Sunteți pe pagina 1din 16

Journal of Sedimentary Research, 2007, v.

77, 324339
Research Article
DOI: 10.2110/jsr.2007.029

HIGH-RESOLUTION FACIES ANALYSES OF MUDSTONES: IMPLICATIONS FOR PALEOENVIRONMENTAL


AND SEQUENCE STRATIGRAPHIC INTERPRETATIONS OF OFFSHORE ANCIENT
MUD-DOMINATED SUCCESSIONS
JOE H. S. MACQUAKER,1 KEVIN G. TAYLOR,2 AND ROB L. GAWTHORPE1
1

School of Earth, Atmospheric and Environmental Sciences, The University of Manchester, Manchester M13 9PL, U.K.
2
Department of Environmental and Geographical Sciences, Manchester Metropolitan University, Manchester, M1 5GD, U.K.

ABSTRACT: The extent of lithofacies variability in fine-grained sedimentary rocks is very poorly known in comparison to that
present in coarser grained clastic and carbonate successions. The absence of this information means that sediments present on
continental shelves are rarely considered as integrated systems because fine-grained facies (such as shales) are mostly excluded
from sophisticated, regional facies models. To shed light upon lithofacies and grain size in shale-dominated successions, so that
they can be incorporated into shelf-wide depositional models, the Blackhawk Formation of the Mancos Shale (Campanian age),
exposed in the Book Cliffs, Utah was investigated using combined field, whole-rock geochemical, optical, and electron optical
methods. These sediments show systematic grain-size variations, are intensely bioturbated, and composed predominantly of
detrital clays (mainly dioctahedral micas), quartz, and feldspar with minor pyrite and organic matter. They are organized into
very thin (, 10 mm), upward-fining, genetic beds; they exhibit both systematic lateral (103 m scale), and vertical (1022 to
100 m scales) lithofacies variability. Preferentially cemented units occur close to sequence boundaries (unconformities). These
cemented units typically contain a very different detrital assemblage (including significant quantities of chlorite) from the rest
of the succession and are located close to levels where there are marked stacking-pattern discontinuities.
Overall, lithofacies variability is interpreted in terms of deposition occurring on an oxic continental shelf, with the coarsergrained facies being deposited in proximal settings (offshore-transition zone), in contrast to the finer-grained units that were
deposited in more distal environments (offshore zone). Storms are interpreted to have been the dominant mechanism dispersing
the sediment. Once deposited, the surface layers of the sediment were intensely reworked by burrowing organisms. Preferential
preservation of ichnogenera from mid- and lower tiers suggest that the depositional environment was energetic and erosion
commonly removed the surface sediment layers, prior to deposition of thin (1022 m) storm beds. The larger scale (100 m)
upward-coarsening units that are composed of stacked beds, are interpreted to be parasequences.
Preferential cementation, at levels where large-scale stacking patterns change, is interpreted to occur at horizons where there
were breaks in sediment accumulation, and bacterial metabolic processes were able to supply sufficient solutes to fill
uncompacted pore space with cement. The presence of a very different detrital assemblage in these units suggests that
winnowing occurred. Despite their relatively innocuous appearance in the field, these surfaces are likely to be significant bypass
surfaces, over which sediment was delivered to the deeper parts of the basin during times of relative lowstand of sea level.

INTRODUCTION AND AIMS

Architectural elements present within fine-grained siliciclastic successions (shales), and controls on their lithofacies variability are poorly
understood, in comparison to our understanding of these processes in
coarse clastic and carbonate successions (e.g., Bohacs 1998; Potter et al.
2005). This lack of understanding exists because most investigations of
this variability have concentrated on interpreting data obtained from
a relatively narrow range of techniques, employing either whole-rock
geochemical (e.g., elemental analyses coupled with mineralogical determinations), component-specific geochemical (e.g., organic carbon content, stable isotopes in carbonates), or paleontological methods. Despite
being apparently fit-for-purpose, and their wide application by
geologists investigating facies variability in fine-grained sedimentary
successions, these datasets are actually less than ideal because key
information such as grain-size, origin of different components, and

Copyright E 2007, SEPM (Society for Sedimentary Geology)

sedimentary textures are typically not directly obtained using these


techniques. The absence of these data inevitably mean that researchers are
forced to make complex and subtle extrapolations, using sometimes
contradictory proxies to interpret many of the fundamental controls on
mudstone variability.
Fine-grained sediments, like the majority of sedimentary rocks,
comprise three distinctive components in varying proportions (e.g.,
Macquaker and Adams 2003). These components are derived from
sedimentary detritus entering the basin (allochthonous or detrital-derived
components), organisms living within the water column as well as in the
surface sediment layers (autochthonous or productivity-derived components), and chemical (diagenesis-derived) components which precipitate
either in the water column, at the sedimentwater interface, or once the
sediment has been buried. Once the various components had been
produced, they were then available to be transported by mechanisms such
as storms, tides, and ice before being deposited in areas where

1527-1404/07/077-324/$03.00

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

325

FIG. 1.Map showing location of the Book


Cliffs in Eastern Utah, and locations of sections of
Mancos Shale that were logged and sampled.

accommodation was available. After deposition, the sediment was then


either resuspended or colonized by burrowing organisms, prior to being
buried and experiencing a variety of diagenetic transformations. Here,
early diagenetic processes were typically mediated by microbial metabolic
processes prior to compaction, and by physicochemical processes deeper
after compaction. To obtain a comprehensive understanding of how these
different processes interacted to produce an individual rock, it is
necessary to have information about bed and laminae thicknesses,
descriptions of any depositional and diagenetic textures present, information about bedding contacts, data on grain-size variability, and
information about what processes generated the constituent grains. Once

FIG. 2. Stratigraphic setting of the studied succession of the Book Cliffs,


Mancos Shale and Blackhawk Formation.

these data have been obtained for an individual sample, it is then possible
to determine the controls on larger-scale temporal and spatial facies
variability (architecture) by comparing genetically related samples within
the context of a well-constrained stratigraphic framework.
Obtaining the crucial data to determine architectural and facies
variability in coarse-grained sedimentary successions is relatively
straightforward (e.g., Hampson et al. 1999; Adams and Bhattacharya
2005; Edwards et al. 2005). On their own, however, field-, geochemical-,
and paleontological-based descriptions of fine-grained successions are
rarely sufficient to perform a critical investigation of facies and
architectural variability in fine-grained strata. In this setting difficulties
arise because fine-grained sediments exhibit relatively little obvious handspecimen-scale variability and are highly susceptible to weathering,
making it hard to observe primary fabrics in these rocks (e.g., Potter et al.
1980). Moreover, relying on descriptions of individual units based upon
features that are readily obtainable in the field (such as color, bedding
thicknesses, and weathering style) is commonly fraught with uncertainly.
Uncertainties arise because these descriptors do not always vary
systematically with key sedimentary parameters, such as grain-size,
proportion of total organic carbon, proportion of calcareous phytoplankton relative to diagenetic carbonate, and proportion of authigenic as
opposed to detrital clay (e.g., Macquaker and Gawthorpe 1993). To
compound these problems further, most field-based descriptions of
mudstones are vague about microtextures present, in as much as they are
beyond the resolution of even the most diligent geologists with hand
lenses!
Recent research on fine-grained sediments, combining field, optical,
electron optical, and geochemical methods, utilizing systematic sampling
techniques and subsequent manufacture of unusually thin (20 mm)
polished thin sections of the rocks, has revealed enormous amounts of
previously unavailable information (e.g., Potter et al. 1980; OBrien and

326

J.H.S. MACQUAKER ET AL.

JSR

FIG. 3. Field photographs showing outcrop


expression of the studied Mancos Shale timeequivalent to the Grassy Member of the Blackhawk Formation, location of stratal surfaces
(mapped in the field) and location of the sampled
sections. A) Thompson Pass. B) Blaze Canyon.
C) Coquina Wash. (SB/ts 5 sequence boundary/transgressive surface). (This figure is reproduced in color in the digital version.)

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

327

FIG. 4.Details of the three logged and sampled Mancos Shale successions, including sample numbers, mudstone lithofacies codes, sand + silt:clay ratios, and total
organic carbon (TOC) data. For full data information, see Table 1.

Slatt 1990; Macquaker and Gawthorpe 1993; Macquaker 1994; Macquaker and Taylor 1996; Macquaker et al. 1998; Macquaker and Howell
1999; Williams et al. 2001; Rohl et al. 2001; Schieber 2003; Macquaker
and Keller 2005; Rohl and Schmid-Rohl 2005). With these methods
geologists have been able to make direct observations of grain-size,
determine the origin of individual components, and make microtextural
observations of the rock without resorting to making unreliable
extrapolations from proxy data (e.g., whole-rock geochemistry). Together
these direct observations, in conjunction with information obtained from
conventional paleontological and geochemical techniques, have allowed
researchers to significantly extend their understanding of architectural
elements present, and controls on lithofacies variability within siliciclastic
mudstone successions. Such studies are important because . 65% (e.g.,
Blatt 1970; Aplin et al. 1999) of the rock record at the Earths surface is
composed of fine-grained sediments. Moreover, these rocks are major
sites of natural carbon sequestration, and their investigation provides
insight into processes occurring elsewhere on continental shelves. Their
study, for instance, enable predictions of facies variability downdip, on
the continental slope, and updip, in the middle and upper shoreface, to be
made.
The main aims of this study are to enhance our understanding of the
scale of architectural elements present, and to investigate the fundamental
controls that underpin temporal and spatial controls on lithofacies
variability in mud-dominated portions of the sediment transport path on
marine continental shelves. To meet this aim we have investigated
lithofacies variability utilizing field relations, optical, electron-optical,
and geochemical techniques in the Upper Cretaceous Mancos Shale
(Grassy Member, Blackhawk Formation) (Fig. 1), exposed in the Book
Cliffs, Utah (Fig. 2) and interpreted these observations within a sequencestratigraphic framework. The Mancos Shale was specifically chosen for
this research because its updip stratigraphy is well constrained and there

is evidence that dynamic bypass in this succession led to the formation of


downdip lowstand shoreface deposits (e.g., Hampson et al. 1999).
Moreover, the excellent exposure means that it is possible to identify,
trace, and sample the same stratigraphic interval at different locations
along the sediment transport path within the mud-dominated parts of this
succession.
GEOLOGICAL AND STRATIGRAPHIC SETTING OF THE MANCOS SHALE

Tectonism during the early Cretaceous in western North America


resulted in the development of a foreland basin and formation of the
Western Interior Seaway (e.g., Burchfiel et al. 1992). By Maastrichtian
times, this epeiric sea linked the polar ocean and the subtropical Gulf of
Mexico (Fig. 1). The Upper Cretaceous succession currently exposed in
the Book Cliffs was deposited along the western margin of this Seaway as
a wedge of eastward-prograding siliciclastic sediment, derived from the
unroofing of the Sevier Fold and Thrust Belt (Fouch et al. 1983) to the
west. Excellent exposures in the Book Cliffs allow detailed studies of
large-scale geometry and stratal architecture, and many stratigraphic,
sedimentological, and diagenetic studies have been published on the
fluvial and shallow marine strata (e.g., Van Wagoner 1995; OByrne and
Flint 1995; Kamola and Huntoon 1995; Hampson et al. 1999; Yoshida
2000; Taylor et al. 2000; Taylor et al. 2002; Taylor et al. 2004; Miall and
Arush 2001; Taylor and Gawthorpe 2003; Pattison 2005).
In this paper we study the downdip strata that are time-equivalent to
the Grassy Member of the Blackhawk Formation (Fig. 2). The Blackhawk Formation is composed of tongues of coastal-plain, fluvial, and
shoreface strata, which interfinger eastwards into the mudstonedominated Mancos Shale (Fig. 2). Within individual members, sequence
boundaries, marine flooding surfaces, and transgressive surfaces can be
traced for tens of kilometers where the sediments are exposed (Kamola

328

J.H.S. MACQUAKER ET AL.

