ArticlePDF Available

Depositional constraints on the Lower Cretaceous stikes quarry dinosaur site: Upper yellow cat member, cedar mountain formation, Utah

Authors:

Abstract

A new mass death assemblage in Lower Cretaceous strata of east-central Utah contains well-preserved skeletons representing an ontogenetic series of individuals of Utahraptor, and at minimum two iguanodont grade ornithischian skeletons. The dinosaurs were entombed in ovoid-lensoidal, fine-grained sandstone sills linked by sandstone pipes and/or dikes and another basal lensoidal mass with scattered and broken iguanodont and sauropod bones and to an underlying gravelly sandstone bed. Exposed in the excavation high-walls are syndepositional normal-faults bounding graded ripple strata. Multiphased fluid over-pressurization in an artesian setting creating the structures. Trapping, killing, and subsequent burial mechanism was generated by variations of pressure in a localized artesian spring system that breached the surface and is the first such mechanism documented with numerous dinosaur victims.
1
RRH: DEPOSITIONAL CONSTRAINTS STIKES QUARRY
LRH: J.I. KIRKLAND ET AL.
Research Article
DOI: http://dx.doi.org/10.2110/palo.2016.041
DEPOSITIONAL CONSTRAINTS ON THE LOWER CRETACEOUS STIKES QUARRY
DINOSAUR SITE: UPPER YELLOW CAT MEMBER, CEDAR MOUNTAIN FORMATION,
UTAH
JAMES I. KIRKLAND,1 EDWARD L. SIMPSON,2 DONALD D. DeBLIEUX,1 SCOTT K.
MADSEN,1 EMILY BOGNER,2 AND NEIL E. TIBERT3
1Utah Geological Survey, PO Box 146100, Salt Lake City, Utah, 84114-6100 USA
2Deparment of Physical Sciences, Kutztown University, Kutztown, Pennsylvania 19530 USA
3Department of Earth and Environmental Sciences, University of Mary Washington,
Fredericksburg, Virginia 22401 (deceased)
Abstract: A new mass death assemblage in Lower Cretaceous strata of east-central Utah contains
well-preserved skeletons representing an ontogenetic series of individuals of Utahraptor, and at
minimum two iguanodont grade ornithischian skeletons. The dinosaurs were entombed in ovoid-
lensoidal, fine-grained sandstone sills linked by sandstone pipes and/or dikes and another basal
lensoidal mass with scattered and broken iguanodont and sauropod bones and to an underlying
gravelly sandstone bed. Exposed in the excavation high-walls are syndepositional normal-faults
bounding graded ripple strata. Multiphased fluid over-pressurization in an artesian setting
creating the structures. Trapping, killing, and subsequent burial mechanism was generated by
variations of pressure in a localized artesian spring system that breached the surface and is the
first such mechanism documented with numerous dinosaur victims.
INTRODUCTION
Taphonomic processes affecting modern vertebrate carcasses have been documented in a
variety of continental settings (Weigelt 1989; Behrensmeyer 1991; Rogers et al. 2007). For
example, the recognition of the taphonomic biases introduced by a mud miring mechanism has
permitted unraveling the population structure in a mass-accumulation of small herbivorous,
theropod dinosaurs (Varricchio et al. 2008).
The Stikes Quarry is a recently discovered concentration of a tightly packed Utahraptor
and iguanodontid skeletons in 125 million year old Lower Cretaceous strata (north of Arches
National Park) in east-central Utah. The quarry is within the upper Yellow Cat Member of the
Cedar Mountain Formation (Fig. 1). This study examines the soft-sediment deformation
structures and strata entombing the Utahraptors and iguanodont grade ornithischians (hereafter
referred to as iguanodonts) proposing the hypothesis that a localized artesian system with a
hydrologic head powerful enough to cause quicksand development, trapping, and the unique
burial of at least eight Utahraptor dromaeosaurids (Fig. 1). This proposed hypothesis will be
tested during the preparation of the large Utahraptor block through more detailed analysis of the
spatial arrangement of the bones and taphonomic data.
GEOLOGICAL SETTING
The Early Cretaceous Cedar Mountain Formation with discrete terrestrial faunas is one of
the most important sources of Lower Cretaceous dinosaurs (Kirkland et al. 1997, 1999, 2015;
Cifelli et al. 1999; Kirkland and Madsen 2007, Kirkland and Farlow 2012). In the stratigraphic
section that includes the Stikes Quarry, the Cedar Mountain Formation unconformably overlies
the Morrison Formation with a hiatus of approximately 20 Ma (Kirkland and Madsen 2007;
2
Hendrix et al. 2015) and in turn is overlain by the Naturita Formation (Young 1960, 1965;
Carpenter 2014).
In the northern Paradox Basin of east-central Utah, the Cedar Mountain Formation is
subdivided into members, from oldest to youngest, Yellow Cat, Poison Strip, and Ruby Ranch
(Kirkland et al. 1997, 1999; Sprinkle et al. 2012; Doelling and Kuehne 2013; Fig. 1). Well-
developed paleosols, spring deposits, lacustrine limestones, mudstones, siltstones, and
sandstones linked to northeast-directed fluvial systems characterize the Yellow Cat Member
(Kirkland et al. 1997; DiCroce and Carpenter 2001; Greenhalgh and Britt 2007; Kirkland and
Madsen 2007; Suarez et al. 2007a, 2007b). The Yellow Cat Member is subdivided into two
informal units. At the Stikes Quarry section the lowest unit is ~ 29 m thick and consists of
mainly stacked paleosols. The upper unit is ~ 14 m thick and dominantly a lacustrine system.
The basal boundary of the upper Yellow Cat Member is placed at the top of a regional bench-
forming calcrete paleosol (Figs. 1, 2; Kirkland and Madsen 2007; McDonald et al. 2010;
Kirkland et al. 2012) that once defined the base of the Cedar Mountain Formation (Aubrey
1998). U-Pb single crystal detrital zircon analysis of the upper Yellow Cat Member yields a
maximum age of approximately latest Barremian to basal Aptian age, 124±2.5 Ma (Greenhalgh
et al. 2006; Britt et al. 2009; Ludvigson et al. 2010), whereas other studies based on detrital
zircons and microfossils suggest considerably older Early Cretaceous dates (Sames et al. 2010;
Sames 2011a, 2011b; Martin-Closas et al. 2013; Hendrix et al. 2015).
Yellow Cat strata were deposited under arid to semi-arid conditions with monsoonal
overprinting (Kirkland et al. 1999; Kirkland and Madsen 2007; Suarez et al. 2014). At the Stikes
locality, the overlying 15 m thick Poison Strip Sandstone preserves northeast-directed, low- to
moderate-sinuosity channels with lateral bars characteristic of distal braided stream systems that
overstepped the lacustrine shoreline sandstones of the upper Yellow Cat Member (Stikes 2007;
Lockley et al. 2015). The Ruby Ranch Member consists of northeast-directed, fluvial “ribbon”
sandstones with overbank deposits containing abundant carbonate nodules and is ~ 30 m thick
across the northern Paradox Basin. Along the west side of Arches National Park, the thickness
more than doubles with a well-developed lacustrine system forming the upper half of the
member (Harris 1980; Kirkland et al. 1997; Kirkland and Madsen 2007; Williams et al. 2009;
Ludvigson et al. 2010; Montgomery 2014).
PALEONTOLOGICAL SETTING
The upper Yellow Cat Member preserves a distinctive fauna dominated by the
polacanthid ankylosaur Gastonia, diverse iguanodontid including Hippodraco, and several
sauropod species including the brachiosaurid Cedarosaurus and a basal macronaran (Tidwell et
al. 1999; Kirkland and Madsen 2007; Britt et al. 2009; McDonald et al. 2010). Utahraptor
ostrommaysorum is the most common theropod (Kirkland et al. 1993; Britt et al. 2009) along
with other small theropods. Additional faunal elements include diverse bony fish, hybodont
sharks, lungfish, turtles, a choristothere, sphenodontids, crocodilians, and a mammal (Kirkland
and Madsen 2007; Britt et al. 2009). Vertebrate trace fossils from the upper Yellow Cat include
the oldest bird tracks in North America (Lockley et al. 2015).
FACIES
The Stikes Quarry is in the upper Yellow Cat Member. Six facies are exposed in the
quarry highwall and laterally along the newly constructed haul road (Fig. 1D, 2): (1) dark gray
shale; (2) lower rippled sandstone; (3) heterolithic red mudstone; (4) bone-bearing lensoidal-
ovoid-shaped sandstone masses; (5) rippled sandstone and mudstone; and (6) upper rippled
sandstone.
3
Dark Gray Shale
Description.—Thin laminations of siltstone are interbedded with the predominant dark
gray shale (N2 to N3). Fish scales and poorly preserved gastropods are present (Fig. 1D). In
addition, microfossils include an assemblage of four nonmarine ostracode species (Fig. 3). The
most common taxon from the Stikes Quarry is Cypridea Bosquet (Sames 2011a, 2011b). In
lesser abundance are well-preserved, small and more ornate specimens of C. (Longispinella)
longispina, and Mongolianella stirlingensis (Loranger 1951).
Interpretation.—This facies developed from tractional deposition followed by suspension
settling in an offshore lacustrine setting (Smoot 1991; Smoot and Olsen 1994; Arenas and Pardo
1999). The ostracode Cypridea confirms this interpretation as a lacustrine setting because this
taxon is regarded as tolerant of drought prone conditions (Horne 2002). Therefore, the
predominance of Cypridea is a strong indication of ephemeral ponding in a lacustrine system
during times of increased aridity in the Stikes Quarry strata (Carbonel et al. 1988; Horne 2002;
Horne and Colin 2005; Trabelsi et al. 2011).
Lower Rippled Sandstone
Description.—The lower rippled sandstone forms the floor of the Stikes Quarry (Figs.
1D, 4, 5). Medium- to fine-grained sandstone is yellowish gray in color (5Y 8/1), ~ 2 m thick,
and laterally continuous for over 200 m. Chert granules are present on horizontal partings and
form laminations throughout this facies. Low-angle, cross beds are present. Rare, green-gray
siltstone and mudstone interlaminations separate beds.
Soft-sediment deformation structures are present throughout the quarry and the haul road.
A broad dome-like structure is centered in the quarry and preserved on the upper bedding plane
of the lower sandstone (Fig. 5A). Internally, the dome has upward arching laminations that are
near parallel to the upper bedding surface (Fig. 5A). Exposed along the haul road, 30 m to the
east of the Stikes Quarry are thin (< 10 cm wide) flame structures within the facies (Fig. 5B),
fluidization, and load features (sensu Lowe, 1975). Connected to the lower rippled sandstone and
crosscutting the overlying heterolithic mudstone facies is a small, 40 cm tall vertical pipe
composed of fine-grained sandstone (Fig. 6).
Interpretation.—The lower rippled sandstone facies in conjunction with the underlying
dark gray shale facies is best interpreted as a prograding clastic lacustrine shoreline deposit
(Fraser and Hester 1977; Bracken 1994; Ilgar and Nemec 2005). Fraser and Hester (1977) report
similar type of sandstone with coarse-grained laminations and wavy horizontal laminations in the
shoreface settings in modern Lake Erie.
The soft-sediment deformation features in this facies are consistent with overpressuring
and fluidized deformation that generate convoluted bedding and cause vertical migration of over-
pressured fluids (Lowe 1975; Allen 1982, 1986; Mills 1983; Maltman 1984; Owen 1987, 1996;
Nichols et al. 1994; Frey et al. 2009). Not all primary structures in the lower sandstone were
obliterated by the loss of internal shear strength and mobilization (Surlyk et al. 2007).
Heterolithic Red Mudstone
Description.—The heterolithic red mudstone facies groups together red sandy mudstones,
red siltstones, red fine-grained sandstones, gray mudstones and gray carbonate nodules. Color
varies from dark gray (N2) to dusky red (5R 3/4). This facies grades from siltstone rich beds at
the base to mudstone dominated at the top (Fig. 7). Gray mudstone is more prevalent on the west
quarry wall and adjacent to the sandstone lenses elsewhere in the facies is dusky red. Carbonate
nodules varying from 10–20 cm and are concentrated in the lower portion of the facies, near the
level of the Utahraptor block.
4
The lens-shaped sandstones preserving the Utahraptor and iguanodont fossils are
encapsulated by this facies (Fig. 4). The heterolithic mudstone and sandy red and gray mudstone
facies together with the underlying and upper rippled sandstone facies extend laterally for more
than 500 m (Fig. 4C). Along the haul road the thickness of the heterolithic mudstone facies
varies from 3.6 to 2.9 m, an approximate 20% reduction in thickness at the center of the quarry
(Fig. 1D). At the eastward termination of the haul road a pebbly mudstone dike cross cuts the
lower siltstones and originates in this facies (Fig. 7).
Interpretation.—This facies is best interpreted as a flood plain deposit (Stear 1978; Miall
1996; Bridge 2003; Ghosh et al. 2006) with possible pedogenic carbonates (Suarez et al. 2014).
Paleosol features appear absent from the quarry area although slickensides are reported by Poole
(2008).
Within a few meters of the surface the permeability of the lower rippled sandstone and
the lower strata of this facies mainly siltstones, are close to equal. These specific conditions
induced fluidization of the lower rippled sandstone and the overlying host sediment together as
recorded in the pebbly mudstone dikes in this facies (see Jonk (2010) for a parallel interpretation
for a similar setting)
Bone-Bearing Lensoidal-Shaped Sandstone
Description.—In the field, four ovoid lens-shaped sandstone masses with bone were
discovered and given identifiers: Poole, iguanodont, Utahraptor, and iguanodont-sauropod (Fig.
