You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236234426

Permeability-porosity relationships in interbedded limestone-dolostone


reservoirs

Article in AAPG Bulletin · January 2006


DOI: 10.1306/08100505087

CITATIONS READS

176 3,265

4 authors:

Stephen Ehrenberg Gregor P. Eberli


none University of Miami
95 PUBLICATIONS 5,209 CITATIONS 613 PUBLICATIONS 10,566 CITATIONS

SEE PROFILE SEE PROFILE

Marzieh Keramati Nojedeh Sadat Seyed Ali Moallemi


Islamic Azad University Research Institute of Petroleum Industry (RIPI)
9 PUBLICATIONS 306 CITATIONS 61 PUBLICATIONS 1,627 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Gregor P. Eberli on 19 March 2019.

The user has requested enhancement of the downloaded file.


Porosity-permeability AUTHORS
S. N. Ehrenberg  Statoil, Stavanger, Nor-
relationships in interlayered way, and Department of Geology, United
Arab Emirates University, P.O. Box 17551, Al
limestone-dolostone reservoirs Ain, United Arab Emirates; sne@uaeu.ac.ae
Steve has a Ph.D. from the University of Cali-
S. N. Ehrenberg, G. P. Eberli, M. Keramati, and fornia at Los Angeles (1978). He has worked
S. A. Moallemi at Shell Development Company’s Bellaire
Research Center (1981 – 1985) and Statoil
(1985–2005) and has recently begun teaching
at the United Arab Emirates University.
ABSTRACT
G. P. Eberli  Rosenstiel School of Marine
Porosity and permeability data from five carbonate platform succes- and Atmospheric Sciences, University of
sions of different settings, ages, and burial depths are examined Miami, 4600 Rickenbacher Cswy., Miami,
to identify overall similarities and differences in the reservoir quality Florida 33149; geberli@rsmas.miami.edu
of interlayered limestones and dolostones. Each succession consists Gregor Eberli received his Ph.D.s from the Swiss
mainly of limestone and dolostone, with subordinate proportions Institute of Technology (ETH), Zürich, Switzer-
of intermediate, partly dolomitized compositions. In the three land, in 1985 and the University of Miami in
deeply buried platforms, the key features are that limestones have 1991. With his colleagues at the Comparative
much lower average porosity than associated dolostones, and Sedimentologic Laboratory, he conducts re-
that limestones and dolostones show little difference in average search in sedimentology, stratigraphy, geo-
permeability-for-given-porosity. In contrast, the shallowly buried chemistry, and petrophysics of modern and
platforms show little difference in average porosity between lime- ancient carbonates. In several projects, he
investigated the influence of sea level changes
stones and dolostones and also display higher average permeability-
on sedimentary architecture. He was an AAPG
for-given-porosity in dolostones than limestones.
Distinguished Lecturer in 1996–1997 and a Joint
These data suggest the following general guidelines for depo- Oceanographic Institutions/U.S. Science Ad-
sitional and diagenetic controls on reservoir architecture in carbon- visory Committee Distinguished Lecturer in
ates consisting of interlayered limestone and dolostone. 1998 – 1999.
(1) Reservoir compartmentalization by the formation of tight
limestone barriers is largely a burial diagenetic process involving cal- M. Keramati  Research Institute of Petro-
leum Industry, National Iranian Oil Company,
cite cementation locally produced by chemical compaction. (2) Both
P.O. Box 1863, Tehran, Iran
the pattern of early dolomitization and the distribution of clay min-
erals ( because they influence the localization of chemical compac- Keramati holds an M. Tech. degree and a Ph.D.
tion) are key factors that determine the distribution of tight lime- in petroleum geology from Roorkee University,
India (1990). He joined the Research Institute of
stone barriers separating flow units. Thus, the pattern of eventual
Petroleum Industry in 1990 as a petroleum ge-
burial compartmentalization follows a template that is hardwired
ologist and has led several research projects in
into the stratigraphic architecture by depositional mineralogy and
petroleum exploration and reservoir studies.
early diagenesis. (3) After burial (2 – 3 km; 1.2 – 1.8 mi), dolostones His area of interest is reservoir geology and
should not be expected to have higher permeability-for-given- characterization. He is currently the head of
porosity than associated limestones. the Institute of Reservoir Studies at the Research
These rules assume dolomitization to occur early relative to Institute of Petroleum Industry.
chemical compaction, which will commonly be the case because
of combined hydrologic and mass-balance constraints. Dolomitiza- S. A. Moallemi  Research Institute of Pe-
troleum Industry, National Iranian Oil Com-
tion and concentration of clay appear linked to cycle and sequence
pany, P.O. Box 1863, Tehran, Iran
Ali holds an M.Sc. degree in sedimentology
(1994) from Azad University of Tehran and is a
Copyright #2006. The American Association of Petroleum Geologists. All rights reserved.
Ph.D. candidate in the Department of Geology
Manuscript received May 14, 2005; provisional acceptance June 1, 2005; revised manuscript received
at Shahid Beheshti University. He is a senior
August 5, 2005; final acceptance August 10, 2005.
DOI:10.1306/08100505087

AAPG Bulletin, v. 90, no. 1 (January 2006), pp. 91 – 114 91


research scientist at the Research Institute of architecture in the present examples and thus may have a useful
Petroleum Industry, working on carbonate degree of predictability, at least in terms of statistical parameters,
reservoir characterization research. His major such as net/gross and probability of flow-unit thickness distribution.
interests include carbonate sequence stra-
tigraphy, depositional systems, and diagenesis.
INTRODUCTION
ACKNOWLEDGEMENTS
A major theme throughout the history of carbonate reservoir re-
The manuscript benefited from suggestions by search has been the effect of dolomitization on reservoir quality
F. J. Lucia, C. T. Feazel, and P. A. Bjørkum. D. A.
(Powers, 1962; Davies, 1979; Allan and Wiggins, 1993; Sun, 1995;
Katz is thanked for the advice and practical
Warren, 2000; Moore, 2001; Braithwaite et al., 2004). Probably
help on the Madison data set. We thank Statoil
the most significant and commonly reported finding has been that
and the National Iranian Oil Company for
permission to publish this article. deeply buried dolostones tend to have higher porosity and per-
meability than associated limestones (Landes, 1946; Hohlt, 1948;
Chilingar and Terry, 1954; Institut Français du Pétrole, 1959; Murray,
1960; Lucia, 1962; Schmidt, 1965; Glover, 1968; Hull and Warman,
1970; Choquette and Steinen, 1980, 1985; Sears and Lucia, 1980;
Schmoker and Halley, 1982; Longman et al., 1983, 1992; Lindsay
and Kendall, 1985; Morgan, 1985; Ruzyla and Friedman, 1985;
Choquette et al., 1992; Amthor et al., 1994; Purser et al., 1994; Qing
and Mountjoy, 1994; Saller and Yaremko, 1994; Watts et al., 1994;
Brown, 1997; Mountjoy and Marquez, 1997; Zempolich and Hardie,
1997; Saller and Henderson, 1998; Eichenseer et al., 1999; Hopkins,
1999, 2004; Mancini et al., 2004; Smith et al., 2004; Westphal et al.,
2004; Zhao et al., 2005). However, quantitative information on the
nature and magnitude of such differences tends to be sparse, and
there is an understandable tendency to focus on the porous dolo-
stones and ignore the limestones.
Despite the preponderance of case studies showing higher po-
rosity in dolostones, a few examples to the contrary have also been
reported. Saller et al. (2004) describe an upper Paleozoic buildup
reservoir from the Horeshoe Atoll complex, Texas, where dolo-
stones are both less porous and more permeable than associated
limestones. The Upper Jurassic Arab reservoirs of Saudi Arabia and
Qatar also show overall decreasing porosity with increasing do-
lomitization (Powers, 1962; Cantrell et al., 2004; Clark et al., 2004).
Thus, the relationships observed in this study cannot be claimed to
be representative of all partially dolomitized carbonate reservoirs,
and appropriate cautions must be recommended.
Furthermore, Schmoker et al. (1985) reported lower average
matrix porosities and permeabilities for dolostone as opposed to
limestone reservoirs in a compilation of data for several thousand
United States carbonate reservoirs. Dolostone reservoirs were also
found to be statistically both deeper and considerably larger in
equivalent oil volume than limestone reservoirs. Schmoker et al.
(1985) suggested that dolomitization may locally improve reser-
voir properties, partly by fracturing, relative to adjacent limestones
that would otherwise make poor reservoirs.
This article reexamines the question of limestone versus
dolostone reservoir quality by comparing petrophysical, petrograph-
ic, and geological data from five shallow-water carbonate platform

92 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 1. Locations of carbonate porosity-permeability data sets studied. The background map shows the global distribution of
petroleum reservoirs with production from carbonate strata (modified from Ehrenberg and Nadeau, 2005).

successions that consist mainly of interbedded lime- three deeply buried data sets as they may have been
stones and dolostones (Figure 1). Three of these cases at an earlier stage before burial diagenesis.
have been deeply buried, whereas two have never Reservoir quality is examined here using three
been deeper than their present range of 0 –700 m types of data display: (1) histograms of porosity fre-
(0– 2296 ft). Comparisons in each data set allow the quency distribution (Figures 2 – 7), (2) porosity-depth
evaluation of the effects of dolomitization, whereas profiles showing the vertical stratigraphic organization
comparisons between the deeper and shallower data of porosity with respect to lithology and depositional
sets may be relevant to understanding the effects of cyclicity (Figures 8 – 12), and (3) permeability-porosity
burial diagenesis. Details of stratigraphy and geologic crossplots (Figures 13 – 16). These simple visual repre-
setting for each example have been described in pre- sentations facilitate a rapid comprehension of overall
vious publications. The purpose of this article is to similarities and differences between complex data sets.
compare these data sets and point out overall similari-
ties and differences of possible general significance.
Our five data sets represent very diverse and dis- METHODS
similar assemblages of depositional facies and strati-
graphic patterns, so it is relevant to question whether The porosity, permeability, and grain-density data pre-
comparison is meaningful. It is also uncertain to what sented here are from 1-in. (2.5-cm) plugs that were
degree the dolostones of each data set represent the drilled from cores and outcrops. The measurements
same precursor lithofacies as the associated limestones. were made by various laboratories but used similar
These are valid concerns for which we have no clear techniques that are more or less industry standard. Data
answers. However, we suggest that these five data sets and analytical techniques are available for the shal-
are each so unique and complexly heterogeneous that low data sets in Anselmetti and Eberli (2001), Melim
the commonalities observed potentially carry broad et al. (2001a), and Ehrenberg et al. (2004a). Analytical
significance for carbonate reservoirs worldwide. Nev- techniques for the deeply buried data sets are similar:
ertheless, a key issue that we discuss further in this ar- porosity by helium injection and weight-volume rela-
ticle is the degree to which the two shallowly buried tionships and permeability by pressure differential dur-
data sets represent valid analogs for interpreting the ing nitrogen flow.

