high-frequency tectonic sequences in the campanian

61
University of New Orleans University of New Orleans ScholarWorks@UNO ScholarWorks@UNO University of New Orleans Theses and Dissertations Dissertations and Theses Spring 5-13-2016 High-frequency tectonic sequences in the Campanian Castlegate High-frequency tectonic sequences in the Campanian Castlegate Formation during a transition from the Sevier to Laramide Formation during a transition from the Sevier to Laramide orogeny, Utah, U.S.A. orogeny, Utah, U.S.A. David B. Cross University of New Orleans, [email protected] Follow this and additional works at: https://scholarworks.uno.edu/td Part of the Geology Commons, Stratigraphy Commons, and the Tectonics and Structure Commons Recommended Citation Recommended Citation Cross, David B., "High-frequency tectonic sequences in the Campanian Castlegate Formation during a transition from the Sevier to Laramide orogeny, Utah, U.S.A." (2016). University of New Orleans Theses and Dissertations. 2133. https://scholarworks.uno.edu/td/2133 This Thesis is protected by copyright and/or related rights. It has been brought to you by ScholarWorks@UNO with permission from the rights-holder(s). You are free to use this Thesis in any way that is permitted by the copyright and related rights legislation that applies to your use. For other uses you need to obtain permission from the rights- holder(s) directly, unless additional rights are indicated by a Creative Commons license in the record and/or on the work itself. This Thesis has been accepted for inclusion in University of New Orleans Theses and Dissertations by an authorized administrator of ScholarWorks@UNO. For more information, please contact [email protected].

Upload: others

Post on 04-Nov-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: High-frequency tectonic sequences in the Campanian

University of New Orleans University of New Orleans

ScholarWorks@UNO ScholarWorks@UNO

University of New Orleans Theses and Dissertations Dissertations and Theses

Spring 5-13-2016

High-frequency tectonic sequences in the Campanian Castlegate High-frequency tectonic sequences in the Campanian Castlegate

Formation during a transition from the Sevier to Laramide Formation during a transition from the Sevier to Laramide

orogeny, Utah, U.S.A. orogeny, Utah, U.S.A.

David B. Cross University of New Orleans, [email protected]

Follow this and additional works at: https://scholarworks.uno.edu/td

Part of the Geology Commons, Stratigraphy Commons, and the Tectonics and Structure Commons

Recommended Citation Recommended Citation Cross, David B., "High-frequency tectonic sequences in the Campanian Castlegate Formation during a transition from the Sevier to Laramide orogeny, Utah, U.S.A." (2016). University of New Orleans Theses and Dissertations. 2133. https://scholarworks.uno.edu/td/2133

This Thesis is protected by copyright and/or related rights. It has been brought to you by ScholarWorks@UNO with permission from the rights-holder(s). You are free to use this Thesis in any way that is permitted by the copyright and related rights legislation that applies to your use. For other uses you need to obtain permission from the rights-holder(s) directly, unless additional rights are indicated by a Creative Commons license in the record and/or on the work itself. This Thesis has been accepted for inclusion in University of New Orleans Theses and Dissertations by an authorized administrator of ScholarWorks@UNO. For more information, please contact [email protected].

Page 2: High-frequency tectonic sequences in the Campanian

High-frequency tectonic sequences in the Campanian Castlegate

Formation during a transition from the Sevier to Laramide orogeny,

Utah, U.S.A.

A thesis

Submitted to the Graduate Faculty of the

University of New Orleans

in partial fulfillment of

the requirements for the degree of

Master of Science

In

Earth and Environmental Sciences

by

David Cross

B.S. University of Kentucky, 2012

May, 2016

Page 3: High-frequency tectonic sequences in the Campanian

ii

Copyright

By David B. Cross

2016

Page 4: High-frequency tectonic sequences in the Campanian

iii

ACKNOWLEDGEMENTS

First and foremost, I thank my thesis supervisor Dr. M. Royhan Gani for his technical,

financial and motivational support. Above all, I am grateful for him ‘taking a chance’ on me as a

graduate student. Without him, I would likely not be in this scenario of obtaining higher

education. I am also indebted to my fellow Kentucky alumnus Dr. Mark Kulp who has provided

technical support and ‘opportunities’ to expand my geoscientific knowledge and experience.

