You are on page 1of 16

Origins of gypsum in AUTHORS

Faith O. Amadi  Department of Geology


deep carbonate reservoirs: and Geological Engineering, University of
Mississippi, University, Mississippi;
Implications for hydrocarbon foamadi@olemiss.edu
Faith Amadi received a master’s degree in fluvial
exploration and production geology from Southeast Missouri State Uni-
versity in May 2004. During the summers of 2005
Faith O. Amadi, R. P. Major, and Lawrence R. Baria and 2006, he worked as an intern for Occi-
dental Petroleum Corporation in Long Beach,
California, and Houston, Texas. He continued
his studies at the University of Mississippi, where
ABSTRACT he received a Ph.D. in carbonate sedimentology,
sequence stratigraphy, sulfate diagenesis, and
Gypsum can form penecontemporaneously in carbonate sedi- basin analysis in May 2009. He is currently
ments, and it may be altered to anhydrite during burial. Some serving as a visiting assistant professor in the
workers have suggested that gypsum dehydration is complete Department of Geology and Geological
at burial depths of 3500 ft (1067 m); however, gypsum has Engineering at the University of Mississippi.
been documented at depths as great as 13,100 ft (3993 m). R. P. Major  Department of Geology and
Two examples of deeply buried gypsum illustrate con- Geological Engineering, University of Mississippi,
trasting origins. Gypsum in the Permian San Andres Forma- University, Mississippi; rpm@olemiss.edu
tion in the Permian Basin of west Texas at a depth of 6000 ft R. P. Major holds a Ph.D. in geology from
(1829 m) occurs as anhydrite nodules with an outer rind of Brown University and was a National Research
gypsum, suggesting that anhydrite was rehydrated to gypsum Council postdoctoral associate with the U.S.
by interaction with interstitial brines. In contrast, gypsum in Geological Survey in Denver. He has taught at
Union College and the University of Colorado
the carbonate beds of the Cretaceous Ferry Lake Anhydrite of at Denver, and he served as a research geologist
the Mississippi Interior Salt Basin at 13,100 ft (3993 m) occurs at the Bureau of Economic Geology, University
as gypsum nodules with a rind of anhydrite, suggesting that of Texas at Austin. He is a former senior geolo-
the gypsum is primary and in the process of dehydration to gist with the Amoco Production Company, and
anhydrite. he has served on the editorial boards of the
These contrasting origins are likely a function of evaporite Journal of Sedimentary Research and the AAPG
Bulletin. Since 1998, he has been a professor
and carbonate stratigraphy. In the Permian Basin, a thick sec- in the Department of Geology and Geological
tion of carbonate rocks is overlain by a relatively thin section of Engineering at the University of Mississippi.
evaporites. In the Mississippi Interior Salt Basin, carbonates are
interbedded with multiple evaporite beds, which inhibit cir- Lawrence R. Baria  Jura–Search Incor-
porated, Flowood, Mississippi;
culation of interstitial brines. The burial temperature, pressure, bearlear@bellsouth.net
and salinity of these formations suggest that gypsum could be
Lawrence Baria, after receiving a B.S. degree
preserved as primary or secondary gypsum. in geology in 1967, pursued a master’s degree in
The presence of gypsum in reservoir rocks can affect po- stratigraphy and sandstone petrology at North-
rosity calculations, especially calculations using neutron po- east Louisiana University and a Ph.D. in carbon-
rosity logs, which measure bound water of hydration in gypsum ate facies, stratigraphy, and sulfate diagenesis
as porosity. Identification of gypsum at the depth of 13,100 ft at the Louisiana State University (LSU). After
leaving LSU, he went to work for Getty Oil Com-
pany in its E&P Research Laboratory, specializing
in Cretaceous and Jurassic stratigraphy. Since
1980, he has been a consulting and exploration
Copyright ©2012. The American Association of Petroleum Geologists. All rights reserved.
geologist active in the Haynesville, Smackover,
Manuscript received November 12, 2010; provisional acceptance January 17, 2011; revised manuscript
received April 21, 2011; final acceptance May 10, 2011.
and Norphlet formations throughout the central
DOI:10.1306/05101110179 and eastern Gulf Coast.

AAPG Bulletin, v. 96, no. 2 (February 2012), pp. 375–390 375


(3993 m) indicates that log estimates of porosity in some deep
ACKNOWLEDGEMENTS carbonate reservoirs containing gypsum could have significant
F. O. Amadi acknowledges partial funding from porosity and water saturation measurement errors.
AAPG and the Gulf Coast Association of Geo-
logical Societies. We thank Crawford Elliott of
Georgia State University for providing the x-ray INTRODUCTION
diffraction analyses. The conventional core-plug
analysis for porosity and permeability was
Several workers have suggested that the complete transition of
conducted by Houston Core Laboratory. The
analysis for water chemistry was conducted gypsum to anhydrite will occur at a depth less than 3500 ft
by Halliburton Laboratory. We thank Seismic (<1067 m), thus leading to the generalization that gypsum is
Micro-Technology, Inc., for providing a Kingdom not expected to exist below this depth (Murray, 1964; Billo,
software grant used for the project. We are 1986; Rosen and Warren, 1990; Kasprzyk, 1995; Williams-
grateful for the encouragement of R. C. Murray, Stroud and Paul, 1997; Kasprzyk and Orti, 1998; Warren, 2006).
F. J. Lucia, and S. C. Ruppel.
However, examples of gypsum at depths greater than 3500 ft
The AAPG Editor thanks the following reviewers
for their work on this paper: Wayne M. Ahr (>1066 m) (Tilly et al., 1982; Bouquillon, 1984; Jowett et al.,
and an anonymous reviewer. 1993; Major and Holtz, 1997; Amadi et al., 2008) exist, thereby
leaving in doubt the origin of gypsum at such great depth. The
possibility of gypsum being present in reservoir rocks deeper
EDITOR’S NOTE than 3500 ft (1066 m) is commonly not considered when
A color version of Figures 1, 4 and 5 may be evaluating deep reservoir rocks. This article evaluates the ori-
seen in the online version of this article. gin of gypsum in deep carbonate reservoirs from two basins
and its implications for hydrocarbon exploration and produc-
tion: gypsum in the Lower Cretaceous Ferry Lake Anhydrite of
the Mississippi Interior Salt Basin and in the San Andres For-
mation of the Permian Basin in west Texas.