Ehrenberg et al. 93
Figure 2. Frequency distribution of DI values estimated from mineralogic analyses for each data set examined. Bar fill indicates
limestone (open), intermediate (gray), and dolostone (black) for each data set examined.

Proportions of calcite, dolomite, and noncarbonate gible contents of noncarbonate minerals. X-ray dif-
minerals have been estimated in each plug sample by fraction analyses were also available for many plugs,
various methods using whatever information is avail- but in all other cases, plug mineralogy was confirmed
able for each sample group. For the Finnmark samples, by either testing for reaction with HCl or from thin-
bulk rock x-ray diffraction analyses, bulk chemical anal- section petrography. For Asmari plugs, calcite, dolo-
yses, and thin sections were made for approximately mite, quartz, clay, and anhydrite contents were esti-
every third plug over most cored intervals, and the mated first from qualitative thin-section examination
results were then interpolated to intervening plug sam- (roughly one per meter, approximately equivalent to
ples using information from grain density, core descrip- every third plug), with reference to grain-density values
tions, and, in cases of particular uncertainty, exami- (available for most cores). These estimates were then
nation of the actual plug samples. For the Madison checked against a set of bulk chemical analyses for 478
samples, relative dominance of calcite and dolomite was of the samples. Finally, these results were then inter-
estimated from grain-density values, assuming negli- polated to intervening plug samples using information

Figure 3. Frequency distribution of plug


porosity measurements in cores from
the Great Bahama Bank (late Miocene–
Pleistocene; boreholes Unda and Clino),
burial depth 33 –670 m (108–2198 ft).
Only platform samples are included.

94 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 4. Frequency distribution of plug
porosity measurements in cores from
carbonate platforms of the Marion Pla-
teau (Miocene; ODP Leg 194, Sites 1193,
1196, and 1199), burial depth 0–650 m
(0–2132 ft).

from grain density and core descriptions. For the Baha- GEOLOGICAL SETTING
mas plugs, the mineralogic data reported in Anselmetti
and Eberli (2001) and Melim et al. (2001a) were used. The locations of the five data sets studied are shown in
For the Marion Plateau plugs, mineral abundances were Figure 1. Miocene–Pleistocene platform-to-slope strata
estimated from a combination of petrographic obser- of the Great Bahama Bank were sampled in boreholes
vations and bulk chemical analyses of selected plugs Unda and Clino at depths of 33–670 m (108–2198 ft)
(Ehrenberg et al., 2004a) and interpolation using grain- below the mud pit in 7 m (23 ft) of water depth (Eberli
density values, shipboard core descriptions, and x-ray et al., 2001; Melim et al., 2001a). Parts of the cores from
diffraction data (Isern et al., 2002). these locations that consist of fine-grained slope facies

Figure 5. Frequency distribution of plug


porosity measurements in cores from
the Finnmark platform (Gipsdalen Group,
Pennsylvanian –Lower Permian, offshore
north Norway), maximum burial depth
2400 –3300 m (7874 – 10,826 ft).

Ehrenberg et al. 95
Figure 6. Frequency distribution of plug
porosity measurements in outcrops and
cores of the Madison Formation (Missis-
sippian, Wyoming and Montana), maxi-
mum burial depth 3300–7500 m (10,826–
24,606 ft).

are excluded from the present comparison involv- United States (Smith et al., 2004; Westphal et al., 2004).
ing platform carbonate facies. Dolomitized platform The maximum burial depth is believed to have been
(reefal) facies occur only in the Unda core at 293 – 3300 – 7500 m (10,827 –24,606 ft). Early dolomitiza-
360-m (961 – 1181-ft) depth, where fabric-preserving tion is interpreted to have occurred by both reflux of
to microsucrosic dolomite is interpreted to have hypersaline brine and circulation of normal-marine sea-
formed shortly after deposition from slightly altered water (Moore, 2001; D. Katz, 2005, personal commu-
seawater (Swart and Melim, 2000; Melim et al., nication), but stable isotope data have been viewed as
2001b). indicating subsequent recrystallization of most Madi-
Miocene platform strata from two carbonate plat- son dolomite in mixed meteoric-marine waters (Budai
forms on the Marion Plateau, offshore northeast Aus- et al., 1987). Later hydrothermal dolomitization oc-
tralia, were sampled at depths of 0 – 650 m (0 – 2133 ft) curred during Laramide thrusting (Katz et al., in press).
in three boreholes drilled by Ocean Drilling Program Data compiled in this study include outcrop samples
(ODP) Leg 194 (Isern et al., 2002; Ehrenberg, 2004a; from the Benbow Mine Road, Shoshone Canyon,
Ehrenberg et al., 2004b). As in the Bahamas cores, do- Sheep Mountain, Wind River Canyon, Casper Moun-
lomitization of the Marion Plateau strata is interpreted tain, Freemont Canyon, Hartville Canyon sections, and
to have occurred by circulation of normal to slightly subsurface samples from the Lysite, Madden Deep, and
altered seawater at low temperature with essentially Beaver Creek cores (Smith et al., 2004; Westphal et al.,
no influence of evaporative concentration (Ehrenberg, 2004).
2004a; Ehrenberg et al., in press). The Oligocene – lower Miocene Asmari Forma-
The Pennsylvanian –Lower Permian Finnmark car- tion is represented by cores from five wells in two
bonate platform was cored in two exploration wells, giant oil fields of southwest Iran (Aqrawi et al., 2004;
offshore north Norway (Ehrenberg et al., 1998a, b, Ehrenberg et al., 2004a). Published information on
2002; Ehrenberg, 2004b). The present burial depth is the Asmari Formation is sparse, but general reviews
1410 –1750 m (4626– 5741 ft), but maximum burial in are given in Hull and Warman (1970) and McQuillan
the early Tertiary was probably 2400 –3300 m (7874 – (1985). The present burial depth in the studied oil
10,827 ft). fields ranges from 1660 to 3700 m (5446 to 12,139 ft),
The Mississippian Madison Formation was sam- but the maximum burial was probably 500 –1500 m
pled in outcrop sections and cores from the Big Horn (1640–4921 ft) deeper. Variable dolomitization of both
and Wind River basins of Wyoming and Montana, the Asmari and Finnmark strata is suggested to have

96 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 7. Frequency distribution of plug
porosity measurements in cores from the
Asmari Formation (Oligocene–Miocene,
Iran) in the (A) Marun field, burial depth
3000–3700 m, (9842 – 12,139 ft) and
(B) Bibi Hakimeh field, burial depth
1660 – 2130 m (5446 – 6988 ft).

occurred mainly by the shallow circulation of evapo- MINERALOGY


ratively concentrated seawater based on the ubiqui-
tous association between dolostones and anhydrite Key parameters used in this study are the dolomiti-
nodules and cements, the common presence of high- zation index (DI), defined as 100 times the ratio of
salinity fluid inclusions in these dolomites, and the dolomite to total carbonate, and the siliciclastic index,
general increase in dolomitization toward shelf-interior defined as the weight percentage of the bulk sample
settings. composed of siliciclastic minerals. The weight percent

Ehrenberg et al. 97
Figure 8. Depth-porosity
profiles in two bore-
holes of the Bahamas
Drilling Project. Distance
between locations is
8.5 km (5.2 mi). Hori-
zontal lines indicate bases
of seismic sequences a
to h (Eberli et al., 2001).
Plotting symbols indicate
the dominant mineralo-
gy and depositional set-
ting (large symbols =
platform; small symbols =
slope). The bar to the
right of the porosity data
shows the division of
each sequence into low-
stand + transgression
(black) and highstand
(white).

anhydrite is another key parameter related to lomitization. Samples with DI greater than 0 but less
porosity in the Finnmark and Asmari data sets. Cal- than 25 are commonly wackestones and packstones,
cium sulfate minerals are not present in the Bahamas in which the mud matrix is partly replaced by dolo-
and Marion Plateau sections because of nonarid depo- mite. The relative paucity of intermediately dolomi-
sitional conditions and are absent from the Madison tized rock has been reported previously (Steidtmann,
outcrop samples because of surficial leaching. Both the 1917; Chave, 1954; Institut Français du Pétrole, 1959;
Finnmark and Asmari cores contain subordinate beds of Schmidt, 1965) and may be a general characteristic
dominantly siliciclastic composition, but samples from of carbonate strata, reflecting some intrinsic aspect of
these beds have been excluded from Figures 2 – 7 and the dolomitization process (Blatt et al., 1980, p. 517).
13–16 because this study concerns only carbonate rocks. This characteristic may largely reflect the process of
Figure 2 shows histograms of DI values from sam- dissolution of the remaining calcium carbonate during
ples for which mineral abundances have been estimated or after partial dolomitization, as discussed below.
from direct analytical control (based on petrography, Only for the Finnmark, Bahamas, and Marion Pla-
x-ray diffraction, bulk chemistry, or some combination teau data sets are the constraints of analytical control
of these, as opposed to mineralogic estimation by inter- tight enough to define an intermediately dolomitized
polation between analyzed samples or from grain den- grouping of plugs throughout each data set (for ex-
sity alone). Most samples in each data set are com- ample, Ehrenberg, 2004b). For the Asmari and Madison
posed mainly of either calcite or dolomite, with partly data, it has proved impractical to consistently distin-
dolomitized samples of intermediate composition (DI = guish intermediately dolomitized samples from either
25–75) being subordinate in all cases. In each data set, more pure limestones or more pure dolostones for the
a large proportion of the samples with DI less than majority of plugs where either a thin section or a bulk-
100 but greater than 75 are dolostones in which the chemical analysis is not available. Throughout the re-
calcite occurs only as late cement, postdating do- mainder of this article, therefore, ‘‘limestone’’ and

98 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 9. Depth-porosity profiles in three boreholes on the Marion Plateau (ODP Leg 194). Site 1193 is on a separate carbonate
platform, separated by 140 km (87 mi) from sites 1196 and 1199, which are separated by only 5 km (3.1 mi) (Isern et al., 2002).
Horizontal lines indicate bases of sequences labeled B1 to C (Ehrenberg et al., in press). Plotting symbols indicate dominant mineralogy.
The bar to the right of the porosity data shows the division of each sequence into transgression (black) and highstand (white).