JSR

FIG. 5.A) Thin-section scan of a representative clay-rich mudstone (Coquina 06). This clay-rich mudstone is thin, relict-bedded, and partially bioturbated (top of bed
marked by dashed line). The thin relict beds overall fine upward, from being relatively silt enriched at their bases to silt depleted at their tops (annotated c). The
diminutive ichnogenera present include Phycosiphon isp. (arrowed), Planolites isp., as well as an indeterminate fauna. B) Low-power optical micrograph of Blaze 6 (clayrich mudstone) illustrating detrital silt grains in a matrix dominated by clay, amorphous organic matter, and pyrite. Note the presence of an agglutinated foraminifer and
the presence of dark, clay-filled burrows towards the bottom right of the micrograph (cfb). These diminutive burrows are attributed to Phycosiphon isp. C) Backscattered
electron micrograph of Coquina 6 (clay-rich mudstone) illustrating prominent agglutinated foraminifer in a matrix of clay with some fine-grained silt (composed mainly
of quartz [low g] and detrital feldspar [intermediate g]). The foraminfer test is composed of quartz grains. The chambers within the foraminifer are infilled with framboidal

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

and Huntoon 1995). The Mancos Shale is composed of very fine-grained


sandstone and mudstones deposited in a shallow, well-oxygenated open
marine shelf (Howell and Flint 2003a; 2003b). OByrne and Flint (1995)
published a high-resolution sequence stratigraphic framework for the
Grassy Member, in which they recognized progradational tongues of
coastal-plain and shoreface strata with two sequence boundaries, marked
by fluvial incision, in the upper part of the member. They did not,
however, extend this framework into the time-equivalent Mancos Shale.
Hampson et al. (1999), Pattison (2005), and Chan et al. (1991) variously
recognized the presence of tidally influenced river channels, fluvialdominated delta fronts, and storm-influenced shoreface deposits within
the Mancos Shale. They correlated these units variously to the Aberdeen
and Kenilworth members of the Blackhawk Formation. Equivalent
deposits have not been identified in the Mancos Shale downdip of the
Grassy Member.
METHODS

In order to meet the aims of this research, a part of the Mancos Shale
that is time-equivalent to the Grassy Member of the Blackhawk
Formation exposed in the Book Cliffs, Utah (Figs. 1, 2) was studied.
The rocks here are well exposed and allow direct correlation of major
stratal surfaces preserved within coastal plain and shoreface successions
updip, with mudstone-dominated strata downdip. The availability of
a detailed stratigraphic framework for the Grassy Member (OByrne and
Flint 1995), allowed us to place the time-equivalent Mancos Shale into
a wider stratigraphic context (Fig 2). Lateral correlations between these
units were achieved through a combination of physical correlation along
the exposure, binocular observation, and photomontage examination of
cliff faces. Detailed sedimentary logs were measured and samples of the
Mancos Shale collected (Figs. 2, 3) along a 20-km-long transect, oriented
at a slightly oblique angle to the main paleosediment transport direction
(oriented SE to ESE). Samples were collected from Thompson Pass
(proximal location), Blaze Canyon (intermediate location path), and
Coquina Wash (distal location). Overall approximately 90 samples were
obtained at vertical intervals of approximately 0.5 m from each of these
localities (Fig. 4).
Unusually thin (20 mm), polished thin sections were prepared from
each sample, in order to facilitate making detailed lithofacies descriptions.
Initially, each thin section was scanned using a flat-bed scanner (Epson
1250) equipped with a 35 mm slide illumination system to obtain details
of 1022 to 1023 m-scale textures present. The sections were then examined
optically at low power (Nikon Labophot Pol) to obtain details of the 1023
to 1024 m-scale textures, and then at higher power using backscattered
electron imagery (JEOL 6400 equipped with a Link 4 Quadrant solid state
backscattered electron detector and Semifore digital framestore) to obtain
details of the 1024 to 1025 m-scale textures. Where mineral identity was
not immediately obvious on the basis of varying optical characteristics
and backscatter coefficients (g), identity was confirmed utilizing semiquantitative energy-dispersive spectrometry (PGT with windowless
detector). The scanning electron microscope was operated at 20 kV and
2.0 nA, at a working distance of 15 mm. Using this combined optical and
electron optical dataset, the grain-size of the silt could be estimated and
the abundance of each component semi-quantified by comparison with

329

published grain abundance charts (e.g., Flugel 1982). Crucially, these


methods allowed us to broadly estimate the proportion of silt and fine
sand present, as well as the proportion of accessory minerals and clay by
difference.
Total organic-carbon (TOC) analyses were performed by the University of Newcastle (U.K.). The total C contents of untreated, powdered
samples were obtained using an induction furnace (Leco C/S analyzer).
Once the total C data had been obtained, a split of each sample was
treated with 6N HCl to remove all carbonate prior to determining the
quantity of residual carbon in the acid-treated samples using the same
furnace. Organic-carbon contents were then determined by difference.
The organic-carbon contents have a precision of 6 0.2%.
RESULTS

Systematic analyses of the thin sections revealed mainly the following


fine-grained lithofacies (nomenclature after Macquaker and Adams
2003):
a)
b)
c)
d)
e)
f)
g)
h)

clay-rich mudstones (Fig. 5A, B, C);


silt-bearing, clay-rich mudstones (Fig. 5D, E, F);
very fine sand- and silt-bearing, clay-rich mudstones (Fig. 6A, B, C);
very fine sand-, silt- and clay-bearing mudstones (Fig. 6D, E, F);
fine-grained muddy sandstones (Fig. 7A, B, C);
carbonate-cement-rich mudstones (Fig. 7D, E, F);
silt- and clay-bearing, carbonate-cement-rich mudstones;
fine sand- and silt-bearing, carbonate cement-rich mudstones.

Descriptions of each sample are summarized in Table 1. Individual


units are either thin-bedded, or relict-bedded (, 10 mm), or intensely
bioturbated (e.g., Figs. 5A, D, 6D, 7A). Where identifiable, individual
units have erosional bases (Fig. 6A), fine upward, and have bioturbated
tops (e.g., Figs. 6A, 7D). Typically the bases of these upward-fining units
are composed of muddy sandstones and very fine sand-, silt-, and claybearing mudstones, in contrast to their tops, which are composed of siltbearing, clay-rich mudstones and clay-rich mudstones (Figs. 6D, 7D). In
addition, dewatering has disrupted some of the original depositional
textures (Fig. 6A).
The diminutive ichnofauna present comprise an assemblage of
Phycosiphon isp., Planolites isp., and Terebellina isp., (e.g., Figs. 5A, D,
6A) in the fine-grained units and Palaeophycus isp., Rhizocorallium isp.,
Teichinus isp., Chondrites isp., and indistinct Thallisanoides isp. (Figs. 6D,
7A) in the coarser-grained units. Escape traces, which completely disrupt
all the laminae within beds, are present in some units (e.g., Fig. 6A).
The very fine-grained sand and silt fractions are composed predominantly of quartz with smaller proportions of detrital feldspar (albite and
plagioclase) and dolomite (Figs. 5C, F, 6C, F, 7C, F). In most samples the
fine-grained matrix is a mixture of illite, mixed layer illitesmectite, and
amorphous organic matter, with minor framboidal pyrite (grain-size range
10 to 30 mm), and macerated woody material (Figs. 5E, F, 6C, E).
Total organic-carbon (TOC) values average 1.3% (range: 0.2 to 2.0%),
with the silt-bearing clay-rich mudstones and clay-rich mudstones being
the most enriched (average 1.6%) compared to the coarser-grained units
and cement-rich facies (average of 1.2% and 0.7%, respectively) (Table 1).

r
pyrite (high g). D) Thin-section scan of Coquina 25, a representative silt-bearing clay-rich mudstone. This unit is intensely bioturbated by a combination of Phycosiphon
isp. (arrowed), Palaeophycus isp. (Pa), and Planolites isp. Most of the original lamina and some of the early Phycosiphon isp. burrows have been completely disrupted by
later burrowing activity. E) Low-power optical micrograph of Blaze 5 (silt-bearing clay-rich mudstone) illustrating silt grains and an agglutinated foraminifer in a matrix
dominated by clay, amorphous organic matter and pyrite. Note the presence of clay-filled burrows and macerated woody material (arrowed w) in the matrix. F)
Backscattered electron micrograph of Coquina 22 (silt-bearing clay-rich mudstone) illustrating prominent agglutinated foraminifer in a matrix of clay, organic matter,
and fine-grained silt (composed of quartz and detrital feldspar). The chambers within the foraminifer are infilled with framboidal pyrite (very high g). Note the presence of
macerated woody material in the matrix (very low g, arrowed w). (This figure is reproduced in color in the digital version.)

330

J.H.S. MACQUAKER ET AL.