PREVIOUS WORK

Conversion of gypsum to anhydrite is influenced by three


factors: (1) intense solar heating in very arid regions at shallow
depth, (2) reaction to high-salinity brines, and (3) burial dia-
genesis (Murray, 1964; Hardie, 1967; Lock et al., 1983; Warren,
1991; Testa and Lugli, 2000; Kendall, 2001).
Deep carbonate reservoirs commonly contain evaporite
nodules (Lowenstein, 1987; Machel and Burton, 1990, 1991;
Machel, 1993; Schreiber and El Tabakh, 2000; Warren, 2006).
These evaporite nodules could have formed from any of these
processes: concretionary growth, syndepositional flowage of
semifluid sediment (Riley and Byrne, 1961), replacement of
displacive gypsum nodules concomitant with partial replace-
ment of the surrounding carbonate sediment (Alsharhan and
Whittle, 1995), subaqueous growth as cement (Langbein, 1987),
growth as floating nodules (Richter-Bernburg, 1985), and com-
paction of bedded gypsum or anhydrite (Langbein, 1987).
Conversion of gypsum to anhydrite with depth has been
extensively studied as a function of temperature, pressure,

376 Origins of Gypsum in Deep Carbonate Reservoirs


water activity, and geotechnical processes. Jowett
et al. (1993) noted that gypsum is very susceptible
to dehydration, producing water and anhydrite when
subjected to elevated temperatures and pressures.
A review of some recent contributions suggests
that conversion will occur at a temperature range
of 95 to 200°C (Ostroff, 1964; Kosztolanyi et al.,
1987; Worden et al., 1997; Kontrec et al., 2002;
Freyer and Voigt, 2004; Vanko and Bach, 2005), a
pressure range of 1 to 14,503 psi (0.007–100 MPa)
(Blount and Dickson, 1969, 1973; Berdugo et al.,
2008), and chlorinity of 145‰ (Butler, 1969; Ali
et al., 2005). Warren (2006) suggested that, with
less saline brines in the pores of buried evaporite
beds, conversion of gypsum to anhydrite may not
occur until burial depths of hundreds of meters. At
such depths, compaction may have greatly de-
pleted intercrystalline porosity in gypsum beds and
nodules. Jowett et al. (1993) proposed that gyp-
sum converts to anhydrite at shallow depths ap-
proximately 1312 ft (∼400 m) when overlain by
low-heat conductors such as shale in a rift envi-
ronment, and at a depth greater than 13,123 ft
(>4000 m) when overlain by high-heat conductors
such as salt in a stable cratonic region.
Major and Holtz (1997) documented second-
ary gypsum at 6000 ft (1800 m) in the Jordan (San
Andres) field of the Permian Basin, west Texas.
Figure 1. San Andres Formation rocks. (A) Nodular anhydrite
They reported a diagenetic alteration of the outer (an) displaying outer edges hydrated to gypsum (g). Note that a
edges of anhydrite nodules to secondary gypsum. halo of light-colored dolomite (diagenetically altered texture)
This diagenetic process is associated with leaching surrounds the partly hydrated nodules. (B) Anhydrite hydrated to
and partial dissolution of dolomite crystals (Figure 1). gypsum. Note that the outer margin of the anhydrite nodule has
Most anhydrite nodules with outer edges altered been hydrated to gypsum (cross-polarized light photomicrograph).
to gypsum are commonly surrounded by a halo of
altered dolomite (Figure 1). The outer parts of an-
hydrite nodules that are rehydrated to gypsum sug- 1997), and the Zechstein Basin (Peryt and Scholle,
gest that gypsum is secondary instead of primary 1996). These basins contain thick sequences of
(Figure 2). Hydration of anhydrite to gypsum in carbonate rocks that are overlain by or interbedded
the Permian Basin has also been reported by Baria with evaporites.
(1976), Longacre (1976), Lucia et al. (1987), and
Tilly et al. (1982).
Gypsum within carbonate rocks in most deep METHODS
basins associated with bedded evaporites is com-
pletely altered to anhydrite. Examples of such basins Six Ferry Lake cores from Hancock County, Mis-
are the Williston Basin (Tanguay and Friedman, sissippi, were used for this study (Figure 3). These
2001), the Permian Basin (Harris and Walker, 1990; cores are not continuous, anhydrite beds were not
Sonnenfeld and Cross, 1993; Major and Holtz, cored, and only the last 15 to 20% of limestone beds

Amadi et al. 377


Figure 2. Rehydration of anhydrite to
secondary gypsum in the San Andres
Formation, resulting from Ca-rich connate
water released from the overlying and
underlying gypsum beds.

were recovered. The cores were carefully stored core plugs were compared with those from po-
and handled during preparation to prevent contact rosity logs and thin-section point counting.
with water. Low-temperature core sample analysis
was used in cleaning the samples to prevent gyp-
sum from losing its water of crystallization (Tilly
et al., 1982). RESULTS
Standard petrographic analysis was conducted
on 40 thin sections to document the diagenetic Evaporites in Ferry Lake limestones of the Mis-
development, grain types, depositional fabrics, and sissippi Interior Salt basin occur at a depth greater
porosity amount and type. A representative subset than 13,000 ft (3962 m) in different forms—as
of 20 thin sections was stained with alizarin red S nodules, replacement of dissolved fossils, and along
and potassium ferricyanide and was point counted styolitic zones (Figure 4). Evaporite nodules occur
using a petrographic microscope equipped with a in variable shapes and sizes ranging from 2 to 10 mm
swift automated stage and counter; 400 points were in diameter. Some sections of the evaporite nod-
counted on each thin section using the Glagoley- ules were dissolved and replaced by carbonate crys-
Chayes method (Chayes, 1956; Van der Plas and tals. In some areas, poikilotopic anhydrite cements
Tobi, 1965). bind carbonate rock grains. The nodules bear no
Limestones were classified using the method relationship either in the direction or the orien-
of Dunham (1962). X-ray diffraction analysis from tation to sedimentary structures in the limestones.
two representative samples selected from the same Evaporite nodules are composed predominantly
samples used for thin sections was performed to of anhydrite crystals and some gypsum crystals in
augment results from thin-section analyses. X-ray both displacive and replacive forms. Anhydrite
diffraction analysis was performed on the pulver- crystals are mainly prismatic (laths), felted crystals,
ized fine fraction (<235 mesh) using a Philips PW fascicular (bundlelike), fibrous radiating, and spe-
1729 x-ray diffractometer. All samples were ana- rulitic. Anhydritization of microgranular gypsum
lyzed with a scanning speed of 0.02° 2q /min range is partial to complete with anhydrite laths oriented
with Cu radiation (40 kV and 35 mA). Porosity and preferentially. The contacts of carbonate with cal-
permeability values were determined from con- cium sulfate are dominantly characterized by an-
ventional core-plug analysis using 10 representative hydrite with fibrous calcite growing into the cleav-
samples randomly selected from the same depths age planes. Numerous fossil molds are cemented by
as samples used for thin sections. The results from blocky anhydrite and, in very few cases, with fibrous

378 Origins of Gypsum in Deep Carbonate Reservoirs


Figure 3. Index map of Hancock County, Mississippi, showing Waveland field and the location of six wells used for petrographic analysis
of Ferry Lake rocks.