‘‘dolostone’’ are defined as referring to carbonate- stones and dolostones have similar average porosity,
dominated rocks with DI < 50 and DI > 50, respectively. although the Marion Plateau dolostones are distinctly
This convention lumps relatively pure limestones to- skewed toward higher frequency at lower porosities.
gether with subordinate numbers of partly dolomitized Porosity histograms comparing limestones and do-
limestones and lumps relatively pure dolostones with lostones of the three deeply buried data sets show
subordinate numbers of dolomite-dominated carbonates important features in common (Figures 5–7), which
of intermediate composition. We feel that such simpli- contrast sharply with the histograms of the shallowly
fication is nevertheless preferable to working with an in- buried strata (Figures 3, 4). In each case, the deeply
termediately dolomitized grouping that would likely in- buried limestones have strongly positively skewed po-
clude significant proportions of both pure limestones and rosity distributions with the highest frequency of val-
pure dolostones. In any case, the uncertainties involved ues, or mode, in the range 0–5%, whereas deeply
in the classification of intermediately dolomitized sam- buried dolostones have more symmetrical distributions
ples are believed to be relatively insignificant with regard with the mode about 8–15% and relatively few values
to the overall differences observed in comparing the lime- less than 5%. For Asmari reservoirs in other oil fields,
stones and dolostones defined by the simple 50% rule. Lees (1933) found no relationship between dolomi-
tization and porosity, whereas Chilingar and Terry
(1954) reported a strong relationship.
POROSITY A degree of uncertainty exists for the samples from
the Madison Formation because surficial weathering
Limestone versus Dolostone appears to have leached anhydrite from the outcrop
samples, which make up the major part of this data
Porosity histograms for the two shallowly buried data set, and deep penetration of meteoric water is reported
sets show only modest differences between limestones to have leached anhydrite throughout the subsurface
and dolostones (Figures 3, 4). In each case, the lime- of the Wind River basin (Moore, 2001). Anhydrite is,

Ehrenberg et al. 99
Figure 10. Depth-
porosity profiles in cores
from two exploration
wells on the Finnmark
platform, penetrating
the Gipsdalen Group
(shallow-water carbonate-
dominated section of se-
quences 1 to 4). Hori-
zontal lines indicate the
bases of sequences 1 to 5.
Plotting symbols indicate
the dominant mineralo-
gy. The bar to the right
of the porosity data
shows the division of
each sequence into
transgression (black)
and highstand (white).

however, commonly present in subsurface Madison petrographic categories: (1) grainstones and mud-poor
samples from the Big Horn basin (Sonnenfeld, 1996). packstones in which formerly abundant intergranular,
Insofar as anhydrite may be more abundant in dolo- intrafossil, and moldic macropores have been mostly
stones than limestones, the leaching could have the filled by coarse calcite cement; and (2) mud-supported
effect of enhancing the porosity contrast between Ma- facies that have lost porosity by tight cementation of
dison limestones and dolostones. the mud matrix (dominant process) by filling of intra-
fossil spaces with coarse calcite spar and by mechanical
compaction.
Pore Types and Mechanisms of Porosity Loss Limestones with porosity greater than 10% com-
prise 21% of this data set. Of these 268 plugs, 57 samples
To better understand the comparisons in Figures 3 – 7, represented by thin sections reveal that most macro-
it is necessary to examine the range of lithofacies and porosity consists of primary intergranular and intra-
diagenetic fabrics characterizing the following key groups fossil pores that have been partly but not completely
from each of the three deeply buried platforms: (1) tight filled by calcite cement.
limestones, (2) porous limestones, and (3) porous do- Dolostones with porosity greater than 10% com-
lostones. It is especially pertinent to enquire why the prise 17% of the Finnmark data set. Of these 215
tight limestones of each data set have low porosity. plugs, 59 samples represented by thin sections con-
tain two dominant pore types: intercrystalline pores
Finnmark Platform between dolomite crystals (replacing matrix and bio-
Limestones with porosity less than 5% comprise 24% of clasts and, to some degree, also infilling relict inter-
this data set. Of these 311 plugs, most of the 87 samples granular pore spaces) and moldic pores representing
represented by thin sections can be grouped into two former bioclasts.

100 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 11. Depth-porosity profiles in three outcrop sections through the Madison Formation. Localities are separated by distances
of 90 –110 km (56 –68 mi) (Smith et al., 2004). Horizontal lines indicate bases of sequences labeled I to V. Plotting symbols indicate
the dominant mineralogy. The bar to the right of the porosity data shows the division of each sequence into transgression (black) and
highstand (white).

Madison Formation exclusively in dolomitized grainstones, where early


Limestones (including partly dolomitized limestones dolomitization inhibited compaction. Moldic porosity
with DI < 50) having porosity less than 5% com- is present in about 57% of dolostones and occurs in
prise 16% of the Madison data set. Thin sections were both tight and porous samples (Westphal et al., 2004).
made from a selection of 18 of these plugs (plus another Intercrystalline porosity is most common in dolomi-
11 limestones with porosity of 5–10%) to examine the tized mudstone and wackestone and peloid grainstone
causes of low porosity. All samples examined are from facies (Smith et al., 2004).
the outcrop sections shown in Figure 11. These samples
are dominantly grainstones to packstones but also in- Asmari Formation
clude a few mudstones to wackestones. A few of the Limestones with porosity less than 5% comprise
samples examined have textures (dolomite rhombs 22% of the plug measurements. Of these 1694 plugs,
with calcite cores or corroded edges) suggesting dedo- 63 samples represented by thin sections consist, in most
lomitization, as described by Budai et al. (1984). In cases, of mud-rich textures that have lost porosity by
most cases, the dominant cause of porosity loss is ce- tight cementation of the mud matrix combined with
mentation by coarse, blocky calcite spar, which post- complete filling of open (not mud-filled) spaces by coarse
dates dolomite, wherever dolomite is also present. calcite spar.
Limestones with porosity greater than 10% com- Limestones with porosity greater than 10% com-
prise only 2% of the Madison plugs. Thin sections were prise 13% of this data set. Of these 980 plugs, 41 sam-
made from 5 of these 26 plugs. In two cases, the po- ples represented by thin sections reveal that most
rosity consists of preserved intergranular macropores. macroporosity consists of primary intergranular and in-
In two other thin sections, impregnation reveals the trafossil pores that have been partly but not completely
porosity to exist mainly as very small pores in unce- filled by calcite cement. Microporous mud matrix also
mented mud matrix, and in the final case, no porosity accounts for a large part of the total porosity.
is visible in thin section and is thus microporosity. Dolostones with porosity greater than 10% com-
Dolostones with porosity greater than 10% com- prise 34% of the Asmari data set. Of these 2618 plugs,
prise 43% of the Madison plugs. Pore types include pri- 110 samples represented by thin sections contain 2 domi-
mary interparticle, moldic, intercrystalline, and fracture nant pore types: intercrystalline pores between dolomite
(Westphal et al., 2004). Primary porosity is preserved crystals and moldic pores representing former bioclasts.

Ehrenberg et al. 101


Figure 12. Depth-porosity profiles in five wells through the Asmari Formation. The vertical scale is the depth below the top Asmari,
corrected for bedding dip and wellbore deviation. Distances between wells from Marun field (MN-68, MN-30, MN-8) are 13 and 51
km (8 and 31 mi). The distance between Bibi Hakimeh wells (BH-18, BH-51) is 11 km (6.8 mi). The distance between Marun wells
and Bibi Hakimeh wells is roughly 140 km (87 mi). Plotting symbols indicate the dominant mineralogy. The bar to the right of the
porosity data shows the suggested positions of transgressive (black) and highstand (white) systems tracts.

