulletin, v. 55, no. 10, p. 1873–1881.
Lucia, F. J., C. Kerans, and R. K. Senger, 1992a, Defining flow units Scholle, P. A., 1977, Chalk diagenesis and its relation to petroleum
in dolomitized carbonate-ramp reservoirs: Society of Petroleum exploration: oil from chalks, a modern miracle?: AAPG Bulletin,
Engineers, SPE 24702, p. 399–406. v. 61, no. 7, p. 982–1009.
Lucia, F. J., C. Kerans, and G. W. Vander Stoep, 1992b, Senger, R. K., F. J. Lucia, C. Kerans, and M. A. Ferris, 1993,
Characterization of a karsted, high-energy, ramp-margin car- Dominant control of reservoir-flow behavior in carbonate
bonate reservoir: Taylor-Link West San Andres unit, Pecos reservoirs as determined from outcrop studies, in B. Linville,
County, Texas: University of Texas at Austin, Bureau of R. E. Burchfield, and T. C. Wesson, eds., Reservoir characteriza-
Economic Geology Report of Investigations No. 208, 46 p. tion III: Tulsa, Oklahoma, PennWell, p. 107–150.

ABOUT THE AUTHOR

F. Jerry Lucia
F. Jerry Lucia retired from Shell
Oil Company in 1985, where he
had worked as a consulting geolog-
ical 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 carbon-
ate reservoirs by applying outcrop
studies to subsurface reservoirs.

También podría gustarte