anhydrite cements. Most gypsum crystals are com-


pletely engulfed by anhydrite. Microfractures are
pervasively plugged by blocky anhydrite.
Ferry Lake bulk rock contains calcite as the most
abundant mineral, with anhydrite and gypsum pre-
sent in lesser amounts. Dolomite is completely ab-
sent from Ferry Lake carbonate rocks. Limestone
constitutes the dominant carbonate-rock lithology,
with minor amounts of very thin layers of terrige-
nous clastic material that grades from fine to me-
dium sand size.
Carbonate facies have similar textures and
structures, although they vary in the amounts and
types of fossils present. Ostracod, mollusk mud-
stone, miliolid ostracod wackestone, and packstone
dominate. Few lenses of oolitic intraclast packstone,
fecal pellets, and peloidal grains are observed. Fecal
pellets were difficult to recognize because of coa-
lesce during compaction.
Four kinds of porosity occur in carbonate beds:
intergranular, intragranular, vuggy, and fracture
porosities. Average porosity from point counting is
Figure 4. (A) Core sample of nodular evaporite from Ferry Lake 9%, with fracture and vuggy porosities dominating
Anhydrite at 13,100 ft (3993 m). (B) Primary gypsum (g) engulfed (Table 1). Porosity calculated from neutron logs
by anhydrite (an) crystals (cross-polarized light photomicrograph). ranges from 9 to 15% (average, 12%), and porosity

Amadi et al. 379


Table 1. Point-Count Data from Ferry Lake Anhydrite Thin Sections Highlighting the Amount of Gypsum and Porosity

Fracture Total Classification


Interparticle Intraparticle Porosity Calcite Anhydrite Gypsum Miliolid Orbitolinid Others Porosity of Dunham
Well Name Porosity (%) Porosity (%) (%) (%) (%) (%) (%) (%) (%) (%) (1962)
Union 32–11 3 2 4 64 13 9 4 1 0 9 Wackestone
Cinque Bambini
Union 32–11 3 1 3 65 17 9 2 0 0 7 Packstone
Cinque Bambini
Union 32–11 3 0 5 62 16 11 1 2 0 8 Packstone
Cinque Bambini
Union 32–11 0 0 2 98 0 0 0 0 0 2 Mudstone
Cinque Bambini
Union 32–11 2 0 3 66 20 9 0 0 0 5 Mudstone
Cinque Bambini
Union 32–11 1 0 0 71 17 11 0 0 0 1 Mudstone
Cinque Bambini
Union 32–11 2 2 5 62 13 11 4 1 0 9 Wackestone
Cinque Bambini
Union 32–11 1 5 6 63 13 2 3 1 6 12 Wackestone
Cinque Bambini
Union 32–11 1 5 6 59 19 8 2 0 1 11 Packstone
Cinque Bambini
Union 32–11 1 3 7 67 15 3 2 2 0 11 Wackestone
Cinque Bambini
Union 32–11 0 2 4 63 14 5 6 2 4 6 Packstone
Cinque Bambini
Union 32–11 0 2 3 70 13 4 2 2 4 5 Wackestone
Cinque Bambini
Hunt 1 Thigpen 0 2 5 79 3 1 2 0 8 7 Mudstone
Hunt 1 Thigpen 2 2 4 72 9 1 3 1 6 8 Wackestone
Hunt 1 Thigpen 1 1 3 75 11 0 2 2 5 5 Mudstone
Saga 20–4 Macc 2 1 5 50 23 17 2 0 0 8 Wackestone
Saga 20–4 Macc 3 0 4 50 20 18 3 1 1 7 Wackestone
Saga 20–4 Macc 0 4 1 65 15 5 2 2 7 5 Wackestone
Saga 20–4 Macc 2 7 3 59 18 3 3 2 3 12 Wackestone
Saga 20–4 Macc 1 5 6 67 12 2 5 1 1 12 Wackestone
Saga 20–4 Macc 1 5 5 54 11 17 3 1 3 12 Packstone
Saga 20–4 Macc 1 2 4 65 14 3 7 2 2 7 Wackestone
Saga 20–4 Macc 0 2 9 71 5 1 7 2 3 12 Wackestone
Saga 20–4 Macc 3 1 5 64 13 9 3 1 0 9 Wackestone
Saga 20–4 Macc 4 1 3 72 9 1 2 2 5 8 Wackestone
Tideway 12–7 1 3 5 65 14 2 3 3 4 9 Wackestone
Jack Lott
Tideway 12–7 1 4 5 67 10 3 5 1 4 10 Wackestone
Jack Lott
Tideway 12–7 2 0 4 53 21 18 1 1 0 6 Wackestone
Jack Lott
Union 27–11 0 2 6 75 9 2 1 0 4 9 Mudstone
Lindner J. R.

380 Origins of Gypsum in Deep Carbonate Reservoirs


Table 1. Continued

Fracture Total Classification


Interparticle Intraparticle Porosity Calcite Anhydrite Gypsum Miliolid Orbitolinid Others Porosity of Dunham
Well Name Porosity (%) Porosity (%) (%) (%) (%) (%) (%) (%) (%) (%) (1962)
Union 27–11 0 2 6 70 10 7 2 0 3 8 Wackestone
Lindner J. R.
Union 27–11 1 1 5 81 2 0 4 1 5 7 Mudstone
Lindner J. R.
Union 27–11 4 2 9 70 8 2 0 1 4 12 Wackestone
Lindner J. R.
Union 27–11 4 1 4 72 3 1 6 4 5 9 Wackestone
Lindner J. R.
Union 27–11 5 1 3 71 3 1 5 4 7 9 Wackestone
Lindner J. R.
Union 27–11 2 2 3 54 16 7 6 4 6 7 Packstone
Lindner J. R.
Union 27–11 1 3 10 68 2 0 6 4 6 12 Wackestone
Lindner J. R.
Union 27–11 1 2 7 65 1 0 12 7 5 9 Wackestone
Lindner J. R.
Spooner 2 Bambini 2 5 6 52 14 12 4 2 3 13 Grainstone
Spooner 2 Bambini 3 6 6 50 15 13 3 2 2 15 Grainstone
Spooner 2 Bambini 2 5 9 52 14 11 3 2 2 16 Grainstone

from core-plug analysis averages 11% ( Table 2). Although anhydrite could possibly be incorrectly
Point counts from 40 representative samples in- identified as gypsum if viewed along the optic
dicate gypsum content ranging from 1 to 18%. axis, such errors in identification cannot account

Table 2. Porosity and Permeability Data from Core-Plug Analyses of Ferry Lake Anhydrite Samples