Although not shown in Figure 7, the porosity distribu- Finnmark Platform


tions of Asmari limestones and dolostones both show an The key petrophysical groupings described in the pre-
overall relationship to depositional texture, with grain- vious section (tight limestones, porous limestones, and
dominated fabrics being somewhat more abundant at porous dolostones) occur clustered in distinct strati-
higher porosities and mud-dominated fabrics making graphic intervals (Figure 10), the significance of which
up a higher percentage of the low-porosity samples. varies in different parts of the section (Ehrenberg,
2004b). In sequence 1, tight limestones (highstand
Stratigraphic Patterns of Dolomitization and Porosity cycle tops) form a series of barriers separating po-
rous dolomitic sandstones (transgressive cycle bases).
Bahamas and Marion Plateau Sequence 2 represents higher accommodation and
These strata are characterized by extreme variability of more distal depositional conditions than the underly-
porosity in narrow depth ranges, although sequence- ing sequence 1, resulting in finer grained, more clay-
scale depth trends are also apparent within some in- rich siliciclastic intervals and paucity of dolomite,
tervals (Figures 8, 9). Limestones and dolostones in the but tight limestones again form the highstand cycle
Bahamas and Marion Plateau cores are segregated into tops.
distinct zones, the occurrence of which has been inter- Sequence 3 began with a major long-term trans-
preted as reflecting the combination of platform ar- gression in which siliciclastic influx was shut off and
chitecture and the hydrology of subsurface seawater a thick, aggradational section of algal buildups was
circulation (Swart and Melim, 2000; Eberli et al., 2001; deposited. These are largely preserved as porous lime-
Ehrenberg, 2004a). stone, reflecting paucity of clay laminations that could

102 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 12. Continued.

serve to initiate stylolitic dissolution. However, bar- and wackestones of a proximal inner-platform setting,
riers are formed by thinner intervals of argillaceous where high porosity reflects preservation of intercrys-
limestones that formed during transgressions marking talline and moldic pores.
the bases of the high-frequency sequences comprising Sequence 4 consists mainly of limestones, which
sequence 3. The late-highstand phase of sequence 3 cor- preserve high porosity in the lower, transgressive part
responds with the deposition of dolomitic mudstones of the sequence, but become tightly calcite-cemented

Figure 13. Semilog


plot of permeability ver-
sus porosity from cores
through two shallowly
buried carbonate suc-
cessions: (A) the Baha-
mas transect and (B) the
Marion Plateau.

Ehrenberg et al. 103


Figure 14. Semilog plot of
permeability versus porosity
from cores through the Finn-
mark platform: (A) limestones
and (B) dolostones.

upward because of increasing frequency of stylolite- bution of both dolomitization and porosity are related
prone clay laminations. Sequence 5 consists of bryozoan- to platform paleogeography in the lower composite
echinoderm grainstone and packstone and is not in- sequence. In the most downdip (seaward) part of the
cluded in the photozoan carbonate strata that are the ramp (Benbow Mine Road and Shoshone Canyon sec-
subject of this article. Uniformly low porosity in tions), dolomitization affected mainly mud-dominated
sequence 5 reflects the combination of frequent stylo- strata, whereas less than 5% of grain-dominated strata
lites and abundant echinoderm sites for syntaxial cal- are dolomitized. Further updip, in the midramp part of
cite cement growth. the platform (Wind River Canyon and Lysite core sec-
tions), there is pervasive dolomitization of all rock types,
Madison Formation whereas the most landward locations (Casper Mountain,
The Madison platform evolved from a ramp profile in Freemont Canyon, Hartville Canyon) have a patchy do-
its lower composite sequence (including sequences I, lomite distribution with both muddy and grainy litholo-
IIa, and IIb of Figure 11) to a flat-topped shelf during gies preserved as limestone in parts of the succession.
the deposition of its upper composite sequence (se- Figure 11 compares porosity and dolomitization
quences III to VI) (Sonnenfeld, 1996). Smith et al. (2004) profiles through three sections from the relatively down-
and Westphal et al. (2004) recognized how the distri- dip (seaward) part of the ramp. Benbow Mine Road and

Figure 15. Semilog plot of per-


meability versus porosity from
the Madison Formation (mainly
outcrop samples): (A) limestones
and (B) dolostones.

104 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Figure 16. Semilog plot of per-
meability versus porosity from
cores through the Asmari plat-
form: (A) limestones of Marun
field, (B) dolostones of Marun
field, (C) limestones of Bibi
Hakimeh field, (D) dolostones
of Bibi Hakimeh field.

Shoshone Canyon are located in the most downdip the Madison data set are not shown in Figure 11 but
position, where sequence I shows strong differentiation are located farther updip ( landward), where se-
into an entirely dolomitized and porous transgressive quences I, IIa, IIb, and lower III are pervasively dolo-
tract and a variably dolomitized highstand tract with mitized. In all localities, sequences IV and V consist
generally low porosity. Sequences IIa and IIb (as defined mostly of tight limestone, where porosity loss has oc-
by D. Katz, 2005, personal communication) have low curred mainly by chemical compaction during burial
dolostone porosity, whereas sequence III has a high do- (Crockett, 1994).
lostone porosity throughout. Sheep Mountain is located
farther updip, in a position transitional to the pervasively
dolomitized midramp setting. Higher proportions of Asmari Formation
grain-dominated strata were dolomitized in this loca- The principal barriers to matrix flow in the Asmari
tion, and there are higher proportions of porous dolo- Formation consist of tight limestones (Figure 12), al-
stone in both the highstand tract of sequence I and though the extent to which these are fractured and,
throughout sequences IIa and IIb. In all three locations thus, open to flow between adjacent porous layers is
in Figure 11, porous dolostone occurs mainly in the largely unknown. Fractures are important for Asmari
muddy, transgressive parts of sequences and cycles oil production (Hull and Warman, 1970; McQuillan,
(Smith et al., 2004). The other localities included in 1985). A striking feature is the development of tight

Ehrenberg et al. 105


limestone horizons underlying surfaces C, D, and E, stones form laterally extensive flow barriers that divide
where core observations and stable isotope analyses re- the formations into stratigraphic compartments. Under-
veal the presence of major subaerial exposure horizons. standing the diagenetic processes within a sequence-
stratigraphic context is therefore important to produc-
Factors Controlling Porosity tion strategy and reservoir modeling ( Watts et al.,
1994; Smith et al., 2004; Westphal et al., 2004).
All of the five carbonate platforms studied show evi-
dence of major heterogeneity in porosity and perme- Origin of Tight Limestone Barriers
ability related to the preburial factors of depositional In the Finnmark Platform strata, geochemical and
texture and eogenetic diagenesis, including the crea- fluid-inclusion data indicate that porosity-occluding
tion of moldic porosity by early leaching. Certainly, calcite cementation of grain-dominated limestones
these factors are solely responsible for the wide hetero- and buildup facies occurred mainly during burial (Eh-
geneity in the Bahamas and Marion Plateau strata. Su- renberg et al., 2002). Furthermore, correspondence of
perimposed on this early heterogeneity, however, the low-porosity intervals with both higher bulk rock alu-
deeply buried platforms display an important, strati- mina contents and greater frequency of stylolites indi-
graphically focused phase of burial calcite cementa- cates that the burial cementation responsible for porosi-
tion that dominates their ultimate petrophysical dif- ty loss was sourced mainly from chemical compaction
ferentiation and resulting reservoir architecture. along stylolites (Ehrenberg, 2004b).
Madison outcrops and cores have not been exam-
Stratigraphic Controls ined in this study to evaluate the presence of stylolites
In each of the deeply buried platforms, the porosity- and their distribution relative to limestone and dolo-
depth profiles show that the tight limestones are clus- stone intervals. Petrographic study of a selection of
tered in distinct zones alternating with more porous low-porosity Madison limestones, however, suggests
intervals. In large parts of the sections, these porosity that cementation by coarse calcite spar is the major
alternations are linked with the high-frequency depo- cause of porosity loss, consistent with an origin by
sitional cyclicity that is a fundamental characteristic of chemical compaction during burial, as also concluded
these strata, and much of the total porosity variance by Crockett (1994) for the tight limestones of Madison
therefore lies within the scale of the high-frequency sequence IV. Late-stage calcite cementation, however,
cycles (Ehrenberg, 2004b). This cycle-scale control of can also be attributed to precipitation following hy-
porosity variation may be a general characteristic of drothermal brecciation during the Laramide orogeny
carbonate reservoirs, whatever the degree of dolomi- (Westphal et al., 2004; Katz et al., in press).
tization (Eisenberg et al., 1994; Jennings et al., 1998). For the Asmari data set, stylolite counts done using
There is commonly strong porosity partitioning be- core photographs show a strong relationship between
tween transgressive and regressive hemicycles, but stylolite frequency and the low-porosity limestone
both the polarity of this partitioning (whether the tops intervals (Ehrenberg, 2004, personal observations).
or bases of cycles are porous) and the facies involved Petrographic examination of low-porosity Asmari lime-
vary greatly both between and within the different stones shows that the dominant causes of porosity loss
platforms. Based principally on the Madison example, in most samples are (1) tight calcite microcementation
the nature of cycle-scale partitioning is believed to de- of lime mud matrix, combined with (2) precipitation
pend on both the location within the platform paleo- of coarse calcite spar. This coarse spar is interpreted
geography and the position within the lower order as a burial cement based on fluid-inclusion data (N. H.
system tracts (Smith et al., 2004). Oxtoby, 2004, personal communication) and because it
Larger scale trends of porosity variation appear re- postdates all associated dolomite. It displays uniformly
lated to both third-order and second-order sequences. dull, commonly sector-zoned cathodoluminescence, as
Thus, groups of successive parasequences tend to have is characteristic of burial-related calcite cements.
overall higher or lower porosity depending on their posi- Spatial coincidence of low-porosity, mud-dominated
tion with respect to transgressive and highstand systems beds with abundant stylolites suggests that chemical
tracts. Once again, however, the polarity of these larger compaction causes cementation of the mud matrix in
scale porosity trends is highly variable from case to case. addition to cementation by coarse spar.
Because of the inferred link between porosity and We suggest that chemical compaction of lime-
stratigraphic cyclicity, it is apparent that the tight lime- stone is, in general, activated by increasing temperature