JSR

FIG. 6. A) Thin-section scan of a representative sand- and silt-bearing, clay-rich mudstone (Thompson Pass 8). This sample is a relict-bedded and partially
bioturbated. The ichnofauna includes Rhizocorallium isp. (Rh), Planolites isp. (Pl), and Phycosiphon isp. (arrowed Ph). An escape trace (dashed outline ET) disrupts one
of the beds. Individual beds have erosional lower contacts (arrow to short dashed lines) and fine upward from being sand-rich at their bases to clay- and silt-enriched
towards their tops. Dewatering has further disrupted some of the erosional contact. B) Low-power optical micrograph of Coquina 9 (sand- and silt-bearing, clay-rich
mudstone) from the top part of an upward-fining unit. Note the presence of clay-rich burrow fills. Scattered, partially crushed agglutinated formaminifers are present in
the matrix. The foraminifer tests are infilled by pyrite. C) Backscattered electron micrograph of the basal portion of an upward-fining unit from Thompson Pass 19 (sandand silt-bearing, clay-rich mudstone). This micrograph illustrates very fine sand grains (composed mainly of quartz with minor feldspar) in a matrix of clay, organic

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

Nearly all samples contain agglutinated foraminifers, including


organisms with uniserial, biserial, trochospiral, and planispiral involute
morphologies (e.g., Figs. 5B, C, E, F, 6B). These foraminifers are quite
well preserved, scattered throughout the matrix, and their tests are
composed predominantly of quartz grains (e.g., Fig. 5C, F). The majority
of the shelter porosity preserved within their tests has been infilled by
framboidal pyrite (e.g., Fig. 5B, C, F); some, however, has been infilled
by authigenic kaolinite.
The composition of the detrital components in the coarse-grained,
carbonate-cemented strata is very different from that of the enclosing
sediments. In particular, these levels contain much greater proportions of
fine-sand-size detrital clay grains, variously composed of chlorite and
muscovite (Fig. 7E, F). Much of this coarse, detrital clay fraction
contains authigenic kaolinite intergrowths (Fig. 7F).
Within each of the three studied successions, there is systematic vertical
lithofacies variability on a 100 m scale. For instance, at Blaze Canyon
successive samples overall fine upward between depths 0 m and 2.0 m
(Fig. 4). They also exhibit upward-coarsening trends on both 1 to 3 m
(small) scales and 5 to 10 m (large) scales (Figs. 4, 8). Additionally, smallscale upward-coarsening units (e.g., from 5.0 to 8.5 m at Thompson Pass,
from 20.0 to 22.5 m at Blaze Canyon (illustrated in Fig. 8A, B, C), and
from 8.0 to 11.0 m at Coquina Wash) are separated from one and another
by noticeably finer-grained intervals (e.g., 8.5 m at Thompson, 22.5 m at
Blaze, and 11.0 m at Coquina Wash) (Fig. 4). To complicate matters
further, the small-scale upward-coarsening units systematically stack into
overall large-scale upward-coarsening successions. These large-scale
upward-coarsening successions are illustrated in Figure 4. (e.g., depths
3.0 and 8.0 m at Coquina Wash, depths 10.0 and 19.0 m at Blaze Canyon,
and depths 8.5 and 13.0 m at Thompson Pass).
Concretionary, cemented units are present either at the tops of these
large-scale upward-coarsening successions (e.g., 19 m at Thompson Pass
and 10.5 m at Coquina Wash), or at horizons where there are significant
grain-size changes (e.g., 2.0 m at the Blaze Canyon location, where there
is abrupt coarsening). In the latter settings, units that contain abundant
coarse detrital clays abruptly overlie silt-bearing clay-rich mudstones
(Fig. 4). Most of these cemented units have high (. 70%) minus-cement
porosities (Table 1) and are composed predominantly of siderite, ferroan
calcite, and nonferroan calcite cements. Granule-size apatite microconcretions are also present associated with these units (Fig. 7E ).
DISCUSSION AND INTERPRETATION

The overall paleogeographic setting, predominant muddy grain-size,


and presence of trace fossils such as Phycosiphon isp., and Planolites isp.,
coupled with the presence of pyrite, indicate that this region of the
Mancos Shale was originally deposited on the distal parts of a clasticdominated, marine continental shelf (see also Fouch et al. 1983; Miall and
Arush 2001; Howell and Flint 2003a).

331

weathering processes. The climate at this time is interpreted to have been


humid (Van Wagoner 1995), because abundant coals and deep
paleoweathering profiles are present in coeval successions updip (Van
Wagoner 1995; Taylor and Gawthorpe 2003). In contrast, the coarsergrained detrital quartz and feldspar components are interpreted to be the
residuum of weathering, and are probably derived ultimately from the
unroofing of the Sevier Fold and Thrust Belt. In addition to siliciclastic
detritus, these sediments also contain some macerated woody material.
This woody debris is likely to be the remains of higher plants that were
living originally in the catchment and transported into the basin following
their demise. Detrital, nonferroan dolomite is a dominant mineral in
fluvial, coastal-plain, and shoreface sandstones in the Book Cliff
successions updip from the Mancos Shale (Taylor et al. 2000) and in
other Upper Cretaceous successions on the western margin of the
Western Interior Seaway (e.g., Crossey and Larsen 1992; McKay et al.
1995). It is likely that this dolomite was derived from the erosion of
dolomite platform strata in the Sevier Fold and Thrust Belt.
Although intense burrowing has destroyed many of the original
depositional fabrics, residual small-scale upward fining is preserved in
some units (Fig. 6A). These upward-fining strata are mostly less than
10 mm thick, and are typically composed of very fine muddy sandstones
and coarser grained mudstones at their bases and bioturbated silt-bearing
clay-rich mudstones towards their tops (e.g., Fig. 7D). Goldring et al.
(1991) and Macquaker and Taylor (1996) described similar structures
from analogous paleogeographic settings (e.g., Cleveland Ironstone
Formation, Staithes Sandstone Formation, exposed on the N.E. coast
of England) that they informally called lam-scram (laminated and
scrambled) units. These authors argued that lam-scram units were
produced by infaunal organisms colonizing the tops of distal storm
deposits between storm events. Given the similarity of the structures
observed here in the Mancos Shale to those from other locations, a similar
interpretation for their origin is invoked.
The presence of bioturbation in the tops of these units indicates that
there was sufficient time between storm events for an infaunal community
to colonize the sediment surface. This colonization overprinted the
original depositional laminae, close to the sedimentwater interface, with
a burrowed fabric, leaving the laminae in deeper tiers undisturbed. Given
this interpretation, the upward-fining lam-scram units conform to the
definition of genetic beds (sensu Campbell 1967), because they are the
products of individual depositional events, separated by breaks in
sedimentation, rather than being component parts of individual depositional units. Many field-based geologists would describe these units
as being laminated, because their parting spacings are , 10 mm, but
given their origin as individual depositional events, we believe that it is
more appropriate to describe them as thin beds. This observation also
means that, unsurprisingly, beds in this part of the Mancos Shale are
much thinner than coeval units deposited in marginal and shallow marine
settings updip (e.g., OByrne and Flint 1995).
Biogenic Components

Clastic Components
The fine-grained detrital clays, which constitute the bulk of sediment
delivered to the study area, were probably derived from reworking of
updip soil profiles, having originally formed in overbank environments by

In addition to detrital components, some autochthonous, biogenic


materials are also present. These include agglutinated foraminifers and
amorphous organic matter. Foraminifers are relatively well preserved and
particularly common in the finer-grained lithofacies. The diversity of

r
matter (note the macerated woody detritus [arrowed]), and pyrite. Kaolinite (low g) fills some of the intergranular porosity (k). D) Thin-section scan of a representative
bioturbated very fine sand-, silt-, and clay-bearing mudstone (Blaze 20). This sample is intensely bioturbated by Planolites isp., and an indeterminate community of
burrowing organisms. Most traces of the depositional laminae have been disrupted. E) Low-power optical micrograph (Blaze 20) of the contact between a mud-filled
burrow and more sandy host sediment in a fine sand-, silt-, and clay-bearing mudstone. F) Backscattered electron micrograph of Blaze 24 (fine sand-, silt-, and claybearing mudstone) illustrating very fine sand grains in a matrix of clay, organic matter, and silt. The sand-size grains are made of quartz (low g). and feldspar
(intermediate g). Pyrite framboids (high g. (arrowed). (This figure is reproduced in color in the digital version.)

332

J.H.S. MACQUAKER ET AL.

JSR

FIG. 7.A) Thin-section scan of a representative bioturbated silt-bearing, very fine-grained muddy sandstone (Thompson Pass 06). The burrowing fauna is dominated
by Chondrites isp. B) Low-power optical micrograph of part of a mud-filled burrow within this fine-grained muddy sandstone. C) Backscattered electron micrograph of
Thompson Pass 06 illustrating very fine-grained muddy sandstone. Individual sand grains are composed of quartz (low g). and feldspar (intermediate g). A pervasive
quartz cement infills the intergranular porosity. D) A thin-section scan of a representative upward-fining, partially bioturbated (dominated by Planolites isp. arrowed),
carbonate-cement-rich mudstone (Coquina Wash 29). Faint residual laminae (arrowed) are preserved towards the bottom of this unit. E) Low-power optical micrograph
of a representative fine sand- and silt-bearing (comprising grains of quartz and altered detrital chlorite, arrowed), carbonate-cement-rich mudstone (Blaze 4). This
photomicrograph was taken close to the margin of the cemented zone. The cements in the matrix comprise authigenic apatite (a) micro-concretions enclosed by zoned