8). Sandstones and siltstones are very light gray (N8) to light greenish gray (5GY 8/1).
Mudstones clasts are dusky green (10GY 3/2) to yellowish gray (5Y 8/1) and up to 4 cm.
Granules and pebbles are present. These sandstones are in sharp contact with the surrounding
heterolithic mudstone facies.
The Poole block was the first encountered during east-side excavation of the upper strata
of the site and was removed in 2005 (Figs. 2, 8). Little was recorded concerning the margins of
this block (Poole 2008). Grain-size difference caused separation at its base from the underlying
Utahraptor block. When prepared, the Poole block yielded a partial juvenile iguanodont skeleton
with some articulated elements and diagnostic elements of an adult Utahraptor (Fig. 9; Poole
2008). Small-scale inclined sandstones were observed in the Poole block.
The ~ 500 kg iguanodont block on the west side contains parts of a large iguanodont and
at least one baby and one juvenile Utahraptor. A thin clastic dike connects this block to the
underlying main Utahraptor block (Fig. 8). The underside of the iguanodont block has a sharp,
irregular vertical contact between sandstone and gray shale of the heterolithic mudstone facies
(Fig. 10A, 10D). Slickenlines are apparent on this vertical contact. Exposed in the block is a
large iguanodont limb bone that extends through the connecting dike into the Utahraptor block
(Fig. 10B).
The plaster field jacket dimensions of the Utahraptor block are 3.2 m by 2.5 m with a
thickness of 0.50 to 0.85 m (Fig. 4). During the excavation of this lensoidal block, the
surrounding unfossiliferous heterolithic mudstone and sandy red and gray mudstone facies was
excavated and its limited extent documented. The absence of the fossiliferous lens from the
excavation walls indicates that the block was not a sheet-form or channel-form geometry but
lensoidal (Figs. 2, 4, 8). The southern limit of the Utahraptor-bearing sandstone mass is the
modern erosion surface.
The Utahraptor lens contains a spectrum of different sized Utahraptor skeletons, a
minimum of two babies (skull length ~ 10 cm), four juveniles, and one adult based on material
exposed at the mass margins and at the top of the block during excavation while reducing block
5
dimensions and placing of the plaster jackets (Fig. 9; Kirkland et al. 2011). Teeth have not
shifted in the aveoli of the 10 currently known jaws from the site (Fig. 9). Several articulated
skeletons have been identified on the exposed portions of the block.
A near-vertical pipe composed of fine-grained sandstone connects the large Utahraptor
block to the iguanodont-sauropod block approximately 75 cm below and to the west that
preserves disarticulated and broken iguanodont and sauropod bones along with an isolated
maxilla of a juvenile Utahraptor (Figs. 2, 4, 8, 11A). The iguanodont-sauropod block was only
partially excavated, covered, placed in a plaster jacket, and was reburied in the quarry for future
research.
The pipe is approximately 0.5 m in diameter and connects to the underside of main
Utahraptor block at its northwestern margin (Fig. 11A). The pipe margins are sharp with
elongate green to light tan mudstone clasts aligned parallel to the long axis and margins of the
pipe (Fig. 11B). The longest clast axis is ~ 4 cm. Small gray and green mudstone clasts are
elongate to rounded and present throughout the interior of the pipe and in the blocks, along with
scattered 3–5 mm isolated chert granules (Fig. 11B, 11C). Chert granules in the pipe are similar
in composition and grain size to those constructing the horizontal partings in the lower rippled-
sandstone facies. The pipe and blocks contain angular fragments of bone (Fig. 11B). Within the
pipe, there are at least two generations of clasts, the more common gray green mudstone rip up
clasts, and medium gray clasts that host gray green rip up clasts (Fig. 11C). The internal clasts in
the gray rip-up clasts have different orientations than the surrounding clasts (Fig. 11B, 11C).
Interpretation.—The geometry of the pipe and dike connecting sill-like structures is
consistent with over pressuring and vertical mobilization of sediment. The relative structureless
nature and grain size of the dike and pipe fill indicates over-pressuring resulting in fluidization of
the underlying lower rippled sandstone and forceful injection of sediment-fluid mixture via a
conduit into the overlying strata (Daley 1971; Owen 1987; Cosgrove 2001; Draganits et al.
2003). The sand-fluid moved vertically following the dropping lithostatic pressure and when
reaching the surface produced sand volcanoes (see Obermeier 1996, 1998). Cylindrical
fluidization pipes are emplaced in shallow depths from 0.5 to 20 m and have diameters
commonly less than 0.5 m, but may be up to 1.5 m (see Gill and Kuenen 1957; Seed 1979;
Walsh et al. 1995; Obermeier 1996; Gallo and Woods 2004). Emplacement processes
incorporated mudstones from the lower rippled sandstone and the heterolithic red mudstones.
Dikes and pipes were the locus of multiple generations of flow as evidenced by the internal
rotated clasts.
Jolly and Lonergan (2002) demonstrated that the more irregular-shaped the sandstone
body, dikes, and sills are the shallower the emplacement. This is exemplified by the irregular
nature of the fossiliferous sills. The intrusions were exposed to surface processes incorporating
vertebrate fossils and producing the inclined tractional structures present in the Poole block.
Bone fragments attest to the turbulent and abrasive nature of fluid flow that generated the
sandstone bodies.
Rippled Sandstone and Mudstone
Description.—The rippled sandstone and mudstone facies is bound by normal faults and
is preserved only in the associated grabens or footwalls. Fine- to very fine-grained sandstones
and siltstones are very light gray (N8) to light greenish gray (5GY 8/1). Mudstones are dusky
green (10GY 3/2). Graded beds range from 2–3 mm to 2–3 cm in thickness. Sandstones have
loaded to irregular bases overlain by graded beds. Dusky green mudstone rip up chips from 1–2
mm to 3 cm occur at the bases.
6
In all the grabens, this facies is deformed into larger-scale synclinal folds that increase in
dip towards the syndepositional fault plane (Fig. 12B). An additional graben filled rippled
sandstone and mudstone facies occurs 15 m from the center of the quarry (Fig. 6). Soft-sediment
fault structures extend to the base of the overlying rippled-sandstone. Normal and reverse slip
faults form graben-like structures (Figs. 2, 4B, 12). The fault planes strike approximately NE-
SW and have steep dips that flatten with depth near the stratigraphic level of the Utahraptor
mass. On the northern and southern ends of the walls are normal faults that have slickensides
indicating near-vertical down-dip movement (Fig. 12). Displacement on the faults is
approximately 1.0 m or less.
Sediment deformation in this facies increases towards the syndepositional faults,
changing from near pristine preservation of the graded beds to complexly folded to homogenized
strata directly adjacent to the faults (Fig. 13). Because fluidization overprinted the ripple and
mudstone facies, no vertical or horizontal trend in thickness is apparent (Fig. 13).
Interpretation.—The syndepositional faults are the result of brittle behavior of the red
mudstone facies. Variation in fault dip is related to the change in rheological properties with
depth (see Price 1966; Davison 1987; Ferrill and Morris 2003). Brittle behavior of sediment is
initiated by a variety of mechanisms including sediment loading, slope failure and seismic
activity (Seilacher 1969; Rider 1978; Field et al. 1982; Grimm and Orange 1997; Strachan 2002).
Emplacement of clastic intrusions have been linked to both polygonal and ring faulting
(Cartwright et al. 2007; Shoulders et al. 2007; Cartwright 2010; Ross et al. 2014).
The restricted nature of the rippled sandstone and mudstone facies to the footwall of the
faults indicates an original position stratigraphically higher. The down-dip movement preserved
this facies while stratigraphically higher remnants were eroded because of the absence in the
hanging wall. Fault slip initiated by sediment fluidization in the hanging wall is evidenced by
increasing soft sediment deformation intensity towards the fault. No apparent growth strata are
developed in the hanging wall sediment.
Graded beds and associated sedimentary structures represent multiple episodes of
decreasing current velocity (Harms et al. 1982; Boguchwal and Southard 1990; Southard and
Boguchwal 1990). Variation in fluid expulsion rates produce graded beds separated by fine-
grained sediment in sand volcanoes (Quigley et al., 2013; Rodríguez-Pascua et al. 2015). In
addition, multiple reactivation of small-scale sand blows occurs in areas experiencing successive
earthquakes (Quigley et al. 2013) or may be related to local unstable topography related to
seismic disturbance (Rodríguez-Pascua et al. 2015). Reactivation or the longer-term exploitation
of soft-sediment deformation conduits by a later event or continuous gas generation and fluid
migration is common (Hurst et al. 2003; Perez-Garcia et al. 2009; Quigley et al. 2013; Bastin et
al. 2015). Quigley et al. (2013) recognized significant subaerial erosion of the features including
vent degradation, rill development, erosion of fine-grained layers, and development of ripples
and microdunes.
Upper Rippled Sandstone
Description.—The upper rippled sandstone is laterally continuous and was traced across
the outcrop for about a kilometer (Figs. 1D, 4D, 15). The medium-grained sandstone has floating
granules and is light gray (N8). Dusky green mudstone (10GY 3/2) ripped chips are present.
Thickness ranges from 0.5 to 1.2 m. Climbing ripple cross laminations, small-scale, trough cross
bedding and parallel laminations are present. This facies exhibits better sorting than all the
others. Petrographically, the lower sandstone and lensoidal bodies are similar to this facies (Fig.
14). The lower sandstone is less sorted than the lensiodal bodies with the upper sandstone having
7
coarse to very fine sands (Fig. 15) and laminae of 2–5 mm angular chert granules that form
horizontal partings.
Interpretation.—The overall fining and thinning upward of the deposit coupled with the
suite of sedimentary structures indicates this bed is the product of waning flows either in a
crevasse splay (see Bristow et al. 1999), sheet-flood (Tooth 1999; Blair 2000; Fisher et al. 2007)
or in the context of the overlying lacustrine beds, a transgressive lacustrine unit (Fraser and
Hester 1977; Bracken 1994; Ilgar and Nemec 2005). This facies was not disturbed by the older
fluidization events.
DISCUSSION—THE STRATIGRAPHIC AND SEDIMENTOLOGIC MODEL
The associated sedimentary facies, synsedimentary faults, suite of soft-sediment
deformation structures, lensiodal bone-bearing blocks, connected by a pipe and dike, coupled
with the near pristine fossil preservation (i.e., preservation of delicate elements and partial
articulation) in the Stikes Quarry eliminates most commonly documented taphonomic
mechanisms for forming bone beds. Facies, facies relationships, and soft-sediment deformation
structures exposed in the Stikes Quarry point to a localized artesian spring system source capable
of generating quicksand in the crater (Fig. 16). This hypothesis will be further constrained and
tested during the preparation of the large Utahraptor block. No matter what the additional data
collection yields, any new hypothesis must incorporate the association of soft-sediment
deformation structures and stratigraphic relationships outlined here.
The generation of soft-sediment deformation structures is often attributed to seismic
activity (Sims 1973, 1975; Allen 1982, 1986; Owen 1987; Pratt 1994; Obermeier 1996; Jones
and Omoto 2000; Moretti 2000; Moretti and Sabato 2007; Montenat et al., 2007; Spalluto et al.
2007; Reicherter et al. 2009; El Taki and Pratt 2012; Loope et al. 2013; Törő and Pratt 2015;
Wizevich et al. 2016). Consensus has centered on a number of criteria that, when recognized in
strata, indicates a high probability that the soft-sediment deformation structures are seismogenic
in origin (seismites sensu Seilacher 1969; Wheeler 2002). These key criteria include: (1)
association of soft-sediment deformed structures with a potential consanguineous fault; (2)
deformation consistent with seismicity supported by presence of field and experimental analogs;
(3) widespread synchronous soft-sediment deformed structures in a bed or layer; (4) spatial
variation in intensity towards the consanguineous fault; (5) elimination of other possible causal
mechanisms; (6) recurrence over time, in the same facies; (7) bracketed by undisturbed
sediments; (8) associated colluvial breccias, conglomerates and massive sandstones; and (9) soft-
sediment deformed structures that cross cut regional and facies boundaries (Sims 1973, 1975;
Obermeier 1996; Rossetti 1999; Jones and Omoto 2000; Ettensohn et al. 2002; Montenat et al.
2007; Moretti and Sabato 2007; Moretti and Sabato 2007; Reichter et al. 2009; Van Loon 2009;
Audemard and Michetti 2011; Hilbert-Wolf and Roberts 2015).
Movement on salt structures in the Arches National Monument during the development
of the Cedar Mountain Formation influenced its overall geometry (Kirkland and Madsen 2007;
Doelling 2010). The Stikes Quarry soft-sediment deformation however is limited to the
immediate vicinity of the bone bed and is not linked with a specific fault and therefore, does not
satisfy enough of the criteria outlined above, making a seismic origin extremely unlikely. In the
absence of these key seismic criteria an alternate mechanism to cause over pressuring, associated
soft-sediment deformation structures and the vertical migration of sand-size material is required.
Localized artesian systems have been documented to breech surficial deposits in fluvial,
delta, and shoreline systems and produce sedimentary features similar to those recorded in the
Stikes Quarry (Guhman and Pederson 1992; Younger and McHugh 1995; Massari et al. 2001;
8
Heubeck 2009; Fig. 16A). In addition to water, crater-like structures on subaerial sediment
surfaces can develop from a combination of gas, water and sediment expulsion (Fig. 16B;
Maxson 1940; Jianhua et al. 2004; Netoff et al. 2010; Livingston et al. 2014; Miller et al. 2015;
Sherrod et al. 2016).