106 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


in the same general way as in quartzose sandstone into the rock as ions in solution. Because of the lower
(Bjørkum et al., 1998) but at much lower temperatures molar volume of dolomite, porosity increases by
(Dunnington, 1967; Meyers and Hill, 1983). Intensity roughly 12% of the volume of the calcium carbonate
of cementation is expected to be proportional to the replaced. This is probably the oldest explanation for
abundance and proximity of stylolites, as has been ob- dolostone porosity but was first advanced cogently by
served in several previous case studies ( Wong and Weyl (1960), who noted that most subsurface water
Oldershaw, 1981; Purser, 1984; Koepnick, 1987). Fur- analyses contain abundant dissolved magnesium
thermore, stylolite frequency and consequent intensity (facilitating the supply of Mg and the removal of
of cementational porosity loss in the limestone intervals Ca) but relatively little bicarbonate (limiting mass
is determined in large degree by the stratigraphic dis- transport of carbonate).
tribution of clay minerals, which act to facilitate and 2. The dissolution-of-undolomitized-calcium-carbonate
localize chemical dissolution in carbonates (Weyl, 1959; model envisions overall porosity increase caused by
Oldershaw and Scoffin, 1967; Marshak and Engelder, the dissolution and export of residual, undolomitized
1985; Choquette and James, 1987; Brown, 1997; Saller calcium carbonate during or after partial dolomiti-
et al., 1999) as in quartz sandstones (Bjørkum, 1996). zation. This dissolution may occur either as an integral
In the three deeply buried examples, porous dolo- part of the dolomitization process (Morrow, 1982,
stone intervals are interlayered with tight limestone 2001; Sun, 1995) or by later meteoric or marine cir-
barriers, where calcite cementation reflects chemical culation unrelated to dolomitization (Moore, 2001).
compaction along stylolites. Most of the dissolved car- 3. The dolomite-cementation model regards the bulk-
bonate released along the stylolites therefore appears volume porosity present in dolostone as being equal
to have precipitated locally within the same beds where to or less than that of the precursor limestone (Lucia
it was generated. Precipitation occurred before the dis- and Major, 1994; Lucia, 2004). Precipitation of do-
solved calcite was able to diffuse outward into the sur- lomite cement from carbonate transported into the
rounding dolostones. The resulting compartmentali- rock in solution counters any porosity gain resulting
zation of carbonate reservoirs by formation of tight from the lower molar volume of dolomite and com-
limestone barriers during burial is a fundamental dif- monly reduces porosity further.
ference from quartzose sandstone reservoirs, where bar-
riers are mainly formed by shaly layers present from the All three models acknowledge that a major component
time of deposition and where quartz cementation (also of dolostone porosity is inherited from the precursor
a temperature-dependent burial process related to limestones. Such inherited pores may be either pre-
stylolitization) tends to result in relatively uniform po- served in their original shapes during fabric-preserving
rosity loss throughout the reservoir interval ( Walder- dolomitization or rearranged during fabric-destructive
haug and Bjørkum, 2003). The spatial association be- dolomitization. The differences between the models
tween stylolite occurrence and calcite cementation in concern the nature of the dominant process that either
the same beds (also noted by Wong and Oldershaw, increases or decreases this inherited pore volume.
1981) suggests that the solutes released by stylolitiza- Unfortunately, it is difficult to devise meaningful
tion tend not to be exported far from their point of tests for distinguishing between the above models,
origin, implying that solute supply, instead of transport either in general or with respect to any particular doc-
or precipitation (as in quartzose sandstones), may be lostone sample or reservoir. This is partly because the
the rate-limiting step for burial cementation of lime- processes involved are not mutually exclusive, and the
stones (Oelkers et al., 1996). models are thus differentiated only on the basis of which
process is dominant. Dolomitization is closed to car-
Origin of Dolostone Porosity bonate mass transport in model 1 and open in models 2
Explanations that have been advanced over the past and 3, but textural evidence is of little help in differ-
45 yr and more for the porosity present in dolostones entiating between these alternatives. Murray (1960)
can be summarized in terms of three ideal models: described evidence for extensive dissolution of resid-
ual, undolomitized calcium carbonate components in
1. The Mole-for-mole-replacement model assumes con- dolostone reservoirs and apparently viewed this as re-
servation of CO2 during the formation of dolomite by sulting from local redistribution instead of export of the
the replacement of roughly half the calcium present in carbonate. Others have viewed similar textural evidence
calcium carbonate by magnesium that is transported and inferred porosity increase by carbonate export

Ehrenberg et al. 107


(Dawans and Swart, 1988; Choquette et al., 1992; Preferential preservation of dolostone porosity may
Vahrenkamp and Swart, 1994; Saller and Henderson, primarily reflect a greater resistance of dolostone to
1998; Eichenseer et al., 1999; Ehrenberg, 2004a, b). In chemical compaction and resulting cementation, as has
this study, textural evidence for dissolution of residual been variously suggested by Choquette and Steinen
calcium carbonate in the Marion Plateau, Finnmark, (1980), Schofield (1984), Ruzyla and Friedman (1985),
and Asmari dolostones includes (1) the common pres- Railsback (1993), Amthor et al. (1994), Purser et al.
ence in the dolostones of molds after bioclast types that (1994), and Brown (1997). Such differential resistance
rarely show signs of dissolution in the associated lime- is tentatively supported in the Finnmark and Asmari
stones (especially large benthic foraminifers and echi- cores by the common presence of stylolites in the tight
noderms) and (2) the occurrence of sucrosic dolomite limestones compared with the relatively infrequent
fabric, which has been interpreted as the product of stylolite occurrence in dolostone intervals. However,
leaching of partly dolomitized lime mud (Murray, 1960; other factors besides carbonate mineralogy could be
Dawans and Swart, 1988). It is impossible to determine, critical for controlling stylolite frequency, in particular
however, whether the dissolved calcite was mainly the distribution of depositional clay. Differential resis-
precipitated locally as dolomite or mainly exported. tance to mechanical compaction is also a possible ex-
A fundamental problem with attempting to test planation for the lower porosities of deeply buried
the relative correctness of the above models is uncer- limestones (Lucia, 2004). Few limestones in the Finn-
tainty regarding the porosity of the precursor lime- mark and Asmari cores, however, show petrographic
stones. Even in fairly large limestone and dolostone data evidence of extensive mechanical compaction, possibly
sets, such as for the Seroe Domi Formation of Bonaire because of effective inhibition by early calcite cement.
(Lucia and Major, 1994) or the Marion Plateau cores
(Figure 4), few real constraints exist regarding the
degree to which the limestones analyzed are statis- PERMEABILITY
tically representative of the limestone precursors of
the presently associated dolostones. Data
Despite these uncertainties, the similar porosity
distributions in limestones and dolostones of the shal- Permeability distributions of associated limestones and
lowly buried strata (Figures 3, 4) are at least broadly dolostones have not been examined independently from
consistent with the conclusion that shallow dolomi- porosity because permeability is fundamentally depen-
tization does not result in significant overall increase dent on total porosity in sedimentary rocks (Nelson,
in carbonate porosity (Lucia and Major, 1994; Lucia, 1994). It is more meaningful to compare different sam-
2004). The contrastingly much higher porosity distri- ple groups in terms of permeability-for-given-porosity
butions of the deeply buried Finnmark, Madison, and and overall permeability-porosity trends, instead of dis-
Asmari dolostones, compared with their respective as- cussing permeability independently from porosity (Eh-
sociated limestones (Figures 5 –7), therefore indicate renberg and Nadeau, 2005). Variations in permeability-
that dolostone porosity was either (1) increased or for-given-porosity depend on factors such as surface area
(2) preferentially preserved during burial diagenesis. per unit volume and tortuosity (Nelson ,1994), which
As described above, paucity of low porosity values are, in turn, determined by rock fabric and pore geome-
in the dolostones of the deeply buried data sets is caused try (Lucia, 1995; Anselmetti et al., 1998).
by the ubiquitous presence of intercrystalline macro- The two data sets from shallowly buried carbonate
porosity and the common presence of moldic pores. platforms both show distinctly higher permeability-
These two types of secondary porosity are interpreted for-given-porosity in dolostones compared with as-
as being predominantly of near-surface origin in the sociated limestones (Figure 13). In each case, this has
studied strata because of a general lack of evidence for been interpreted as reflecting the replacement of
dissolution during burial. For example, the dolostones microporous lime mud by dolomite having intercrys-
commonly contain both coarse dolomite cement and talline macroporosity (Melim et al., 2001a; Ehren-
subsequent (postdolomite) coarse calcite cement but berg, 2004a). In contrast, the three deeply buried
rarely, if ever, show signs of late dissolution of these data sets do not show major differences between
cements. Consequently, it is the relative preservation of limestones and dolostones in terms of either overall
dolostone porosity during burial that is important for trends of permeability-porosity correlation or average
producing the contrasts shown by Figures 5–7. permeability-for-given-porosity (Figures 14 – 16).

108 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Trends of permeability-porosity correlation (diagonal zation. Similarly, it may not be correct to infer from
lines in Figures 14–16) were calculated after excluding the comparison in Figure 13 that dolomitization was
samples with permeability of 0.01 md because these are necessarily an effective agent for permeability increase
assumed to represent noneconomic zones whose con- in the shallowly buried data sets. This would not be
tribution to the overall porosity-permeability relation- true if, for example, the dolomitized strata in these
ship is of no interest. For the Asmari trends, samples with cases were generally more reefal or coarser grained than
porosity less than 4% were also excluded because many the presently associated limestones.
of these have exceptionally high permeabilities, believed All that can be reasonably concluded from the
to reflect the presence of fractures in the plugs measured. present comparisons, therefore, is the purely empirical
recommendation that dolostones should not be ex-
Interpretation pected to have a better permeability-porosity rela-
tionship than associated limestones in deeply buried
It might easily be concluded from the comparisons reservoirs. Nevertheless, a possible explanation for the
in Figures 13 – 16 that dolomitization has improved obvious contrast between the shallowly buried and
porosity-permeability relationships in the shallowly deeply buried data sets (Figures 13–16) is that lime-
buried platforms but had no such effect in the deeply stones already having a lower permeability at shallow
buried strata. These comparisons are potentially mis- depth are systematically prone to severe porosity and
leading, however, because it is, in most cases, im- permeability loss during burial, such that they end up
possible to know the degree to which the dolostones with essentially zero reservoir quality in the deeper
were originally similar to the associated limestones. subsurface and are thus effectively removed from the
Therefore, it may not be warranted to conclude from comparisons in Figures 14–16. The result would be
Figures 14 – 16 that dolomitization does not improve that the deeply buried carbonates with greater than 4%
permeability. This would not be true if, for example, porosity and greater than 0.01-md permeability include
the dolostones in the deeply buried examples were only those limestones that had relatively high perme-
systematically muddier, lower permeability sediments ability to begin with, along with nearly all of the do-
to begin with than the strata that escaped dolomiti- lostones (Figure 17).