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

forms present in these units suggests that the environment was favorable
and they represent an equilibrium community. There is no evidence to
suggest that these organisms were either simply colonizing the sediment
surface or utilizing an exaerobic metabolic mechanism (i.e., able to
tolerate episodic anoxia). It is possible, however, that they were able to
tolerate slightly lower than normal marine pore-water oxygen concentrations and were utilizing dysaerobic respiratory mechanisms.
Once deposited, the sediment was disrupted by diminutive burrowing
organisms. Much of this infauna has an indeterminate affinity. Planolites
isp., Phycosiphon isp., Chondrites isp., and Terebellina isp., however, have
been identified in the more fine-grained, clay-rich portions of beds, while
diffuse Thallasinoides isp., Palaeophycus isp., Teichichus isp., and
Rhizocorallium isp. are present in the coarse-grained facies at the bases
of beds. There is also evidence of a variety of crosscutting relations, with
Thallasinoides isp. and Planolites isp. being interpreted as occupying the
mid- to upper layers, Phycosiphon isp. and Palaeophycus isp. living in
mid-tiers, and Chondrites isp. colonizing the deep tiers (e.g., Ekdale and
Bromley 1991; Goldring et al. 1991). The presence of this tiering also
indicates that sediment supply was episodic, and that there was sufficient
time for burrowing organisms to develop a stable community in the
sediment between episodes of sediment delivery. That Phycosiphon isp.
traces are so common, have colonized mud, and directly underlie erosion
surfaces (e.g., Fig. 6A) suggests that erosion may have removed both the
mixed-layer close to the sedimentwater interface and the upper burrow
tier prior to deposition of the overlying unit. These data suggest that the
organism responsible for producing Phycosiphon isp. was an opportunistic colonizer of muddy sediment (see also Goldring et al. 1991). Overall,
the presence of this infauna indicates that the substrate, at least in the
mid- and higher sediment tiers, was oxic and firm (rather than soupy) at
the time of deposition. The presence of Chondrites isp., a putative
symbiont (e.g., Bromley 1996), in deeper tiers of the more muddy units
suggests, however, that a sulfidicoxic interface may have existed in the
pore waters fairly close to the sediment surface. Such near-surface
sulfidicoxic interfaces commonly occur within Recent organic-rich shelf
sediments (e.g., Aller et al. 1986; Chanton et al. 1987; Canfield et al.
1993).
Variability of Vertical Stacking Patterns
Analyses of successive samples from all of the sampled sites reveal
larger-scale stacking variability (Fig. 8). While it is tempting to interpret
the 1 to 3 m thick upward-coarsening strata as thick beds, this simplistic
interpretation is likely to be wrong, because individual genetic beds here
are thin (typically , 10 mm thick; see above). Given this observation, we
believe the 1 to 3 m thick upward-coarsening units are stacked beds (or
bedsets) (sensu Campbell 1967). In this context, the existence of much
finer-grained strata between individual upward-coarsening packages is
significant, in as much as their presence suggests that the length of the
sediment transport path increased significantly over these intervals. The
simplest explanation for all these observations is that that water depths
rapidly deepened at these levels, although updip diversion of clastic
sediment supply, perhaps in response to channel switching, is possible.
Given the significant role that changes in relative sea level are known to
exert on updip facies variability elsewhere in the Mid-Cretaceous Seaway
(e.g., Van Wagoner 1995), however, we prefer the first of these two
explanations. We, therefore, interpret these upward-coarsening strata to
be the distal expression of parasequences, separated from one another by

333

much finer-grained units, i.e., marine flooding surfaces (cf. Van Wagoner
et al. 1990). The fact that so many of the exposures are weathered,
coupled with the subtle expression of much of this variability in the field,
means that we are unable to distinguish unequivocally between the
various environmental processes (namely tectonically driven subsidence
and eustatic sea level) that might have forced these grain-size changes
(compare Miall and Arush 2001 with Howell and Flint 2003b). With this
caveat in mind, and bearing in mind the widespread and pervasive nature
of the observed stratal architectures, it is likely that regional controls were
operating to control lithofacies variability. We therefore prefer the option
that eustatic sea-level change was probably a significant driver on facies
variability on this shelf.
Further investigation of successive 1 to 3 m upward-coarsening units
(Fig. 4) indicates they too stack into overall upward-coarsening packages.
Typically each of these larger-scale packages is 5 to 10 m thick and is
composed of three or four individual parasequences that individually
coarsen upward. Using analogous logic that allowed us to interpret the
smaller upward-coarsening units as parasequences, then these large-scale
upward-coarsening units are interpreted to be prograding parasequence
sets (cf. Van Wagoner et al. 1990).
In offshore marine settings, thin overall, upward-fining successions are
unlikely be parasequences because in these environments these units
typically coarsen upward (e.g., Van Wagoner et al. 1990). Given the
upward fining observed here, our best estimate is that these units are
a retrogradationally stacked succession of thin parasequences (each
, 0.5 m thick) that overall fine upward. The sampling strategy that we
adopted in this study, however, was not at sufficient vertical resolution to
demonstrate this pattern unequivocally.
Diagenetic Components
Preferentially cemented units occur noticeably at those surfaces
correlated to amalgamated sequence boundaries and transgressive
surfaces updip. In the Mancos Shale these surfaces mark either levels
where large-scale stacking patterns change or where there are discontinuities in stacking patterns. In addition to these significant features, these
cemented units also possess distinct compositional and textural properties. For instance, their detrital mineralogies are very different from the
surrounding mudstones, and their pore spaces have been infilled with
diagenetic ferroan carbonate (variously siderite, ferroan calcite, and/or
ferroan dolomite) and sulfide (pyrite) cements. Petrographic evidence,
such as high minus-cement porosity, indicates that cementation took
place early and certainly prior to significant compaction. The surviving
cement compositions suggest that they precipitated in response to
increased porewater bicarbonate, iron, and sulfide activities, probably
as a consequence of bacterially mediated metabolic processes (e.g.,
Macquaker and Taylor 1996; Taylor and Macquaker 2000). Certainly,
ferroan carbonate and pyrite cements are well-known products of
bacterial sulfate reduction, iron (III) reduction, and methanogenesis in
analogous (e.g., Coleman 1985; Curtis et al. 1986; Raiswell 1988; Taylor
and Curtis 1995) and closely related depositional environments (Chan
1992; Taylor et al. 2002). The existence of high minus-cement porosities,
and the pre-compaction textures, indicates that significant volumes of
cement were precipitated early in the pore space. This evidence suggests
that the pores in these units were either subjected to prolonged input from
the products of bacterial metabolic activity or associated with enhanced
fluid flow in the subsurface, to facilitate such large volumes of pore space

r
ferroan carbonates. F) Backscattered electron micrograph of a representative carbonate cement-rich mudstone (Coquina 8) illustrating very fine-grained sand and silt
(composed of quartz (intermediate g), K-feldspar, and altered chlorite grains arrowed) in a matrix of fine-grained clay (low g). Zoned, ferroan carbonate cements have
infilled much of the intergranular porosity (high g). Minor framboidal pyrite is also present (very high g). (This figure is reproduced in color in the digital version.)

Brief Description

Silt-bearing, clay-rich mudstones


Blaze 28
Bioturbated (Phycosiphon isp., and Terrebellina isp.) silt-bearing clay-rich mudstone
Blaze 27
Bioturbated (Phycosiphon isp., Planolites isp., and Terrebellina isp.) silt-bearing clay-rich
mudstone
Blaze 15
Bioturbated (Planolites isp., and Phycosiphon isp.) silt-bearing clay-rich mudstone.
Blaze 11
Bioturbated (Phycosiphon isp., Planolites isp.) silt-bearing clay-rich mudstone
Blaze 10
Silt-bearing, clay-rich mudstone
Blaze 09
Bioturbated (Palaeophycus isp., Phycosiphon isp.) silt-bearing clay-rich mudstone.
Blaze 08
Bioturbated (Planolites isp, Phycosiphon isp.) silt-bearing, clay-rich mudstone.
Blaze 07
Bioturbated (Phycosiphon isp., Terrebellina isp.) silt-bearing clay-rich mudstone
Blaze 05
Bioturbated (Phycosiphon isp., Terrebellina isp.) silt-bearing clay-rich mudstone.
Blaze 02
Bioturbated (Planolites isp., Phycosiphon isp., Terrebellina isp.) silt-bearing, clay-rich mudstones
Coquina 01 Bioturbated (Phycosiphon isp., Planolites isp., Terebellina isp.) silt-bearing clay-rich mudstone
Coquina 02 Bioturbated (Planolites isp., Phycosiphon isp.) silt-bearing clay-rich mudstone
Coquina 03 Bioturbated (Planolites isp., Phycosiphon isp., Chondrites isp.) silt-bearing clay-rich mudstone
Coquina 04 Bioturbated (Terebellina isp., Phycosiphon isp., Chondrites isp.) silt-bearing clay-rich mudstone
with abundant woody fragments.
Coquina 05 Bioturbated (Phycosiphon isp., Terebellina isp.) silt-bearing clay-rich mudstone.
Coquina 07 Bioturbated (Terebellina isp., Phycosipon isp., Planolites isp.) silt-bearing clay-rich mudstone
Coquina 10 Bioturbated (Planolites isp.) silt-bearing clay-rich mudstone
Coquina 11 Bioturbated (Planolites isp., Phycosiphon isp., Terebellina isp.) silt-bearing clay-rich mudstone
Coquina 16 Bioturbated, (Terebellina isp., Phycosiphon isp., Planolites isp.) silt-bearing clay-rich mudstone
Coquina 19 Bioturbated, (Phycosiphon isp., Planolites isp.) silt-bearing, clay-rich mudstone
Coquina 21 Bioturbated (Phycosiphon isp.) silt-bearing clay-rich mudstone
Coquina 22 Bioturbated (Terebellina isp., Phycosiphon isp. Planolites isp.) silt-bearing clay-rich mudstone .
Coquina 24 Bioturbated (Phycosiphon isp., Planolites isp.) silt-bearing, clay-rich mudstone
Coquina 25 Bioturbated (Planolites isp., Phycosiphon isp., Paleophycus isp.) silt-bearing clay-rich mudstone.
Coquina 26 Bioturbated (Phycosiphon isp., Planolites isp.) silt-bearing, clay-rich mudstone
Thompson
Bioturbated (Palaeophycus isp., Planolites isp., Chondrites isp.) silt-bearing, clay-rich mudstone
Pass 24
Thompson
Bioturbated (Chondrites isp., Phycosipon isp.) silt-bearing clay-rich mudstone
Pass 23
Thompson
Bioturbated (Planolites isp., Phycosiphon isp., Chondrites isp.) silt-bearing clay-rich mudstone
Pass 22
Thompson
Bioturbated silt-bearing clay-rich mudstone
Pass 18
Thompson
Bioturbated silt-bearing, clay-rich mudstone
Pass 16
Thompson
Bioturbated (Palaeophycus isp. Chondrites isp.) silt-bearing, clay-rich mudstone.
Pass14

Clay-rich mudstones
Blaze 13
Clay-rich mudstone with Terrebellina isp. and Phycosiphon isp.
Blaze 06
Bioturbated (Planolites isp., Terrebellina isp.) clay-rich mudstone with a small proportion of
very fine silt.
Blaze 03
Bioturbated (Phycosiphon isp.) clay-rich mudstone
Coquina 06 Relict-bedded and bioturbated (Phycosiphon isp., and Planolites isp.) clay-rich mudstone
Coquina 18 Bioturbated (Planolites isp.) clay-rich mudstone with some silt
Coquina 23 Bioturbated (Planolites isp., and Phycosiphon isp.) clay-rich mudstone
Coquina 27 Bioturbated (Phycosiphon isp., Planolites isp., and Chondrites isp.) clay-rich mudstone with
a small proportion of silt
Coquina 28 Bioturbated (Phycosiphon isp., Planolites isp., and Terebellina isp.) clay-rich mudstone with
a small proportion of silt.