Permeability (md)
Depth Net Confining Porosity Grain Density
Well Name (ft) Stress (psig) (%) Klinkenberg Kair* bair* (psi) b* (ft–1) a* ( mm) (g/cm3)
Union 32–11 Cinque 13,333 Ambient 11.66 N/A* 0.379 N/A* N/A* N/A* 2.695
Bambini
Saga 20–4 Macc 13,550 800 10.57 0.276 0.389 9.75 1.28E+12 1.11E+03 2.706
Saga 20–4 Macc 13,463 800 10.91 0.433 0.602 9.00 7.25E+11 9.94E+02 2.711
Saga 20–4 Macc 13,360 800 13.15 0.858 1.12 6.85 1.44E+11 3.94E+03 2.695
Tideway 12–7
Jack Lott 13,351 800 12.91 1.05 1.33 5.90 3.02E+11 1.01E+02 2.718
Union 27–11 Lindner J. R. 13,130 Ambient 12.85 N/A* 2.70 N/A* N/A* N/A* 2.683
Union 27–11 Lindner J. R. 13,370 800 9.86 0.521 0.695 7.61 1.69E+11 2.78E+02 2.688
Spooner 2 Bambini 13,380 800 13.32 67.4 92.4 6.35 8.83E+08 1.92E+02 2.698
Spooner 2 Bambini 13,385 800 9.10 0.099 0.141 11.00 6.62E+12 2.07E+03 2.693
Spooner 2 Bambini 13,399 800 8.40 0.313 0.396 6.25 3.71E+12 3.69E+03 2.701
*Kair = permeability to air; bair = Klinkenberg slip factor; b = Forcheimer inertial resistance factor; a = a factor equal to the production of b and K∞; K∞ = equivalent
nonreactive liquid permeability; N/A = not applicable.

Amadi et al. 381


for almost 18% gypsum point counted from some Primary Gypsum
of the Ferry Lake evaporite nodules.
The chlorinity of producing water from the Staircase cleavage planes of most anhydrite laths
Ferry Lake Anhydrite in Waveland field ranges from are filled with calcite crystal overgrowths, suggesting
114 to 127‰, with measured bottom-hole pressure that most nodules were formed before compaction.
(BHP) values of 6200 to 6500 psi (43–45 MPa) and The presence of calcite overgrowths is an indica-
bottom-hole temperature (BHT) range of 110 to tion that anhydrite laths predate calcite. Blocky an-
137°C. The salinities, BHT, and BHP of San An- hydrite crystals formed as a replacement of primary
dres Formation reservoirs of the Permian Basin are gypsum that precipitated in the fractured zones
based on 161 samples of brine water that yielded a during burial diagenesis. Partly deformed and dis-
chlorinity range of 50 to greater than 200‰, BHP solved limestone grains are pervasively filled by an-
of 1661 to 1944 psi (11–13 MPa), and BHT of 131 hydrite that has microcrystalline to coarse-crystalline
to 140°C (Dutton and Orr, 1986; Dutton, 1987). textures.
During anhydritization of gypsum, fractured
zones served as pathways to drain hypersaline fluid
DISCUSSION from gypsum. As sulfate-rich water drained through
fractured zones, anhydrite precipitated as a blocky
Depositional Environment and fibrous crystal. Anhydrite cements are present
as blocky laths and elongate fiber-shaped forms
The presence of normal marine-salinity fossils and displaying random orientations. The blocky crys-
abundant burrows in mudstones and wackestones talline microstructure occurs mainly as a secondary
indicates that Ferry Lake limestones were depos- fabric, characteristic of transformation of the gyp-
ited in a low-energy shallow-water marine environ- sum precursor, whereas most of the felted fabrics
ment shoreward of the bioherms and grainstone are inherited from an initial early anhydrite texture.
banks at the platform edge. These bioherms caused These cements, found in intercrystalline, interpar-
the restriction necessary for sufficient concentra- ticle, and moldic pores, are free of inclusions. Sparry
tion of seawater to deposit gypsum. The fossils are anhydrite replaces ooids and peloids and fills pores
indicative of a normal marine environment and between calcite crystals. Most of the nodular an-
were transported during small-scale transgressive hydrite represents growth of gypsum by displace-
sequence events. ment of soft sediments after deposition of the host
Evaporite nodules in these rocks indicate a material and is considered to be an early diagenetic
hypersaline environment. The shapes of the nod- feature.
ules are slightly horizontally elongated, suggesting The orientation of anhydrite laths with refer-
that the nodules were precipitated contempora- ence to gypsum and the nature of anhydrite en-
neously with deposition of limestone beds. Some gulfing gypsum are indications of gradual replace-
nodules are nearly isometric, with no indication of ment of primary gypsum by anhydrite (Figure 5).
compaction, suggesting that they formed after the Anhydrite enclosed and in touch with calcite crys-
limestone beds were compacted and as a replace- tals indicates that Ferry Lake gypsum nodules may
ment of skeletal grains or as cements in molds. have been formed concomitant with soft calcite
Circulation of normal marine waters into the sediment (prelithification). Displacive growths of
lagoon may have been induced by periodic small- the nodules and preservation of primary lithologic
scale transgressive-regressive cycles formed during features are evidence for transformation of gypsum
high-frequency eustatic sea level fluctuations. Hyper- to anhydrite in the nodules beginning during early
saline and normal marine environments alternated burial diagenesis.
as the Ferry Lake was deposited, resulting in the X-ray diffraction and petrographic observa-
cyclic depositional pattern of relatively deeper wa- tions confirm the presence of gypsum at 13,000 ft
ter limestone and shallow-water evaporites. (3962 m) (Figures 5, 6). The presence of gypsum