Figure 17. Hypothetical model for the effect of burial diagenesis on permeability-porosity relationships in interlayered limestones
(circles) and early-formed dolostones (diamonds). At shallow burial depth (A), the limestones and dolostones do not have notably
different porosity distribution, but all dolostones have high permeability (black diamonds), whereas the limestones are divisible
into strata with high permeability (black circles) and low permeability (green circles). After deep burial (B), the limestones initially
having low permeabilities have been most strongly affected by burial diagenesis, resulting in postburial permeabilities of 0.01 md
or less. Limestones initially having high permeabilities have been less severely affected because of more grain-supported textures
relatively resistant to permeability loss. As a result, those limestones and dolostones having greater than 0.01 md permeability in
the deeper subsurface do not have notably different overall porosity-permeability relationships.

Ehrenberg et al. 109


Figure 18. Cartoon illustrating concepts of differential porosity development in interlayered limestone-dolostone reservoirs.
Responses of five different lithologies are schematically shown as beds or intervals: (1) limestone prone to stylolite development
because of depositional clay distribution; (2) limestone with little potential for stylolitization (minor depositional clay content, assumed
to be necessary for the initiation of stylolite development); (3) early-formed dolostone with high initial porosity; (4) dolostone having
low to medium porosity because of early dolomite cementation; (5) dolostone with extensive early calcium-sulfate cementation. Before
burial deeper than about 1 km (0.6 mi) (left), these depositionally heterogeneous, shallow-water carbonate strata have widely varying
but generally high porosity. Dolomitized layers may have overall higher or lower porosity than limestones, depending on various factors
that are, for the most part, not subject to testing with empirical data (absence of carbonate mass transport during dolomitization, amount
of dolomite cement imported, amount of calcium carbonate removed after partial dolomitization). After deeper burial, the stratigraphic
variation in porosity loss is shown by the difference (shaded) between the before and after porosity curves in the far-right column.
The most severe porosity loss has occurred in limestone beds prone to chemical compaction and associated local calcite cemen-
tation (1). Certain limestone intervals, however, retain primary porosity because of limited chemical compaction (2). Stylolite
formation also causes porosity loss in dolostone intervals but to lesser degree than in most limestones. Dolostones with both high (3)
and low (4) early porosity form flow units, although some intervals with especially low early porosity and beds cemented by calcium
sulfate (5) form barriers. The ultimate architecture of barriers and flow units is thus largely dependent on the distribution of prestylolite
dolomitization, as well as clay and calcium sulfate.

CONCLUSIONS buried examples and three deeply buried examples.


The results support the often-reported pattern of
Compilations of porosity-permeability data have been preferential porosity preservation in dolostones during
examined to compare associated limestones and dolo- burial cementation. A conceptual model for the key
stones from five carbonate platforms of widely varying processes involved is shown in Figure 18, illustrat-
ages and geographic settings, including two shallowly ing the hypothetical porosity response of different

110 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


lithologies to burial diagenesis. Flow barriers consist- in the Madison Group, Wyoming and Utah overthrust belt:
ing of tight limestone are created during burial by AAPG Bulletin, v. 71, p. 909 – 924.
Cantrell, D. L., P. Swart, and R. M. Hagerty, 2004, Genesis and
chemical compaction and resulting cementation. This characterization of dolomite, Arab-D reservoir, Ghawar field,
process is stratigraphically focused by factors related Saudi Arabia: GeoArabia, v. 9, p. 11 – 36.
to sequence architecture, including the occurrence Chave, K. E., 1954, Aspects of the biogeochemistry of magnesium:
2. Calcareous sediments and rocks: Journal of Geology, v. 62,
of intervals with clay laminations ( promoting chemi- p. 287 – 599.
cal compaction and associated cementation) and the Chilingar, G. V., and R. D. Terry, 1954, Relationship between po-
distribution of early dolomitization (promoting po- rosity and chemical composition of carbonate rocks: The Pe-
troleum Engineer, v. 26, p. 341 – 342.
rosity preservation during burial). Despite higher per-
Choquette, P. W., and N. P. James, 1987, Diagenesis 12. Diagenesis
meability because of higher porosity, however, the in limestones: 3. The deep burial environment: Geoscience
deeply buried dolostones do not, in general, have higher Canada, v. 14, p. 3 – 35.
permeability-for-given-porosity than associated lime- Choquette, P. W., and R. P. Steinen, 1980, Mississippian non-
supratidal dolomite, Ste. Genevieve Limestone, Illinois basin:
stones (Figures 14 – 16). Evidence for mixed water dolomitization, in D. H. Zenger, J. B.
Dunham, and R. L. Ethington, eds., Concepts and models of
dolomitization: SEPM Special Publication 28, p. 163 – 196.
Choquette, P. W., and R. P. Steinen, 1985, Mississippian oolite and
REFERENCES CITED non-supratidal dolomite reservoirs in the Ste. Genevieve For-
mation, North Bridgeport field, Illinois basin, in P. O. Roehl and
Allan, J. R., and W. D. Wiggins, 1993, Dolomite reservoirs — P. W. Choquette, eds., Carbonate petroleum reservoirs: New
Geochemical techniques for evaluating origin and distribution: York, Springer-Verlag, p. 207 – 225.
AAPG Short Course Note Series 36, 129 p. Choquette, P. W., A. Cox, and W. J. Meyers, 1992, Characteristics,
Amthor, J. E., E. W. Mountjoy, and H. G. Machel, 1994, Regional- distribution and origin of porosity in shelf dolostones: Burlington-
scale porosity and permeability variations in Upper Devo- Keokuk Formation (Mississippian), U.S. mid-continent: Journal
nian Leduc buildups: Implications for reservoir development of Sedimentary Petrology, v. 62, p. 167 – 189.
and prediction in carbonates: AAPG Bulletin, v. 78, p. 1541 – Clark, D., J. Heaviside, and K. Habib, 2004, Reservoir properties of
1559. Arab carbonates, Al Rayyan field, offshore Qatar, in C. J. R.
Anselmetti, F. S., and G. P. Eberli, 2001, Sonic velocity in Braithwaite, G. Rizzi, and G. Darke, eds., The geometry and
carbonates — A combined product of depositional lithology petrogenesis of dolomite hydrocarbon reservoirs: Geological
and diagenetic alterations, in R. N. Ginsburg, ed., Subsurface Society (London) Special Publication 235, p. 193 – 232.
geology of a prograding carbonate platform margin, Great Crockett, J., 1994, Porosity evolution of the Madison Limestone,
Bahama Bank: Results of the Bahamas Drilling Project: SEPM Mississippian: Wind River basin, Wyoming: M.S. thesis, Loui-
Special Publication 70, p. 193 – 216. siana State University, Baton Rouge, Louisiana, 103 p.
Anselmetti, F. S., S. Luthi, and G. P. Eberli, 1998, Quantitative Davies, G. R., 1979, Dolomite reservoir rocks: Processes, controls,
characterization of carbonate pore systems by digital image porosity development, in C. H. Moore, ed., Geology of car-
analysis: AAPG Bulletin, v. 82, p. 1815 – 1836. bonate porosity: AAPG Continuing Education Course Note
Aqrawi, A. A. M., M. Keramati, N. Pickard, G. Darke, A. Moallemi, Series 11, p. C1 – C17.
and T. A. Svånå, 2004, Dolomitization of the subsurface Asmari Dawans, J. M., and P. K. Swart, 1988, Textural and geochemical
Formation (Oligocene – lower Miocene), SW Iran: Suggested alternations in late Cenozoic Bahamian dolomites: Sedimen-
models and reservoir quality controls (abs.): GeoArabia, v. 9, tology, v. 35, p. 385 – 403.
p. 40 – 42. Dunnington, H. V., 1967, Aspects of diagenesis and shape change in
Bjørkum, P. A., 1996, How important is pressure in causing dis- stylolitic limestone reservoirs, in Proceedings of the Seventh
solution of quartz in sandstones?: Journal of Sedimentary Re- World Petroleum Congress, Mexico City: New York, Elsevier,
search, v. 66, p. 147 – 154. v. 2, p. 339 – 352.
Bjørkum, P. A., E. H. Oelkers, P. H., Nadeau, O. Walderhaug, and Eberli, G. P., F. S. Anselmetti, J. A. M. Kenter, D. F. McNeill, and
W. M. Murphy, 1998, Porosity prediction in quartzose sand- L. A. Melim, 2001, Calibration of seismic sequence stratigra-
stones as a function of time, temperature, depth, stylolite fre- phy with cores and logs, in R. N. Ginsburg, ed., Subsurface
quency, and hydrocarbon saturation: AAPG Bulletin, v. 82, geology of a prograding carbonate platform margin, Great
p. 637 – 648. Bahama Bank: Results of the Bahamas Drilling Project: SEPM
Blatt, H., G. Middleton, and R. Murray, 1980, Origin of sedi- Special Publication 70, p. 241 – 265.
mentary rocks: New Jersey, Prentice Hall, 782 p. Ehrenberg, S. N., 2004a, Porosity and permeability in Miocene car-
Braithwaite, C. J. R., G. Rizzi, and G. Darke, eds., 2004, The bonate platforms on the Marion Plateau, offshore NE Australia:
geometry and petrogenesis of dolomite hydrocarbon reservoirs: Relationships to stratigraphy, facies and dolomitization, in
Geological Society (London) Special Publication 235, 413 p. C. J. R. Braithwaite, G. Rizzi, and G. Darke, eds., The geometry
Brown, A., 1997, Porosity variation in carbonates as a function of and petrogenesis of dolomite hydrocarbon reservoirs: Geologi-
depth: Mississippian Madison Group, Williston basin, in J. A. cal Society (London) Special Publication 235, p. 233 – 253.
Kupecz, J. Gluyas, and S. Bloch, eds., Reservoir quality predic- Ehrenberg, S. N., 2004b, Factors controlling porosity in Upper
tion in sandstones and carbonates: AAPG Memoir 69, p. 29 – 46. Carboniferous – Lower Permian carbonate strata of the Barents
Budai, J. M., K. C. Lohmann, and R. M. Owen, 1984, Burial dedolo- Sea: AAPG Bulletin, v. 88, p. 1653 – 1676.
mitization in the Mississippian Madison Limestone, Wyoming Ehrenberg, S. N., and P. H. Nadeau, 2005, Sandstone versus car-
and Utah thrust belt: Journal of Sedimentary Petrology, v. 54, bonate petroleum reservoirs: A global perspective on porosity-
p. 276 – 288. depth and porosity-permeability relationships: AAPG Bulletin,
Budai, J. M., K. C. Lohmann, and J. L. Wilson, 1987, Dolomitization v. 89, p. 435 – 445.