Sample

15.0
10.0
15.0
12.5
10.0
20.0
12.0
10.0
25.0
20.0
15.0
15.0
10.0
10.0
10.0
12.5
12.5
10.0
15.0
12.5
10.0
10.0
10.0
20.0
15.0
15.0
25.0
25.0
25.0

7.5
2.5
7.5
5.0
2.5
5.0
2.0
5.0
5.0
2.5
7.5
5.0
2.5
1.0
2.5
5.0
5.0
5.0
5.0
1.0
0.0
5.0
1.0
2.5
0.0
2.5
5.0
5.0
5.0

7.5

0.0

10.0
12.5

7.5
7.5
7.5
7.5
7.5

2.5
0.0
2.5
0.0
0.0

5.0
7.5

%
7.5
7.5

Silt

%
0.0
1.0

Fine
sand

67.5

0.0

0.0

0.0

0.0

0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0

0.0

0.0
0.0
0.0
0.0
0.0

%
0.0
0.0

Nannoplankton

0.0

0.0

0.0

0.0

0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
2.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0

0.0

0.0
0.0
0.0
0.0
0.0

%
0.0
0.0

Authigenic
carbonate

0.5

0.5

0.5

2.0

1.0

1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

1.0
1.0
1.0
1.0
1.0
1.0
1.0
0.5
1.0
1.5
1.0
1.0

2.0
1.0

1.0

1.0
2.0
1.0
1.0
1.0

%
1.0
1.0

1.0

1.0

1.0

1.0

1.0

1.0
1.0
1.0
1.0
1.0
1.5
1.0
1.0
1.0
1.0
1.0
1.0

1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

1.0
1.0

1.0

1.0
1.0
1.0
1.0
1.0

%
1.0
1.0

0.5

0.5

0.5

0.5

0.5

1.0
1.0
1.0
1.0
1.0
0.5
0.5
0.5
0.5
1.0
1.0
0.5

0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
1.0
1.0
1.0

1.0
1.0

0.5

0.5
1.0
0.5
0.5
0.5

%
0.5
0.5

Foraminifer
Visible
material Pyrite om

0.5

0.5

0.5

0.5

0.5

0.0
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.0
0.0
0.5

0.0
0.0
0.0
0.0
0.0
0.0
0.5
0.0
0.5
0.5
0.5
0.5

0.5
0.5

0.5

0.0
0.5
0.5
0.5
0.5

%
0.5
0.5

Authigenic
Clays

1.6

1.9

1.8

1.6

1.7
1.7
1.4
1.3
1.9
1.7
1.4
1.8
2.0

1.9
1.8
1.8

1.7
1.3
1.1
1.4

1.3

1.7

1.1
1.8
1.5
2.1
1.8

%
1.6
1.3

TOC

J.H.S. MACQUAKER ET AL.

67.5

67.5

78.5

82.0

84.5
85.5
84.0
79.0
79.0
81.5
77.0
83.5
87.0
82.0
86.0
74.5

75.0
85.0
75.0
78.0
85.0
72.5
83.0
83.0
67.0
73.5
74.0
76.5

80.5
76.5

89.5

87.5
88.0
87.0
89.5
89.5

%
89.5
88.5

Clay

TABLE 1.Semiquantitative descriptions of each sample analyzed.

334

JSR

1.0
5.0
2.5
7.5

Bioturbated silt-bearing, clay-rich mudstone

Bioturbated (Planolites isp., Phycosiphon isp.) silt and dolomite cement, bearing clay-rich mudstone

Bioturbated (Planolites isp., Chondrites isp.) silt-bearing clay-rich mudstone

Bioturbated (Planolites isp., Palaeophycus isp., and Chondrites isp.) silt bearing, clay-rich mudstone

20.0
15.0
10.0
15.0
15.0
15.0
20.0
15.0
30.0
30.0
20.0
30.0
30.0
20.0
20.0

10.0
10.0
15.0
10.0
10.0
10.0
10.0
25.0
10.0
10.0
10.0
15.0
10.0
10.0
10.0

35.0
30.0

15.0

10.0

30.0
20.0

10.0
20.0
35.0
25.0
20.0

15.0
15.0
30.0
20.0
15.0

15.0

10.0

20.0

33.0
48.5

68.5

68.0

58.0

51.0

67.0

58.0

57.5

67.0
57.5

71.5
72.0

71.5

73.0

72.5

67.5

72.0

72.0
63.0
33.5
53.5
63.0

75.5

85.0

63.5

82.5

77.0

75.5

Clay

0.0
0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0
0.0

0.0
0.0

0.0

0.0

0.0

0.0

0.0

0.0
0.0
0.0
0.0
0.0

0.0

0.0

0.0

0.0

0.0

0.0

Nannoplankton

0.0
0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0

0.0
0.0

0.0
0.0

0.0

0.0

0.0

0.0

0.0

0.0
0.0
0.0
0.0
0.0

0.0

0.0

10.0

0.0

0.0

0.0

Authigenic
carbonate

0.0
0.0

0.0

0.0

0.0

0.5

0.5

0.0

0.5

1.0
0.5

1.0
1.0

1.0

0.5

0.5

1.0

1.0

1.0
0.5
0.0
0.0
0.5

0.5

1.0

0.0

0.0

1.0

0.0

1.0
1.0

1.0

1.0

1.0

0.5

1.0

1.0

1.0

1.0
1.0

1.0
1.0

1.0

1.0

1.0

1.0

1.0

1.0
1.0
1.0
1.0
1.0

1.0

1.0

1.0

1.0

1.0

1.0

0.5
0.5

0.5

0.5

0.5

2.5

1.0

0.5

0.5

0.5
0.5

1.0
0.5

1.0

0.5

1.0

0.5

1.0

1.0
0.5
0.5
0.5
0.5

0.5

0.5

0.5

0.5

0.5

0.5

Foraminifer
Visible
material Pyrite om

0.5
0.0

0.0

0.5

0.5

0.5

0.5

0.5

0.5

0.5
0.5

0.5
0.5

0.5

0.0

0.0

0.0

0.0

0.0
0.0
0.0
0.0
0.0

0.0

0.0

0.0

0.0

0.5

0.5

Authigenic
Clays

1.9

0.9

1.0

1.3

1.1

0.6

0.9

1.2
1.4

1.8
1.6

1.3

1.0

1.5

1.4

1.4

1.1
1.3

1.3

2.0

1.8

1.6

1.1

TOC

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

Sand-, silt-, and clay-bearing mudstones


Blaze 25b
Bioturbated, fine and very fine sand-, silt- and clay-bearing mudstone
Blaze 25a
Bioturbated, fine and very fine sand-, silt- and clay-bearing mudstone

Sand- and silt-bearing, clay-rich mudstones


Blaze 26
Bioturbated (Phycosiphon isp.) fine sand- and silt-bearing, clay-rich mudstone
Blaze 23
Bioturbated, very fine sand- and silt-bearing, clay-rich mudstone.
Blaze 22
Bioturbated (Palaeophycus isp.) fine and very fine sand- and silt-bearing, clay-rich mudstone.
Blaze 18
Bioturbated (Phycosiphon isp.), very fine sand- and silt-bearing, clay-rich mudstone.
Blaze 17
Bioturbated (Phycosiphon isp. and Terrebellina isp.) very fine sand- and silt-bearing, clay-rich
mudstone.
Blaze 16
Bioturbated (Phycosiphon isp., Planolites isp.) very fine sand and silt-bearing, clay-rich
mudstones
Blaze 14
Bioturbated (Phycosiphon isp., Palaeophycus isp., Teichichus isp., and Planolites isp.) very fine
sand- and silt-bearing, clay-rich mudstones
Blaze 12
Thin and relict-bedded, partially bioturbated (Planolites isp., Phycosiphon isp.) very fine sandand silt-bearing, clay-rich mudstone.
Blaze 01
Bioturbated (Planolites isp., Phycosiphon isp.) very fine sand- and silt-bearing, clay-rich
mudstone
Coquina 09 Bioturbated (Terebellina isp., Phycosiphon isp., and Planolities isp.) very fine sand- and siltbearing clay-rich mudstone
Coquina 12 Bioturbated (Planolites isp., Phycosiphon isp.) very fine sand- and silt-bearing, clay-rich mudstone.
Coquina 17 Bioturbated (Terebellina isp., and Planolites isp.) silt- and very fine sand-bearing clay-rich
mudstone
Coquina 20 Bioturbated silt- and fine sand-bearing, clay-rich mudstone
Thompson
Bioturbated very fine sand- and silt-bearing, clay-rich mudstone
Pass 21
Thompson
Bioturbated very fine sand- and silt-bearing, clay-rich mudstone
Pass 19
Thompson
Bioturbated (Rhizocorallium isp., Planolites isp., and Palaeophycus isp.) very fine sand- and siltPass 15
bearing clay-rich mudstone
Thompson
Bioturbated (Phycosiphon isp., and Planolites isp.) very fine sand- and silt-bearing, clay-rich
Pass 12
mudstone
Thompson
Relict-bedded, bioturbated (Chondrites isp., and Tecichnus isp.) sand- and silt-bearing, clay-rich
Pass 11
mudstone. Individual beds upward-fine are disrupted by dewatering. Some macerated wood
Thompson
Bioturbated (Terebellina isp.? Teichicnus isp., Planolites isp., and Chondrites isp.) very fine sandPass 09
and silt-bearing clay-rich mudstone
Thompson
Relict-bedded, partially bioturbated (Rhizocorallium isp., Chondrites isp., and Planolites isp.)
Pass 08
sand- and silt-bearing clay-rich mudstone.Beds have erosional lower contacts and upward-fine.
Thompson
Bioturbated (Planolites isp.) very fine sand- and silt-bearing clay-rich mudstone
Pass 01

15.0

5.0
15.0

20.0

Silt

2.5

Thompson
Pass 13
Thompson
Pass 07
Thompson
Pass 05
Thompson
Pass 04
Thompson
Pass 03
Thompson
Pass 02

Fine
sand

Bioturbated (Planolites isp., Chondrites isp., Palaeophycus isp., Teichichnus isp.) silt bearing
clay-rich mudstone. Chondrites isp. growing within other burrow fills.
Bioturbated (Planolites isp., Chondrites isp.) silt-bearing clay-rich mudstone

Brief Description

Sample

TABLE 1.Continued.