382 Origins of Gypsum in Deep Carbonate Reservoirs


not exclusively depend on depth; regional tectonic
activity, geothermal gradient, and availability of
diagenetic fluids have substantial functions (Lock
et al., 1983; Warren, 1991; Jowett et al., 1993).
The presence of Ferry Lake gypsum at 13,000 ft
(3962 m) can be explained as follows. The gradual
process of anhydritization of gypsum nodules con-
stitutes one of the pathways. This process starts
after precipitation of gypsum, followed by an in-
crease in burial depth, a rise in ambient tempera-
ture, and an increase in lithostatic pressure. As these
processes continue, gypsum begins to convert to
anhydrite. Conversion to anhydrite was initiated at
the rim and gradually progressed toward the core
as compaction and dehydration continued. These
processes caused anhydrite laths to encase gypsum
crystals (Figure 7). Such configuration character-
ized most gypsum and anhydrite crystal contacts.
This interpretation suggests that the conversion
process from precursor gypsum nodules in the
Ferry Lake limestones is still in progress.
The gypsum crystals in the nodules are all
engulfed by anhydrite laths, which are more pref-
erentially oriented to those of gypsum and sub-
parallel to the host bedding planes. This demon-
strates that the process of anhydritization is still in
progress, with the rims of the evaporite nodules
consisting of anhydrite that has replaced gypsum.
In some nodules, porphyroblastic gypsum crystals
are partly replaced by anhydrite. Rehydration (re-
placement of anhydrite by secondary gypsum)
would have destroyed gypsum pseudomorph fab-
rics in the anhydrite laths. Instead, primary gyp-
sum fabrics are well preserved in most anhydrite
Figure 5. Ferry Lake rocks showing partly replaced gypsum (g) laths, indicating that gypsum textures were not de-
crystals completely engulfed by anhydrite (an), indicating gradual stroyed because they were replaced by anhydrite
replacement from the rim closing in to the core. (A) Blocky gyp- and providing evidence of a gypsum origin of the
sum crystal almost replaced by anhydrite lath. (B) Prismatic anhydrite nodules. The gypsum pseudomorphs also
gypsum and anhydrite with anhydrite lath preferentially oriented
and in touch with the host calcite rock. (C) Pseudomorphs of
indicate a volume-for-volume replacement of gyp-
gypsum precursor can be traced in the anhydrite. The gypsum sum for anhydrite (Figure 5).
pseudomorph also suggests a volume-for-volume replacement of Burial conversion of gypsum to anhydrite is fa-
the gypsum to anhydrite (cross-polarized light photomicrographs). cilitated in systems that are high in Ca++ and SO4–.
The source of Ca++ in a closed system is commonly
the dolomitization of calcium carbonate. Dolo-
at this depth is significant because of what has mitization of precursor calcite increases the Ca++/
been previously published regarding gypsum. The Mg++ ratio in the system. Both petrographic and
total transformation of gypsum to anhydrite does x-ray diffraction analysis did not show evidence of

Amadi et al. 383


Figure 6. X-ray patterns of carbonate samples from the Ferry Lake Anhydrite at a depth of 13,100 ft (3993 m). The patterns indicate
that calcite is dominant, followed in abundance by anhydrite then gypsum. d = diffraction peak.

Figure 7. Gradual and direct anhy-


dritization of original nodular gypsum in
Ferry Lake rocks. Gypsum conversion to
anhydrite started from the outer rim of
gypsum nodules and progressed gradually
toward the core.

384 Origins of Gypsum in Deep Carbonate Reservoirs


the Ferry Lake carbonate being dolomitized, there- ward dip direction. This lithologic sequence indi-
by creating a system that is relatively starved of cates environmental changes from a very restricted
Ca++. It may be that conversion of gypsum nod- lagoonal condition to a normal marine environ-
ules to anhydrite at early burial depth was shut off ment. The Stuart City reef (Lock et al., 1983) was
once the system could not generate sufficient Ca++. a pivotal factor in separating the Ferry Lake depo-
The loss of primary porosity at this early stage of center from the open marine environment together
burial may have inhibited the flow of fluids rich with tectonic subsidence of the depocenter, which
in Ca++. promoted small-scale high-frequency transgressive-
regressive cycles (Amadi et al., 2009). These fac-
Subsurface Environments of Primary and tors led to the unique rhythmic stacking patterns
Secondary Gypsums (carbonate-anhydrite cycles) of the Ferry Lake,
thereby establishing an important stratigraphic
Quantitative evaluation of diagenetic environment marker that has been used as a datum for correla-
of both Ferry Lake and San Andres formations tion, interpretation of seismic reflectors, and struc-
suggests environments that could support gypsum tural mapping throughout the Lower Cretaceous
preservation. This conclusion is especially true of of the Gulf of Mexico Basin.
the Ferry Lake, with BHT (110–137°C), BHP (6200– Comparing the Ferry Lake stratigraphy with
6500 psi [43–45 MPa]), and chlorinity (114–127‰) stratigraphies of other basins with carbonate and
comparable to established values where gypsum evaporite sections is pertinent to understanding the
converts to anhydrite (Ostroff, 1964; Blount and presence of primary gypsum in Ferry Lake lime-
Dickson, 1969, 1973; Butler, 1969; Kosztolanyi stone. These basins contain predominantly thick
et al., 1987; Worden et al., 1997; Kontrec et al., sequences of carbonate rocks that are overlain by
2002; Freyer and Voigt, 2004; Ali et al., 2005; Vanko relatively thin anhydrite beds. Stratigraphic se-
and Bach, 2005; Berdugo et al., 2008). The chlo- quences of these carbonate-evaporite basins con-
rinity and BHT from the San Andres Formation trast with the stratigraphic sequence of the Ferry
indicate that complete anhydritization and rehy- Lake Anhydrite. The Ferry Lake comprises mul-
dration of gypsum nodules are possible. tiple cycles of small-scale carbonate-evaporite se-
Because sampling of gypsum and anhydrite quences (Figure 8). Carbonate beds from the Ferry
from the Permian Basin is commonly limited to Lake are interbedded with several anhydrite beds
hydrocarbon-bearing reservoirs, migration of hy- with thicknesses ranging from 4 to 60 ft (1–18 m).
drocarbons into those reservoirs and dilution or The Upper Permian (Guadalupian) San Andres
expulsion of early diagenetic connate brines may Formation of the Permian Basin comprises a thick
also be a factor in the conversion of anhydrite to sequence of dolostone that is capped by relatively
gypsum below 3500 ft (1067 m). In fact, gypsum thin evaporite beds. In some sections, the Upper
rinds around some anhydrite nodules are com- Permian of the Zechstein Basin of Poland con-
monly surrounded by an outer rind of calcite and sists of more than 328 ft (>100 m) of carbonate
elemental sulfur. This diagenetic continuum from rocks that are overlain by massive evaporite ex-
nodule core to outer rind has been attributed to ceeding 984 ft (300 m) in thickness (Peryt, 1989).
thermal and organic processes caused by hydro- The stratigraphy of the Ordovician Red River
carbon migration (Baria, 1976; Longacre, 1976). Formation of the Williston Basin contains a thick
section of carbonate rocks overlain by anhydrite
Stratigraphy and Gypsum beds (Montgomery, 1997; Tanguay and Friedman,
2001).
The lithology of the Lower Cretaceous Ferry Lake The stratigraphic patterns from these three
Anhydrite of the Mississippi Interior Salt Basin formations contrast with that of the Ferry Lake,
varies in thickness from an evaporite-dominated which has multiple cycles of anhydrite and carbon-
facies to a carbonate-dominated facies in the sea- ate beds. Cyclic anhydrite beds may have possibly

Amadi et al. 385


Figure 8. Generalized stratigraphic sections from three different basins with formations similar to the Ferry Lake Anhydrite. Note the
dolostone in all of these sequences except the Ferry Lake. The Ferry Lake stratigraphy contains more small-scale carbonate-evaporite
cycles compared with the other basins that contain only one or two anhydrite beds. Multiple small-scale cycles of carbonate-evaporite
parasequences are instrumental in inhibiting free flow of diagenetic fluids in the Ferry Lake, which is in contrast with the other formations.