We propose the following sequence of events for the Stikes Quarry sedimentologic and
stratigraphic setting (Fig. 17): (1) Development of an artesian setting: the lower sandstone (the
aquifer), sealed by the underlying dark gray shale and overlying heterolithic mudstone,
(aquitards), and elevation generated either by local elevation variation, halokinesis, lowering of
base level linked surrounding lake levels, or a combination (Fig. 1E); (2) Up-arching of the
heterolithic mudstone as pressure built initiating mobilization and fluidization of sediment, the
lower rippled sandstone; (3) Intrusion sediment-rich fluid from the lower rippled sandstone into
the overlying heterolithic mudstone and breaching; (4) Elutriation of fines from the slurry to
form mud-sand volcanoes at the surface breech; (5) Construction of a mud-sand volcano with
unsteady flow creating the vertically stacked waning flow deposits; (6) Formation of over-
pressured water-filled crater with sand left by elutriation of fines recorded in the iguanodont-
sauropod block/sill that is an attritional bone accumulation similar to the Cleveland-Lloyd
Quarry (Gates 2005) and the Mygatt-Moore Quarry (Kirkland and Armstrong 1992; Kirkland et
al. 2005); (7) Multiple generations of horizontal sands at the base of the crater separated by
partial crater infill, Utahraptor block followed by either the iguanodont or the Poole blocks; tops
of sand intrusions are exposed subaqueously or subaerially for an extended period of time; (8)
Deformation of crater walls by normal faulting preserved the rippled sandstone and mudstone
facies on the footwall block—the relative timing of the deformation is linked the formation of
crater sands so which phase is uncertain; and (9) Destruction of elevated portion of the crater by
erosion, sediment infilling the crater and subsequent burial by the upper rippled sandstone.
Documented examples of artesian systems producing soft-sediment deformation
structures similar to the Stikes Quarry include: (1) the modern Dismal River in Nebraska (Fig.
16A); (2) Holocene mire in England; and (3) modern crater features on the Lake Powell delta
(Fig. 16B).
Guhman and Pederson (1992) recognized near vertical, large conduits that parallel the
Dismal River in the Sand Hills of Nebraska (Fig. 16A). A drop in base level caused rapid
incision of the Dismal River into the underlying confined aquifer releasing pressure by
producing the large diameter pipes. The conduit craters contain roiling water with suspended fine
sands creating spatially variable quicksand conditions. A significant pressure gradient caused the
robust vertical flow in the artesian spring system even though the surrounding area’s elevation is
very limited. These spring-generated craters are up to 10 m in diameter with estimated depths of
~ 44 m with a water level up to 3 m below the rim level. Over time, the flow rate and surface
area of vertical flow diminished (Fig. 16A). Flow rates in the Dismal River craters are non-
steady and subject to variations in air pressure and water recharge. During low-flow rates the
sands on the periphery of the pipe reestablish grain-to-grain contact and are capable of
supporting weight; variations in flow rate changes the width of this marginal firm sand band, but
quicksand conditions were still present away from the margins (D.T. Pederson personal
communication 2015). In the Stikes Quarry the Utahraptor block is the largest and therefore
reflects the greatest magnitude and probable longest pressure release event. The iguanodont and
Poole sills record resurgent events.
On a smaller scale, Younger and McHugh (1995) illustrate both conical and cylindrical
sand bodies preserved in a Holocene mire in East Yorkshire, UK. They attribute the injected
9
sand bodies to the establishment of vigorous artesian springs. Structureless sands and preserved
trees record the development and onset of quicksand conditions caused by the over pressuring of
the artesian system. These Holocene sand bodies have similar geometry to pipes and sandstone
lensoidal sills present in the Stikes Quarry. The lateral extent of the known soft-sediment vertical
disruption in the Stikes Quarry is less than 50 m and therefore is comparable to both the boiling
springs and this mire size. The 3-m depth from the youngest disrupted surface to the Utahraptor
lensoidal bone bed comparable to the Holocene mire dimensions but not to the Dismal River
boiling sand pipe depth of 44 m.
At the modern Hite delta formed by the Colorado River entering Lake Powell, Netoff et
al. (2010) recognized a localized artesian system charged additionally with locally generated
methane that deformed the confining muddy delta top sediments into dome-like features that
initially developed subaqueously (Fig. 16B). As water levels in Lake Powell rapidly diminished,
the domes were breeched followed by the development of craters and mud volcanoes. These
fluid expulsion features continued to discharge fluid and methane for more than seven months
(Netoff et al. 2010). After a period of time the associated sediment surface collapsed creating a
crater that had a central vent (Fig.16; Livingston et al. 2014, 2015; Sherrod et al. 2016). These
craters develop ponds with rising lake levels that are exploited by flora and fauna (Miller et al.
2015; Sherrod et al. 2016). Sherrod et al. (2016) used geophysical techniques to demonstrate the
change in thickness in the overlying confining sediment and the presence of pipe-like features.
Some breeched the overlying sediments whereas others did not. These soft-sediment deformation
structures are akin to those observed in the Stikes Quarry.
Additional insights into the trapping mechanism of the Stikes Quarry may be gleaned
from the Pleistocene Rancho La Brea tar deposits (Woodard and Marcus 1973). The Ranch La
Brea deposit was thought to trap animals in erupting viscous hydrocarbons followed by the
luring of predators to the site, thereby producing a predator trap (Merriam 1911; Stock 1965).
The predator trap was then subjected to convective overturn by rising hydrocarbons that created
“pit wear” on bone. Woodard and Marcus (1973) examined the historical records of excavation,
internal stratigraphy, and drill hole data proposing that the fill formed by fluvial incursion into
pits or craters delivering a diverse suite of transported, single vertebrate elements to the pit. This
infilling model has crude vertical stratigraphy. More recent taphonomic examination has
reinvigorated the tar-based predator trap model based on the absence of a good definable internal
stratigraphy and mixed ages of the isolated single element bones (Akersten el al. 1983; Spencer
et al. 2003; Friscia et al. 2008). The disarticulated nature and pit wear on the bones requires that
the tar pit taphonomy be unique, and may be the result of increased bone compaction through
time (Friscia et al. 2008). The Stikes Quarry trapping mechanism must be different because of
articulated remains.
Miring in mud is a well-recognized mechanism for killing dinosaurs (Hungerbühler 1998;
Varricchio et al. 2008). In the upper Yellow Cat Member, the miring of Nedcolbertia is inferred
because of the bias of preserving only lower limbs and tails (Kirkland et al. 1998). Why were the
dinosaurs at the Stikes locality trapped in the sands and not in the adjoining muds? Bryant and
Miall (2010) report two articulated vertebrates (protosuchid crocodylomorphs; Lucas et al. 2005)
they inferred to be killed and buried in a liquefaction event. Khaldoun et al. (2005) conducted a
series of controlled experiments involving a mixture of fine sand, clays, and salt water to create a
natural colloidal gel that resists flowage equivalent to quicksand. A colloidal gel differs from
typical quicksand in that the absence grain-to-grain contact is minimized by surface chemistry. If
a colloidal gel developed, then the movement by a body, such as a struggling dinosaur, alters the
10
colloidal gel to a two-phase liquid and solid sand that induces sinking and trapping. The wet sand
phase has a high enough viscosity to trap the animals. Natural upwelling may not be of sufficient
flow rate to convert the wet sand back into a colloidal gel. These potential conditions coupled
with vigorous vertical flow would preferentially trap smaller, weaker dinosaurs over larger,
healthy animals. Clays are present in the sandstone masses containing the Utahraptor skeletons
but are not required to enhance trapping ability. Saline groundwater at least periodically
increases the probability of generating quicksand gel, based upon the Khaldoun et al. (2005)
experiments. Periodic variations in salinity may have been present from four lines of evidence:
(1) movement on the salt structures during Cedar Mountain Formation time (Kirkland and
Madsen 2007); (2) semi-arid to arid setting with monsoons (Kirkland et al. 1999; Suarez et al.
2007a, 2007b, 2014; Arens and Harris 2015); (3) barite crystals in the upper Yellow Cat
lacustrine sediments (Kirkland and Madsen 2007); and (4) sandstone masses being hosted in a
red mudstone with carbonate nodules, as documented in the Ruby Ranch Member (Ludvigson et
al. 2010).
Articulated theropods trapped in saturated, trampled sediments have been reported from
the Jurassic of China (Eberth et al. 2010). The Shisugou Formation deposits are associated with
extensive soft-sediment deformation attributed to vertical loading. The Stikes Quarry exhibits no
evidence of vertical loading but extensive mobilization and injection from underlying
stratigraphic levels.
SUMMARY
Previously reported mechanisms that generate bone beds are not sufficient to explain the
linkage by dikes and pipes of sill-like clustered bone beds and other related soft-sediment
deformation features discovered in the Lower Cretaceous, upper Yellow Cat Member of the
Cedar Mountain Formation. The Stikes Quarry deformed strata encases three dinosaur-bearing
lensoidal masses the largest of which contains a minimum of three baby, four juvenile and one
adult Utahraptor skeletons. The suite of sedimentary structure in the Stikes Quarry is consistent
with the development of a non-steady state artesian spring system that breached a mud layer.
Vigorous eruption rates develop quicksand conditions on the crater floor that trapped Utahraptor
dinosaurs and potentially other organisms.
The initial artesian eruption produced a crater with an attritional accumulation of
disarticulated and fragmented bones. The crater was subsequently reactivated up to three
additional times and trapped the various dinosaurs. Fluctuating pressure through time caused by
recharging the aquifer produced smaller springs that continued to trap the dinosaurs.
The Stikes Quarry death trap provides a stratigraphic and sedimentologic model that can
further be tested during preparation of additional material and provides an important new
mechanism attributed for trapping and preserving vertebrates.
ACKNOWLEDGMENTS
Matt Stikes is thanked for reporting this site to the Utah Geological Survey. Excavations
were conducted under Utah state permits # 2005-346, 2008-368, 2011-402, and 2014-438.
Thanks to all the volunteers, students and scientists that visited the site and aided in the
excavation. Special thanks are due to Utah Friends of Paleontology members D. Gray and P.
Policelli. Students and friends of the Washington University paleontology program contributed
mightily to the early phases of excavation. Numerous scientists have visited and worked with us
on the site over the years providing an invaluable sounding board for some of the ideas presented
here, including K. Poole, T. Birthisel, R. Chapman, R. Gaston, J. Foster, G. Hunt, J. Cavin, R.
Hernandez-Rivera, J. Choiniere, M. Loewen, M. Madsen, C. Suarez, M. Suarez, M. Joeckel, G.
11
Ludgvigson, and A. Moller. J. Cross of Cross Marine Projects, D. Harrison of High Desert
Excavation, H. Firm, and D. Brummel provided critical logistic support in bringing the massive
9-ton fossiliferous plaster jacket down off the cuesta with funding by the Utah Geological
Survey, Utah Friends of Paleontology, Andrew Leitch, and Jim Cross. Kutztown University
Research Committee funded participation of KU students and staff; E. Heness, S. Ireland, K.
Livingston, E. Laub and M. Malenda are thanked for their help in the field. Reviews by E.
Roberts. M. Chan, B. Gates, M. Hayden, D. Loope, M. Lowe, B. Sames, L. Sherrod, and M.
Hylland are appreciated.
REFERENCES
AKERSTEN, W.A., SHAW, C.A., and JEFFERSON, G.T., 1983, Rancho La Brea: status and future:
Paleobiology, v. 9, p. 211–217.
ALLEN, J.R.L., 1982, Sedimentary structures: their character and physical basis, volume 2:
Elsevier, Amsterdam, 663 p.
ALLEN, J.R.L., 1986, Earthquake magnitude-frequency, epicentral distance, and soft sediment
deformation in sedimentary basins: Sedimentary Geology, v. 46, p. 67–75.
ARENAS, C. and PARDO, G. 1999, Latest Oligocene–Late Miocene lacustrine systems of the
north-central part of the Ebro Basin (Spain): sedimentary facies model and
palaeogeographic synthesis: Palaeogeography, Palaeoclimitology, Palaeoecology, v. 151,
p. 127–148.
ARENS, N.C. and HARRIS, E.B., 2015, Paleoclimatic reconstruction of the Albian–Cenomanian
transition based on a dominantly angiosperm flora from the Cedar Mountain Formation,
Utah, USA: Cretaceous Research, v. 53, p. 140–152.
AUBREY, W.M., 1998, A newly discovered, widespread fluvial facies and unconformity marking
the Upper Jurassic/Lower Cretaceous boundary, Colorado Plateau, in Carpenter, K., Chure,
D., and Kirkland, J.I. (eds.), The Upper Jurassic Morrison Formation—An
Interdisciplinary Study, Part I: Modern Geology, v. 22, p. 209–233.
AUDEMARD, M.F.A. and MICHETTI, A.M., 2011, Geological criteria for evaluating seismicity
revisited: forty years of paleoseismic investigations and the natural record of past
earthquakes in Audemard, M.F.A., Michetti, A.M., and McCalpin, J.P. (eds.), Geological
Criteria for Evaluating Seismicity Revisited: Forty Years of Paleoseismic Investigations
and the Natural Record of Past Earthquakes: Geological Society of America Special
Publication 479, p. 1–21.
BASTIN, S.H., QUIGLEY, M.C. and BASSETT, K., 2015, Paleoliquefaction in Christchurch, New
Zealand: Geological Society of America, v. 127, p. 1348–1365.
BEHRENSMEYER, A.K., 1991, Terrestrial vertebrate accumulations, in Allison, P.A. and Briggs,
D.E.G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil Record: Plenum,
New York, p. 291–331.