JSR
335

25
24
20
19

25.0
20.0
25.0
25.0

Bioturbated, fine and very fine sand-, silt- and clay-bearing mudstone
Bioturbated very fine sand-, silt- and clay-bearing mudstones
Bioturbated (Phycosiphon isp., and Planolites isp.) very fine sand-, silt- and clay-bearing mudstones
Bioturbated (Phycosiphon isp.) very fine sand-, silt- and clay-bearing mudstone

Carbonate-cement-rich mudstones
Blaze 04
Pelleted fine sand- and silt-bearing, siderite cement-rich mudstone
20.0
Coquina 8
Bioturbated (Planolites isp. and Phycosiphon isp.) carbonate cement-rich mudstones
2.5
Coquina 13 Carbonate cement-rich mudstone
0.0
Coquina 14 Carbonate cement-rich mudstone
1.0
Coquina 15 Bioturbated silt- and clay-bearing carbonate cement-rich mudstone
0.0
Coquina 29 Thin, relict-bedded, partially bioturbated (Planolites isp.), carbonate-cement-rich mudstone.
0.5
Indiviudal beds upward-fine and have bioturbated tops.
Thompson
Bioturbated (Chondrites isp.) fine-grained sand- and silt-bearing, carbonate cement-rich mudstone 20.0
Pass 20
Thompson
Bioturbated (Chondrites isp., and Planolites isp.) silt-bearing carbonate-cement-rich mudstone 1.0
Pass 10

60.0

Fine
sand

Brief Description

Fine-grained muddy sandstone


Thompson
Bioturbated (Chondrites isp.) silt-and clay-bearing, very fine-grained muddy sandstone.
Pass 06

Blaze
Blaze
Blaze
Blaze

Sample

TABLE 1.Continued.

4.0
9.0
6.0
8.0
18.5
7.5
0.1
5.0

10.0
10.0

18.0

38.5
48.5
43.5
43.5

Clay

15.0
5.0
2.5
5.0
10.0
7.5

20.0

35.0
30.0
30.0
30.0

Silt

0.0

0.0

0.0
0.0
0.0
0.0
0.0
0.0

0.0

0.0
0.0
0.0
0.0

Nannoplankton

83.5

69.9

55.0
80.0
90.0
85.0
70.0
80.0

0.0

0.0
0.0
0.0
0.0

Authigenic
carbonate

0.0

0.0

0.0
0.0
0.0
0.0
0.0
2.0

0.0

0.0
0.0
0.0
0.0

0.5

0.0

0.5
1.0
0.5
0.5
1.0
1.0

1.0

1.0
1.0
1.0
1.0

0.0

0.0

0.5
0.5
0.5
0.5
0.5
1.0

0.5

0.5
0.5
0.5
0.5

Foraminifer
Visible
material Pyrite om

0.0

0.0

5.0
2.0
0.5
0.0
0.0
0.5

0.5

0.0
0.0
0.0
0.0

Authigenic
Clays

0.4

0.6

0.6
1.0
0.3
0.6
0.7
1.3

0.2

1.2
0.9

TOC

336
J.H.S. MACQUAKER ET AL.

JSR

FIG. 8.Low-power optical micrographs illustrating small-scale upward


coarsening at Blaze Canyon between sample depths 20.0 and 22.0 m (Fig. 4).
The proportion of sand and silt increases from the bottom to the top samples. A)
Bioturbated, fine sand-, silt-, and clay-bearing mudstone with a sand + silt/clay
ratio of 1.6 (Blaze 25). B) Bioturbated, fine sand-, silt-, and clay-bearing mudstone
with a sand + silt/clay ratio of 1.0 (Blaze 24). C) Bioturbated, fine sand-, and siltbearing clay-rich mudstone with a sand + silt/clay ratio of 0.6 (Blaze 23). (This
figure is reproduced in color in the digital version.)

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

being infilled with cement. If these cements were indeed precipitating in


response to bacterial metabolic activity, then it is most likely that these
surfaces were located close to the sedimentwater interface for prolonged
periods (e.g., Raiswell 1987; Macquaker and Taylor 1996; Taylor et al.
2000; Taylor and Macquaker 2000). Together, these factors strongly
suggest that at these levels cementation was indeed intimately associated
with long breaks in sediment accumulation. Of course the presence of
cemented mudstones at major stratal surfaces, and in particular
associated with sequence boundaries and transgressive surfaces, has also
been documented elsewhere (e.g., Macquaker and Taylor 1996; Taylor
and Macquaker 2000; Taylor et al. 2002). In these settings, as with the
Mancos Shale, the significant controlling element at these levels is that
there was enough time to transport sufficient solutes derived from the
metabolic processes of bacteria to the precipitation sites to fill most of the
uncompacted pore space with cement.
Processes on Stratal Surfaces
Given that these cemented mudstones are located at sequence
boundaries, their very different detrital mineralogical composition,
compared with the rest of the succession, is significant (these units, in
addition to containing an assemblage comprising quartz, feldspars, finegrained detrital dioctahedral micas, and detrital dolomite also contain,
detrital-grained muscovite and abundant chlorite). The existence of so
much coarse, dense, clay detritus at these levels suggests that dynamic
bypass was also occurring on these surfaces. Under these conditions, less
dense and finer detritus (e.g., soil-derived dioctadedral micas and quartz
silt) was carried farther down the sediment transport path, leaving
a coarse, dense, winnowed lag on these surfaces. Of course, it is possible
that this pattern may have resulted from a brief provenance shift at this
level. Such an interpretation, however, does seem unnecessarily complex,
in as much as bypass processes are commonly invoked as having occurred
at sequence boundaries, both in other mudstone successions (see
Macquaker 1994; Bohacs 1998) generally, and here in the Mancos Shale
(OByrne and Flint 1995; Hampson et al. 1999), although not previously
recognized at this level.
Interestingly, at the Thompson Pass and Coquina Wash localities the
cemented units at the Lower Grassy Sequence Boundary overlie an
upward coarsening succession, whereas at Blaze Canyon they overlie an
upward fining succession (Fig. 4). The presence of upward fining at the
Blaze Canyon location below this level suggests that erosion has removed
the underlying highstand systems tract at this particular locality. At Blaze
Canyon the existence of significant erosion and dynamic bypass at this
level is not obvious from field investigations alone. This confirms the
general observations that mud-on-mud erosion surfaces, sequence
boundaries, and bypass surfaces are cryptic and difficult to observe in
such settings, unless they have been preferentially cemented. No doubt
similar surfaces, with associated erosion, are common in analogous
depositional settings elsewhere, and sedimentologists should look out for
them.
Inferred maximum flooding and marine flooding surfaces here, unlike
some other mudstone-dominated successions (e.g., the Cleveland
Ironstone Formation in northeast England see Macquaker and Taylor
1996), are marked neither by preferential cementation nor by any unusual
enrichment in organic carbon contents. Instead, units at these levels are
clay-rich mudstones at Coquina Wash and Blaze Canyon and silt-bearing
clay-rich mudstones at Thompson Pass. In this part of the Mancos Shale
the maximum flooding surfaces are therefore cryptic. This suggests that
although there may have been a short break in supply of detrital sediment
at these levels, the breaks were not of sufficient duration to allow (because
the rate of supply of solutes from bacterial metabolic processes were too
slow) significant volumes of cement to precipitate in the uncompacted
sediment pore spaces and a preferentially cemented layer to develop.

337

Spatial and Temporal Variability


The lateral variability observed from relatively coarse muddy
sandstones and sand- and silt-bearing clay-rich mudstones, with an
infaunal assemblage comprising Palaeophycus isp., Chondrites isp.,
Planolites isp., and Rhizocorallium isp. at Thompson Pass, through to
finer grained silt-bearing clay-rich mudstones, and clay-rich mudstones,
with an infaunal assemblage comprising Phycosiphon isp., Planolites isp.,
and Terebellina isp. at Coquina Wash, suggests a significant distal shift in
depositional environments, from lower shoreface in the more proximal
locations (Thompson Pass), to offshore-transition zones in the more
distal locations (Coquina Wash) (see also Goldring et al. 1991; Taylor
and Gawthorpe 2003; Macquaker and Taylor 1996). This lateral
variability is consistent with the vertical variability and suggests that
the sediments in this succession are indeed genetically related and that it is
reasonable to use sequence stratigraphic principles to predict their
variability.
Most of the succession, with the exception of the carbonate-cemented
horizons, contains between 1 and 2% total organic carbon. The most
distal samples from Coquina Wash contain, on average, the most TOC
(1.5%). Typically, the finest-grained facies, i.e., clay-rich mudstones and
silt-bearing clay-rich mudstones, are the most enriched in organic matter,
although the lithofacies TOC trends at any one location are not that
significant. The absence of unusually high organic-matter contents is
unsurprising given the extent of infaunal colonization of the sediment. Its
relative abundance in the finest-grained mudstone facies does suggest that
burial efficiencies were maximized in these locations and/or some form of
organic matter adsorption on to clays was occurring (compare with
Tyson 1995; Kennedy et al. 2002; Tyson 2005) to result in the sediment
being enriched in organics. Certainly, there is no evidence of enrichment
of organics being a response to the existence of bottom-water anoxia in
this part of the succession.
Wider Implications
This work suggests that during deposition of the Mancos Shale the
distal continental shelf was a dynamic environment and strongly tied to
more proximal shoreface settings. Assuming that these observations are
representative of other shelf settings then:

Distal shelf environments are energetic, even at times of relatively high


sea-level stands, and are not low-energy systems where sediment input
is derived mainly from pelagic rain. In these settings the sediment
water interface is subject to significant reworking and sediment inputs
from a number of sources.
The data from this study reinforce the fact that the whole shelf
depositionaldiagenetic system in marginal marine systems is closely
integrated and it is an artificial device to separate it on the basis of
rock type. This observation has major implications for those seeking
to model sediments on continental shelves from source to sink.
Despite the fact that the sedimentwater interface in these environments was dynamic and the surface sediment layers were burrowed,
significant quantities of organic matter were still preserved. This
emphasizes the fact that bottom-water anoxia is not a prerequisite for
organic-matter preservation and implies that efficiencies of organicmatter burial are probably controlled by rates of sediment accumulation.
CONCLUSIONS

