You are on page 1of 15

SEDIMENTOLOGY AND FRACTAL-BASED ANALYSIS OF PERMEABILITY DATA, JOHN HENRY

MEMBER, STRAIGHT CLIFFS FORMATION (UPPER CRETACEOUS), UTAH, U.S.A.

JAMES W. CASTLE, FRED J. MOLZ, SILONG LU,* AND CYNTHIA L. DINWIDDIE**


Departments of Geological Sciences and Environmental Engineering & Science,
Clemson University, Clemson, South Carolina 29634-0919, U.S.A.
e-mail: jcastle@clemson.edu

ABSTRACT: Interpretation of depositional environments combined in recent studies (e.g., Weber 1982; Weber and van Geuns 1990; Hurst and
with field measurement of permeability for a portion of the Upper Rosvoll 1991; Kerans et al. 1994; White and Barton 1999). However, the
Cretaceous Straight Cliffs Formation near Escalante, Utah, provides small-scale variation (at a sample spacing of approximately 20 cm or less)
new results for understanding and modeling facies-dependent perme- of permeability is not well understood. This study examines the character-
ability variations. Offshore, transition–lower shoreface, upper shore- istics of small-scale permeability variation along horizontal and vertical
face, and foreshore environments are interpreted for part of the John transects through two sandstone facies that differ both depositionally and
Henry Member on the basis of outcrop investigation. Using a newly lithologically.
designed drillhole minipermeameter probe, permeability was measured Collection of field permeability data in the context of a sedimentological
for two of the facies within this unit: lower-shoreface bioturbated sand- understanding aids in conditioning computer models to outcrop data, which
stone and upper-shoreface cross-bedded sandstone. Approximately 500 can improve subsurface characterization and flow simulation. Several pre-
permeability measurements at a sample spacing of 15 cm were made vious studies have reported values of permeability measured from outcrops
along four vertical profiles and three horizontal transects on a 6 m 3 (e.g., Stalkup and Ebanks 1986; Goggin 1988; Tyler et al. 1991; Goggin
21 m outcrop. et al. 1992; Davis et al. 1993; Fisher et al. 1993; Barton 1994; Liu et al.
Permeability ranges from 41 to 1,675 millidarcies (md) in the bio- 1996; Dutton and Willis 1998; Willis 1998; Willis and White 2000). Data
turbated sandstone facies, which is massive-bedded, moderately to from these outcrop investigations have confirmed the importance of un-
well-sorted, and very fine- to fine-grained. In contrast, permeability derstanding sediment depositional processes in predicting and modeling the
ranges from 336 to 5,531 md in the moderately to moderately well permeability distribution. However, some of the previous methods used to
sorted, fine- to medium-grained, cross-bedded sandstone facies. A high measure outcrop permeability can introduce problems in data quality
degree of variability in permeability of the cross-bedded facies is caused by a poor seal between the instrument and the outcrop surface.
caused by small-scale variations in grain size and structure related to These problems are summarized by Dinwiddie (2001). In order to obtain
depositional processes. The geometric mean permeability in the bio- permeability values of sufficiently high quality for the current investigation,
turbated sandstone is 253 md, compared with 1,395 md in the cross- a new field permeameter, called the drillhole minipermeameter, was de-
bedded facies. signed and built (Dinwiddie et al. 2000; Dinwiddie 2001; Molz et al. 2002;
While the facies-dependent differences in permeability (k) are ap- Dinwiddie et al. 2003). This instrument was used successfully to collect
parent and related to depositional and biological processes, fractal- the field permeability data for our study.
based statistical analysis of the horizontal ln(k) increments yields near-
ly identical results for the bioturbated facies and the cross-bedded fa- Geologic Setting
cies, possibly suggesting an underlying statistical commonality in the
formation of both facies. Increment distributions from both facies ap- The study area is located approximately 10 km northwest of the town
pear similar with peaking around the mean. Ln(k) increments from of Escalante, Utah, where the Straight Cliffs Formation crops out along the
the smaller vertical data set appear similar also, but with a variance northeastern margin of the Kaiparowits Plateau. The Straight Cliffs For-
approximately 3.7 times larger than the horizontal value. Variance mation is underlain by the Upper Cretaceous Tropic Shale and overlain by
scaling analyses of horizontal and vertical data both yield a Hurst co- the Upper Cretaceous Wahweap Formation (Peterson 1969). A wide range
efficient near 0.26, which is characteristic of negative spatial correla- of fluvial to marine depositional settings, including both wave-dominated
tion of the increments. The methodology developed herein offers a po- and tide-dominated environments, are represented within the Straight Cliffs
tential high-resolution alternative to existing methods for understand- Formation (Peterson 1969; Shanley and McCabe 1991, 1993; Shanley et
ing and characterizing subsurface properties. al. 1992; Hettinger et al. 1993; Hettinger 1995; Hettinger et al. 1996; Ste-
phen and Dalrymple 2002). In general, alluvial and coastal-plain deposits
INTRODUCTION
in the formation grade eastward into marine deposits (Shanley et al. 1992;
Hettinger et al. 1996).
Objectives and Significance The Straight Cliffs Formation represents deposition near the western
edge of the Western Interior seaway, which covered most of modern-day
The objectives of this investigation are to interpret the sedimentology of central and midwestern North America during Middle to Late Cretaceous
a portion of the John Henry Member of the Straight Cliffs Formation (Up- time (Peterson 1969; Jordan 1995). Foreland basin subsidence in this area,
per Cretaceous) near Escalante in southern Utah (Figs. 1, 2) and to char- which is east of the Sierra Nevada arc and the Sevier Thrust belt, was
acterize the nature of small-scale permeability variations. An understanding caused by flexural loading in response to uplift associated with subduction
of the permeability distribution in sedimentary facies and the relationship of the Pacific plate under the North American plate and with compression
to depositional processes are key to predicting subsurface flow patterns and caused by extension in eastern North America (Jordan 1981, 1995).
the occurrence of hydrocarbons and groundwater, as has been emphasized The Straight Cliffs Formation is divided into four members (Peterson
1969): the Tibbet Canyon Member, Smoky Hollow Member, John Henry
* Present address: Tetra Tech, Inc., 2110 Powers Ferry Road, Suite 202, Atlanta,
Member, and Drip Tank Member (Fig. 2). The Early Coniacian to Late
Georgia 30339, U.S.A. Santonian John Henry Member, which includes the interval studied, con-
** Present address: Southwest Research Institute, Center for Nuclear Waste Reg- tains coal-bearing continental strata to the west that grade eastward into
ulatory Analyses, San Antonio, Texas 78238, U.S.A. nearshore-marine deposits, including shoreface sandstones (Shanley and

JOURNAL OF SEDIMENTARY RESEARCH, VOL. 74, NO. 2, MARCH, 2004, P. 270–284


Copyright q 2004, SEPM (Society for Sedimentary Geology) 1527-1404/04/074-270/$03.00
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 271

FIG. 1.—Location map.

McCabe 1991; Hettinger et al. 1996). The John Henry Member is underlain to the informal stratigraphic unit called the ‘‘B sandstone’’ by Hettinger et
by coal-bearing coastal-plain and braided-river deposits of the Smoky Hol- al. (1996). The base of this unit in some of the previously studied outcrops
low Member and overlain by fluvial sandstone of the Drip Tank Member is marked by a pebble bed interpreted as a transgressive lag on a maximum
(Peterson 1969, Hettinger 1995). flooding surface (Shanley et al. 1992; Hettinger et al. 1993; Hettinger
On the basis of correlation with outcrops studied by Hettinger et al. 1995). Because of the lateral continuity and excellent quality of outcrops,
(1993) several kilometers to the south, the interval of our study corresponds the flooding surface can be traced into our study area from the outcrops to

FIG. 2.—Stratigraphic divisions of the Straight


Cliffs Formation for the study area, eastern part
of Kaiparowits Plateau. The ‘‘A sandstone’’
through ‘‘G sandstone’’ represent informal
stratigraphic units after Hettinger et al. (1996).
Position of age boundaries is approximate. After
Peterson (1969), Eaton (1991), Shanley and
McCabe (1991), Hettinger et al. (1993), and
Hettinger et al. (1996).
272 J.W. CASTLE ET AL.

TABLE 1.—Depositional features in outcrops studied.