Ehrenberg et al. 111


Ehrenberg, S. N., E. B. Nielsen, T. A. Svånå, and L. Stemmerik, Katz, D. A., G. P. Eberli, P. K. Swart, L. B. Smith, Jr., and J. Kislak,
1998a, Depositional evolution of the Finnmark carbonate plat- in press, Hydrothermal brecciation associated with calcite
form, Barents Sea: Results from wells 7128/6-1 and 7128/4-1: precipitation and permeability destruction in Mississippian
Norsk Geologisk Tidsskrift, v. 78, p. 185 – 224. carbonate reservoirs, Montana and Wyoming: AAPG Bulletin.
Ehrenberg, S. N., E. B. Nielsen, T. A. Svånå, and L. Stemmerik, Koepnick, R. B., 1987, Distribution and permeability of stylolite-
1998b, Diagenesis and reservoir quality of the Finnmark bearing horizons within a Lower Cretaceous carbonate reser-
carbonate platform, Barents Sea: Results from wells 7128/6-1 voir in the Middle East: Society of Petroleum Engineers Paper
and 7128/4-1: Norsk Geologisk Tidsskrift, v. 78, p. 225 – 251. 14173, SPE Formation Evaluation, v. 2, p. 137 – 142.
Ehrenberg, S. N., N. A. H. Pickard, T. A. Svånå, and N. H. Oxtoby, Landes, K. K., 1946, Porosity through dolomitization: AAPG
2002, Cement geochemistry of photozoan carbonate strata Bulletin, v. 30, p. 305 – 318.
(Upper Carboniferous – Lower Permian), Finnmark carbonate Lees, G. M., 1933, Reservoir rocks of Persian oil fields: AAPG
platform, Barents Sea: Journal of Sedimentary Research, v. 72, Bulletin, v. 17, p. 229 – 240.
p. 95 – 115. Lindsay, R. F., and C. G. St. C. Kendall, 1985, Depositional facies,
Ehrenberg, S. N., J. M. McArthur, M. F. Thirlwall, N. A. H. diagenesis, and reservoir character of Mississippian cyclic
Pickard, and G. V. Laursen, 2004a, Strontium-isotope dating carbonates in the Mission Canyon Formation, Little Knife
of the Asmari Formation (Oligocene – Miocene) in Ahwaz, field, Williston basin, North Dakota, in P. O. Roehl and P. W.
Bibi Hakimeh and oil fields, southwestern Iran (abs.): Geo- Choquette, eds., Carbonate petroleum reservoirs: New York,
Arabia, v. 9, p. 61. Springer-Verlag, p. 175 – 190.
Ehrenberg, S. N., G. P. Eberli, and G. L. Bracco Gartner, 2004b, Longman, M. W., T. G. Fertal, and J. S. Glennie, 1983, Origin and
Data report: Porosity and permeability of Miocene carbonate geometry of Red River dolomite reservoirs, western Williston
platforms on the Marion Plateau, ODP Leg 194, in F. S. basin: AAPG Bulletin, v. 67, p. 744 – 771.
Anselmetti, A. R. Isern, P. Blum, and C. Betzler, eds., Pro- Longman, M. W., T. G. Fertal, and J. R. Stell, 1992, Reservoir
ceedings of the Ocean Drilling Program, Scientific Results, performance in Ordovician Red River Formation, Horse Creek
v. 194, available at: http://www-odp.tamu.edu/publications and South Horse Creek fields, Bowman County, North Dakota:
/194_SR/007/007.htm. AAPG Bulletin, v. 76, p. 449 – 467.
Ehrenberg, S. N., J. M. McArthur, and M. F. Thirlwall, in press, Lucia, F. J., 1962, Diagenesis of a crinoidal sediment: Journal of
Growth, demise, and dolomitization of Miocene carbonate Sedimentary Petrology, v. 32, p. 848 – 865.
platforms on the Marion Plateau, offshore NE Australia: Journal Lucia, F. J., 1995, Rock-fabric/petrophysical classification of car-
of Sedimentary Research. bonate pore space for reservoir characterization: AAPG Bulle-
Eichenseer, H. T., F. R. Walgenwitz, and P. J. Biondi, 1999, tin, v. 79, p. 1275 – 1300.
Stratigraphic control on facies and diagenesis of dolomitized Lucia, F. J., 2004, Origin and petrophysics of dolostone pore space,
oolitic siliciclastic ramp sequences (Pinda Group, Albian, in C. J. R. Braithwaite, G. Rizzi, and G. Darke, eds., The
offshore Angola): AAPG Bulletin, v. 83, p. 1729 – 1758. geometry and petrogenesis of dolomite hydrocarbon reser-
Eisenberg, R. A., P. M. Harris, C. W. Grant, D. J. Goggin, and E. J. voirs. Geological Society (London) Special Publication 235,
Conner, 1994, Modeling reservoir heterogeneity within outer p. 141 – 155.
ramp carbonate facies using an outcrop analog, San Andres For- Lucia, F. J., and R. P. Major, 1994, Porosity evolution through
mation of the Permian basin: AAPG Bulletin, v. 78, p. 1337 – hypersaline reflux dolomitization, in B. Purser, M. Tucker, and
1359. D. Zenger, eds., Dolomites, a volume in honour of Dolomieu:
Glover, J. E., 1968, Significance of stylolites in dolomitic lime- International Association of Sedimentologists Special Publica-
stones: Nature, v. 217, p. 835 – 836. tion 21, p. 325 – 341.
Hohlt, R. B., 1948, The nature and origin of limestone porosity: Mancini, E. A., T. A. Blasingame, R. Archer, B. J. Panetta, J. C.
Colorado School of Mines Quarterly, v. 43, no. 4, 51 p. Llinás, C. D. Haynes, and D. J. Benson, 2004, Improving
Hopkins, J. C., 1999, Characteristics of reservoir lithologies recovery from mature oil fields producing from carbonate
within sub-unconformity pools: Pekisko Formation, Medicine reservoirs: Upper Jurassic Smackover Formation, Womack
River field, Alberta, Canada: AAPG Bulletin, v. 83, p. 1855 – Hill field (eastern Gulf Coast, U.S.A.): AAPG Bulletin, v. 88,
1870. p. 1629 – 1651.
Hopkins, J. C., 2004, Geometry and origin of dolomudstone Marshak, S., and T. Engelder, 1985, Development of cleavage in
reservoirs: Pekisko Formation (Lower Carboniferous), western limestones of a fold-thrust belt in eastern New York: Journal of
Canada, in C. J. R. Braithwaite, G. Rizzi, and G. Darke, eds., Structural Geology, v. 7, p. 345 – 359.
The geometry and petrogenesis of dolomite hydrocarbon res- McQuillan, H., 1985, Fracture-controlled production from the
ervoirs: Geological Society (London) Special Publication 235, Oligo-Miocene Asmari Formation in Gachsaran and Bibi
p. 349 – 366. Hakimeh fields, southwest Iran. in P. O. Roehl and P. W.
Hull, C. E., and H. R. Warman, 1970, Asmari oil fields of Iran, Choquette, eds., Carbonate petroleum reservoirs: New York,
in M. T. Halbouty, ed., Geology of giant petroleum fields: Springer-Verlag, p. 511 – 523.
AAPG Memoir 14, p. 428 – 437. Melim, L. A., F. S. Anselmetti, and G. P. Eberli, 2001a, The
Institut Français du Pétrole, 1959, Relations entre mode de gise- importance of pore type on permeability of Neogene carbo-
ment et propriétés physicochimiques des dolomies: Revue de nates, Great Bahama Bank, in R. N. Ginsburg, ed., Subsurface
l’Institute Français du Pétrole, v. 14, p. 475 – 518. geology of a prograding carbonate platform margin, Great
Isern, A. R., et al., 2002, Proceedings of the Ocean Drilling Pro- Bahama Bank: Results of the Bahamas Drilling Project: SEPM
gram, Initial Reports: Ocean Drilling Program, Texas A&M Special Publication 70, p. 217 – 238.
University, College Station, Texas, v. 194, 88 p. Melim, L. A., P. K. Swart, and R. G. Maliva, 2001b, Meteoric and
Jennings, J. W., S. C. Ruppel, and W. B. Ward, 1998, Geostatistical marine-burial diagenesis in the subsurface of Great Bahama
analysis of petrophysical data and modeling of fluid-flow Bank, in R. N. Ginsburg, ed., Subsurface geology of a prograd-
effects in carbonate outcrops: Society of Petroleum Engineers ing carbonate platform margin, Great Bahama Bank: Results of
Annual Technical Conference and Exhibition, New Orleans, the Bahamas Drilling Project: SEPM Special Publication 70,
September 27 – 30, SPE Paper 49025, 16 p. p. 137 – 161.