hampered free circulation of diagenetic fluids rich (305 m) or more with anhydrite beds at the top of
in Ca++ that could have enhanced complete con- the sequence. Such stratigraphic settings and stack-
version of gypsum nodules to anhydrite. Conversely, ing patterns probably affect the rate of anhydriti-
conversion of gypsum to anhydrite was complete zation of gypsum (Figure 8).
in the other formations partially because of fewer Carbonate beds in the three basins discussed
impermeable anhydrite beds and thick porous car- above are partially to completely dolomitized, in-
bonate beds. These stratigraphic settings would have creasing the Ca+/Mg++ ratio in an enclosed system.
promoted free flow of diagenetic fluids. Although Such elevated levels of Ca++ and free circulation
secondary gypsum occurs in the San Andres For- of diagenetic fluids in an enclosed system will in-
mation, the general geometry and orientation of crease the chances of gypsum conversion to anhy-
nodule conversion differs from those of the Ferry drite. Stratigraphic sequence patterns will likely af-
Lake Anhydrite. fect the general circulation of diagenetic fluids that
The Ferry Lake evaporite nodules, with anhy- are capable of transforming gypsum to anhydrite.
drite rinds engulfing gypsum, are enclosed within
the carbonate beds. Multiple stacked anhydrite
beds may have restricted circulation of diagenetic IMPLICATIONS FOR EXPLORATION
fluids that could have completed anhydritization AND PRODUCTION
of gypsum nodules. Although the thicknesses of
anhydrite beds in the Ferry Lake are less than those Effects of Gypsum on Porosity Logs
in other basins, the Ferry Lake contains more cycles
of carbonate-anhydrite sequences, whereas in the Gypsum in reservoir rocks has a significant effect
other basins with carbonate-evaporite sequences, on neutron- and density-log responses and little ef-
carbonate beds are thicker, ranging as much as 1000 ft fect on acoustic-log response (Tilly et al., 1982; Lucia

386 Origins of Gypsum in Deep Carbonate Reservoirs


Figure 9. The effect of bulk gypsum in
limestone-reservoir porosity measure-
ments using wireline porosity logs (mod-
ified from Holtz and Major, 2004).

et al., 1987; Holtz and Major, 2004). Hydrogen is close to that of anhydrite, 50 ms/ft. Every cubic
contained in chemically bound water in gypsum is millimeter of gypsum will generate approximately
a major factor causing error in neutron porosity log. 0.02% porosity, which is generally insignificant
Tilly et al. (1982) noted that most of the other (Lucia et al., 1987).
minerals found in hydrocarbon reservoirs are com-
pletely devoid of hydrogen. Effects of Gypsum on Ferry Lake Carbonate
Hydrogen is assumed to be found in reservoir Porosity Logs
rocks as water and hydrocarbons. Water and hydro-
carbons fill the pores of reservoir rocks; neutron The average porosity results from point counts
logs measure the hydrogen ion content of the rock (9%), core-plug analysis (11%), and values calcu-
and convert those measurements to porosity. For- lated from neutron and density logs (13%) suggest
mations containing clay are an exception because a 2–4 difference in porosity percent unit. Appar-
some clays contain chemically bound water. Neu- ent variation in porosity values obtained from vari-
tron logs do not have the capability to differentiate ous porosity tools is attributed to 18% gypsum.
hydrogen that is present in gypsum, hydrocarbons, Holtz and Major (2004) calculated that 20% bulk-
or water. Lucia et al. (1987) noted that, for every volume gypsum will generate approximately 13%
cubic millimeter of gypsum in reservoir rocks, neu- error in porosity calculations (Figure 9). Based on
tron logs record 0.49% porosity. As a result, the re- the porosity-log–error data calculation of Holtz and
sponse of neutron logs to this chemically bound Major (2004). The Ferry Lake limestone is esti-
hydrogen causes a higher apparent porosity. mated to contain approximately 7% bulk-volume
Gypsum has a density of 2.35 g/cm3, which is gypsum (Figure 9). Seven percent bulk-volume gyp-
much lower than those of calcite, dolomite, and sum accounts for the 4% porosity error recorded in
anhydrite. Thus, a density log calibrated to lime- the neutron log relative to point-count results.
stone would read approximately 0.3% porosity for Both x-ray diffraction results and petrographic
every cubic millimeter of gypsum because of dif- analysis indicate that clay minerals are not present
ferences in mineral densities (Lucia et al., 1987). in the Ferry Lake limestone. The apparent discrep-
Acoustic logs measure the transit time of sound ancy from porosity measurements must be attrib-
waves through rock, and porosity is calculated by uted to the presence of gypsum. Such porosity mea-
assuming an average transit time for the matrix and surement errors associated with nodular gypsums
fluids. The transit time of gypsum is 53 ms/ft, which at great depth are commonly overlooked because