Depositional
Environment Thickness Lithology Physical Features Biological Features
Foreshore 9–14 m Horizontally laminated, moderately Common horizontal lamination; minor low-angle cross- Minor to common shell fragments; rare burrows
well sorted fine- to medium-grained bedding and planar-tabular cross-bedding in lower part in interval, becoming more com-
sandstone mon upward (Skolithos, Planolites, and possi-
bly Psilonichnus)
Upper Shoreface 6–8 m Cross-bedded, moderately to moderate- Common trough cross-bedding and low-angle cross-bed- Common plant fragments and common shell
ly well sorted fine- to medium- ding; minor horizontal lamination and planar-tabular fragments on bedding planes; rare horizontal
grained sandstone cross-bedding; lower contact is erosional and laterally and vertical burrows
continuous
Transition-Lower Shoreface 10–15 m Moderately to well sorted, very fine- to Massive, bioturbated intervals; minor, resistant sandstone Abundant horizontal and vertical burrows includ-
fine-grained sandstone; overall up- interbeds (15–100 cm thick) containing low-angle ing Ophiomorpha, Planolites, and Thalassino-
ward increase in grain size cross-bedding and hummocky cross-stratification ides; minor Ophiomorpha, Palaeophycus, and
Skolithos in hummocky cross-stratified beds;
minor plant fragments; minor shell fragments
Offshore 9–12 m Interlaminated to thinly interbedded Minor ripple cross-lamination, lenticular bedding, and Common, small plant fragments; minor, small
shale, siltstone, and very fine- to flaser bedding horizontal and vertical burrows; minor shell
fine-grained sandstone fragments

the south studied by Hettinger et al. (1993). Above this surface, Hettinger a seal between the probe and the sides of the hole. Pressurized nitrogen
et al. (1993) interpreted offshore marine shale coarsening upward to shore- was then injected into the formation. Once steady-state conditions were
face sandstone. These marine strata represent the lowermost part of the reached, the in situ permeability was calculated from injection pressure,
highstand systems tract of the A-sequence in the John Henry Member (Het- flow rate, and knowledge of the system geometry (Dinwiddie 2001; Din-
tinger et al. 1993; Hettinger et al. 1996). widdie et al. 2003). As soon as measurements were completed, the seal
was broken by turning the torque wheel and the probe was removed. Three
Methods steady-state measurements were performed at different gas pressures for
each drillhole.
Approximately 425 m of outcrop section were described from 14 sites Advantages of the approach described above include the reduction of
within the study area. Gamma-ray profiles were obtained from each outcrop questionable permeability measurements caused by weathering of the out-
using a hand-held scintillometer. On the basis of the sequences of grain crop surface, a superior sealing mechanism around the air-injection zone,
size and sedimentary structures, along with comparison to modern and and the potential for making measurements at multiple depths below the
ancient analogs, depositional environments were interpreted. outcrop surface.
One of the outcrops, with a flat, nearly vertical face measuring approx-
imately 21 m across and 6 m high, was selected for study of small-scale
variation in permeability. A total of 515 permeability measurements from LITHOLOGY AND DEPOSITIONAL ENVIRONMENTS
two sandstone facies were made in triplicate at a sample spacing of 15 cm
along three horizontal transects (380 measurements) and four vertical pro- Offshore
files (135 measurements). To obtain an approximation of the mineralogy
Description.—The lower 9–12 m of the interval studied consists of in-
and fabric in each of the two sandstone facies, thin sections were prepared
terlaminated to thinly interbedded shale, siltstone, and very fine- to fine-
and examined from 32 measurement points. Sodium cobaltinitrite was ap-
grained sandstone (Table 1; Figs. 3, 4). Because of the low resistance of
plied to each thin section as a stain to aid in the petrographic identification
of potassium feldspar. Each thin section was point-counted (300 points this shaly interval to weathering, it tends to form vegetated slopes and is
each) for detrital mineral content, matrix, cements, and porosity. Secondary poorly exposed in most of the study area. Common, small plant fragments
porosity was identified following the criteria proposed by Schmidt and and minor, small, horizontal and vertical burrows are present in the shale
McDonald (1979). Grain size was estimated by measuring the maximum and siltstone. The sandstone beds contain minor ripple lamination and mi-
length of grains of approximately average size in each thin section. Sorting nor shell fragments. Minor lenticular bedding and flaser bedding are pre-
was estimated following methods of Beard and Weyl (1973) and Longiaru sent.
(1987). At the base of this interval, a pebble-conglomerate lag directly overlies
A newly designed drillhole minipermeameter probe (Dinwiddie et al. the marine flooding surface at the lower contact of the ‘‘B sandstone.’’
2000; Dinwiddie 2001; Molz et al. 2002; Dinwiddie et al. 2003) was used The pebble conglomerate, which ranges in thickness from 7 to 25 cm in
for the in situ measurement of permeability. The averaging volume of the the study area, contains poorly sorted pebbles in a fine-grained sandstone
probe is highly localized, and is thus amenable to determining small-scale matrix. Rare to minor shell fragments and plant fragments are present in
permeability variations (Molz et al. 2003). The probe was built for insertion the sandstone matrix.
into cylindrical holes drilled with a standard 5/8 inch (1.59 cm) masonry Interpretation.—On the basis of comparison with studies of modern
drill bit to a depth of 10.16 cm. The equipment used includes a nitrogen sediments and ancient examples (Leithold 1989; Swift et al. 1991; Davis
tank, flowmeter, pressure transducer, leak-proof tubing, and the probe. A and Byers 1993; Johnson and Baldwin 1996), the lower interval of inter-
commercially available mini-permeameter control system can be used with laminated to thinly interbedded shale, siltstone, and very fine-grained sand-
the new probe to detect flow rate and pressure and to interface with a stone is interpreted as representing deposition in an offshore marine envi-
portable computer that performs various tasks such as data acquisition, ronment. The thin sandstone beds and laminae are interpreted as the bases
transducer calibration, and permeability computation. of storm deposits, which may be analogous to laminated ‘‘storm-sand lay-
In preparation for testing, a cylindrical hole was drilled into the outcrop ers’’ that have been described from modern shelf muds (Reineck and Singh
surface beyond the zone of weathering and perpendicular to the rock face. 1972). The intervening shale and siltstone beds are thought to represent the
The drillhole was vacuumed to remove any dust or loose sand that remained muddy top of the underlying storm-deposited sand combined with mud
in the hole after drilling. The distal end of the probe was inserted into the deposited from events too weak to transport sand out from the nearshore
hole, and a torque wheel was turned to expand a rubber sleeve to provide environments.
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 273

FIG. 4.—Outcrop description at location 5. The upward coarsening suggests shore-


line progradation. The marine flooding surface that corresponds to the base of the
interval studied (Fig. 2) is present below the exposure at this location. The rectangle
on the location map corresponds to the outline of the study area shown on Figure
1. See Figure 3 for legend.

Transition–Lower Shoreface
Description.—Approximately 10–15 m of moderately to well sorted,
very fine- to fine-grained sandstone gradationally overlies the offshore fa-
cies of the interval studied (Table 1; Figs. 3, 4, 5). Massive-bedded, bio-
turbated intervals alternate with much thinner interbeds of hummocky
cross-stratified sandstone. Overall, grain size gradually increases upward
FIG. 3.—Outcrop description at location 1. A sedimentologic profile, natural gam- within this interval.
ma radiation curve, and interpretation are shown. The gradual upward decrease in Abundant horizontal and vertical burrows (Fig. 6A) are present within
gamma radiation reflects the decrease in clay content relative to sand. The lower
part of the interval studied is poorly exposed because of slope-forming shale. The the bioturbated intervals, which contain minor clay. Diameter of the bur-
marine flooding surface, which has been interpreted as the maximum flooding sur- rows, which are identified as Ophiomorpha and Planolites, ranges from
face of the A-sequence of the John Henry Member (Shanley et al. 1992; Hettinger approximately 0.5 to 1.5 cm, with lengths up to 15 cm. Most of the burrows
1995), was traced into our study area from outcrops studied by Hettinger et al. are gently curving to sinuous, and some are lined with pellets. Similarly
(1993) to the south. The top of the interval studied is above the top of the outcrop sized, branching burrows that occur in clusters near bedding contacts are
at this location. The rectangle on the location map corresponds to the outline of the
study area shown on Figure 1.
identified as Thalassinoides (Fig. 6B).
Rare mud clasts and minor fragments of plant and shell material are
present within the bioturbated intervals. However, internal sedimentary
structures are not apparent within most of the burrowed sand, probably
274 J.W. CASTLE ET AL.