BRACKEN, B.R., 1994, Syn-rift lacustrine beach and deltaic sandstone reservoirs—pre-salt
(Lower Cretaceous) of Cabinia, Angola, West Africa, in Lomando, A.J., Schreiber, B.C.,
and Harris, P.M. (eds.), Lacustrine Reservoirs and Depositional Systems: SEPM Core
Workshop No. 19, p. 173–200.
BLAIR, T.C., 2000, Sedimentology and progressive tectonic unconformities of the sheetflood-
dominated Hells Gate alluvial fan, Death Valley, California: Sedimentary Geology, v.
132, p. 233–262.
12
BOGUCHWAL, L.A. and SOUTHARD, J.B., 2002, Bed configurations in steady unidirectional water
flows, Part 1, scale model study using fine sands: Journal of Sedimentary Research, v. 60,
p. 649–657.
BRIDGE, J.S., 2003, Rivers and Floodplains: Processes and Sedimentary Record: Blackwell
Science, Oxford, 491 p.
BRISTOW, C.S., SKELLY, R.L., AND ETHRIDGE, F.G., 1999, Crevasse splays from rapidly aggrading,
sand-bed, braided Niobrara River, Nebraska: effect of base-level rise: Sedimentology, v.
46, p. 1029–1047.
BRITT, B.B., EBERTH, D.A., SCHEETZ, R., GREENHALGH, B.W., and STADTMAN, K.L., 2009,
Taphonomy of debris-flow hosted dinosaur bonebeds at Dalton Wells, Utah (Lower
Cretaceous, Cedar Mountain Formation, USA): Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 280, p. 1–22.
BRYANT, G. and MIALL, A., 2010, Diverse products of near-surface sediment mobilization in an
ancient eolianite: outcrop features of the early Jurassic Navajo Sandstone: Basin
Research, v. 22, p. 578–590.
CARBONEL, P., COLIN, J-P., DANIELOPOL, D.L., LOFFLER, H. AND NEUSTUEVA, I., 1988,
Paleoecology of limnic ostracodes: a review of some major topics: Paleogeography,
Paleoclimatology, Paleoecology, v. 62, p. 413–461.
CARPENTER, K., 2014, Where the sea meets the land: the unresolved Dakota Problem in Utah, in
MacLean, J.S., Biek, R.F., and Huntoon, J.E. (eds.), Geology of Utah’s Far South: Utah
Geological Society Special Publication 43, p. 357–372.
CARTWRIGHT, J., 2010, Regionally extensive placement of sandstone intrusions: a brief review:
Basin Research, v. 22, p. 502–516.
CARTWRIGHT, J., HUUSE, M., and APLIN, A., 2007, Seal bypass systems: AAPG Bulletin, v. 91,
p. 1141–1166.
CIFELLI, R.L., NYDAM, R.L., GARDNER, J.D., WEIL, A., EATON, J.G., KIRKLAND, J.I., AND
MADSEN, S.K., 1999, Medial Cretaceous vertebrates from the Cedar Mountain
Formation, Emery County, the Mussentuchit local fauna, in Gillette, D. (ed.), Vertebrate
Paleontology in Utah: Utah Geological Survey Miscellaneous Publication 99-1, p. 219–
242.
COSGROVE, J.W., 2001, Hydraulic fracturing during the formation and deformation of a basin: a
factor in dewatering of low-permeability sediments: American Association of Petroleum
Geologists Bulletin, v. 85, p. 737–748.
DALEY, B., 1971, Diapiric and other deformational structures in an Oligocene argillaceous
limestone: Sedimentary Geology, v. 6, p. 29–51.
DAVISON, I., 1987, Normal fault geometry related to sediment compaction and burial: Journal of
Structural Geology, v. 9, p. 393–401.
DICROCE, T. and CARPENTER, K., 2001, New ornithopod from the Cedar Mountain Formation
(Lower Cretaceous) of eastern Utah, in Tanke, D.H. and Skrepnick, M.W. (eds.),
Mesozoic Vertebrate Life: Indiana University Press, Bloomington, Indiana, p. 183–196.
DOELLING, H.H., 2010, Geology of Arches National Park, in Sprinkel, D.A., Chidsey, T.C., Jr., and
Anderson, P.B. (eds.), Geology of Utah’s Parks and Monuments: Utah Geological
Association Publication 28, third edition, p. 11–36.
DOELLING, H.H. and KUEHNE, P., 2013, Geologic maps of the Klondike Bluffs, Mollie Hogans,
and the Windows section 7.5' quadrangles, Grand County, Utah: Utah Geological
Survey, booklet for Map 258DM, Map 259DM, Map 260DM, 21 p., 3 plates.
13
DRANGANTIS, E., GRASEMENN, B., and SCHMID, H.P., 2003, Fluidization pipes and spring pits in
Gondwanan barrier-island environment: groundwater phenomenon, palaeoseismicity or a
combination of both?, in Van Rensbergen, P., Hillis, R.R., Maltman, A.J., and Morley,
C.K. (eds.), Subsurface Mobilization: Geological Society of London Special Publication
216, p. 109–121.
EBERTH, D.A., XING, X., and CLARK, D.M., 2010, Dinosaur death pits from the Jurassic of
China: PALAIOS, v. 25, p. 112–125.
EL TAKI, H. and PRATT, B.R., 2012, Syndepositional tectonic activity in an epicontinental basin
revealed by deformation of subaqueous carbonate laminites and evaporates: seismites in
Red River strata (Upper Ordovician) of southern Saskatchewan, Canada: Bulletin of
Canadian Petroleum Geology, v. 60, p. 37–58.
ETTENSOHN, F.R., RAST, N. and BRETT, C.E., (eds.), 2002, Ancient seismites: Geological
Society of America Special Paper 359, 190 p.
FERRILL, D.A. and MORRIS, A.P., 2003, Dilational normal faults: Journal of Structural Geology,
v. 25, p. 183–196.
FIELD, M.E., GARDNER, V., JENNINGS, A.E. and EDWARDS, B.D., 1982, Earthquake-induced
sediment failure on a 0.25° slope, Klamath River Delta, California: Geology, v. 10, p.
542–546.
FISHER, J.A., NICHOLS, G.J., and WALTHAM, D.A., 2007, Unconfined flow deposits in distal
sectors of fluvial distributary systems: examples from the Miocene Luna and Huesca
Systems, northern Spain: Sedimentary Geology, v. 195, p. 55–73.
FRASER, G.S. and HESTER, N.C., 1977, Sediments and sedimentary structures of a beach-ridge
complex, southwestern shore of Lake Michigan: Journal of Sedimentary Petrology, v.
47, p. 1187–1200.
FREY, S.E., GINGRAS, M.K., and DASHTGARD, S.E., 2009, Experimental studies of gas-escape
and water-escape structures: Journal of Sedimentary Research, v. 79, p. 808–816.
FRISCIA, A.R., VAN VALKENBERG, B., SPENCER, I., and HARRIS, J., 2008, Chronology and
spatial distribution of large mammal bones in pit 91, Rancho La Brea: PALAIOS, v. 23,
p. 35–42.
GALLO, F. and WOODS, A.W., 2004, On steady homogeneous sand-water flows in a vertical
conduit: Sedimentology, v. 51, p. 195–210.
GATES, T.A., 2005, The Late Jurassic Cleveland-Lloyd Dinosaur Quarry as a drought-induced
assemblage: PALAIOS, v. 20, p. 363–375.
GHOSH, P., SARKAR, S., and MAULIK, P., 2006, Sedimentology of a muddy alluvial deposit
Triassic Denwa Formation, India: Sedimentary Geology, v. 191, p. 3–36.
GILL, W.D. and KUENEN, P.H., 1957, Sand volcanoes on slumps in the Carboniferous of County
Clare, Ireland: Quarterly Journal of the Geological Society of London, v. 113, p. 441–
460.
GREENHALGH, B.W. and BRITT, B.B., 2007, Stratigraphy and sedimentology of the Morrison-
Cedar Mountain boundary, east-central Utah, in Willis, C.G., Hylland, M.D., Clark,
D.L., and Chidsey, T.C., Jr. (eds.), Central Utah: Diverse Geology of a Dynamic
Landscape: Utah Geological Association Special Publication 36, p. 81–100.
GREENHALGH, B.W., BRITT, B.B., and KOWALLIS, B.J., 2006, New U-Pb age control for the
lower Cedar Mountain Formation and an evaluation of the Morris Formation/Cedar
Mountain Formation Boundary, Utah: Geological Society of American Abstracts with
Programs, v. 38, no. 7, p. 52.
14
GRIMM, K.A. and ORANGE, D.L., 1997, Synsedimentary fracturing, fluid migration, and
subaqueous mass wasting intrastratal microfractured zones in laminated diatomaceous
sediments, Miocene Monterey Formation, California, USA: Journal of Sedimentary
Research, v. 67, p. 601–613.
GUHMAN, A.I. and PEDERSON, D.T., 1992, Boiling sand springs, Dismal River, Nebraska:
agents for the formation of vertical cylindrical structures and geomorphic change:
Geology, v. 20, p. 8–10.
HARMS, J.C., SOUTHARD, J.B., and WALKER, R.G., 1982, Structures and sequences in clastic
rocks: SEPM Short Course 9, 251 p.
HARRIS, D.R., 1980, Exhumed paleochannels in the Lower Cretaceous Cedar Mountain
Formation near Green River, Utah: Brigham Young University Geology Studies, v. 27,
p. 51–66.
HENDRIX, B., MOELLER, A., LUDVIGSON, G.A., JOECKEL, R.M., and KIRKLAND, J.I., 2015, A
new approach to date paleosols in terrestrial strata: a case study using U-Pb zircon ages
for the Yellow Cat Member of the Cedar Mountain Formation of eastern Utah:
Geological Society of American Abstracts with Programs, v. 47, no. 7, p. 597.
HEUBECK, C., 2009, Geröllührende sedimentäre Gänge der Eisenach–Formation
(Oberrotiegend): modifizierte artesische Injekionen am fuß alluvialer Fächer?:
Zeitschrift der Deutschen Gesellschaft für Geowissenschaften, v. 160, p. 41–56.
HILBERT-WOLF, H.L., SIMPSON, E.L., SIMPSON, W.S., TINDALL, S.E., and WIZEVICH, M.C.,
2009, Insights into syndepositional fault movement in a foreland basin; trends in
seismites of the Upper Cretaceous, Wahweap Formation, Kaiparowits Basin, Utah,
USA: Basin Research, v. 21, p. 856–871.
HILBERT-WOLF, H.L. and ROBERTS, E.M., 2015, Giant seismites and megablock uplift in the
East African Rift: evidence for the late Pleistocene large magnitude earthquakes: PLOS
ONE, v. 10 (6) e0129051. doi:10.1371/journal.pone.0129051.
HORNE, D.J., 2002, Ostracod biostratigraphy and palaeoecology of the Purbeck Limestone
Group in southern England: Special Papers in Palaeontology, v. 6, p. 53–70.
HORNE, D.J. and COLIN, J.-P. 2005, The affinities of the ostracod genus Cypridea Bosquet,
1852, and its allies, with consideration of implications for the phylogeny of nonmarine
Cypridoidea ostracodes: Revue de Micropaléontologie, v. 48, p. 25–35.
HUNGERBÜHLER, A., 1998, Taphonomy of the prosauropod dinosaur Sellosaurus, and its
implications for carnivore faunas and feeding habits in the Late Triassic:
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 143, p. 1–29.
HURST, A., CARTWRIGHT, J., HUUSE, M., JONK, R., SCHWAB, A., DURANTI, D., AND CRONIN, B.,
2003, Significance of large-scale sand injectites as long-term fluid conduits: evidence
from seismic data: Geofluids, v. 3, p. 263–274.
ILGAR, A. and NEMEC, W., 2005, Early Miocene lacustrine deposits and sequence stratigraphy of
the Ermenek Basin, Central Taurides, Turkey: Sedimentary Geology, v. 173, p. 233–
275.
JIANHUA, Z., ZHIFENG, W., GUANMIN, W., XIBIN, W., HONGBO, L., and XIAOHUA, S., 2004,
Air-discharge pits on the Yellow River delta plain: Sedimentary Geology, v. 170, p. 1–
20.
JOLLY, R.J.H. and LONERGAN, L., 2002, Mechanisms and controls on the formation of sand
intrusions: Journal of the Geological Society of London, v. 159, p. 605–617.
15
JONES, A.P. and OMOTO, K., 2000, Towards establishing criteria for identifying trigger
mechanisms for soft-sediment deformation: a case study of late Pleistocene lacustrine
sands and clays, Onikobe and Nakayamadaira Basins, northeastern Japan:
Sedimentology, v. 47, p. 1211–1226.
JONK, R., 2010, Sand-rich injectites in the context of short-lived and long-lived fluid flows:
Basin Research, v. 22, p. 603–621.
KHALDOUN, A., EISERT, E., WEGDAM, G.H., and BONN, D., 2005, Liquefaction of quicksand
under stress: Nature, v. 437, p. 635.
KIRKLAND, J.I. and ARMSTRONG, H.J., 1992, Taphonomy of the Mygatt-Moore Quarry, middle
Brushy Basin Member, Morrison Formation (Upper Jurassic) western Colorado: Journal
of Vertebrate Paleontology, v. 12, supplement to no. 3, [Abstracts], p. 40A.
KIRKLAND, J.I., BRITT, B., BURGE, D.L., CARPENTER, K., CIFELLI, R., DECOURTEN, F., EATON,
J., HASIOTIS, S., and LAWTON, T., 1997, Lower to middle Cretaceous dinosaur faunas of
the central Colorado Plateau: a key to understanding 35 million years of tectonics,
sedimentology, evolution, and biogeography, in Link, P.K. and Kowallis, B.J. (eds.),
Mesozoic to Recent Geology of Utah: Brigham Young University Geology Studies 42,
p. 69–103.