112 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


Meyers, W. J., and B. E. Hill, 1983, Quantitative studies of com- Saller, A. H., J. A. D. Dickson, and F. Matsuda, 1999, Evolution and
paction in Mississippian skeletal limestones, New Mexico: distribution of porosity associated with subaerial exposure in
Journal of Sedimentary Petrology, v. 53, p. 231 – 242. upper Paleozoic platform limestones, west Texas: AAPG
Moore, C. H., 2001, Carbonate reservoirs porosity evolution and Bulletin, v. 83, p. 1835 – 1854.
diagenesis in a sequence stratigraphic framework: Amsterdam, Saller, A. H., S. Walden, S. Robertson, R. Nims, J. Schwab, H.
Elsevier, 444 p. Hagiwara, and S. Mizohata, 2004, Three-dimensional seismic
Morgan, W. A., 1985, Silurian reservoirs in upward-shoaling cycles imaging of an upper Paleozoic ‘‘reefal’’ buildup, Reinecke field,
of the Hunton Group, Mt. Everettte and Southwest Reeding west Texas, United States, in G. P. Eberli, J. L. Massaferro, and
fields, Kingfisher County, Oklahoma, in P. O. Roehl and P. W. R. F. Sarg, eds., Seismic imaging of carbonate reservoirs and
Choquette, eds., Carbonate petroleum reservoirs: New York, systems: AAPG Memoir 81, p. 107 – 122.
Springer-Verlag, p. 107 – 120. Schmidt, V., 1965, Facies, diagenesis, and related reservoir
Morrow, D. W., 1982, Diagenesis 1. Dolomite — Part 1: The chem- properties in the Gigas beds (Upper Jurassic), northwestern
istry of dolomitization and dolomite precipitation: Geoscience Germany, in L. C. Pray and R. C. Murray, eds., Dolomitization
Canada, v. 9, p. 5 – 13. and limestone diagenesis; a symposium: SEPM Special Pub-
Morrow, D. W., 2001, Distribution of porosity and permeability in lication 13, p. 124 – 168.
platform dolomites: Insight from the Permian of west Texas: Schmoker, J. W., and R. B. Halley, 1982, Carbonate porosity versus
Discussion: AAPG Bulletin, v. 85, p. 525 – 529. depth: A predictable relation for south Florida: AAPG Bulletin,
Mountjoy, E. W., and X. M. Marquez, 1997, Predicting reservoir v. 66, p. 2561 – 2570.
properties in dolomites: Upper Devonian Leduc buildups, Schmoker, J. W., K. B. Krystinik, and R. B. Halley, 1985, Selected
deep Alberta basin, in J. A. Kupecz, J. Gluyas, and S. Bloch, characteristics of limestone and dolomite reservoirs in the
eds., Reservoir quality prediction in sandstones and carbonates: United States: AAPG Bulletin, v. 69, p. 733 – 741.
AAPG Memoir 69, p. 267 – 306. Schofield, K., 1984, Are pressure solution, neomorphism and
Murray, R. C., 1960, Origin of porosity in carbonate rocks: Journal dolomitization genetically related? in Stylolites and associated
of Sedimentary Petrology, v. 30, p. 59 – 84. phenomena: Relevance to hydrocarbon reservoirs: Abu Dhabi
Nelson, P. H., 1994, Permeability-porosity relationships in sedi- National Reservoir Research Foundation Special Publication,
mentary rocks: The Log Analyst, May-June, p. 38 – 62. p. 183 – 202.
Oelkers, E. H., P. A. Bjørkum, and W. M. Murphy, 1996, A petro- Sears, S. O., and F. J. Lucia, 1980, Dolomitization of northern
graphic and computational investigation of quartz cementation Michigan Niagara reefs by brine refluxion and freshwater/
and porosity reduction in North Sea sandstones: American seawater mixing, in D. H. Zenger, J. B. Dunham, and R. L.
Journal of Science, v. 296, p. 1 – 28. Ethington, eds., Concepts and models of dolomitization:
Oldershaw, A. E., and T. P. Scoffin, 1967, The source of ferroan and SEPM Special Publication 28, p. 215 – 235.
non-ferroan calcite cements in the Halkin and Wenlock Smith, L. B. Jr., G. P. Eberli, and M. Sonnenfeld, 2004, Sequence
limestones: Geological Journal, v. 5, p. 309 – 320. stratigraphic and paleogeographic distribution of reservoir-
Powers, R. W., 1962, Arabian Upper Jurassic carbonate reservoir quality dolomite, Madison Formation, Wyoming and Montana,
rocks, in W. E. Ham, ed., Classification of carbonate rocks — A in G. M. Grammer, P. M. Harris, and G. P. Eberli, eds.,
symposium: AAPG Memoir 1, p. 122 – 192. Integration of modern and outcrop analogs in reservoir
Purser, B. H., 1984, Stratiform stylolites and the distribution of modeling: AAPG Memoir 80, p. 67 – 92.
porosity: Examples from the Middle Jurassic limestones of the Sonnenfeld, M. D., 1996, Sequence evolution and hierarchy within
Paris basin, in Stylolites and associated phenomena: Relevance the lower Mississippian Madison Limestone of Wyoming, in
to hydrocarbon reservoirs: Abu Dhabi National Reservoir M. W. Longman and M. D. Sonnenfeld, eds., Paleozoic systems
Research Foundation Special Publication, p. 203 – 217. of the Rocky Mountain region: Rocky Mountain Section SEPM,
Purser, B. H., A. Brown, and D. M. Aissaoui, 1994, Nature, origins p. 165 – 192.
and porosity in dolomites, in B. Purser, M. Tucker, and D. Steidtmann, E., 1917, Origin of dolomite as disclosed by stains and
Zenger, eds., Dolomites, a volume in honour of Dolomieu: other methods: Geological Society of America Bulletin, v. 28,
International Association of Sedimentologists Special Publica- p. 431 – 450.
tion 21, p. 283 – 308. Sun, S. Q., 1995, Dolomite reservoirs: porosity evolution and
Qing, H., and E. W. Mountjoy, 1994, Formation of coarsely reservoir characteristics: AAPG Bulletin, v. 79, p. 186 – 204.
crystalline, hydrothermal dolomite reservoirs in the Presqu’ile Swart, P. K., and L. A. Melim, 2000, The origin of dolomites
barrier, Western Canada sedimentary basin: AAPG Bulletin, in Tertiary sediments from the margin of the Great Ba-
v. 78, p. 55 – 77. hama Bank: Journal of Sedimentary Research, v. 70, p. 738 –
Railsback, L. B., 1993, Lithologic controls on morphology of 748.
pressure-dissolution surfaces (stylolites and dissolution seams) Vahrenkamp, V. C., and P. K. Swart, 1994, Late Cenozoic dolo-
in Paleozoic carbonate rocks from the mideastern United mites of the Bahamas: Metastable analogs for the genesis of
States: Journal of Sedimentary Petrology, v. 63, p. 513 – 522. ancient platform dolomites, in B. Purser, M. Tucker, and D.
Ruzyla, K., and G. M. Friedman, 1985, Factors controlling porosity Zenger, eds., Dolomites a volume in honour of Dolomieu:
in dolomite reservoirs of the Ordovician Red River Formation, International Association of Sedimentologists Special Publica-
Cabin Creek field, Montana, in P. O. Roehl and P. W. tion 21, p. 133 – 153.
Choquette, eds., Carbonate petroleum reservoirs: New York, Walderhaug, O., and P. A. Bjørkum, 2003, The effect of stylolite
Springer-Verlag, p. 39 – 58. spacing on quartz cementation in the Lower Jurassic Stø
Saller, A. H., and N. Henderson, 1998, Distribution of porosity and Formation, southern Barents Sea: Journal of Sedimentary Re-
permeability in platform dolomites: Insight from the Permian search, v. 73, p. 146 – 156.
of west Texas: AAPG Bulletin, v. 82, p. 1528 – 1550. Warren, J., 2000, Dolomite: Occurrence, evolution and economi-
Saller, A. H., and K. Yaremko, 1994, Dolomitization and porosity cally important associations: Earth-Science Reviews, v. 52,
development in the middle and upper Wabamun Group, p. 1 – 81.
southeast Peace River arch, Alberta, Canada: AAPG Bulletin, Watts, N. R., M. P. Coppold, and J. L. Douglas, 1994, Application
v. 78, p. 1406 – 1430. of reservoir geology to enhanced oil recovery from Upper

Ehrenberg et al. 113


Devonian Nisku reefs, Alberta, Canada: AAPG Bulletin, v. 78, Devonian Kaybob reef complex, Alberta, Canada: Journal of
p. 78 – 101. Sedimentary Petrology, v. 51, p. 507 – 520.
Westphal, H., G. P. Eberli, L. B. Smith, G. M. Grammer, and Zempolich, W. G., and L. A. Hardie, 1997, Geometry of dolomite
J. Kislak, 2004, Reservoir characterization of the Mississippian bodies within deep-water resedimented oolite of the Vajont
Madison Formation, Wind River basin, Wyoming: AAPG Bul- Limestone, Venetian Alps, Italy: Analogs for hydrocarbon res-
letin, v. 88, p. 405 – 432. ervoirs created through burial dolomitization, in J. A. Kupecz,
Weyl, P. K., 1959, Pressure solution and the force of crystalliza- J. G. Gluyas, and S. Bloch, eds., Reservoir quality prediction in
tion — A phenomenological theory: Journal of Geophysical sandstones and carbonates: AAPG Memoir 69, p. 127 – 162.
Research, v. 64, p. 2001 – 2025. Zhao, W., P. Luo, G. Chen, H. Cao, and B. Zhang, 2005, Origin and
Weyl, P. K., 1960, Porosity through dolomitization: Conservation- reservoir rock characteristics of dolostones in the Early Triassic
of-mass requirements: Journal of Sedimentary Petrology, v. 30, Feixianguan Formation, NE Sichuan basin, China: Significance
p. 85 – 90. for future gas exploration: Journal of Petroleum Geology, v. 28,
Wong, P. K., and A. Oldershaw, 1981, Burial cementation in the p. 83 – 100.

114 Porosity-Permeability Relationships of Interlayered Limestone-Dolostone Reservoirs


View publication stats

You might also like