Amadi et al. 387


of the assumption that anhydritization of gypsum hydration of anhydrite after the core was recovered
is normally complete above 3500 ft (1067 m). and while the core was in storage. We cannot find
this assertion in print, but it continues to be part
of informal conversations and manuscript review
CONCLUSIONS comments.
This is unlikely because rehydration would re-
Gypsum can exist at depths greater than 3500 ft quire saturation with water, not simply a humid
(>1067 m). The depth at which gypsum completely environment. Moreover, this transformation would
converts to anhydrite differs from region to region. probably take a very long time by human measures.
Lithostatic pressure, local geothermal gradient, and The geometry of gypsum in slabbed core also in-
pore brine fluid availability are significant factors dicates that it was not formed during core storage.
in converting gypsum to anhydrite. Gypsum in the Were hydration of anhydrite to occur during stor-
Ferry Lake Anhydrite is a consequence of cyclic age, the anhydrite would form initially on the slab-
stratigraphy, tectonic setting, and diagenetic en- bed surface and recrystallization would proceed
vironment. Limestones in the Ferry Lake are not from this surface toward the interior of the core.
dolomitized, starving the system of the Ca++ that This geometry is not present in any of the cores we
could have promoted the conversion of gypsum to have examined. The most convincing data indi-
anhydrite. Gypsum in the Ferry Lake is interpreted cating that gypsum in subsurface cores is present
to be primary in origin, whereas gypsum in the San in situ are that gypsum is present in cores that have
Andres Formation is interpreted to be rehydrated just been recovered and have not yet been placed
anhydrite. Evidences from petrography, x-ray dif- in storage.
fraction analysis, wellbore environment conditions,
and stratigraphy explain the preservation of gypsum
in both the Ferry Lake and San Andres formations. REFERENCES CITED
Recognizing gypsum in evaporite nodules at
Ali, Y. A., A. A. Dradir, and R. M. El Sheikh, 2005, Evolution
13,100 ft (3993 m) in Ferry Lake limestones and at
of the saline and hypersaline brines in the solar ponds
6000 ft (1829 m) in San Andres Formation dolo- from Lake Quaroun, Faiyoum, Egypt: Egyptian Journal
stones suggests that deep carbonate reservoir rocks of Aquatic Research, v. 31, p. 213–225.
should be evaluated using multiple porosity mea- Alsharhan, A. S., and G. L. Whittle, 1995, Carbonate-evaporite
sequence of the Late Jurassic, southern and southwest-
surement tools. Using only one porosity tool, es- ern Arabian Gulf: AAPG Bulletin, v. 79, p. 1608–1630.
pecially neutron and density logs, to measure the Amadi, F. O., R. P. Major, and L. R. Baria, 2008, An unusual
porosity of deep carbonate reservoir rocks can have deep occurrence of gypsum in the Ferry Lake Anhy-
significant porosity and water saturation measure- drite, south Mississippi Salt Basin: Implication for hydro-
carbon exploration (abs.): AAPG Annual Convention,
ment errors. This study recommends that multiple v. 17, p. 7.
porosity measurements, such as thin-section pe- Amadi, F. O., R. P. Major, and L. R. Baria, 2009, Sequence
trography and core-plug analysis, be incorporated in stratigraphy and geochemical analysis of the Ferry Lake
Anhydrite, northeastern Gulf of Mexico: Implications
the evaluation of porosity in deep carbonate res-
for hydrocarbon potential (abs.): AAPG Annual Conven-
ervoir rocks. Moreover, special low-temperature tion, v. 18, p. 9.
core analysis should be used when gypsum is present Baria, L. R., 1976, Middle Guadalupian coastal carbonate
to minimize the loss of bound water from gypsum. complex: CO3-SO4 diagenesis and timing of hydrocar-
bon migration (abs.): AAPG Bulletin, v. 60, p. 647.
Berdugo, I., E. Romero, M. Saaltink, and M. Albis, 2008, On
the behavior of the CA-SO4-H2O system: Academia
POSTSCRIPT Colombiana de Ciencias Exactas, v. 32, p. 545–557.
Billo, S. M., 1986, Petrology and kinetics of gypsum-anhydrite
transitions: Journal of Petroleum Geology, v. 10, p. 73–
When discussing gypsum in subsurface cores dur-
86.
ing the course of many years, some scientists have Blount, C. W., and F. W. Dickson, 1969, The solubility of
suggested to us that gypsum may have formed by anhydrite CaSO 4 in NaCl-H 2 O from 100 to 450°C

388 Origins of Gypsum in Deep Carbonate Reservoirs


and 1 to 1000 bars: Geochimica et Cosmochimica Acta, of anhydrous calcium sulfate into calcium-sulfate dihy-
v. 33, p. 227–245. drate in aqueous solutions: Journal of Crystal Growth,
Blount, C. W., and F. W. Dickson, 1973, Gypsum-anhydrite v. 240, p. 203–211.
equilibria in systems CaSO4 and CaCO3-NaCl-H2O: Kosztolanyi, C., M. J. Mullis, and M. Weidmann, 1987, Mea-
The American Mineralogist, v. 58, p. 323–331. surements of the phase-transformation temperature of
Bouquillon, A., 1984, Stratigraphie, paléoenvironnement et gypsum anhydrite, included quartz, by microthermome-
diagenèse dans le Primaire sédimentaire des forages pro- try and Raman microprobe techniques: Chemical Geol-
fonds du Nord de la France: Diplôme d’Etudes Appro- ogy, v. 61, p. 19–28.
fondies, Universite de Lille, Lille, France, 53 p. Langbein, R., 1987, The Zechstein sulfates: The state of the
Butler, G. P., 1969, Modern evaporite deposition and geo- art, in T. M. Peryt, ed., The Zechstein facies in Europe:
chemistry of coexisting brines, the sabkha, trucial coast, Lecture Notes in Earth Sciences, v. 10, p. 143–188.
Arabian Gulf: Journal of Sedimentary Petrology, v. 39, Lock, B. E., B. K. Darling, and I. D. Roy, 1983, Marginal
p. 70–89. marine evaporites, Lower Cretaceous of Arkansas: Gulf
Chayes, F., 1956, Petrographic modal analysis: New York, Coast Association of Geological Societies Transactions,
John Wiley, 113 p. v. 33, p. 145–152.
Dunham, R. J., 1962, Classification of carbonate rocks ac- Longacre, S. A., 1976, Middle Guadalupian coastal carbon-
cording to depositional texture, in W. E. Ham, ed., Clas- ate complex: Stratigraphy and environmental recon-
sification of carbonate rocks: A symposium: AAPG Mem- struction (abs.): AAPG Bulletin, v. 60, p. 693–694.
oir 1, p. 62–84. Lowenstein, T. K., 1987, Evaporite depositional fabrics in the
Dutton, A. R., 1987, Origin of brine in the San Andres For- deeply buried Jurassic Buckner Formation, Alabama:
mation, evaporite confining system, Texas Panhandle Journal of Sedimentary Research, v. 57, p. 108–116.
and eastern New Mexico: Geological Society of America Lucia, F. J., C. R. Hocott, and G. W. Vander Stoep, 1987,
Bulletin, v. 99, p. 103–112. Characterization of the Grayburg reservoir, University
Dutton, A. R., and E. D. Orr, 1986, Hydrogeology and Lands Dune field, Crane County, Texas: Austin, Texas,
hydrochemical facies of the San Andres Formation in University of Texas at Austin, Bureau of Economic Geol-
eastern New Mexico and the Texas Panhandle: Austin, ogy Report of Investigations 168, 98 p.
Texas, University of Texas at Austin, Bureau of Eco- Machel, H. G., 1993, Anhydrite nodules formed during deep
nomic Geology Report of Investigations 157, 58 p. burial: Journal of Sedimentary Research, v. 63, p. 659–
Freyer, D., and W. Voigt, 2004, The measurement of sulfate 662.
mineral solubilities in the Na-K-Ca-Cl-SO4-H2O system Machel, H. G., and E. A. Burton, 1990, Burial diagenetic
at temperatures of 100, 150 and 200°C: Geochimica et sabkha-like gypsum and anhydrite nodules: Journal of
Cosmochimica Acta, v. 68, p. 307–318. Sedimentary Research, v. 61, p. 394–405.
Hardie, L. A., 1967, The gypsum-anhydrite equilibrium at Machel, H. G., and E. A. Burton, 1991, Factors governing
one-atmosphere pressure: The American Mineralogist, cathodoluminescence in calcite and dolomite and their
v. 52, p. 171–200. implications for studies of carbonate diagenesis, in C. E.
Harris, P. M., and S. D. Walker, 1990, McElroy field: Devel- Barker and O. C. Kopp, eds., Luminescence microscopy
opment geology of a dolostone reservoir, Permian Basin, and spectroscopy: Qualitative and quantitative applica-
west Texas, in D. G. Bebout and P. M. Harris, eds., Geo- tions: SEPM Short Course 25, p. 9–25.
logic and engineering approaches in evaluation of San Major, R. P., and M. H. Holtz, 1997, Predicting reservoir
Andres/Grayburg reservoirs: Permian Basin: Austin, quality at the development scale: Methods for quantify-
Texas, University of Texas at Austin Bureau of Economic ing remaining hydrocarbon resource in diagenetically
Geology, p. 275–296. complex carbonate reservoirs, in J. A. Kupecz, J. Gluyas,
Holtz, M. H., and R. P. Major, 2004, Integrated geological and S. Bloch, eds., Reservoir quality prediction in sand-
and petrophysical characterization of Permian shallow- stones and carbonates: AAPG Memoir 69, p. 231–248.
water dolostone: Society of Petroleum Engineers Reser- Montgomery, S. L., 1997, Ordovician Red River “B”: Hori-
voir Evaluation and Engineering, v. 7, p. 47–58. zontal oil play in the southern Williston Basin: AAPG
Jowett, E. C., L. M. Cathles, and B. W. Davis, 1993, Predict- Bulletin, v. 81, p. 519–532.
ing depths of gypsum dehydration in evaporitic sedimen- Murray, R. C., 1964, Origin and diagenesis of gypsum and
tary basins: AAPG Bulletin, v. 77, p. 402–413. anhydrite: Journal of Sedimentary Petrology, v. 34,
Kasprzyk, A., 1995, Gypsum-to-anhydrite transition in the p. 512–523.
Miocene of southern Poland: Journal of Sedimentary Re- Ostroff, A. G., 1964, Conversion of gypsum to anhydrite in
search, v. A65, p. 348–357. aqueous salt solution: Geochimica et Cosmochimica
Kasprzyk, A., and F. Orti, 1998, Paleogeographic and burial Acta, v. 28, p. 1363–1372, doi:10.1016/0016-7037
control on the anhydrite genesis: The Badenian Basin (64)90154-1.
in the Carpathian foredeep (southern Poland, western Peryt, T. M., 1989, Basal Zechstein in southwestern Poland:
Ukraine): Sedimentology, v. 45, p. 889–907. Sedimentation, diagenesis, and gas accumulation, in R. W.
Kendall, A. C., 2001, Late diagenetic calcitization of anhy- Boyle, A. C. Brown, C. W. Jefferson, E. C. Jowett, and
drite from the Mississippian of Saskatchewan, western R. V. Kirkham, eds., Sediment-hosted stratiform copper
Canada: Sedimentology, v. 48, p. 29–55. deposits: Geological Association of Canada Special Pa-
Kontrec, J., K. Damir, and L. Brecevic, 2002, Transformation per 36, p. 103–625.