KIRKLAND, J.I., BRITT, B.B., WHITTLE, C.H., MADSEN, S.K., and BURGE, D.L., 1998, A small
coelurosaurian theropod from the Yellow Cat Member of the Cedar Mountain Formation
(Lower Cretaceous, Barremian) of eastern Utah, in Lucas, S.G., Kirkland, J.I., and
Estep, J.W. (eds.), Lower and Middle Cretaceous Terrestrial Ecosystems: New Mexico
Museum Natural History and Science Bulletin 14, p. 239–248.
KIRKLAND, J.I., BURGE, D., and GASTON, R., 1993, A large dromaeosaur (Theropoda) from the
Lower Cretaceous of eastern Utah: Hunteria, v. 2, no. 10, p. 1–16.
KIRKLAND, J.I., CIFELLI, R., BRITT, B., BURGE, D.L., DECOURTEN, F., EATON, J., and PARRISH,
J.M., 1999, Distribution of vertebrate faunas in the Cedar Mountain Formation, east-
central Utah, in Gillette, D. (ed.), Vertebrate Paleontology in Utah: Utah Geological
Survey, Miscellaneous Publication 99-1, p. 201–217.
KIRKLAND, J.I., DEBLIEUX, D., MADSEN, S.K., and HUNT, G.J., 2012, New dinosaurs from the
base of the Cretaceous in eastern Utah suggest that the “so-called” basal Cretaceous
calcrete in the Yellow Cat Member of the Cedar Mountain Formation, while not
marking the Jurassic–Cretaceous unconformity represents evolutionary time: Journal of
Vertebrate Paleontology, v. 18, Supplement, October 2012, Program and Abstracts, p.
121–122.
KIRKLAND, J.I. and FARLOW, J.O., 2012, Dinosaurs and geological time, in Brett-Surman,
M.K., Holtz, T.R., and Farlow, J.O. (eds.), The Complete Dinosaur: Indiana University
Press, Bloomington, p. 224–245.
KIRKLAND, J.I., LOEWEN, M., DEBLIEUX, D., MADSEN, S.K., and CHOINIERE, J., 2011, New
theropod cranial material from the Yellow Cat Member, Cedar Mountain Formation,
Stikes Quarry, north of Arches National Park, east-central, Utah: Journal of Vertebrate
Paleontology, v. 17, Supplement, November 2011, Program and Abstracts, p. 137.
KIRKLAND, J.I. and MADSEN, S.K., 2007, The Lower Cretaceous Cedar Mountain Formation of
eastern Utah: the view up an always interesting learning curve, in Lund, W.R. (ed.),
Field Guide to Geological Excursions in southern Utah: Geological Society of America
Rocky Mountain Section 2007 Annual Meeting, Utah Geological Association
Publication 35, p. 1–108.
16
KIRKLAND, J.I., SCHEETZ, R.D., and FOSTER, J.R., 2005, Jurassic and Lower Cretaceous
dinosaur quarries of western Colorado and eastern Utah, in Rishaed, G. (ed.), 2005
Rocky Mountain Section of the Geological Society of America field trip guidebook,
Grand Junction Geological Society Trip 402, p. 1–26.
KIRKLAND, J.I., YOU, H.-L., ALCALA, L., and LOEWEN, M., 2015, A near-continuous, well-
dated sequence of Cretaceous terrestrial faunas: mid-Cretaceous faunal change in the
Northern Hemisphere as viewed from Utah: Journal of Vertebrate Paleontology, v. 36,
Supplement, Program with Abstracts, p. 121–122.
LIVINGSTON, K.M., BOGNER, E., IRELAND, S.M., SIMPSON, E.L., BETTS, T.A., and LAUB, E.,
2014, The geomorphic evolution of shallow-sourced methane produced mud volcanoes:
Lake Powell, Utah: Geological Society of America Abstracts with Programs, v. 46, no.
6, p. 766–767.
LIVINGSTON, K.M., BOGNER, E., SIMPSON, E.L., MALENDA, M., SHERROD, L., BETTS, T.A., and
LAUB, E., 2015, The proposed evolution of shallow-sourced methane mud volcano
geomorphology: Lake Powell, Hite Utah: Geological Society of America Abstracts with
Programs, v. 47, no. 6, p. 587.
LOCKLEY, M.G., BUCKLEY, L.G., FOSTER, J.R., KIRKLAND, J.I., and DEBLIEUX, D.D., 2015,
First report of bird tracks (Aquatilavipes) from the Cedar Mountain Formation (Lower
Cretaceous), eastern Utah: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 420,
p. 150–162.
LOOPE, D.B., ELDER, J.F., ZLOTNIK, V.A., KETTLER, R.M., and PEDERSON, D.T., 2013, Jurassic
earthquake sequence recorded in multiple generations of sand blows, Zion National
Park, Utah: Geology, v. 41, p. 1131–1134.
LORANGER, D.M., 1951, Useful Bairmore microfossil zone in central and southern Alberta,
Canada: American Association of Petroleum Geologists Bulletin, v. 35, p. 2348–2367.
LOWE, D.R., 1975, Water-escape structures in coarse-grained sediments: Sedimentology, v. 22,
p. 157–204.
LUDVIGSON, G.A., JOECKEL, R.M., GONZÁLEZ, L.A., GULBRANSON, E.I., RASBURY, E.T., HUNT,
G.J., KIRKLAND, J.I., and MADSEN, S., 2010, Correlation of Aptian–Albian carbon
isotope excursions in continental strata of Cretaceous foreland basin of eastern Utah:
Journal of Sedimentary Research, v. 80, p. 955–974.
LUCAS, S. G., HECKERT, A.B., and TANNER, L.H., 2005, Arizona’s Jurassic fossil vertebrates and
the age of the Glen Canyon Group, in Heckert, A.B. and Lucas, S.G. (eds.), Vertebrate
Paleontology in Arizona: New Mexico Museum of Natural History and Science Bulletin
29, p. 95–104.
MALTMAN, A., 1984, On the term ‘soft-sediment deformation’: Journal of Structural Geology, v.
6, p. 589–592.
MARTÍN-CLOSAS C., SAMES B., SCHUDACK M.E., 2013, Charophytes from the upper
Berriasian of the Western Interior Basin of the United States: Cretaceous Research, v. 46,
p. 11–23.
MASSARI, F., GHIBAUDO, G., D’AESSANDRO, A., and DAVAUD, E., 2001, Water-upwelling pipes
and soft-sediment-deformation structures in lower Pleistocene calcarenites (Salento,
southern Italy): Geological Society of America Bulletin, v. 113, p. 545–560.
MAXSON, J.H., 1940, Gas pits in non-marine sediments: Journal of Sedimentary Petrology, v. 10,
p. 142–145.
17
MCDONALD, A.T., KIRKLAND, J.I., DEBLIEUX, D.D., MADSEN, S.K., CAVIN, J., MILNER, A.R.,
and PANZARIN, L., 2010, New basal iguanodonts from the Cedar Mountain Formation of
Utah and the evolution of thumb-spiked dinosaurs: PloS One 5(11): e14075.
doi:10.137journal.pone0014075.
MERRIAM, J.C., 1911, The fauna of Ranch La Brea, Part 1, occurrence: University of California
Memoir 1, p. 192–213.
MIALL, A.D., 1996, The Geology of Fluvial Deposits: Sedimentary Facies, Basin Analysis and
Petroleum Geology: Springer-Verlag, Berlin, 582 p.
MILLER, K., SHERROD, L., BOGNER, E., HIGGINS, R., LIVINGSTON, K., MALENDA, M.
MORGANO, K., SIMPSON, E.L., SIMPSON, W.S., SNYDER, E., and VALES, D., 2015,
Geometry of small-scale fluid/gas conduits: Lake Powell, Hite Utah: Geological Society
of America Abstracts with Programs, v. 47, n. 6, p. 588.
MILLS, P.C., 1983, Genesis and diagnostic value of soft-sediment deformation structures—a
review: Sedimentary Geology, v. 101, p. 69–83.
MONTENAT, C., BARRIER, P., OTT DESTEVOU, P., and HIBSCH, C., 2007, Seismites: an attempt
at critical analysis and classification: Sedimentary Geology, v. 196, p. 5–30.
MONTGOMERY, E.H., 2014, Limnogeology and chemostratigraphy of carbonates and organic
carbon from the Cedar Mountain Formation (CMF), eastern Utah: Unpublished M.S.
thesis, University of Texas at San Antonio, San Antonio, 68 p.
MORETTI, M., 2000, Soft-sediment deformation structures interpreted as seismites in middle-late
Pleistocene aeolian deposits (Apulian foreland, southern Italy): Sedimentary Geology, v.
135, p. 167–179.
MORETTI, M. and SABATO, L., 2007, Recognition of trigger mechanisms for soft-sediment
deformation in the Pleistocene lacustrine deposits of the Sant’Arcangelo Basin (southern
Italy): seismic shock vs. overloading: Sedimentary Geology, v. 196, p. 31–45.
NETOFF, D., BALDWIN, C.T., and DOHRENWEND, J., 2010, Non-seismogenic origin of fluid/gas
escape structures and lateral spreads in the recently exposed Hite delta, Lake Powell,
Utah—preliminary findings, in Carney, S.M., Tabet, D.E., and Johnson, C.L. (eds.),
Geology of south-central Utah: Utah Geological Association Publication 39, p. 61–92.
NICHOLS, R.J., SPARKS, S.J., and WILSON, C.J.N., 1994, Experimental studies of the fluidization
of layered sediments and the formation of fluid escape structures: Sedimentology, v. 41,
p. 233–253.
OBERMEIER, S.F., 1996, Use of liquefaction-induced feature for the paleoseismic analysis—An
overview of how seismic liquefaction features can be distinguished from other features
and how their regional distribution and properties of source-sediment can be used to
infer the location and strength of Holocene paleo-earthquakes: Engineering Geology, v.
44, p. 1–76.
OBERMEIER, S.F., 1998, Liquefaction evidence for strong earthquakes in Holocene and latest
Pleistocene ages in the states of Indiana and Illinois, USA: Engineering Geology, v. 50,
p. 227–254.
OWEN, G., 1987, Deformational processes in unconsolidated sands, in Jones, M.E. and Preston,
R.M.F. (eds.), Deformation of Sediment and Sedimentary Rocks: Geological Society of
London Special Publication 29, p. 137–146.
OWEN, G., 1996, Experimental soft-sediment deformation: structures formed by liquefaction of
unconsolidated sands and some ancient examples: Sedimentology, v. 43, p. 279–293.
18
PECK, R.E., 1941, Lower Cretaceous Rocky Mountain nonmarine microfossils: Journal of
Paleontology, v. 15, p. 285–304.
PEREZ-GARCIA, C., FESEKER, T., MIENERT, J., and BERNDT, C., 2009, The Håkon mud volcano:
330000 years of focused fluid flow activity at the Barents Sea slope: Marine Geology, v.
262, 105–115.
POOLE, K.E., 2008, A new specimen of iguanodont dinosaur from the Cedar Mountian
Formation, Grand County, Utah: Unpublished M.S. thesis, Washington University, Saint
Louis, 56 p.
PRATT, B.R., 1994, Seismites in the Mesoproterozoic Altyn Formation (Belt Supergroup),
Montana: a test for tectonic control on peritidal carbonate cyclicity: Geology, v. 22, p.
1091–1094.
PRICE, N.J., 1966, Fault and joint development in brittle and semi-brittle rock: Pergamon Press,
Oxford, 176 p.
QUIGLEY, M.C., BASTIN, S., and BRADLEY, B.A., 2013, Recurrent liquefaction in Christchurch,
New Zealand, during the Canterbury earthquake sequence: Geology, v. 41, p. 419–422.
REICHERTER, K., MICHETTI, A.M., and SILVA BARROSO, P.G., 2009, Palaeoseismology:
historical and prehistorical records of earthquake ground effects for seismic hazard
assessment, in Reicherter, K., Michetti, A.M., and Silva, P.G. (eds.), Palaeoseismology:
Historical and Prehistorical Records of Earthquake Ground Effects for Seismic Hazard
Assessment: The Geological Society, London, Special Publication 316, p. 1–10.
RIDER, M.H., 1978, Growth faults in Carboniferous of western Ireland: Geological Society of
America Bulletin, v. 62, p. 2191–2213.
RODRÍGUEZ-PASCUA, M.A., SILVA, P.G., PÉREZ-LÓPEZ, R., GINER-ROBLES, J., MARTÍN-
GONZÁLEZ, F., and DEL MORAL, B., 2015, Polygenic sand volcanoes: on the features of
liquefaction processes generated by a single event (2012 Emilia Romagna 5.9 Mw
earthquake, Italy): Quaternary International, v. 357, p. 329–335.
ROGERS, R.R., EBERTH, D.A., and FIORILLO, A.R. (eds.), 2007, Bonebeds—genesis, analysis,
and paleobiological significance: University of Chicago Press, Chicago, 499 p.
ROSS, J.A., PEAKALL, J., and KEEVIL, G.M., 2014, Facies and flow regimes of sandstone-hosted
columnar intrusions: insights from pipes of Kodachrome Basin State Park:
Sedimentology, v. 61, p. 1764–1792.
ROSSETTI, D.F., 1999, Soft-sediment deformation structures in late Albian to Cenomanian
deposits, São Luís Basin, northern Brazil: evidence for paleoseismicity: Sedimentology,
v. 46, p. 1065–1081.