While at UNO, I perhaps learned most from Dr. Ioannis Georgiou in various process-oriented

sedimentological courses. His teaching of sedimentary processes which were my Achilles heels,

tremendously facilitated better understanding of the ancient record during data collection for this

thesis. I am very grateful for Corey Hinyup and CanUSA for their generation donation of

training and rope access equipment that was utilized collecting data for this thesis. I thank Corey

Hinyup, Everett Leslie, Godspower Onyenanu and Hiranya Sahoo for their assistance with field

data collection. I extend additional gratitude to my family and especially Mark and Cindy Cross

for their unrelenting patience and moral support throughout my tenure with academia.

Page 5: High-frequency tectonic sequences in the Campanian

iv

TABLE OF CONTENTS

List of Figures ..................................................................................................................................v

Abstract .......................................................................................................................................... vi

Introduction ......................................................................................................................................1

Regional geology ..........................................................................................................................4

The Cordilleran orogenic system..............................................................................................5

Campanian stratigraphy of central and eastern Utah ................................................................8

Methods ......................................................................................................................................12

Results ............................................................................................................................................14

Facies ..........................................................................................................................................14

Strike-variability of stratigraphy ................................................................................................15

Paleocurrent analysis ..................................................................................................................17

Sandstone petrography ...............................................................................................................19

Sandstone mineralogy and provenance ..................................................................................20

Sandstone diagenesis ..............................................................................................................21

Discussion ......................................................................................................................................23

Manifestation of deformation-style in stratigraphic surfaces .....................................................23

Manifestation of deformation-style in drainage patterns and provenance .................................27

Origin of orogeny .......................................................................................................................29

Conclusions ....................................................................................................................................35

References ......................................................................................................................................36

Appendix ........................................................................................................................................42

Regional geology (extended) .....................................................................................................42

Pre-cordillera sedimentary cover and tectonics ......................................................................42

The Cordilleran orogenic system (extended) .........................................................................43

Campanian stratigraphy of central and eastern Utah (extended) ............................................45

Results (extended) ......................................................................................................................47

Facies (extended) ....................................................................................................................47

Discussion (extended) ................................................................................................................52

Manifestation of deformation-style in stratigraphic surfaces (extended) ...............................52

Vita .................................................................................................................................................54

Page 6: High-frequency tectonic sequences in the Campanian

v

LIST OF FIGURES

Figure 1: Regional geology of western North America ...................................................................2

Figure 2: Regional geology and study area in Utah .........................................................................3

Figure 3: Campanian stratigraphy in central and eastern Utah ........................................................9

Figure 4: Along-strike sequence stratigraphy of Castlegate Formation ........................................16

Figure 5: Paleocurrent analysis in Castlegate Formation ..............................................................18

Figure 6: Quartz-Feldspar-Lithics diagram of Castlegate Formation sandstones ........................19

Figure 7: Thin section photo micrographs ....................................................................................20

Figure 8: Stratigraphic-architecture analysis of Castlegate Formation at Straight Canyon ...........23

Figure 9: Stratigraphic-architecture analysis of Castlegate Formation at Deadman’s Creek ........24

Figure 10: 3-D sequence stratigraphic correlation of Castlegate Formation .................................25

Figure 11: Late Cretaceous paleogeography of the Cordilleran orogenic system .........................30

Figure 12: Late Cretaceous cross-section through the Cordilleran orogenic system ....................31

Page 7: High-frequency tectonic sequences in the Campanian

vi

ABSTRACT

Though stratigraphic correlations are abundant in the Cordilleran basin-fill, they rarely

include along-strike transects providing a spatio-temporal sense of deformation, sediment-supply

and subsidence. A new, high-resolution, regional strike-correlation of the Castlegate Formation

reveals progressive northward-growth of the San Rafael Swell during two embryonic episodes of

Laramide-style deformation in central Utah. The intrabasinal deformation-events produced

gentle lithospheric-folding punctuated by erosional-truncation of upwarped regions. The earliest

episode occurred at 78 Ma in the southern San Rafael Swell likely causing soft-sediment

deformation and stratal-tilting. Following this the alluvial-plain was leveled and rapid, extensive-

progradation took place.