FIG. 5.—Outcrop description at location 2.


Locations of the horizontal transects along which
permeability was measured are indicated (X, H,
D). The marine flooding surface, which is
present at the base of the interval studied, is
covered at this location. The rectangle on the
location map corresponds to the outline of the
study area shown on Figure 1. See Figure 3 for
legend.

because of destruction by burrowing. Rare ripple cross-lamination is pre- Duke 1985; Walker and Plint 1992). In a study of modern beach to offshore
sent, particularly near the tops of beds. deposits along the California coast, Howard and Reineck (1981) described
As many as eight stratified sandstone beds, each 15–100 cm in thickness, bioturbated sediments with parallel lamination, which they interpreted as
are present within the transition–lower shoreface interval. These beds are probable hummocky bedding. Ancient examples containing HCS have been
clearly defined by sharp upper and lower contacts and by generally lighter interpreted as forming between storm wave base and fair-weather wave
color than the bioturbated sandstone, particularly on weathered surfaces base in the upper offshore to lower shoreface zone (e.g., Krassay 1994;
(Fig. 6C). The lower contact is erosional into the underlying sandstone. Colquhoun 1995; Castle 2000). The upward increase in grain size in the
The beds contain a very low content of clay and, because of carbonate interval that we studied suggests possible shoreline progradation. In a pro-
cement, tend be more resistant to weathering than the bioturbated sand- grading upward-coarsening shoreline succession in Upper Cretaceous strata
stone. Internal stratification is horizontal to hummocky, with minor ripple of the Spanish Pyrenees, Ghibaudo et al. (1974) described similar shoreface
cross-lamination. Rare to minor vertical and horizontal burrows, including deposits of bioturbated sandstones with remnant sandstone beds having
Ophiomorpha, Palaeophycus, and Skolithos are present in the hummocky preserved lamination resembling HCS.
cross-stratified beds. Some of the hummocky beds can be traced laterally Deposition in a lower shoreface setting is consistent with occurrence of
along the outcrop for more than 2,700 m as they pinch out and then re- the trace fossils Ophiomorpha, Planolites, and Thalassinoides (Frey and
appear in the same stratigraphic position. Pemberton 1985; Pemberton et al. 1992; Pemberton and MacEachern 1995;
Interpretation.—The presence of hummocky cross stratification (HCS) Pemberton et al. 2001). Ophiomorpha, Palaeophycus, and Skolithos have
suggests that deposition was influenced by storm waves (Harms et al. 1975; been reported in other hummocky cross-stratified Upper Cretaceous sand-
Dott and Bourgeois 1982; Swift et al. 1983; Walker 1984; Brenchley 1985; stones interpreted as storm deposits (Frey 1990; Frey and Howard 1990).
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 275

FIG. 6.—Outcrop photographs. A) Burrows (predominantly Ophiomorpha) in lower-shoreface bioturbated sandstone. Much of this facies is extensively bioturbated, with
individual burrow structures commonly indistinct. From outcrop section 5, 14 m (Fig. 4). B) Thalassinoides burrows near base of sandstone bed in lower-shoreface
bioturbated sandstone. Width of exposure shown is 38 cm. From outcrop section 5, 13 m (Fig. 4). C) Hummocky cross-stratified sandstone bed within the lower shoreface
interval. Skolithos burrows (vertical) and Palaeophycus burrows (inclined) are present. Length of scale 5 6 cm. From outcrop section 5, 18 m (Fig. 4). D) Cross-bedded
upper-shoreface sandstone. From 1 m above base of upper-shoreface sandstone exposed 300 m north of outcrop section 1 (Fig. 3). Length of scale near left edge of
photograph is 5 cm.

Except for within the hummocky cross-stratified beds, most of the original defined and laterally continuous throughout the study area. The contact is
sedimentary structures, including interbedding between mud and sand, in erosional, showing approximately 1.5 m of relief among the outcrops stud-
the transition–lower shoreface interval were probably obliterated by bio- ied. The contact is directly overlain by trough cross-bedded, horizontally
turbation. In the lower-shoreface and transition zone, the extent of biotur- laminated, or low-angle cross-bedded sandstone. Lag deposits were not
bation, and hence the preservation of storm-generated beds, depends on the observed on the contact.
magnitude and frequency of storms and the overall rate of sedimentation Interpretation.—This interval is interpreted as upper shoreface, with the
(Reading and Collinson 1996). The hummocky cross-stratified beds pre- abundant trough cross-beds representing subaqueous migration of dunes.
served between the massive bioturbated intervals in our study area may Grain size and sedimentary structures are similar to those of modern, cross-
record wave processes that were more intense than during deposition of stratified shoreface deposits (Howard and Reineck 1981; Heron et al. 1984;
the bioturbated beds. Shipp 1984; Greenwood and Mittler 1985; Swift et al. 1991) and to ancient
upper-shoreface deposits (Swift et al. 1987; Walker and Plint 1992; Col-
Upper Shoreface quhoun 1995; Johnson and Baldwin 1996). The previous studies demon-
strated that sand deposition in this setting is controlled by a combination
Description.—Cross-bedded, moderately well sorted, fine- to medium- of storm and fair-weather processes. The sharp basal contact of the upper-
grained sandstone sharply overlies the lower shoreface interval (Table 1; shoreface facies is interpreted as a surf diastem, which is defined by Swift
Figs. 3, 4, 5). Intervals of trough cross-bedded sandstone 1–1.5 m thick et al. (2003) as a surface cut by repeated storm scour in the surf zone and
(Fig. 6D) alternate with beds of low-angle cross-bedded sandstone 0.1–0.3 identified in outcrop as the erosional base of the lowest high-angle cross-
m thick. Minor horizontal lamination, minor planar-tabular cross-bedding, bed set in a shoreface succession.
and rare ripple cross-lamination are also present. Common shell debris and
plant fragments occur on bedding planes, and rare horizontal to vertical Foreshore
burrows are present within the sandstone beds. In the outcrops studied, this
interval ranges in thickness from 6 to 8 m. Description.—An interval of predominantly horizontally laminated,
The basal contact with the underlying lower-shoreface interval is sharply moderately well sorted fine- to medium-grained sandstone, approximately
276 J.W. CASTLE ET AL.

FIG. 7.—Locations of vertical profiles and horizontal transects along which permeability was measured by drilling into the outcrop face, section 2 (Figure 5). The strike
of the outcrop face is approximately parallel to strike of the paleoshoreline, which has been interpreted previously as northwest–southeast (Peterson 1969). Each dot on
the diagram corresponds to a permeability measurement. Approximate positions of rock overhang and soil cover are shown.