SAMES, B., 2011a, Early Cretaceous Cypridea bosquet 1852 in North America and Europe, in
Sames, B. (ed.), Taxonomic Studies in Early Cretaceous Nonmarine Ostracoda of North
America: Micropaleontology, v. 57, p. 345–431.
SAMES, B., 2011b, Combined references for taxonomic studies in Early Cretaceous nonmarine
Ostracoda of North America, in Sames, B. (ed.) Taxonomic studies in Early Cretaceous
nonmarine Ostracoda of North America: Micropaleontology, v. 57, p. 455–465.
SAMES, B., CIFELLI, R.L., and SCHUDACK, M.E., 2010, The nonmarine Lower Cretaceous of the
North American Western Interior foreland basin: new biostratigraphic results from
ostracod correlations and early mammals, and their implications for paleontology and
geology of the basin—an overview: Earth Science Reviews, v. 101, p. 207–224.
19
SEED, H.B., 1979, Soil liquefaction and cyclic mobility evaluation for level ground during
earthquakes: Journal Geotechnical Engineering American Society of Civil Engineering,
v. 94, p. 1055–1122.
SEILACHER, A., 1969, Fault graded beds interpreted as seismites: Sedimentology, v. 13, p. 155–
159.
SHERROD, L., SIMPSON, E.L., HIGGINS, R., MILLER, K., MORGANO, K., SNYDER, E., and VALES,
D., 2016, Subsurface structure of water-gas escape features revealed by ground-
penetrating radar and electrical resistivity tomography, Glen Canyon National
Recreation Area, Lake Powell Delta, Utah, USA: Sedimentary Geology, in press,
http://dx.doi.org/10.10116/j.sedgeo.2016.02.005
SHOULDERS, S.J., CARTWRIGHT, J., and HUUSE, M., 2007, Large-scale conical sandstone
intrusions and polygonal fault systems in Tranche 6, Faroe-Shetland Basin: Marine and
Petroleum Geology, v. 24, p. 173–188.
SIMS, J.D., 1973, Earthquake-induced structures in sediments of Van Norman Lake, San
Fernando, California: Science, v. 182, p. 161–163.
SIMS, J.D., 1975, Determining earthquake recurrence intervals from deformational structures in
young lacustrine sediments: Tectonophysics, v. 29, p. 144–152.
SMOOT, J.P., 1991, Sedimentary facies and depositional environments of early Mesozoic
Newark basins, eastern North America: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 84, p. 369–423.
SMOOT, J.P. and OLSEN, P.E., 1994, Climatic cycles as sedimentary controls of the rift-basin
lacustrine deposits in the Early Mesozoic Newark basin based on continuous core, in
Lomando, A.J., Schreiber, B.C., and Harris, P.M. (eds.), Lacustrine Reservoirs and
Depositional Systems: SEPM Core Workshop No. 19, p. 201–237.
SOUTHARD, J.B. and BOGUCHWAL, L.A., 1990, Bed configurations in steady unidirectional
flows, part 1: synthesis of flume data: Journal of Sedimentary Research, v. 60, p. 658–
679.
SPALLUTO, L., MORETTI, M., FESTA, V., and TROPEANO, M., 2007, Seismically induced slumps
in lower-Maastrichtian peritidal carbonates of the Apulian Platform (southern Italy):
Sedimentary Geology, v. 196, p. 81–98.
SPENCER, L.M., VAN VALKENBURGH, B., and HARRIS, J.M., 2003, Taphonomic analysis of
large mammals recovered from the Pleistocene Rancho La Brea tar seeps: Paleobiology,
v. 29, p. 561–575.
SPRINKEL, D.A., MADSEN, S.K., KIRKLAND, J.I., WAANDERS, G.L., and HUNT, G.J., 2012,
Cedar Mountain and Dakota Formations around Dinosaur National Monument—
evidence of the first incursion of the Cretaceous Western Interior Seaway into Utah:
Utah Geological Survey Special Study 143, 21 p., 6 appendices.
STEAR, W.M., 1978, Sedimentary structures related to fluctuating hydrodynamic conditions in
flood plain deposits of the Beaufort Group near Beaufort West, Cape: Geological
Society of South Africa, Transactions, v. 81, p. 393–399.
STIKES, M.W., 2007, Fluvial facies and architecture of the Poison Strip Sandstone, Lower
Cretaceous, Cedar Mountain Formation, Grand County, Utah: Utah Geological Survey,
Miscellaneous Publication 06-02, 84 p.
STOCK, C., 1965, Ranch La Brea—a record of Pleistocene life in California, sixth edition: Los
Angeles County Museum Series 20, Paleontology, p. 1–81.
20
STRACHAN, L.J., 2002, Slump-initiated and controlled syndepositional sandstone
remobilization: an example form the Namurian of County Clare, Ireland:
Sedimentology, v. 49, p. 25–41.
SUAREZ, C.A., GONZALAS, L.A., LUDVIGSON, G.A., KIRKLAND, J.I., CIFELLI, R.A., and KOHN,
M.J., 2014, Multi-taxa isotopic investigation of paleohydrology in the Lower Cretaceous
Cedar Mountain Formation, eastern Utah, U.S.A.: deciphering effects of the Nevadaplano
Plateau on regional Climate: Journal of Sedimentary Research, v. 84, p. 975–987.
SUAREZ, C.A., SUAREZ, M.B., TERRY, D.O., JR., and GRANDSTAFF, D.E., 2007a, Rare earth
element geochemistry and taphonomy of the Early Cretaceous Crystal Geyser dinosaur
quarry, east-central Utah: PALAIOS, v. 22, p. 500–512.
SUAREZ, M.B., SUAREZ, C.A., KIRKLAND, J.I., GONZÁLEZ, L.A., GRANDSTAFF, D.E., and TERRY,
D.O., JR., 2007b, Sedimentology, stratigraphy, and depositional environment of the crystal
geyser dinosaur quarry, east-central Utah: PALAIOS, v. 22, p. 513–527.
SURLYK, F., GJELBERG, J., and NOE-NYGAARD, N., 2007, The Upper Jurassic Hareelv Formation
of East Greenland: a giant sedimentary injection complex, in Hurst, A. and Cartwright, J.
(eds.), Sand Injectites: Implications for Hydrocarbon Exploration and Production:
American Association of Petroleum Geologists Memoir 87, p. 141–149.
TIDWELL, V.C., CARPENTER, K., and BROOKS, W., 1999, New sauropod from the Lower Cretaceous
of Utah, USA: Oryctos, v. 2, p. 21–37.
TOOTH, S., 1999, Floodouts in Central Australia, in Miller, A.J. and Gupta, A. (eds.), Varieties of
Fluvial Form: Wiley and Sons, London, p. 219–247.
TÖRŐ, B. and PRATT, B.R., 2015, Eocene paleoseismic record of the Green River Formation, Fossil
Basin, Wyoming, U.S.A.: implications of synsedimentary deformation structures in
lacustrine carbonate mudstones: Journal of Sedimentary Research, v. 85, p. 855–884.
TRABELSKI, K., COLIN, J.-P., TOUIR, J., AND SOUSSI, M., 2011, Cypridea Bosquet, 1852
(Ostracoda) in the early Albian of Tunisia: Journal of Micropalaeontology, v. 30, p. 187–
188.
VAN LOON, A.J., 2009, Soft-sediment deformation structures in siliciclastic sediments: an
overview: Geologos, v. 15, p. 3–55.
VARRICCHIO, D.J., SERENO, P.C., XIJIN, Z., LIN, T., WILSON, J.A., and LYON, G.H., 2008, Mud-
trapped herd captures evidence of distinctive dinosaur sociality: Acta Palaeontologica
Polonica, v. 53, p. 567–578.
WALSH, T.J., COMBELLICK, R.A., and BLACK, G.L., 1995, Liquefaction features form a
subduction earthquake; preserved examples from the 1964 Alaska earthquake:
Washington State Division of Geology and Earth Resources, Report of Investigations
32, 80 p.
WEIGELT, J., 1989, Recent vertebrate carcasses and their paleobiological implications: Schaefer,
J., (translator), University of Chicago Press, Chicago, 188 p.
WHEELER, R.L., 2002, Distinguishing seismic from aseismic soft-sediment structures: criteria
from seismic-hazard analysis, in Ettensohn, F.R., Rast, N., and Brett, C.E. (eds.),
Ancient Seismites: Geological Society of America Special Paper 359, p. 1–11.
WILLIAMS, R.M.E., IRWIN, R.P., III, and ZIMBELMAN, J.R., 2009, Evaluation of paleohydrologic
models for terrestrial inverted channels: implications for application to Martian sinuous
ridges: Geomorphology, v. 107, p. 300–315.
21
WIZEVICH, M.C., SIMPSON, E.L., HILBERT-WOLF, H.L., and TINDALL, S.E., 2016,
Characteristics and formation of unusual, large-scale, fault proximal seismites in the
Upper Cretaceous Wahweap Formation, Utah: Sedimentology, in press.
WOODARD, G.D. and MARCUS, L.F., 1973, Rancho La Brea deposits: a re-evaluation from
stratigraphic and geologic evidence: Journal of Paleontology, v. 47, p. 54–69.
YOUNG, R.G., 1960, Dakota Group of Colorado Plateau: Bulletin of the American Association of
Petroleum Geologists, v. 44, p. 156–194.
YOUNG, R.G., 1965, Type section of the Naturita Formation: Bulletin of the American
Association of Petroleum Geologists, v. 49, p. 1512–1531.
YOUNGER, P.L. and MCHUGH, M., 1995, Peat development, sand cones and palaeohydrology of
a spring-fed mire in East Yorkshire, UK: The Holocene, v. 5, p. 59–67.
Received 1 April 2016; accepted 19 July 2016.
FIGURE CAPTIONS:
FIG. 1.—Quarry. A) Stratigraphic framework at the Stikes Quarry. The ridge is now called
Utahraptor ridge. B) Locality of the Stikes Quarry modified from Kirkland and Madsen (2007).
Lake Madsen section is red dot. C) Measured section in the upper Yellow Cat Member. Fossil
bone is the stratigraphic position of the Stikes Quarry. D) Panorama. Note the lateral continuity
of quarry strata. Circled track hoe sits on haul road. Yellow vertical bar is ~ 14 m thickness. E)
Panel constructed along the haul road. Aquifer and aquicludes are labeled. Stikes Quarry is not
shown in detail (See Fig. 4).
FIG. 2.—Photomosaic of the Stikes Quarry March 2015. Haul road on the right was developed to
remove large plaster jacket. All blocks are removed except for the lowest iguanodon-sauropod
block—sand bags are covering plaster jacket which was buried during reclamation. Figure on
right is 1.75 m tall.
FIG. 3.—Nonmarine Ostracoda from the lower gray shale in the upper Yellow Cat Member
below the Stikes Quarry. A) Mongolianella stirlingensis (Loranger 1951). B) Candona
albertensis (Loranger 1951). C) Cypridea (Pseudocypridina) inornata. D) Cypridea
(Longispinella) longispina (Peck 1941).
FIG. 4.—A) Stikes Quarry in May 2014. Person on the right is 1.8 m tall. B) Same photograph as
in A with stratigraphy, soft-sediment deformation and excavated blocks labeled. C) Stratigraphy
south of the Stikes Quarry. Note the lateral continuity of the strata and the apparent absence of
soft-sediment deformation. People are circled. Uppermost figure is 1.75 m tall.
FIG. 5.—Field photographs. A) Domal form present in lower sandstone in the Stikes Quarry.
Scale is 5 cm. B) Convolute bedding in lower sandstone near turn in haul road. Scale is 5 cm.
FIG. 6.—Field photographs of soft-sediment deformation structures along the haul road, March,
2015. Hammer is 31 cm in length.
FIG. 7.—Soft sediment deformation at contact of the lower rippled sandstone and the heterolithic
mudstone and sandy red and gray mudstone along haul road. A) Base of the heterolithic
mudstone and sandy red and gray mudstone. Note the fining upward sequence from siltstone to
more mudstone dominated facies from right to left. Field book is 18.5 cm in length. B)
Enlargement showing the pebbly mudstone intrusion cross cutting the siltstones and mudstones.
Scale is 5 cm.
FIG. 8.—Cartoons of the relationship of the sill-like sandstone masses. A) Map view
reconstruction of position of clastic pipe and dike, and fossiliferous lensoidal-ovoidal masses. B)
View of clastic pipe connecting to the Utahraptor block.
22
FIG. 9.—Utahraptor growth series. AC: Utahraptor dentaries growth series in left-lateral view.
Natural History Museum of Utah–Vertebrate Paleontology Collection. A, juvenile,
UMNH.VP.20497 as found in field; B, adult, UMNH.VP.20501; and C, juvenile,
UMNH.VP.20487. D, E) Premaxillary growth series. D, “Baby” Utahraptor left premaxillary,
UMNH.VP.20488; and E, Adult Utahraptor left premaxillary, Utah State University Eastern,
Prehistoric Museum CEUM 184v.300 from the type locality of Utahraptor (Gaston Quarry) 5.7
miles (9.1 km) nearly due east (Kirkland et al., 1993) in left-lateral view. Note the intact teeth
(damage due to post-burial breakage or collecting effects). Size roughly reflects the age of the
Utahraptors.
FIG. 10.—Photographs of the base of the iguanodont block. A) Vertical photograph of the
underside of the iguanodont block. Note the sharp contact between the sandstone and gray shale.
Scale is 15 mm. B) Iguanodont limb bone (brown color). Located in upper part of block in A. C)
Interbedding of sandstone and gray mudstone. Scale is 15 cm. D) Vertical contact between
sandstone and gray mudstone.