Combined field observations and optical, electron optical, and


geochemical analyses reveal a great deal of subtle vertical and spatial
lithofacies variability in the Mancos Shale that is not readily apparent
from field studies alone. By use of these techniques eight lithofacies can be

338

JSR

J.H.S. MACQUAKER ET AL.

identified readily based on different proportions of components derived


from detrital, productivity, and chemical processes that each contains.
These are: a) clay-rich mudstones, b) silt-bearing, clay-rich mudstones, c)
very fine sand- and silt-bearing, clay-rich mudstones, d) very fine sand-,
silt- and clay-bearing mudstones, e) fine-grained muddy sandstones, f)
carbonate-cement-rich mudstones, g) silt- and clay-bearing, carbonatecement-rich mudstones, and h) fine-sand- and silt-bearing, carbonate
cement-rich mudstones.
Textural analyses of the thin sections reveal that individual genetic beds
are mostly very thin (, 10 mm) and were probably the distal products of
storm events. There was sufficient time for the sediment, once deposited,
to be extensively colonized by a diminutive macrofauna prior to the next
storm event. The presence of this infauna indicates that the porewaters
close to the sedimentwater interface were oxic. The dominance of midto lower-tier ichnogenera at the tops of these thin beds suggests that the
mixed-layer and upper faunal tiers were reworked prior to deposition of
the overlying storm units. The absence of these upper tiers suggests that
conditions at the sedimentwater interface were at least episodically
erosive, and indicates that this outer-shelf environment was much more
dynamic than low-resolution field observations suggest.
In spite of not being obvious in the field, analyses of successive vertical
samples from any one location indicate that individual beds stack
systematically into parasequences (1 to 3 m thick) and that the
parasequences themselves stack into systems tracts (5 to 10 m thick).
This larger-scale variability also manifests itself spatially, in as much as
the samples from Coquina Wash are overall significantly finer grained
than those from the same interval at Thompson Pass. Changing
accommodation availability in response to widespread changes in relative
sea level and sediment-dispersal patterns are likely to be responsible for
the substantial, yet largely cryptic (at hand specimen scales), lithofacies
variability preserved in these mudstones.
The presence of early pre-compaction cements (mainly iron-rich
carbonates with minor apatite) at the sequence boundaries is interpreted
to indicate that these particular units were subject to prolonged
bacterially mediated and predominantly anaerobic, diagenetic processes
(Fe-reduction, sulfate reduction, and methanogenesis). Moreover, the
existence of stacking pattern discontinuities below some of these
cemented horizons (e.g., at the Lower Grassy Sequence Boundary at
Blaze) and the presence of a completely different coarse-grained detrital
assemblage at these levels (dominated by detrital chlorite and muscovite),
suggests that erosion and dynamic bypass were occurring at these levels,
prior to the sediment being cemented during the subsequent stillstand and
following transgression. The variability observed downdip reinforces the
sequence stratigraphic interpretations that have already been made updip,
and provides concrete evidence of the dynamic bypass necessary to
generate the lowstand fans of OByrne and Flint (1996). These
observations also vividly demonstrate that field studies alone do not
have sufficient resolution to describe the facies variability present in these
successions and that the distal parts of continental shelves may be
dynamic environments.
These data indicate that in the field the most easily recognizable
architectural elements in this part of the Mancos Shale are the 5 to 10 m
thick units which are interpreted variously to be parasequence sets and
systems tracts in addition to the cemented sequence boundaries over
which bypass occurred. So, while it is not easy to do conventional
sequence stratigraphic analyses using field investigations alone, it is
possible to perform these analyses using combined optical, electron
optical, geochemical, and field data, because these provide details of the
cryptic variability occurring at individual bed and bedset scales.
Moreover, when these high-resolution studies are done in the context of
a well defined stratigraphic framework, they confirm existing interpretations of updip and downdip facies variability and reinforce the fact that
sediments on distal parts of continental shelves form part of a continuum

of sediments that exist in marginal marine environments. Together, these


observations suggest that high-resolution studies of mudstones have
a bright future as part of continuing research to investigate the controls
on variability of these sediments as researchers continue to investigate
sediments from their sources to their ultimate sinks.
ACKNOWLEDGMENTS

We would like to thank Kevin Bohacs, Majorie Chan, and Associate Editor
Richard Yuretich for thoughtful reviews of this paper. We also wish to
acknowledge the remarkable support we get from Harri Williams and David
Oliver, who made the polished thin sections of these clay-rich sediments that
ultimately make this type of work possible. The School of Earth,
Environmental and Atmospheric Sciences at the University of Manchester
and the Department of Environmental and Geographical Sciences at the
Manchester Metropolitan University are also thanked for their continuing
logistic support. Finally, where would the Journal be without the support of
John Southard, Colin North, Melissa Lester (and of course Kitty Milliken),
our paper is much better for your input, thank you.
REFERENCES

ADAMS, M.M., AND BHATTACHARYA, J.P., 2005, No change in fluvial style across
a sequence boundary, Cretaceous Blackhawk and Castlegate Formations of Central
Utah, U.S.A.: Journal of Sedimentary Research, v. 75, p. 10381051.
ALLER, R.C., MACKIN, J.E., AND COX, R.T., 1986, Diagenesis of Fe and S in Amazon
inner shelf muds: apparent dominance of Fe reduction and implications for the genesis
of ironstones: Continental Shelf Research, v. 6, p. 263289.
APLIN, A.C., FLEET, A.J., AND MACQUAKER, J.H.S., 1999, Muds and mudstones: physical
and fluid flow properties in mudstones at a basin scale, in Aplin, A., Fleet, A., and
Macquaker, J., eds., Muds and Mudstones: Physical and Fluid-Flow Properties:
Geological Society of London, Special Publication 158, p. 17.
BLATT, H., 1970, Determination of mean sediment thickness in the crust: a sedimentological method: Geological Society of America, Bulletin, v. 81, p. 255266.
BOHACS, K.M., 1998, Contrasting expressions of depositional sequences in mudrocks
from marine to non-marine environs, in Schieber, J., Zimmerle, W., and Sethi, P.S.,
eds., Shales and Mudstones: Stuttgart, Schweizerbartsche Verlagsbuchhandlung, p.
3378.
BROMLEY, R.G., 1996, Trace Fossils; Biology and Taphonomy (Second Edition):
London, Taylor and Francis, 361 p.
BURCHFIEL, B.C., COWAN, D.S., AND DAVIS, G.A., 1992, Tectonic overview of the
Cordilleran orogen in the western United States, in Burchfiel, B.C., Lipman, P.W.,
and Zoback, M.L., eds., The Cordilleran Orogen: Conterminous U.S.: Geological
Society of America, The Geology of North America, v. G-3, p. 407480.
CAMPBELL, C.V., 1967, Lamina, laminaset, bed and bedset: Sedimentology, v. 8,
p. 726.
CANFIELD, D.R., THAMDRUP, B., AND HANSEN, J.W., 1993, The anaerobic degradation of
organic matter in Danish coastal sediments: iron reduction, manganese reduction and
sulfate reduction: Geochimica et Cosmochimica Acta, v. 57, p. 38673883.
CHAN, M.A., 1992, Oolitic ironstone of the Cretaceous Western Interior Seaway, eastcentral Utah: Journal of Sedimentary Petrology, v. 62, p. 693705.
CHAN, M.A., NEWMAN, S.L., AND MAY, F.E., 1991, Deltaic and shelf deposits in the
Cretaceous Blackhawk Formation and Mancos Shale, Grand County, Utah: Utah
Geological Survey, Miscellaneous Publication, 91-6, 83 p.
CHANTON, J.P., MARTENS, C.S., AND GOLDHABER, M.B., 1987, Biogeochemical cycling in
an organic-rich coastal marine basin, 7. Sulfur mass balance, oxygen uptake and
sulfide retention: Geochimica et Cosmochimica Acta, v. 51, p. 11871199.
COLEMAN, M.L., 1985, Geochemistry on diagenetic non-silicate minerals: kinetic
considerations: Royal Society (London), Philosophical Transactions, v. A315, p.
3956.
CROSSEY, L.J., AND LARSEN, D., 1992, Authigenic mineralogy of sandstone intercalated
with organic-rich mudstones: integrating diagenesis and burial history of the
Mesaverde Group, Piceance Basin, NW Colorado, in Houseknecht, D.W., and
Pittman, E.D., eds., Origin, Diagenesis, and Petrophysics of Clay Minerals in
Sandstones: SEPM, Special Publication 47, p. 125144.
CURTIS, C.D., COLEMAN, M.L., AND LOVE, L.G., 1986, Pore-water evolution during
sediment burial from isotope and mineral chemistry of calcite, dolomite and siderite
concretions: Geochimica et Cosmochimica Acta, v. 50, p. 23212334.
EDWARDS, C.M., HOWELL, J.A., AND FLINT, S.S., 2005, Depositional and stratigraphic
architecture of the Santonian Emery Sandstone of the Mancos Shale: implications for
Late Cretaceous evolution of the western interior foreland basin of Central Utah,
U.S.A.: Journal of Sedimentary Research, v. 75, p. 280299.
EKDALE, A.A., AND BROMLEY, R.G., 1991, Analysis of composite ichnofabrics: an
example in uppermost Cretaceous chalk of Denmark: Palaios, v. 6, p. 232249.
FLUGEL, E., 1982, Microfacies analysis of limestones: Berlin, Springer, 633 p.