9–14 m thick, overlies the upper shoreface interval (Table 1; Figs. 3, 5). PERMEABILITY STRUCTURE
Minor low-angle cross-bedding and planar-tabular cross-bedding are pre-
sent. Typically, beds of horizontally laminated and low-angle cross-bedded Data
sandstone 0.3–3.0 m thick alternate with beds of planar tabular cross-bed- Intrinsic permeability was measured in two facies at outcrop section 2
ded sandstone 0.15–0.3 m thick. A few beds within this interval contain a (Fig. 5): lower-shoreface bioturbated sandstone and upper-shoreface cross-
concentration of common to abundant shell fragments, and minor shell bedded sandstone. In the bioturbated sandstone, 353 measurements of per-
fragments are present on cross-bed foresets in other beds. In most outcrops meability were made along two horizontal transects and in the lower parts
studied, grain size shows a slight upward decrease; minor shale interbeds of four vertical profiles (Fig. 7). For the cross-bedded facies, 162 perme-
are present in the upper part of the interval in some outcrops. ability measurements were made along one horizontal transect and in the
Although burrows are rare in the horizontally laminated medium-grained upper parts of the four vertical profiles. The permeability values for all
sandstone of this interval, minor to common vertical and horizontal burrows measurement points were reported by Current (2001) and are shown graph-
are present in the upper, finer-grained part of the interval. The burrows, ically in Figures 8, 9, and 10.
which are identified as Skolithos, Planolites, and possibly Psilonichnus, are Permeability values range from 41 to 1,675 millidarcies (md) in the
approximately 0.5–1.0 cm in diameter and up to 4 cm long. bioturbated sandstone facies. The arithmetic mean permeability value in
The basal contact of this interval varies from sharp to gradational. The this facies is 276 md, and the geometric mean is 253 md. The range of
upper contact, which marks the top of the interval studied, is sharp and permeability variation in this facies along both the horizontal transects and
overlain by medium-grained sandstone containing abundant oyster shells the vertical profiles is generally less than one order of magnitude. Grain
and fragments and abundant mud clasts. This bed ranges from about 0.75 size in the bioturbated facies also shows little variation laterally and ver-
to 1.5 m in thickness. tically.
Interpretation.—Horizontally laminated and low-angle cross-bedded Permeability in the cross-bedded facies ranges from 336 to 5,531 md,
sandstone of this interval is interpreted as representing beach-face deposi- with an arithmetic mean permeability equal to 1,746 md and a geometric
tion in a foreshore setting. The occurrence of horizontal lamination, low- mean of 1,395 md. Permeability variation exceeding 1,000 md is not un-
angle cross-bedding, shell fragments, upward fining, and burrows are con- common among nearby sample locations. At a vertical position of 5 to 7
sistent with descriptions of modern and ancient foreshore deposits (Clifton sample locations above the contact between the bioturbated facies and the
et al. 1971; Hill and Hunter 1976; Hunter et al. 1979; Howard and Reineck cross-bedded facies, a high-permeability peak can be correlated among the
1981; Heron et al. 1984; Short 1984; Clifton 1988; Reading and Collinson vertical profiles (Fig. 8). Permeability at this position ranges from 1,924
1996). md in profile V-2 to 5,126 md in profile V-3.
Internal structures, finer grain size, and burrowing in the upper part of
this interval indicate lower-energy deposition than represented by the un- Geologic Controls
derlying horizontally laminated foreshore deposits. From position in the Compared to the bioturbated facies, permeability values are greater in
overall succession and from comparison with other examples (e.g., Swift the cross-bedded facies because of coarser grain size, greater primary po-
et al. 1991; Walker and Plint 1992), the upper part of this interval may rosity, and smaller percentages of clay matrix and cement (Table 2). On
represent backshore deposition. Bioturbated sediments, similar to those in the basis of petrographic observation, primary pores are larger and better
the upper part of the interval studied, have been described from modern connected in the cross-bedded facies than in the bioturbated sandstone,
backshore environments (Hill and Hunter 1976; Frey et al. 1984; Frey and which is related to coarser grain size in the cross-bedded facies. Small
Pemberton 1987). On the basis of comparison with similar contacts and amounts of grain-moldic secondary porosity have formed by dissolution of
associated lithologies from other ancient successions (Bergman and Walker feldspar grains in both sandstone facies. Although the percentage of moldic
1987; Pattison and Walker 1992; Bergman 1994; Arnott et al. 1995), the porosity is greater in the bioturbated facies than in the cross-bedded facies,
sharp upper contact is interpreted as a wave-cut surface overlain by a lag its contribution to permeability is small because of low connectivity among
of reworked material. grain-moldic pores. The growth of minor, pore-lining chlorite cement has
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 277

FIG. 8.—Vertical permeability profiles. Most permeability values in the cross-bedded sandstone are greater than values in the bioturbated sandstone. Intersections of
horizontal transects D, H, and X are shown.

FIG. 9.—Horizontal permeability transects.


Permeability values vary over a greater range in
the cross-bedded sandstone (transect X) than in
the bioturbated sandstone (transects H and D).
Intersections of vertical profiles V1, V2, V3, and
V4 are shown.
278 J.W. CASTLE ET AL.

FIG. 10.—Horizontal permeability transects H


and D (bioturbated sandstone) plotted at an
expanded scale. Intersections of vertical profiles
V1, V2, V3, and V4 are shown.

reduced permeability in both sandstone facies. From qualitative petrograph- icantly, so the statistical parameters are not unique, but rather are a function
ic observations from both facies, the individual samples with the highest of vertical position. The same observation applies to the horizontally dis-
permeability values contain the largest amount of porosity and the least tributed data shown in Figure 10. Dividing these data sets into a left half
amount of matrix and cement. and right half yields m 5 330 md and s 2 5 17,628 md 2 for the left half,
Field permeability measurements from lower-shoreface bioturbated sand- but m 5 222 md and s 2 5 7,393 md 2 for the right half, again indicating
stone and upper-shoreface cross-bedded sandstone demonstrate facies-de- dependence of statistical parameters on position. Strictly speaking, it is not
pendent variations in permeability. Because of lithologic homogenization appropriate to perform a statistical analysis of spatial data that are not
by extensive burrowing, the bioturbated sandstone exhibits a small range stationary in space or temporal data that are not stationary in time. Even
of permeability variation over a scale of several meters. In contrast, per- though means, variances, and (possibly) other parameters come out of such
meability in the cross-bedded facies varies by more than an order of mag- analyses, what they represent is not clear, because the results depend on
nitude over distances as small as 30 cm, which is caused predominantly the spatial or temporal variable in an unknown way. So how should one
by variations in grain size and sorting related to the depositional processes proceed?
(Current 2001). The primary textural properties are preserved better in the A traditional approach is to attempt to detrend the data (e.g., Gelhar
cross-bedded facies than in the bioturbated facies, where they have been 1993). Using this procedure, one attempts to identify the large-scale trends
modified by burrowing. in the data and remove them, so that what is left is stationary. Then sta-
tistical analyses are performed on the detrended data. This approach is often
STATISTICAL ANALYSES OF PERMEABILITY STRUCTURE not satisfactory, however, because the detrending process is highly subjec-
Like virtually all small-scale permeability (k) data collected during the tive. A more rigorous approach, increasingly supported by measurements,
past decade (Painter 2001; Lu et al. 2002), the k or ln(k) data from our is to accept the fact that most data from natural sedimentary systems are
study appear irregular and spatially nonstationary in a statistical sense. Ir- nonstationary, and to make use of the theory of nonstationary stochastic
regular means that the slope of k versus distance changes abruptly in a processes with stationary increments (Lu et al. 2002). The mathematical
variable ‘‘saw-tooth’’ fashion (a discontinuous first derivative; Fig. 10), basis for this theory was developed during the first half of the past century
and nonstationary means that statistical parameters computed from the data, (Feller 1971, and earlier editions) and first applied to petroleum geology
such as mean and variance, depend on position. For example, if each of during the late 1980s (Hewitt 1986) and to hydrogeology during the early
the four plots shown in Figure 8 is divided into a top half and a bottom 1990s (Neuman 1990; Molz and Boman 1993).
half, then one can calculate a mean (m) and variance (s 2) for each half. In the following analysis of the data collected in our study, the basic
Proceeding from V-1 to V-4 across the top half and averaging the results, property will be ln(k), which was measured at a number of points a known
one computes a m of 1,237 md and a s 2 of 901,356 md 2. Similar analysis distance apart. Then each pair of points defines an increment of ln(k) [
of the bottom half of the data yields m 5 336 md and s 2 5 52,544 md 2. Dhln(k) 5 ln(k(x 1 h))—ln(k(x)) that is associated with the constant mea-
As apparent from visual inspection of the data, these values differ signif- surement separation, h. Thus n points along a transect where ln(k) is mea-

TABLE 2.—Texture, composition, and fabric of facies tested for permeability in the John Henry Member. Data are averaged from petrographic counting of 300 points
per sample.