FIG. 11.—Field and slab photographs. A) Dike connecting to the underside of main Utahraptor
block. B) Slab of dike stratification. Note the presence of numerous ripped up mudstones from
dike walls. Scale in cm. C) Enlargement of reworked intraclast containing differently oriented
ripped up chips highlighted in yellow. Scale in mm.
FIG. 12.—Field photographs of soft-sediment deformation. A) Rippled-sandstone cut by fault.
Both sides of the fault plane are highlighted in black dash with sense of motion shown in yellow
arrow. Photo taken May 2014. B) Soft sediment normal faults. Figure is 1.65 m tall. C) Near
vertical fault with hanging wall syncline. Hand is for scale.
FIG. 13.—Field photographs taken in March, 2015 of hanging wall strata and associated increase
in deformation proximal to the syndepositional fault. This is the same block as shown in Fig.
12A but reduced in size by construction of the haul road (see Fig. 1D). A) Graded beds of fine-
grained sandstone to green mudstone. Scale is in cm. B) Folded and fluidized graded beds. Scale
is in cm. C) Homogenized fine-grained sandstones with mudstone ripped-up chips. All scales are
in cm.
FIG. 14.—Photomicrographs. A) Lower sandstone facies. B) Clasts in the sill structures (blocks).
C) Mud matrix in the Utahraptor block. D) Upper rippled sandstone.
FIG. 15.—Slab and field photographs of upper rippled sandstone. A) Outcrop photograph 10 m
east of Stikes Quarry. Note the thinning-upward bedding made of rippled sandstones. Scale is in
cm. B) Slab of upper rippled sandstone. Note the mudstone draped ripple bedforms. Scale is in
cm.
FIG. 16.—Photomosaics of fluid-gas escape structures. A) Boiling sand spring, Nebraska. Spring
is flooded by high-water event. Note the geometry of the pit and reduction in the size of the
overpressured zone. Pit is ~ 20 meters across. B) Collapse crater partially filled with mudstone at
Lake Powell Delta, Glen Canyon National Recreation Area, Hite, Utah. Various geomorphic
features are labeled. Figure is 1.8 m tall.
FIG. 17.—Cartoon reconstruction of the Stikes Quarry. Dinosaurs illustrated by Julius Csotonyi.
... The Yellow Cat Member of the terrestrial Cedar Mountain Formation is known worldwide for its diverse fossil vertebrates, and particularly its dinosaur fossils [1,2]. It is dominated by alluvial mudrocks with prominent, stacked paleosols [2][3][4]. ...
... It is dominated by alluvial mudrocks with prominent, stacked paleosols [2][3][4]. Palustrine to lacustrine sediments, including nodular to bedded carbonates ("calcretes") are locally prominent, and fluvial sandstones are comparatively minor components [1,3,4]. Previous research carried out by some of us over more than a decade [3][4][5][6][7] has established the value of comprehensive stable-isotope geochemistry in the Cedar Mountain Formation, especially when coupled with radiometric dating, yet our geochronologic framework the formation is still incomplete. ...
... The Yellow Cat Member is particularly intriguing-and even problematic-because: (1) its depositional extent was limited to a part of southeastern Utah by ancient salt tectonics, (2) its dinosaur faunas suggest that the unit records a large part of Early Cretaceous time, (3) its well-developed Vertisol-like paleosols are likely to record long intervals of geologic time within a few meters or less of sediments [1][2][3][4], and, therefore, (4) it is thought to be a "time-rich" unit [4]. We present a new composite δ 13 C chemostratigraphic profile at Utahraptor Ridge, Utah (Figure 1), where the Stikes Quarry has been excavated over a period of years for spectacular dinosaur fossils, including the remains of the eponymous dromaeosaurid Utahraptor [1,2]. ...
Article
Full-text available
The Early Cretaceous Yellow Cat Member of the terrestrial Cedar Mountain Formation in Utah, USA. has been interpreted as a “time-rich” unit because of its dinosaur fossils, prominent paleosols, and the results of preliminary chemostratigraphic and geochronologic studies. Herein, we refine prior interpretations with: (1) a new composite C-isotope chemostratigraphic profile from the well-known Utahraptor Ridge dinosaur site, which exhibits δ13C features tentatively interpreted as the Valanginian double-peak carbon isotope excursion (the so-called “Weissert Event”) and some unnamed Berriasian features; and (2) a new cryptotephra zircon eruption age of 135.10 ± 0.30/0.31/0.34 Ma (2σ) derived from the CA-ID-TIMS U-Pb analyses of zircons from a paleosol cryptotephra. Our interpretations of δ13C features on our chemostratigraphic profile, in the context of our new radiometric age, are compatible with at least one prior age model for the “Weissert Event” and the most recent revision of the Cretaceous time scale. Our results also support the interpretation that the Yellow Cat Member records a significant part of Early Cretaceous time.
... However, their origins are frequently complex, and drought, fire, or other catastrophic events can force aggregations of normally non-or minimally-gregarious taxa Gates, 2005). Attritional traps such as tar seeps (Stock, 1972), quicksand/mires (Kirkland et al., 2016), or sinkholes (Martin & Gilbert, 1978) can also accumulate non-time associated individuals. ...
... Quicksand/Liquefaction: Quicksand and seismic liquefaction have been hypothesized as mechanisms for spatial concentration of theropods either through mass kills or time attritional accumulations (Eberth, Xing & Clark, 2010;Kirkland et al., 2016). Both result in formation of contorted/vertically disrupted bedforms and other highly distinctive sedimentary structures that typically have meter scale relief. ...
Article
Full-text available
Tyrannosaurids are hypothesized to be gregarious, possibly parasocial carnivores engaging in cooperative hunting and extended parental care. A tyrannosaurid (cf. Teratophoneus curriei ) bonebed in the late Campanian age Kaiparowits Formation of southern Utah, nicknamed the Rainbows and Unicorns Quarry (RUQ), provides the first opportunity to investigate possible tyrannosaurid gregariousness in a taxon unique to southern Laramidia. Analyses of the site’s sedimentology, fauna, flora, stable isotopes, rare earth elements (REE), charcoal content and taphonomy suggest a complex history starting with the deaths and transport of tyrannosaurids into a peri-fluvial, low-energy lacustrine setting. Isotopic and REE analyses of the fossil material yields a relatively homogeneous signature indicating the assemblage was derived from the same source and represents a fauna living in a single ecospace. Subsequent drying of the lake and fluctuating water tables simultaneously overprinted the bones with pedogenic carbonate and structurally weakened them through wet-dry cycling. Abundant charcoal recovered from the primary bone layer indicate a low temperature fire played a role in the site history, possibly triggering an avulsion that exhumed and reburied skeletal material on the margin of a new channel with minimal transport. Possible causes of mortality and concentration of the tyrannosaurids include cyanobacterial toxicosis, fire, and flooding, the latter being the preferred hypothesis. Comparisons of the RUQ site with other North American tyrannosaur bonebeds (Dry Island-Alberta; Daspletosaurus horneri -Montana) suggest all formed through similar processes. Combined with ichnological evidence, these tyrannosaur mass-burial sites could be part of an emerging pattern throughout Laramidia reflecting innate tyrannosaurid behavior such as habitual gregariousness.
... The lack of ornithischians in the Sao Khua Formation is possibly consistent with the ecological composition of Valanginian/Hauterivian aged assemblages from Gondwana, for example, the "wood beds" of the upper Kirkwood Formation of South Africa and the Bajada Colorado Formation of South America, which are dominated by various-sized theropods and multiple cohabiting species of sauropods (Tucker et al. 2022). This is in contrast to those pre-Barremian dinosaur assemblages of Laurasian landmasses from Europe and North America, which have a diverse ornithischian record including ankylosaurians and iguanodontians and which shared habitats with other sauropod and theropod dinosaurs (Norman 2010;Kirkland et al. 2016;Tucker et al. 2022). By comparison with contemporaneous formations in Asia, ornithischian remains are still somewhat limited (Tucker et al. 2022). ...
Article
Full-text available
Ornithischian dinosaurs have been discovered in Thailand, Laos, and Malaysia. These bird-hipped herbivores remain relatively rare by comparison with saurischian dinosaurs. In the Late Jurassic, stegosaurs and basal neornithischians from Thailand showed similarities to Middle-Late Jurassic taxa from China. Ornithischians appeared in the fossil record again during the late Early Cretaceous (Aptian-Albian) of Thailand and Laos. They are represented by non-hadrosaurid iguanodontians and basal ceratopsians. A few specimens have been reported from poorly dated Early Cretaceous rocks of Malaysia. Here, we illustrate the diversity of ornithischian assemblages in Southeast Asia and discuss their palaeobiogeographical implications.
... Studies of the CMF (Berriasian-Albian) on the east side of the San Rafael Swell have provided important δ 13 C and δ 18 O chemostratigraphic data, U-Pb radiometric dates, vertebrate faunas, and palynomorph assemblages. It has also provided the means for interpreting Cretaceous ter-restrial paleoclimates, depositional systems, and paleoenvironments in present western North America (e.g., Kirkland et al., 1997Kirkland et al., , 2016aKirkland et al., , 2016bLudvigson el al., 2010Ludvigson el al., , 2015C. Suarez et al., 2014C. ...
Article
Full-text available
Previously published anomalous whole-rock stable isotopic values from the Poison Strip Sandstone Member of the Cretaceous Cedar Mountain Formation (CMF) of eastern Utah are of uncertain origins. This study investigated the diagenetic history and the processes responsible for these anomalous data. Accordingly, we integrated photomicroscopic techniques including polarized light microscopy, epifluorescence and cathodoluminescence (CL) imaging, micromilling of stable isotope samples, and fluid-inclusion heating and freezing measurements to this end. The key observations involved the microscopic mapping of calcite cement stratigraphy using CL imaging to permit the analysis of stable isotopes of calcite cements that crystallized during early and late diagenesis. The mapping of calcite cement zones of sufficient submillimeter size to mill out and isolate microgram-sized stable isotope samples enabled this discrimination. Early diagenetic calcite cements have the most positive δ18O values (-10 to -8.5‰ Vienna Pee Dee Belemnite [VPDB]) of all components. The pattern of δ13C and δ18O variation in this early diagenetic cement indicates affinities with early meteoric diagenesis previously documented in published literature on the CFM. The late diagenetic calcite cements yield the most negative δ18O values (-18 to -16‰ VPDB). We interpret the late diagenetic cements to be responsible for the anomalously low whole-rock δ18O values previously reported from the Poison Strip Sandstone Member. Our discoveries of bitumen in late-stage pore fillings and liquid petroleum in the fluid inclusions of late diagenetic calcite cements of the Poison Strip Sandstone Member explain the lower whole-rock organic matter δ13C values and anomalous Δ13C values reported from the unit. Comparatively lower carbonate δ18O and organic δ13C values originally derived from whole-rock analyses of samples from the Poison Strip Sandstone Member resulted from high-temperature basinal diagenesis (hydrothermal circulation and/or petroleum migration), rather than the alternative interpretation of early diagenesis related to a Cretaceous paleoclimatic perturbation. Our results are illustrative of methods to resolve the long-standing geologic problem of discriminating and characterizing products of early vs. late diagenesis in terrigenous clastic sedimentary strata.
... Unfortunately, no detailed taphonomical study was conducted in the Yellow Cat Quarry, thus the mass mortality origin of this material is assumed based on the following features: (1) the ankylosaur individuals were discovered in a thin siltstone layer (Kirkland et al., 2008); (2) the skeletal parts were situated close to each other within an area of approximately 30 m 2 (Kinneer et al., 2016); (3) the bonebed is clearly dominated by Gastonia, whereas other vertebrate remains are only subordinate (Kirkland et al., 1999); (4) the bone-bearing strata were deposited in an ephemeral lake or pond under arid to semiarid conditions with monsoonal overprinting (Kirkland et al., 2016), which conditions often result in MDAs during the seasonal drought events (e.g. Conybeare and Haynes, 1984;Rogers, 1990;Fiorillo et al., 2000;Gates, 2005;Varrichio et al., 2008). ...
Article
Gregarious behaviour of large bodied herbivorous dinosaurs, such as ceratopsians, hadrosaurs and sauropods, has received much attention due to their iconic mass death assemblages (MDAs). Yet, social lifestyle of ankylosaurs, a highly specialized group of armoured herbivores that flourished predominantly during the Cretaceous Period, remains largely ambiguous. Whereas most ankylosaurs are found as isolated individuals, which may suggest a dominantly solitary lifestyle, the few examples of ankylosaur MDAs indicate that some members of this clade could have been gregarious. In this review, we assess taphonomic history, ontogenetic composition of the MDAs, defence system and other comparative anatomical attributes, and inferred habitat characteristics of ankylosaurs; aspects that may indicate and/ or influence group formation in extant herbivores and can also be studied in fossils. We show that the ankylosaurian gross anatomy, such as their heavy armour, barrel-shaped body and usually stocky limbs, combined with the rarity of their MDAs and multiple parallel trackways, all suggest a solitary adult life with efficient anti-predator defence system, limited agility, and confined foraging range. However, characteristics of the known MDAs of Pinacosaurus, Gastonia, and the Iharkút nodosaurids evaluated in this study imply that at least some ankylosaurs formed groups. Nevertheless, we found no common and consistent set of features to explain why these particular ankylosaurs were gregarious. While inefficient anti-predator defence along with likely higher agility of juvenile Pinacosaurus living in open habitats could account for their gregarious behaviour, such ontogenetic, anatomical and habitat features are not combined either in Gastonia or in the Iharkút nodosaurid MDAs. Instead, members of each MDA likely had their own specific conditions driving them to form relatively small herds, indicating a more complex social structuring in ankylosaurs than previously acknowledged. Studying morphological and functional disparity within Ankylosauria may help explain the repertoire of their social behaviour. Our holistic approach shows that combining palaeontological and biological information is essential and can provide new insights into the behavioural ecology of long extinct vertebrates.