JSR

MANCOS SHALE SEDIMENTOLOGY AND STRATIGRAPHY

FOUCH, T.D., LAWTON, T.F., NICHOLS, D.J., CASHION, W.B., AND COBBAN, W.A., 1983,
Patterns and timing of synorogenic sedimentation in Upper Cretaceous rocks of
central and northeast Utah, in Reynolds, M.W., and Dolly, E.D., eds., Mesozoic
Paleogeography of West-Central United States: SEPM, Rocky Mountain Section,
Symposium 2, p. 305334.
GOLDRING, R., POLLARD, J.E., AND TAYLOR, A.M., 1991, Anconichnus horizontalis:
a pervasive ichnofabric-forming trace fossil in post-Paleozoic offshore siliciclastic
facies: Palaios, v. 6, p. 250263.
HAMPSON, G.J., HOWELL, J.A., AND FLINT, S.F., 1999, A sedimentological and sequence
stratigraphic re-interpretation of the Upper Cretaceous Prairie Canyon Member
(Mancos B) and associated strata, Book Cliffs area, Utah, U.S.A.: Journal of
Sedimentary Research, v. 69, p. 414433.
HOWELL, J.A., AND FLINT, S.F., 2003a, Tectonic Settings, stratigraphy and sedimentology of the Book Cliffs, in Coe, A.L., ed., Sedimentary Record of Sea-Level Change:
Cambridge University Press and Open University, p. 135157.
HOWELL, J.A., AND FLINT, S.F., 2003b, Sequences and systems tracts in the Book Cliffs,
in Coe, A.L., ed., Sedimentary Record of Sea-Level Change: Cambridge University
Press and Open University, p. 179197.
KAMOLA, D.L., AND HUNTOON, J.E., 1995, Repetitive stratal patterns in a foreland basin
sandstone and their possible tectonic significance: Geology, v. 23, p. 177180.
KENNEDY, M.J., PEVEAR, D.R., AND HILL, R.J., 2002, Mineral surface control of organic
carbon in black shale: Science, v. 295, p. 657660.
MACQUAKER, J.H.S., 1994, A lithofacies study of the Peterborough Member, Oxford
Clay Formation (Jurassic), U.K.: an example of sediment bypass in a mudstone
succession: Geological Society of London, Journal, v. 151, p. 161172.
MACQUAKER, J.H.S., AND ADAMS, A.E., 2003, Maximizing information from finegrained sediments: an inclusive nomenclature for mudstones: Journal of Sedimentary
Research, v. 73, p. 735744.
MACQUAKER, J.H.S., AND GAWTHORPE, R.L., 1993, Mudstone lithofacies in the
Kimmeridge Clay Formation, Wessex Basin, southern England: implications for the
origin and controls of the distribution of mudstones: Journal of Sedimentary
Petrology, v. 63, p. 11291143.
MACQUAKER, J.H.S., AND HOWELL, J.K., 1999, Small-scale (, 5.0 m) vertical heterogeneity in mudstones: implications for high-resolution stratigraphy in siliciclastic
mudstone successions: Geological Society of London, Journal, v. 156, p. 105112.
MACQUAKER, J.H.S., AND KELLER, M.A., 2005, Mudstone sedimentation at high
latitudes: ice as transport medium for mud and supplier of nutrients: Journal of
Sedimentary Research, v. 75, p. 696709.
MACQUAKER, J.H.S., AND TAYLOR, K.G., 1996, A sequence stratigraphic interpretation
of a mudstone-dominated succession: the Lower Jurassic Cleveland Ironstone
Formation, U.K.: Geological Society of London, Journal, v. 153, p. 759770.
MACQUAKER, J.H.S., GAWTHORPE, R.L., TAYLOR, K.G., AND OATES, M.J., 1998,
Heterogeneity, stacking patterns and sequence stratigraphic interpretation in distal
mudstone successions: examples from the Kimmeridge Clay Formation, U.K, in
Schieber, J., Zimmerle, W., and Sethi, P.S., eds., Shales and Mudstones: Stuttgart,
Schweizerbartsche Verlagsbuchhandlung, p. 163186.
MCKAY, J.L., LONGSTAFFE, F.J., AND PLATT, A.G., 1995, Early diagenesis and its
relationship to depositional environment and relative sea-level fluctuations (Upper
Cretaceous Marshybank Formation, Alberta and British Columbia): Sedimentology,
v. 42, p. 161190.
MIALL, A.D., AND ARUSH, M., 2001, The Castlegate Sandstone of the Book Cliffs, Utah:
sequence stratigraphy, paleogeography, and tectonic controls: Journal of Sedimentary
Research, v. 71, p. 53548.
OBRIEN, N.R., AND SLATT, R.M., 1990, Argillaceous Rock Atlas: New York, SpringerVerlag, 141 p.
OBYRNE, C.J., AND FLINT, S.S., 1995, Sequence, parasequence and intra-parasequence
architecture of the Grassy Member, Blackhawk Formation, Book Cliffs, Utah, USA,
in Van Wagoner, J.C., and Bertram, G.T., eds., Sequence Stratigraphy of Foreland
Basin Deposits: Outcrop and Subsurface Examples from the Cretaceous of North
America: American Association of Petroleum Geologists, Memoir 64, p. 225255.
OBYRNE, C.J., AND FLINT, S.S., 1996, Interfluve sequence boundaries from the Grassy
Member, Book Cliffs, Utah: criteria for recognition and implications for subsurface
correlation, in Aitken, J.F., and Howell, J.A., eds., High Resolution Sequence
Stratigraphy: Innovations and Applications: Geological Society of London, Special
Publication 104, p. 208220.
PATTISON, S.A.J., 1995, Sequence stratigraphic significance of sharp-based lowstand
shoreface deposits, Kenilworth Member, Book Cliffs, Utah: American Association of
Petroleum Geologists, Bulletin, v. 79, p. 444462.

339

POTTER, P.E., MAYNARD, J.B., AND PRYOR, W.A., 1980, Sedimentology of Shale: New
York, Springer-Verlag, 306 p.
POTTER, P.E., MAYNARD, J.B., AND DEPETRIS, P.J., 2005, Mud and mudstones,
Introduction and Overview: New York, Springer, 297 p.
RAISWELL, R., 1987, Non-steady state microbial diagenesis and the origins of concretions
and nodular limestones, in Marshall, J.D., ed., Diagenesis of Sedimentary Sequences:
Geological Society of London, Special Publication 36, p. 4154.
RAISWELL, R., 1988, Chemical model for the origin of minor limestoneshale cycles by
anaerobic methane oxidation: Geology, v. 16, p. 641644.
ROHL, H.-J., AND SCHMID-ROHL, A., 2005, Lower Toarcian (Upper Liassic) black shales
of the Central European epicontinental basin: a sequence stratigraphic case study
from the SW German Posidonia Shale, in Harris, N., ed., The Deposition of OrganicCarbon-Rich Sediments: Models, Mechanisms, and Consequences: SEPM, Special
Publication 82, p. 165189.
ROHL, H.-J., SCHMID-ROHL, A., OSCHMANN, W., FRIMMEL, A., AND SCHWARK, L., 2001,
The Posidonia Shale (Lower Toarcian) of SW-Germany: an oxygen-depleted
ecosystem controlled by sea level and palaeoclimate: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 165, p. 2752.
SCHIEBER, J., 2003, Simple gifts and buried treasuresimplications of finding
bioturbation and erosion surfaces in black shales: SEPM, The Sedimentary Record,
v. 1, p. 48.
TAYLOR, K.G., AND CURTIS, C.D., 1995, Stability and facies association of early
diagenetic mineral assemblages: an example from a Jurassic ironstone--mudstone
succession, U.K.: Journal of Sedimentary Research, v. 65, p. 358368.
TAYLOR, K.G., AND GAWTHORPE, R.L., 2003, Basin-scale dolomite cementation of
shoreface sandstones in response to sea-level fall: Geological Society of America,
Bulletin, v. 115, p. 12181229.
TAYLOR, K.G., AND MACQUAKER, J.H.S., 2000, Spatial and temporal distribution of
authigenic minerals in continental shelf sediments: implications for sequence
stratigraphic analysis, in Glenn, C., et al., eds., Marine Authigenesis: Microbial to
Global: SEPM, Special Publication 66, p. 309323.
TAYLOR, K.G., GAWTHORPE, R.L., CURTIS, C.D., MARSHALL, J.D., AND AWWILLER, D.N.,
2000, Carbonate cementation in a sequence-stratigraphic framework: Upper
Cretaceous sandstones, Book Cliffs, Utah--Colorado: Journal of Sedimentary
Research, v. 70, p. 360372.
TAYLOR, K.G., SIMO, J.A., YOCUM, D., AND LECKIE, D.A., 2002, Stratigraphic
significance of ooidal ironstones from the Cretaceous Western Interior Seaway: the
Peace River Formation, Alberta, Canada, and the Castlegate Sandstone, Utah,
U.S.A.: Journal of Sedimentary Research, v. 72, p. 316327.
TAYLOR, K.G., GAWTHORPE, R.L., AND FANNON-HOWELL, S.F., 2004, Basin-scale
diagenetic alteration of shoreface sandstones in the Upper Cretaceous Spring Canyon
and Aberdeen members, Blackhawk Formation, Book Cliffs, Utah: Sedimentary
Geology, v. 172, p. 99115.
TYSON, R.V., 1995, Sedimentary Organic Matter: Organic Facies and Palynofacies:
London, Chapman and Hall, 615 p.
TYSON, R.V., 2005, The productivity versus preservation controversy: cause, flaws,
and resolution, in Harris, N., ed., The Deposition of Organic-Carbon-Rich Sediments:
Models, Mechanisms, and Consequences: SEPM, Special Publication 82, p. 1733.
VAN WAGONER, J.C., 1995, Sequence stratigraphy and marine to nonmarine facies
architecture of foreland basin strata, Book Cliffs, Utah, USA, in Van Wagoner, J.C.,
and Bertram, G.D., eds., Sequence Stratigraphy of Foreland Basin Deposits: Outcrop
and Subsurface Examples from the Cretaceous of North America: American
Association of Petroleum Geologists, Memoir 64, p. 137223.
VAN WAGONER, J.C., MITCHUM, R.M., CAMPION, K.M., AND RAHMANIAN, V.D., 1990,
Siliciclastic Sequence Stratigraphy in Well Logs, Cores, and Outcrops: Concepts for
High Resolution Correlation of Time and Facies: American Association of Petroleum
Geologists, Methods in Exploration Series 7, 55 p.
WILLIAMS, C.J., HESSELBO, S.P., JENKYNS, H.C., AND MORGANS-BELL, H.S., 2001, Quartz
silt in mudrocks as a key to sequence stratigraphy (Kimmeridge Clay Formation, Late
Jurassic, Wessex Basin, UK): Terra Nova, v. 13, 449 p.
YOSHIDA, S., 2000, Sequences and facies architecture of the upper Blackhawk Formation
and the Lower Castlegate Sandstone (Upper Cretaceous), Book Cliffs, Utah, U.S.A.:
Sedimentary Geology, v. 136, p. 239276.
Received 18 December 2005; accepted 30 September 2006.

S-ar putea să vă placă și