**Proportion of Detrital Proportion of Fabric Components (%)


Framework Grains (%)
Number of Mean Grain *Mean Lithic Framework Clay Calcite Chlorite Primary Secondary
Facies Samples Size (mm) Sorting Quartz Feldspar Fragments Grains Matrix Cement Cement Porosity Porosity
Upper Shoreface 12 0.23 3.4 65. 2.4 32. 67. 1.0 3.0 2.3 25. 1.8
(Cross-bedded Sandstone)
Lower Shoreface 20 0.11 4.8 80. 3.5 16. 65. 2.9 5.2 2.5 21. 3.3
(Bioturbated Sandstone)
* Mean sorting estimated using the methods of Beard and Weyl (1973) and Longiaru (1987): 1 5 very poor; 2 5 poor; 3 5 moderate; 4 5 moderately well; 5 5 well; 6 5 very well.
** Feldspar mineralogy is orthoclase. Lithic grains are predominantly chert, with minor sedimentary and volcanic rock fragments. The higher proportion of detrital quartz to lithic grains in the bioturbated sandstone than
in the cross-bedded sandstone is probably related to grain size difference between the two facies.
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 279

FIG. 11.—Empirical frequency distribution for the horizontal ln(k) increments (‘‘Sample Frequency’’) along with that of the best fitting Gaussian distribution (‘‘Normal
Distribution’’). The data are from horizontal transects D and H in the bioturbated sandstone facies. The empirical distribution was obtained by first computing a cumulative
distribution function and then differentiating this curve numerically. A Chi-Square test made on the empirical data directly suggested that the Gaussian fit was significant
at only the 51% confidence level, with apparently non-Gaussian peaking observed around the mean.

sured define n-1 ln(k) increments. All statistical analyses are then performed called the ‘‘fractal/facies model,’’ which is a hypothesis that Dhln(k) fre-
on the increment distributions, the set of values of Dhln(k), and not on the quency distributions within a single facies will be Gaussian. The often
ln(k) distributions directly. According to theory, such distributions should observed Levy-like distributions would then result because of mixing data
have a mean of zero that is independent of h and a variance that increases from several different facies, which would be equivalent to superimposing
with h as a power function (Molz et al. 1997). Traditionally, this power- several Gaussian distributions, in general of different variance. Thus, the
law dependence is written as s 2(h) 5 s 2(1)h 2H, with H being called the approach advocated by Lu et al. (2002) would require that data gathered
Hurst coefficient. Furthermore, property distributions that are constructed from each facies of a multi-facies structure be analyzed separately. To the
from sets of stationary increments, with and without correlation of the extent practical, this approach was followed in the analysis of the present
increments, display scaling properties characteristic of a group of self-affine dual-facies data set.
stochastic processes (to distinguish from self-similar when H 5 ½) known As shown in Figure 11, the ln(k) increment frequency distribution from
as stochastic fractals (Molz et al. 1997). The basic analysis procedure then the horizontal transects of the bioturbated sandstone displays some aspects
is to: of a Gaussian distribution of mean 0.0013 and variance 0.122; however, a
chi-square test suggests that the hypothesis of a Gaussian fit is significant
(1) compute Dhln(k) frequency distributions for a variety of h values;
at only the 51% confidence level. This result is likely due to the poor fit
(2) assuming that such distributions are Gaussian, compute the mean
in the vicinity of the mean, where one would expect the comparison to be
(should be zero) and variance for the various Dhln(k) sets;
relatively good for a true Gaussian fit. The frequency distribution of the
(3) plot s 2 versus h on double logarithmic paper; and
smaller horizontal data set from the cross-bedded facies, shown in Figure
(4) if the plot is a straight line, determine the slope of the line, which
12, displays more noise, but behavior in the vicinity of the mean is similar.
enables one to calculate H.
In addition, the best-fit Gaussian has a mean of 0.0014 and a variance of
In practice, the variance calculation is highly sensitive to the relatively 0.127, nearly indistinguishable from the bioturbated facies. This variance
small number of data points for the larger lags, so a lot of noise is produced. result was unexpected, and it indicates that in a statistical ln(k) increment
This has led to the development of less data-sensitive techniques such as sense the two facies are not distinguishable in terms of mean and variance,
rescaled range analysis and dispersional analysis (Lu et al. 2002). These at least in the horizontal direction. The combined increment data are shown
latter two techniques are simply mathematical surrogates for the use of in Figure 13, which is a plot of all the horizontal Dhln(k) with h 5 15 cm.
direct variance scaling plots to determine H, which is discussed later. Having a mean of 0.0013 and a variance of 0.123, this combined plot is a
When such analyses were applied previously, most scaling requirements good example of the appearance of a spatially stationary stochastic process,
were met (Painter 2001), but the resulting frequency distributions for as required by theory.
Dhln(k) usually appeared non-Gaussian, with a distinct resemblance to the Analysis of all the vertical Dhln(k) data resulted in the frequency distri-
Levy distribution, which has undesirable statistical properties (Painter bution and best-fit Gaussian shown in Figure 14, with the qualitative result
1995; Liu and Molz 1997). However, later studies showed that such dis- not unlike that shown in Figure 12. Here the mean is 0.033 and the sig-
tributions, while appearing Levy-like, did not exhibit true Levy behavior nificantly larger variance is 0.454. The individual means to two decimal
in the tails of the distributions (Lu and Molz 2001; Painter 2001). This is places are 0.03 for both facies, with a variance of 0.33 for the cross-bedded
fundamental, because it is the tail behavior that gives Levy distributions sandstone and 0.46 for the bioturbated sandstone. Thus in the vertical di-
their famous property of infinite variance. rection there is a significant difference in variance between facies. In sed-
In order to explain the appearance of Levy distributions as possible ar- imentary deposits it is common for variability perpendicular to bedding to
tifacts of the data analysis-procedure, Lu et al. (2002) presented what they differ from that along bedding, so the factor of 3.7 increase in the variance
280 J.W. CASTLE ET AL.

FIG. 12.—Empirical frequency distribution for the horizontal ln(k) increments from transect X in the cross-bedded sandstone facies. The best-fit Gaussian is essentially
the same as that for transects D and H (Fig. 11), indicating that ln(k) increments from both facies display similar statistical structure.

of the vertical increments compared to the horizontal increments is not ysis to all of the horizontal data and all of the vertical data. The results
surprising, and can be viewed as a type of statistical anisotropy, as dis- yielded Hurst coefficients of 0.24 and 0.28 for the horizontal and vertical
cussed in Lu et al. (2003). data, respectively. (It may be of interest to note that both results are near
When applied to individual data sets, direct variance scaling tests are 0.25, a value inferred by Neuman (1990) on the basis of an analysis of
generally quite noisy (Mandelbrot and Wallis 1969), as mentioned previ- large-scale tracer tests and the resulting observation of scale dependence
ously. In addition, the results described in the previous paragraph indicate of hydrodynamic dispersivity for a large variety of materials. Given the
that the two facies are nearly statistically indistinguishable in a Dhln(k) limitations of our data set from the perspectives of size and rock type, this
sense in the horizontal direction. Therefore, we applied dispersional anal- may be a coincidence.) Straight lines of slopes 2H (half slopes of 0.24 and

FIG. 13.—Ln(k) increments for all of the horizontal data as a function of position. This plot provides a good example of the ‘‘appearance’’ of a portion of a spatially
stationary stochastic process. These numbers represent the statistically meaningful data extracted from the irregular, nonstationary, property distribution. That one should
be able to do this is the basis of the stochastic fractal theory.
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 281

FIG. 14.—Empirical frequency distribution for the vertical ln(k) increments (profiles shown in Fig. 8). Here the best-fit Gaussian has a variance that is 3.7 times larger
than the horizontal value (Fig. 13). This is a type of statistical anisotropy that was discussed by Lu et al. (2003).

0.28, respectively) were superimposed on the direct variance plots deter- peaking around the mean that seems to appear in all the computed ln(k)
mined independently as shown in Figures 15 and 16 for the horizontal and increment frequency distributions and causes the plots to display an ap-
vertical data, respectively, with the result confirming that the H values parent non-Gaussian shape.
resulting from the dispersional analysis are reasonable. Magnitudes of H
in the vertical and horizontal directions are similar, although the ln(k) in- CONCLUSIONS
crement variances are quite different. Very few data sets contain both hor-
izontally distributed and vertically distributed data, so additional measure- From sedimentological study of strata within the ‘‘B sandstone’’ of the
ments are needed to see whether similar results hold for different sedi- John Henry Member, the following depositional environments are inter-
mentary formations. In particular, one would like to better understand the preted: offshore, transition–lower shoreface, upper shoreface, and fore-

FIG. 15.—Double natural logarithm plot of the variance of the increments (var) as a function of lag (h) for all horizontal data. The straight line was computed independently
of the variance using dispersional analysis (Lu et al. 2003). The Hurst coefficient (H) of 0.24 is one-half the slope of the line, and indicates that the increments are
negatively correlated.
282 J.W. CASTLE ET AL.