A second episode, at 75 Ma, where deformation was focused in the northern San Rafael

Swell, also caused sediment-liquefaction and erosional beveling. The stratal-tilting and sediment-

liquefaction is attributed to seismicity induced by basal-traction between a subducting flat-slab

and continental-lithosphere. The south-to north time-transgression of uplift is spatio-temporally

consistent with NE-propagation of an oceanic-plateau subducted shallowly beneath the region.

Key Words: Castlegate Sandstone, San Rafael Swell, Wasatch Plateau, Laramide Orogeny,

Sevier Orogeny, Cordilleran foreland basin

Page 8: High-frequency tectonic sequences in the Campanian

1

INTRODUCTION

Clues useful to understanding the timing and style of orogenesis can be found within a

foreland basin-fill (Miall and Arush, 2001; Ettensohn, 2004; Aschoff and Steel, 2011). Foreland

stratigraphic architecture can provide insight about tectonic conditions as it reflects the balance

of sediment supply and accommodation change, which can be driven by tectonism (Martinsen

and Helland-Hansen, 1995). It is impractical, however, to make regional predictions on the basis

of stratal stacking pattern alone (Martinsen and Helland-Hansen, 1995), as crustal upwarping and

associated downwarping yield contrasting yet coeval stacking pattern (Krystinik and DeJarnett,

1995). Foreland basins form viscoelastically with often poorly defined shelf breaks in response

to superposed isostatic and flexural subsidence patterns (Jordan, 1981; Martinsen and Helland-

Hansen, 1995). With increasing subsidence and sediment supply towards the orogenic-load,

foreland basins have a characteristic asymmetric cross-sectional profile skewed towards the

orogen (Martinsen et al., 1993). Along-strike, foreland basin subsidence and sediment supply can

also vary if there is tectonic-diachroneity along the length of an orogeny (Ettensohn, 2004),

localized intra-foreland uplift (Heller et al., 1993), or multiple sediment sources (Lawton and

Bradford, 2011). Bounding surface, not stratal stacking pattern (e.g. systems tract), correlation is

useful for resolving tectonic timing and style, which can be constrained temporally using tools

such as biostratigraphy (Krystinik and DeJarnett, 1995).

Investigating orogenic style and timing is a possibility for the tectonically-sensitive

nonmarine deposits in the North American Cordilleran foreland basin (Fig. 1). Particularly

noteworthy are the unconformity-bounded, coarse-grained, anomalously progradational and

extensive sheet-like deposits (Figs. 2 and 3) of the Upper Cretaceous (Campanian) Castlegate

Sandstone in Utah (Krystinik and DeJarnett, 1995; Aschoff and Steel, 2011). It is well-known

Page 9: High-frequency tectonic sequences in the Campanian

2

that coarse-

grained detritus is

delivered to

foreland basins in

pulses but the

timing and

mechanism of

which are highly

contested. Early

research (Spieker,

1946) attributed

the extensive and

rapid Castlegate

Sandstone

progradation to

reflect

synchronous

loading and

erosion in the

Sevier fold-and-

thrust belt.

Subsequent basin-fill architecture models (Heller et al., 1988; Blair and Bilodeau, 1988;

Flemings and Jordan, 1990) suggested that coarse-grained sediment-pulses to foreland basins

Figure 1. A. Late Cretaceous

paleogeographic reconstruction of

North America modified from Gani et

al. (2015). B. Map displaying the

present day location of major

Cordilleran orogenic elements in

western North America (modified

from DeCelles and Coogan, 2006).