Amadi et al. 389


Peryt, T. M., and P. A. Scholle, 1996, Regional setting and Sedimentary Geology, v. 130, p. 249–268, doi:10.1016
role of meteoric water in dolomite formation and diagen- /S0037-0738(99)00118-9.
esis in an evaporite basin: Studies in the Zechstein (Perm- Tilly, H. P., B. J. Gallagher, and T. D. Taylor, 1982, Methods
ian) deposits of Poland: Sedimentology, v. 43, p. 1005– for correlating porosity data in a gypsum-bearing carbon-
1023, doi:10.1111/j.1365-3091.1996.tb01516.x. ate reservoir: Journal of Petroleum Technology, v. 34,
Richter-Bernburg, G., 1985, Zechstein-Anhydrite: Fazies p. 2449–2454.
und genese: Geology of Jahrbuch Reihe A, v. 85, p. 3–82. Van der Plas, L., and A. C. Tobi, 1965, A chat for judging the
Riley, C. M., and J. V. Byrne, 1961, Genesis of primary struc- reliability of point-counting results: American Journal of
tures in anhydrite: Journal of Sedimentary Petrology, Science, v. 263, p. 87–90, doi:10.2475/ajs.263.1.87.
v. 31, p. 553–559. Vanko, D. A., and W. Bach, 2005, Heating and freezing ex-
Rosen, M. R., and J. K. Warren, 1990, The origin and signifi- periments on aqueous fluid inclusions in anhydrite: Rec-
cance of groundwater-seepage gypsum from Bristol Dry ognition and effects of stretching and the low-temperature
Lake, California, U.S.A.: Sedimentology, v. 37, p. 983– formation of gypsum: Chemical Geology, v. 223, p. 35–
996, doi:10.1111/j.1365-3091.1990.tb01840.x. 45, doi:10.1016/j.chemgeo.2004.11.021.
Schreiber, B. C., and M. El Tabakh, 2000, Deposition and Warren, J. K., 1991, Sulfate-dominated sea-marginal and
early alteration of evaporites: Sedimentology, v. 47, platform evaporative settings, in J. L. Melvin, ed., Evap-
p. 215–238, doi:10.1046/j.1365-3091.2000.00002.x. orites, petroleum and mineral resources: Developments
Sonnenfeld, M. D., and T. A. Cross, 1993, Volumetric par- in sedimentology: Amsterdam, Netherlands, Elsevier,
titioning and facies differentiation within the Permian v. 50, p. 477–533.
upper San Andres Formation of Last Chance Canyon, Warren, J. K., 2006, Evaporites: Sediments, resources and
Guadalupe Mountains, New Mexico, in R. G. Loucks and hydrocarbons: Berlin, Germany, Springer, 601 p.
J. F. Sarg, eds., Advances and applications of carbonate Williams-Stroud, S., and J. Paul, 1997, Initiation and growth
sequence stratigraphy: AAPG Memoir 57, p. 435–474. of gypsum piercement structures in the Zechstein Basin:
Tanguay, H. L., and G. M. Friedman, 2001, Petrophysical fa- Journal of Structural Geology, v. 19, p. 897–907, doi:10
cies of the Ordovician Red River Formation, Williston .1016/S0191-8141(97)00017-5.
Basin, U.S.A.: Carbonate and Evaporites, v. 16, p. 71– Worden, R. H., P. C. Major, and A. E. Fallick, 1997, Sulfur
92, doi:10.1007/BF03176227. cycle in buried evaporites: Geology, v. 25, p. 643–646,
Testa, G., and S. Lugli, 2000, Gypsum-anhydrite transforma- doi:10.1130/0091-7613(1997)025<0643:SCIBE
tions in Messinian evaporites of central Tuscany (Italy): >2.3.CO;2.

390 Origins of Gypsum in Deep Carbonate Reservoirs

You might also like