FIG. 16.—Double natural logarithm plot of the variance of the increments (var) as a function of lag (h) for all vertical data. The straight line was computed independently
of the variance using dispersional analysis (Lu et al. 2003). The Hurst coefficient (H) of 0.28 is one-half the slope of the line, and indicates negative increment correlation.

shore. Shoreline progradation is indicated by upward coarsening in grain Jinjun Wang, who performed the statistical calculations underlying Figures 11
size and by the sequence of sedimentary structures, which indicate an up- through 16. We express our sincere appreciation to David Dominic, Donald Swift,
and Gary Weissmann, who reviewed the manuscript and provided many valuable
ward increase in wave energy. Closely spaced field permeability measure- suggestions that significantly improved the paper.
ments made in the lower-shoreface bioturbated sandstone and upper-shore- This research was supported by the U.S. Department of Energy, Fossil Energy
face cross-bedded sandstone demonstrate facies-dependent variations and Oil Technology Program, through the National Petroleum Technology Office under
similarities that are caused by lithologic differences related to depositional contract number DE-AC26-98BC15119.
processes. The work conducted by C.L. Dinwiddie as described in this article was initiated
As illustrated by our study of the John Henry Member, sedimentary and completed at Clemson University when she was a Ph.D. candidate, before the
beginning of her current employment with the CNWRA.
deposits are typically characterized by distinct facies that may change
abruptly in both the vertical and horizontal directions, with facies structure
relating to the depositional processes. Lu et al. (2002) proposed the fractal/ REFERENCES
facies model as a new approach for quantifying the underlying structure of
ARNOTT, R.W.C., HEIN, F.J., AND PEMBERTON, S.G., 1995, Influence of the ancestral Sweetgrass
geological heterogeneity on the basis of an understanding of facies. Rather arch on sedimentation of the Lower Cretaceous Bootlegger Member, north-central Montana:
than using Levy probability density functions (PDFs) to describe ln(k) in- Journal of Sedimentary Research, v. B65, p. 222–234.
crement distributions, the fractal/facies hypothesis proposes Gaussian BARTON, M.D., 1994, Outcrop characterization of architecture and permeability structure in
fluvial–deltaic sandstones within a sequence stratigraphic framework, Cretaceous Ferron
PDFs, with a (potentially) different PDF for each facies (Lu et al. 2002). Sandstone, Utah [unpublished Ph.D. dissertation]: University of Texas at Austin, 260 p.
However, statistical analysis of data from the bioturbated and cross-bedded BEARD, D.C., AND WEYL, P.K., 1973, Influence of texture on porosity and permeability of
sandstone facies of the John Henry Member, while providing evidence for unconsolidated sand: American Association of Petroleum Geologists, Bulletin, v. 57, p. 349–
non-Levy behavior of ln(k) increments, particularly the tail behavior, yield- 369.
BERGMAN, K.M., 1994, Shannon Sandstone in Hartzog Draw–Heldt Draw Fields (Cretaceous,
ed similar Dhln(k) frequency distributions in both facies in the horizontal Wyoming, USA) reinterpreted as lowstand shoreface deposits: Journal of Sedimentary Re-
direction. Variance scaling of Dhln(k), as determined by the value of the search, v. B64, p. 184–201.
Hurst coefficient, also appeared consistent across facies. These results may BERGMAN, K.M., AND WALKER, R.G., 1987, The importance of sea-level fluctuations in the
formation of linear conglomerate bodies: Carrot Creek Member of Cardium Formation,
suggest some type of underlying commonality in the formation of these Cretaceous western interior seaway, Alberta, Canada: Journal of Sedimentary Petrology, v.
two facies that is not obvious from visual inspection. Additional data and 57, p. 651–665.
statistical analysis are needed to fully determine the applicability of the BRENCHLEY, P.J., 1985, Storm-influenced sandstone beds: Modern Geology, v. 9, p. 369–396.
CASTLE, J.W., 2000, Recognition of facies, bounding surfaces, and stratigraphic patterns in
stochastic fractal model (nonstationary stochastic functions with stationary foreland-ramp successions: An example from the Upper Devonian, Appalachian basin,
increments) as well as the fractal/facies hypothesis, because non-Gaussian U.S.A.: Journal of Sedimentary Research, v. 70, p. 896–912.
peaking in the vicinity of the mean appeared in all empirical frequency CLIFTON, H.E., 1988, Sedimentologic approaches to paleobathymetry, with applications to the
Merced Formation of central California: Palaios, v. 3, p. 507–522.
distributions and Chi-square tests for a normal fit were not definitive. CLIFTON, H.E., HUNTER, R.E., AND PHILLIPS, R.L., 1971, Depositional structures and processes
in the non-barred high-energy nearshore: Journal of Sedimentary Petrology, v. 41, p. 651–
ACKNOWLEDGMENTS 670.
COLQUHOUN, G.P., 1995, Siliciclastic sedimentation on a storm- and tide-influenced shelf and
This study would not have been possible without the unselfish contributions of shoreline: the Early Devonian Roxburgh Formation, NE Lachlan Fold Belt, southeastern
Australia: Sedimentary Geology, v. 97, p. 69–98.
Bob Hettinger, who introduced us to the area, guided us in the field, and provided CURRENT, C.L., 2001, Characterization of geologic controls on permeability and their incor-
valuable discussion. We thank Nicole McMahan and Joey Koon for their untiring poration into a three-dimensional geologic model of the Temblor Formation, Coalinga, Cal-
efforts in describing outcrops and tracing beds. Caitlin Current and Robert Bridges ifornia [unpublished Master’s thesis]: Clemson University, 113 p.
assisted with collecting the outcrop permeability data. We gratefully acknowledge DAVIS, H.R., AND BYERS, C.W., 1993, The role of bottom currents and pelagic settling in the
the contribution of Chris Hepler, who diligently point counted the thin sections, and deposition of shale in an oxygen-stratified basin: case study of the Mowry Shale (Cretaceous)
SEDIMENTOLOGY AND FRACTAL ANALYSIS, UPPER CRETACEOUS, UTAH 283