... Megapnosaurus (Raath, 1977), the basal neotheropod Liliensternus (MB R.2175), the noasaurid Masiakasaurus ( Carrano et al., 2002; Figure 26.5), the spinosaurid Spinosaurus (NHMUK VP R.16421), adult individuals of the dromaeosaurid Utahraptor (UMNH VP 20501; Kirkland et al., 2016, figure 9B) and the troodontid Daliansaurus (Shen et al., 2017a, figures 3 and 4A). Among Tyrannosauroidea, the proceratosaurid Proceratosaurus (Rauhut et al., 2010; Figure 26.6) and a single specimen of the basal pantyrannosaurian Dilong (IVPP V11979) also show this condition. ...
Article
Full-text available
Isolated theropod teeth are some of the most common fossils in the dinosaur fossil record and are continually reported in the literature. Recently developed quantitative methods have improved our ability to test the affinities of isolated teeth in a repeatable framework. But in most studies, teeth are diagnosed on qualitative characters. This can be problematic because the distribution of theropod dental characters is still poorly documented, and often restricted to one lineage. To help in the identification of isolated theropod teeth, and to more rigorously evaluate their taxonomic and phylogenetic potential, we evaluated dental features in two ways. We first analyzed the distribution of 34 qualitative dental characters in a broad sample of taxa. Functional properties for each dental feature were included to assess how functional similarity generates homoplasy. We then compiled a quantitative data matrix of 145 dental characters for 97 saurischian taxa. The latter was used to assess the degree of homoplasy of qualitative dental characters, address longstanding questions on the taxonomic and biostratigraphic value of theropod teeth, and explore the major evolutionary trends in the theropod dentition. In smaller phylogenetic datasets for Theropoda, dental characters exhibit higher levels of homoplasy than non-dental characters, yet they still provide useful grouping information and optimize as local synapomorphies of smaller clades. In broader phylogenetic datasets, the degree of homoplasy displayed by dental and non-dental characters is not significantly different. Dental features on crown ornamentations, enamel texture, and tooth microstructure have significantly less homoplasy than other dental features and can be used to identify many theropod taxa to ‘family’ or ’sub-family’ level, and some taxa to genus or species. These features should, therefore, be a priority for investigations seeking to classify isolated teeth. Our observations improve the taxonomic utility of theropod teeth and in some cases can help make isolated teeth useful as biostratigraphic markers. This proposed list of dental features in theropods should, therefore, facilitate future studies on the systematic paleontology of isolated teeth.
... Joeckel et al. (2017) identified well-developed alluvial palaeosols through most of a section of the Yellow Cat Member at locality 2 of the present paper (Fig. 1a). Locality 3 (Fig. 1a) is Utahraptor Ridge or Stikes Quarry, a dinosaur fossil site at which there is evidence of an ancient artesian spring system that figured prominently in the site's taphonomic origins (Kirkland et al. 2016a). The new palynoflora described in this study was collected from locality 4 (Fig. 1a). ...
Article
Full-text available
This paper presents breakthroughs in the chronostratigraphy of the heretofore poorly constrained Yellow Cat Member of the Cedar Mountain Formation, which is an important record of terrestrial environments, ecosystems and global change in the ancient North American Cordilleran foreland. Zircon populations from 10 stratigraphic horizons in the Yellow Cat Member yield youngest single-grain ages ranging from 142.5+2.7 to 133.7+2.7 Ma (Berriasian–late Valanginian); those from one mudstone palaeosol yield a robust Concordia Age of 136.3+1.3 (Valanginian). Additionally, a new palynoflora – one of a few to be published from the Cedar Mountain Formation – is assigned to the middle Berriasian to early Hauterivian stages, based on the presence of Foraminisporis wonthaggiensis and Trilobosporites sp. cf. T. canadensis, and the absence of F. asymmetricus, Appendicisporites spp. and angiosperms. Furthermore, these chronostratigraphic data allow us to interpret part of the so-called ‘Weissert Event’ C-isotope excursion (Valanginian) in a new C-isotope profile through a palaeosol-bearing alluvial succession in the Yellow Cat Member. This research extends a firm understanding of the formation further back into the Early Cretaceous than was the case previously (except for ostracod biostratigraphy) and sets the stage for future advancements.
Article
We describe the sedimentology, geochronology, and geochemistry of the Early Cretaceus Sao Khua Formation of the Khorat Basin, northeastern Thailand, and report a temporal range adjustment for its dinosaurian assemblage. Facies analysis and architectural studies reveal that sedimentation occurred within a floodplain setting fed by large meandering bedload-rich channels. Interfluve areas comprised freshwater lakes and emergent areas subject to pedogenic modification. Multiple paleosol types are identified and geochemistry is indicative of a stable humid subtropical climate regime. Based on radiometric dating of detrital zircons (via LA-ICP-MS), we interpret that the middle part of the Sao Khua Formation was deposited no later than 133.8 (±1.8) Ma (late Valanginian), and grain ages collected from the overlying lowermost Phu Phan Fm constrain sedimentation of the upper part of the Sao Khua Formation to no earlier than 132.4 (±2.0) Ma (early Hauterivian). In consideration of the Early Cretaceous regional tectonic framework, we interpret that youthful igneous zircon grains are derived from the adjacent South China-Vietnam South Borneo Volcanic Arc. We establish that the entombed dinosaur biota (including members of the Ornithomimosauria, Spinosauridae, Megaraptora, and Somphospondylia) is ~5–9 million years older than previously recognized and that these records are among the oldest known globally for these clades. Constraining the age of the Sao Khua Formation indicates that the shift from sauropod-dominated, ornithischian depauperate ecosystems of the Sao Khua Formation to iguanodontian-rich ecosystems of the Khok Kruat Formation occurred sometime between the early Hauterivian and Aptian on the Khorat Plateau.
Article
The image of the highly intelligent, pack-hunting raptor has become engrained in scientific literature and popular works alike. First proposed to explain the relatively common co-occurrence of the large-bodied iguanodontian Tenontosaurus tilletti and the wolf-sized Deinonychus antirrhopus from the Lower Cretaceous of North America, a canid-like social hunting structure has become the standard depiction of dromaeosaurs in popular works over the last three decades. This reconstruction is, however, problematic largely due to the fact that highly coordinated hunting strategies are rarely observed in modern archosaurs. This has led to the alternative hypothesis that D. antirrhopus was more analogous to agonistic reptilian carnivores, like the Komodo dragon (Varanus komodoensis). Among the many differences between these two analogs is how social and asocial organisms rear their young, producing a diagnostic pattern based on the presence or absence of ontogenetic dietary changes. In order to test for dietary changes through growth, stable carbon and oxygen isotope (δ¹³C, δ¹⁸O) analysis was performed on tooth carbonate from small (<4.5 mm crown height) and large (>9 mm crown height) D. antirrhopus specimens from two microsites from the Lower Cretaceous Cloverly (Montana) and Antlers (Oklahoma) formations. Teeth from goniopholidid crocodylians and Tenontosaurus tilletti from the Cloverly Formation were also tested for comparison. The results show that the Cloverly goniopholidids, like their modern counterparts, went through a distinct transition in diet as they grew. The smallest teeth were the relatively most enriched in ¹³C (mean = −9.32‰; n = 5), the medium-sized teeth were the most-depleted in ¹³C (mean = −10.56‰; n = 5), and the largest teeth were intermediate (mean = −10.12‰; n = 6). These factors are characteristic of the dietary shifts seen in modern asocial reptiles. D. antirrhopus showed this same pattern in tooth samples collected from both rock units, with small teeth being the more enriched in ¹³C (mean = −8.99‰; n = 10) and the large teeth being more depleted in ¹³C (mean = −10.38‰; n = 10). These differences suggest that juvenile and adult D. antirrhopus from both formations likely consumed different prey. Hypothetical food sources, such as T. tilletti, are close to the ¹³C isotopic signal of adult D. antirrhopus, consistent with the hypothesized trophic relationship (predator-prey) between these two species. Juvenile D. antirrhopus had a diet more enriched in ¹³C, likely composed of smaller-bodied, but trophically-higher species. Taken together, these data add to the growing evidence that D. antirrhopus was not a complex social hunter by modern mammalian standards.
Article
Full-text available
Although only recognized as a discrete stratigraphic unit since 1944, the Cedar Mountain Formation represents tens of millions of years of geological and biological history on the central Colorado Plateau. This field guide represents an attempt to pull together the results of recent research on the lithostratigraphy, chronostratigraphy, sequence stratigraphy, chemostratigraphy, and biostratigraphy of these medial Mesozoic strata that document the dynamic and complex geological history of this region. Additionally, these data provide a framework by which to examine the history of terrestrial faunas during the final breakup of Pangaea. In fact, the medial Mesozoic faunal record of eastern Utah should be considered a keystone in understanding the history of life across the northern hemisphere. Following a period of erosion and sediment bypass spanning the Jurassic–Cretaceous boundary, sedimentation across the quiescent Colorado Plateau began during the Early Cretaceous. Thickening of these basal Cretaceous strata across the northern Paradox Basin indicate that salt tectonics may have been the predominant control on deposition in this region leading to the local preservation of fossiliferous strata, while sediment bypass continued elsewhere. Thickening of overlying Aptian strata west across the San Rafael Swell provides direct evidence of the earliest development of a foreland basin with Sevier thrusting that postdates geochemical evidence for the initial development of a rain shadow.
Chapter
Full-text available
The first occurrence of an iguanodontian from the Lower Cretaceous Poison Strip Member of the Cedar Mountain Formation, Utah, is described and named. This new taxon is represented by a well-preserved ilium, femora, tibiae, and vertebrae, as well as other material. The fe-mora are typical for an ornithopod, but the ilium has a short, horizontal postacetabular process that is functionally an antitrochanter.
Article
Earthquakes along convergent plate boundaries commonly occur in sequences that are complete within 1 yr, and may include 8–10 events strong enough to generate sand blows. Dune crossbeds within the Jurassic Navajo Sandstone of Utah (western United States) enclose intact and truncated sand blows, and the intrusive structures that fed them. We mapped the distribution of more than 800 soft-sediment dikes and pipes at two small sites. All water-escape structures intersect a single paleo-surface, and are limited to the upper portion of the underlying set of cross-strata and the lower portion of the overlying set. A small portion of one set of crossbeds that represents ~1 yr of dune migration encloses eight generations of eruptive events. We interpret these superimposed depositional and deformational structures as the record of a single shock-aftershock earthquake sequence. The completeness and temporal detail of this paleoseismic record are unique, and were made possible when sand blows repeatedly erupted onto lee slopes of migrating dunes. Similar records should be sought in modern dunefields with shallow water tables.
Article
Examination of the silty sediments in the lower Van Normal reservoir after the 1971 San Fernando, California earthquake revealed three zones of deformational structures in the 1-m-thick sequence of sediments exposed over about 2 km² of the reservoir bottom. These zones are correlated with moderate earthquakes that shook the San Fernando area in 1930, 1952, and 1971. The success of this study, coupled with the experimental formation of deformational structures similar to those of the Van Norman reservoir, led to a search for similar structures in Pleistocene and Holocene lakes and lake sediments in other seismically active areas. Thus, studies have been started in Pleistocene and Holocene silty and sandy lake sediments in the Imperial Valley, southeastern California; Clear Lake, in northern California; and the Puget Sound area of Washington. The Imperial Valley study has yielded spectacular results: five zones of structures in the upper 10 m of Late Holocene sediments near Brawley have been correlated over an area of approximately 100 km², using natural outcrops. These structures are similar to those of the Van Norman reservoir and are interpreted to represent at least five moderate to large earthquakes that affected the southern Imperial Valley area during Late Holocene time. The Clear Lake study has provided ambiguous results with respect to determination of earthquake recurrence intervals because the cores studied are in clayey rich in organic material sediments that have low liquefaction potential. A study of Late Pleistocene varved glacio-lacustrine sediments has been started in the Puget Sound area of Washington, and thirteen sites have been examined. One has yielded 18.75 m of sediments that contains 1,804 varves and fourteen deformed zones interpreted as being caused by earthquake, because they are identical to structures formed experimentally by simulated seismic shaking. Correlation of deformational structures with seismic events is based on: (1) proximity to presently active seismic zones; (2) presence of potentially liquéfiable sediments; (3) similarity to structures formed experimentally; (4) small-scale internal structures within deformed zones that suggest liquefaction; (5) structures restricted to single stratigraphie intervals; (6) zones of structures correctable over large areas; and (7) absence of detectable influence by slopes, slope failures, or other sedimentological, biological, or deformational processes. © 1975, Elsevier Scientific Publishing Company
Article
I am most grateful to my colleagues in the ASCE Geotechnical Division for their invitation to prepare this state-of-the-art paper on soil liquefaction during earthquakes. It is a great compliment to many colleagues and friends who through their original ideas, cooperation, and invaluable critiques have helped to raise the state of knowledge in this field over the past decade to a condition where engineers can practice in an extremely difficult area with some reasonable level of confidence.
Article
Represents an update and expansion of a set of course notes previously published in 1975. Topics covered include: bed configurations; deposition and stratification; stratification sequences; fluvial deposits and facies models; conglomerates - emphasising fluvial and alluvial fan environments; stratification and sequence in prograding shoreline deposits; shallow marine environments - a comparison of some ancient and modern environments.-C.T.Baldwin