of Wyoming, in Caldwell, W.G.E., and Kaufmann, E.G., eds., Evolution of the Western sequence: Comparison with low energy sequence: American Association of Petroleum Ge-
Interior Basin: Geological Association of Canada, Special Paper 39, p. 177–188. ologists, Bulletin, v. 65, p. 807–830.
DAVIS, J.M., LOHMANN, R.C., PHILLIPS, F.M., WILSON, J.L., AND LOVE, D.W., 1993, Architecture HUNTER, R.E., CLIFTON, H.E., AND PHILLIPS, R.L., 1979, Depositional processes, sedimentary
of the Sierra Ladrones Formation, central New Mexico: Depositional controls on the per- structures, and predicted vertical sequences in barred nearshore systems, southern Oregon
meability correlation structure: Geological Society of America, Bulletin, v. 105, p. 998– coast: Journal of Sedimentary Petrology, v. 49, p. 711–726.
1007. HURST, A., AND ROSVOLL, K.J., 1991, Permeability variations in sandstones and their relationship
DINWIDDIE, C.L., 2001, A new small drillhole minipermeameter probe for in-situ measurement: to sedimentary structures, in Lake, L.W., Carroll, H.B., Jr., and Wesson, T.C., eds., Res-
Design, theoretical analysis, operation, and performance characteristics [unpublished Ph.D. ervoir Characterization II: San Diego, Academic Press, p. 166–196.
dissertation]: Clemson University, 156 p. JOHNSON, H.D., AND BALDWIN, C.T., 1996, Shallow clastic seas, in Reading, H.G., ed., Sedi-
DINWIDDIE, C.L., MOLZ, F.J., LU, S., CASTLE, J.W., AND MURDOCH L.C., 2000, The new drillhole mentary Environments; Processes, Facies and Stratigraphy, Third Edition: Oxford, U.K.,
permeameter probe: Design, theoretical analysis, operation and performance characteristics Blackwell Science, p. 232–280.
(abstract): Geological Society of America, Abstracts with Programs, v. 32, no. 7, p. A-515. JORDAN, T.E., 1981, Thrust loads and foreland basin evolution, Cretaceous, western United
DINWIDDIE, C.L., MOLZ, F.J., AND CASTLE, J.W., 2003, A new small drillhole minipermeameter States: American Association of Petroleum Geologists, Bulletin, v. 65, p. 2506–2520.
probe for in-situ permeability measurement: fluid mechanics and geometrical factors: Water JORDAN, T.E., 1995, Retroarc foreland and related basins, in Busby, C.J., and Ingersoll, R.V.,
Resources Research, v. 39, no. 7, 1178, doi:10.1029/2001WR001179. Tectonics of Sedimentary Basins: Cambridge, U.K., Blackwell Science, p. 331–362.
DOTT, R.H., JR., AND BOURGEOIS, J., 1982, Hummocky stratification: significance of its variable KERANS, C., LUCIA, F.J., AND SENGER, R.K., 1994, Integrated characterization of carbonate ramp
bedding sequences: Geological Society of America, Bulletin, v. 93, p. 663–680. reservoirs using Permian San Andreas Formation analogs: American Association of Petro-
DUKE, W.L., 1985, Hummocky cross-stratification, tropical hurricanes, and intense winter leum Geologists, Bulletin, v. 78, p. 181–216.
storms: Sedimentology, v. 32, p. 167–194. KRASSAY, A.A., 1994, Storm features of siliciclastic shelf sedimentation in the mid-Cretaceous
DUTTON, S.P., AND WILLIS, B.J., 1998, Comparison of outcrop and subsurface sandstone per- epeiric seaway of northern Australia: Sedimentary Geology, v. 89, p. 241–264.
meability distribution, Lower Cretaceous Fall River Formation, South Dakota and Wyoming: LEITHOLD, E.L., 1989, Depositional processes on an ancient and modern muddy shelf, northern
Journal of Sedimentary Research, v. 68, p. 890–900. California: Sedimentology, v. 36, p. 179–202.
EATON, J.G., 1991, Biostratigraphic framework for the Upper Cretaceous rocks of the Kaipa- LIU, H.H., AND MOLZ, F.J., 1997, Comment on ‘‘Evidence for non-Gaussian scaling behavior
rowits Plateau, southern Utah, in Nations, J.D., and Eaton, J.G., eds., Stratigraphy, Depo- in heterogeneous sedimentary formations’’ by Scott Painter: Water Resources Research, v.
sitional Environments, and Sedimentary Tectonics of the Western Interior Seaway: Geolog- 33, p. 907–908.
ical Society of America, Special Paper 260, p. 47–63. LIU, K., BOULT, P., PAINTER, S., AND PATERSON, L., 1996, Outcrop analog for sandy braided
FELLER, W., 1971, An Introduction to Probability Theory and Its Application: New York, John stream reservoirs: permeability patterns in the Triassic Hawksbury Sandstone, Sydney Basin,
Wiley, 670 p. Australia: American Association of Petroleum Geologists, Bulletin, v. 80, p. 1850–1866.
FISHER, R.S., BARTON, M.D., AND TYLER, N., 1993, Quantifying reservoir heterogeneity through LONGIARU, S., 1987, Visual comparators for estimating the degree of sorting from plane and
outcrop characterization: 1. Architecture, lithology, and permeability distribution of a land- thin section: Journal of Sedimentary Petrology, v. 57, p. 791–794.
ward-stepping fluvial-deltaic sequence, Ferron Sandstone (Cretaceous), central Utah: Uni- LU, S., AND MOLZ, F.J., 2001, How well are hydraulic conductivity variations approximated by
versity of Texas, Bureau of Economic Geology, topical report no. GRI-93–0022, 132 p. additive stable processes?: Advances in Environmental Research, v. 5, p. 39–45.
FREY, R.W., 1990, Trace fossils and hummocky cross-stratification, Upper Cretaceous of Utah: LU, S., MOLZ, F.J., FOGG, G.E., AND CASTLE, J.W., 2002, Combining stochastic facies and fractal
Palaios, v. 5, p. 203–218. models for representing natural heterogeneity: Hydrogeology Journal, v. 10, p. 475–482.
FREY, R.W., AND HOWARD, J.D., 1990, Trace fossils and depositional sequences in a clastic LU, S., MOLZ, F.J., AND LIU, H.H., 2003, An efficient, three-dimensional, anisotropic, fractional
shelf setting, Upper Cretaceous of Utah: Journal of Paleontology, v. 64, p. 803–820. Brownian motion and truncated fractional Levy motion simulation algorithm based on suc-
FREY, R.W., AND PEMBERTON, S.G., 1985, Biogenic structures in outcrops and cores. 1. Ap- cessive random additions: Computers and Geosciences, v. 29, p. 15–25.
proaches to ichnology: Bulletin of Canadian Petroleum Geology, v. 32, p. 72–115. MANDELBROT, B.B., AND WALLIS, J.R., 1969, Robustness of the rescaled range R/S in the mea-
FREY, R.W., AND PEMBERTON, S.G., 1987, The Psilonichnus ichnocoenose, and its relationship surement of noncyclic long run statistical dependence: Water Resources Research, v. 5, p.
to adjacent marine and nonmarine ichnocoenoses along the Georgia coast: Bulletin of Ca- 967–988.
nadian Petroleum Geology, v. 35, p. 333–357. MOLZ, F.J., AND BOMAN, G.K., 1993, A stochastic interpolation scheme in subsurface hydrology:
FREY, R.W., CURRAN, H.A., AND PEMBERTON, S.G., 1984, Tracemaking activities of crabs and Water Resources Research, v. 29, p. 3769–3774.
their environmental significance: The ichnogenus Psilonichnus, Journal of Paleontology, v. MOLZ, F.J., DINWIDDIE, C.L., AND WILSON, J.L., 2003, A physical basis for calculating instrument
58, p. 333–350. spatial weighting functions in homogeneous systems: Water Resources Research, v. 39, no.
GELHAR, L.W., 1993, Stochastic Subsurface Hydrology: Englewood Cliffs, Prentice-Hall, 390 p. 4, 1096, doi:10.1029/2001WR001220.
GHIBAUDO, G., MUTTI, E., AND ROSELL, J., 1974, Le spiagge fossili delle Arenarie di Aren MOLZ, F.J., LIU, H.H., AND SZULGA, J., 1997, Fractional Brownian motion and fractional Gauss-
(Cretacico superiore) nella valle Noguera Ribagorzana (Pirenei centro-meridionali, Province ian noise in subsurface hydrology: A review, presentation of fundamental properties, and
di Lerida e Huesca, Spagna): Società Geologica Italiana, Memorie, v. 13, p. 497–537. extensions: Water Resources Research, v. 33, p. 2273–2286.
GOGGIN, D.J., 1988, Geologically-sensible modeling of the spatial distribution of permeability MOLZ, F.J., MURDOCH, L.C., DINWIDDIE, C.L., AND CASTLE, J.W., 2002, Small drill-hole, gas
in eolian sandstone deposits: Page Sandstone (Jurassic), northern Arizona [unpublished mini-permeameter probe: U.S. Statutory Invention Registration H2052 H, U.S. Patent and
Ph.D. dissertation]: University of Texas at Austin, 417 p. Trademark Office, Washington, D.C.
GOGGIN, D.J., CHANDLER, M.A., KOCUREK, G., AND LAKE, L.W., 1992, Permeability transects of NEUMAN, S.P., 1990, Universal scaling of hydraulic conductivities and dispersivities in geologic
eolian sands and their use in generating random permeability fields: Society of Petroleum media: Water Resources Research, v. 26, p. 1749–1758.
Engineers, Formation Evaluation, v. 7, p. 7–16. PAINTER, S., 1995, Random fractal models of heterogeneity: The Levy-stable approach: Math-
GREENWOOD, B., AND MITTLER, P.R., 1985, Vertical sequence and lateral transitions in the facies ematical Geology, v. 27, p. 813–830.
of a barred nearshore environment: Journal of Sedimentary Petrology, v. 55, p. 366–375. PAINTER, S., 2001, Flexible scaling model for use in random field simulation of hydraulic
HARMS, J.C., SOUTHARD, J.B., SPEARING, D.R., AND WALKER, R.G., 1975, Depositional environ- conductivity: Water Resources Research, v. 37, p. 1155–1163.
ments as interpreted from primary sedimentary structures and stratification sequences: So- PATTISON, S.A.J., AND WALKER, R.G., 1992, Deposition and interpretation of long, narrow sand-
ciety of Economic Paleontologists and Mineralogists, Short Course Notes, no. 2, 161 p. bodies underlain by a basinwide erosion surface: Cardium Formation, Cretaceous western
HERON, S.D., JR., MOSLOW, T.F., BERELSON, W.M., HERBERT, J.R., STEELE, G.A., III, AND SUSMAN, interior seaway, Alberta, Canada: Journal of Sedimentary Petrology, v. 62, p. 292–309.
K.R., 1984, Holocene sedimentation of a wave-dominated barrier-island shoreline: Cape PEMBERTON, S.G., AND MACEACHERN, J.A., 1995, The sequence stratigraphic significance of trace
Lookout, North Carolina: Marine Geology, v. 60, p. 413–434. fossils: Examples from the Cretaceous foreland basin of Alberta, Canada, in Van Wagoner,
HETTINGER, R.D., 1995, Sedimentological descriptions and depositional interpretations, in se- J.C., and Bertram, G.T., eds., Sequence Stratigraphy of Foreland Basin Deposits: Outcrop
quence stratigraphic context, of two 300- meter cores from the Upper Cretaceous Straight and Subsurface Examples from the Cretaceous of North America: American Association of
Cliffs Formation, Kaiparowits Plateau, Kane County, Utah: U.S. Geological Survey, Bulletin Petroleum Geologists, Memoir 64, p. 429–475.
2115-A, 32 p. PEMBERTON, S.G., MACEACHERN, J.A., AND FREY, R.W., 1992, Trace fossil facies models: En-
HETTINGER, R.D., MCCABE, P.J., AND SHANLEY, K.W., 1993, Detailed facies anatomy of trans- vironmental and allostratigraphic significance, in Walker, R.G., and James, N.P., eds., Facies
gressive and highstand systems tracts from the Upper Cretaceous of southern Utah, U.S.A., Models: Response to Sea Level Change: Geological Association of Canada, p. 47–72.
in Weimer, P. and Posamentier, H.W., eds., Siliciclastic Sequence Stratigraphy: Recent PEMBERTON, S.G., SPILA, M., PULHAM, A.J., SAUNDERS, T., MACEACHERN, J.A., ROBBINS, D., AND
Developments and Applications: American Association of Petroleum Geologists, Memoir SINCLAIR, I.K., 2001, Ichnology and sedimentology of shallow to marginal marine systems:
58, p. 235–257. Ben Nevis and Avalon reservoirs, Jeanne d’Arc basin: Geological Association of Canada,
HETTINGER, R.D., ROBERTS, L.N.R., BIEWICK, L.R.H., AND KIRSCHBAUM, M.A., 1996, Preliminary Short Course Volume 15, 343 p.
investigations of the distribution and resources of coal in the Kaiparowits Plateau, southern PETERSON, F., 1969, Four new members of the Upper Cretaceous Straight Cliffs Formation in
Utah: U.S. Geological Survey, Open File Report 96–539, 72 p. the southeastern Kaiparowits region, Kane County, Utah: U.S. Geological Survey, Bulletin
HEWITT, T.A., 1986, Fractal distributions of reservoir heterogeneity and their influence on fluid 1274-J, 28 p.
transport: Society of Petroleum Engineers, SPE Paper No. 15386, 15 p. READING, H.G., AND COLLINSON, J.D., 1996, Clastic coasts, in Reading, H.G., ed., Sedimentary
HILL, G.W., AND HUNTER, R.E., 1976, Interaction of biological and geological processes in the Environments: Processes, Facies and Stratigraphy, Third Edition: Oxford, U.K., Blackwell
beach and nearshore environments, northern Padre Island, Texas, in Davis, R.A., Jr., and Science, p. 154–231.
Ethington, R.L., eds., Beach and Nearshore Sedimentation: Society of Economic Paleontol- REINECK, H.E., AND SINGH, I.B., 1972, Genesis of laminated sand and graded rhythmites in
ogists and Mineralogists, Special Publication 24, p. 169–187. storm-sand layers of shelf mud: Sedimentology, v. 18, p. 123–128.
HOWARD, J.D., AND REINECK, H.E., 1981, Depositional facies of high energy beach to offshore SCHMIDT, V., AND MCDONALD, D.A., 1979, Texture and criteria of secondary porosity in sand-
284 J.W. CASTLE ET AL.

stones, in Scholle, P.A., and Schluger, P.R., eds., Aspects of Diagenesis: Society of Eco- SWIFT, D.J.P., PHILLIPS, S., AND THORNE, J.A., 1991, Sedimentation on continental margins, IV:
nomic Paleontologists and Mineralogists, Special Publication 26, p. 209–225. Lithofacies and depositional systems, in Swift, D.J.P., Tillman, R.W., and Thorne, J.A.,
SHANLEY, K.W., AND MCCABE, P.J., 1991, Predicting facies architecture through sequence stra- eds., Shelf Sand and Sandstone Bodies: Geometry, Facies, and Sequence Stratigraphy: In-
tigraphy—An example from the Kaiparowits Plateau, Utah: Geology, v. 19, p. 742–745. ternational Association of Sedimentologists, Special Publication 14, p. 89–152.
SHANLEY, K.W., AND MCCABE, P.J., 1993, Alluvial architecture in a sequence-stratigraphic SWIFT, D.J.P., PARSONS, B.S., FOYLE, A., AND OERTEL, G.F., 2003, Between beds and sequences:
framework: A case history from the Upper Cretaceous of southern Utah, USA, in Flint, stratigraphic organization at intermediate scales in the Quaternary of the Virginia coast,
S.S., and Bryant, I.D., eds., The Geological Modelling of Hydrocarbon Reservoirs and Out- U.S.A.: Sedimentology, v. 50, p. 81–111.
crop Analogues: International Association of Sedimentologists, Special Publication 15, p. TYLER, N., BARTON, M.D., AND FINLEY, R.J., 1991, Outcrop characterization of flow unit and
21–56. seal properties and geometries, Ferron Sandstone, Utah: Society of Petroleum Engineers,
SHANLEY, K.W., MCCABE, P.J., AND HETTINGER, R.D., 1992, Tidal influence in Cretaceous fluvial SPE Paper No. 22670, p. 127–134.
strata from Utah, USA: A key to sequence stratigraphic interpretation: Sedimentology, v. WALKER, R.G., 1984, Shelf and shallow marine sands, in Walker, R.G., ed., Facies Models,
39, p. 905–930. Second Edition: Geological Association of Canada, Geoscience Canada, Reprint Series 1,
SHIPP, R.C., 1984, Bedforms and depositional sedimentary structures of a barred nearshore p. 141–170.
system, eastern Long Island, New York: Marine Geology, v. 60, p. 235–259. WALKER, R.G., AND PLINT, A.G., 1992, Wave- and storm-dominated shallow marine systems,
SHORT, A.D., 1984, Beach and nearshore facies: southeast Australia: Marine Geology, v. 60, in Walker, R.G., and James, N.P., eds., Facies Models: Response to Sea Level Change:
p. 261–282. Geological Association of Canada, p. 219–238.
STALKUP, F.I., AND EBANKS, W.J., JR., 1986, Permeability variation in a sandstone barrier island– WEBER, K.J., 1982, Influence of common sedimentary structures on fluid flow in reservoir
tidal delta complex, Ferron Sandstone (Lower Cretaceous), southern Utah: Society of Petro- models: Journal of Petroleum Technology, v. 34, p. 665–672.
leum Engineers, SPE Paper No. 15532, 8 p. WEBER, K.J., AND VAN GEUNS, L.C., 1990, Framework for constructing clastic reservoir simu-
STEPHEN, K.D., AND DALRYMPLE, M., 2002, Reservoir simulations developed from an outcrop lation models: Journal of Petroleum Technology, v. 42, p. 1248–1253, 1296–1297.
WHITE, C.D., AND BARTON, M.D., 1999, Translating outcrop data to flow models, with appli-
of incised valley fill strata: American Association of Petroleum Geologists, Bulletin, v. 86,
cations to the Ferron Sandstone: Society of Petroleum Engineers, Reservoir Evaluation &
p. 797–822. Engineering, v. 2, p. 341–350.
SWIFT, D.J.P., FIGUEIREDO, A.G., JR., FREELAND, G.L., AND OERTEL, G.F., 1983, Hummocky cross- WILLIS, B.J., 1998, Permeability structure of a compound valley fill in the Cretaceous Fall
stratification and megaripples: a geological double standard?: Journal of Sedimentary Pe- River Formation of South Dakota: American Association of Petroleum Geologists, Bulletin,
trology, v. 53, p. 1295–1317. v. 82, p. 206–227.
SWIFT, D.J.P., HUDELSON, P.M., BRENNER, R.L., AND THOMPSON, P., 1987, Shelf construction in WILLIS, B.J., AND WHITE, C.D., 2000, Quantitative outcrop data for flow simulation: Journal of
a foreland basin: storm beds, shelf sandbodies, and shelf-slope depositional sequences in Sedimentary Research, v. 70, p. 788–802.
the Upper Cretaceous Mesaverde Group, Book Cliffs, Utah: Sedimentology, v. 34, p. 423–
457. Received 24 February 2003; accepted 28 August 2003.

You might also like