CNTB—Central Nevada thrust belt;

FC—Franciscan complex; GVF—

Great Valley forearc; IB—Idaho

batholith; LFTB—Luning-

Fencemaker thrust belt; SAF—San

Andreas fault; SFTB—Sevier fold-

and-thrust-belt; SNB—Sierra Nevada

batholith.

Page 10: High-frequency tectonic sequences in the Campanian

3

rather reflect post-tectonic isostatic adjustments such that coarse-grained detritus becomes

trapped proximally where subsidence is exceptionally high as thrust sheets are stacking.

However, in a 2-D seismic line transecting the Sevier fold-and-thrust belt and proximal

Cordilleran foreland basin, Horton et al. (2004) revealed rapid, lengthy Castlegate Sandstone

progradation coeval with movement along the Charleston-Nebo thrust and triangle zone (Fig.

2A). Therefore, somehow Cordilleran foreland basin subsidence was reduced enabling

Castlegate Sandstone progradation despite active loading in the adjacent Sevier belt.

With the knowledge that thrust-loading increases accommodation and inhibits

progradation, there must have been additional but accommodation-reducing forces responsible

for the sheet-like distribution and anomalous progradation of the Castlegate Sandstone. Van

Wagoner (1995) favored a falling eustatic sea-level hypothesis, which is unlikely because the

Late Cretaceous had greenhouse conditions associated with low-amplitude sea-level fluctuations

that would

have been

overprinted

by

autogenic

or tectonic

processes

(Vakarelov

et al.,

2006).

Many

Figure 2. A. Map showing generalized distribution of Cordilleran orogenic elements, Upper

Cretaceous strata, and Sevier thrust-fault systems in Utah. CN—Charleston-Nebo thrust; CR—

Canyon Range thrust; GU—Gunnison thrust; PV—Pavant thrust; PX—Paxton thrust; TZ—

triangle zone. The map was modified from DeCelles and Coogan (2006) and Lawton and

Bradford (2011). B. Shaded-relief map of the study area showing locations of measured sections

and line of correlated sections of Figure 4.

Page 11: High-frequency tectonic sequences in the Campanian

4

(Lawton, 1986; Van Wagoner, 1995; Miall and Arush, 2001; Aschoff and Steel, 2011) have

hinted at the possibility of incipient Laramide deformation driving the accommodation-reduction

based on architecture, provenance, and paleocurrent deflection but critical evidence supporting

this hypothesis in the form of stratigraphic surfaces with tilting is missing.

The aim of this study is to address two key questions regarding Cordilleran foreland basin

stratigraphy: (1) how is deformation style and timing manifested in sequence stratigraphic

surfaces and, (2) what caused the transition from Sevier-to Laramide-style deformation?

Answering these questions necessitates a high-resolution regional correlation and sequence

stratigraphic analysis of the basin-fill, and a spatio-temporal kinematic understanding of the

Cordilleran orogenic system. A new, regional (70-km-long), proximal, and basin-parallel (NE-

SW oriented) correlation is presented here in the eastern Wasatch Plateau and northwestern Book

Cliffs. The correlation is age-constrained with published radiometrically-dated ammonite

biostratigraphy.

Regional Geology

Upper Cretaceous sedimentary units in the Book Cliffs and Wasatch Plateau were

deposited on the western margin of the Cretaceous Western Interior Seaway during Campanian

(Fig. 1A) (Spieker, 1949). As evidenced by flora and fauna fossils, central and eastern Utah was

humid and sub-tropical during Campanian (Kauffman, 1977). The climate and tectonic

conditions yielded an exceptionally-high sediment supply (Aschoff and Steel, 2011). Late

Cretaceous experienced a greenhouse climate, and in western North America majority of the

region’s annual precipitation likely occurred mid-year from west verging monsoons (Fricke et

al., 2010). Sediments were mostly sourced from the Sevier fold-and-thrust belt of the Cordilleran

orogenic system and transported in a general eastward (basin-transverse) direction, filling the

Page 12: High-frequency tectonic sequences in the Campanian

5

north-south elongate retroarc Cordilleran foreland basin (Fig. 1B) (Kauffman, 1977).

Episodically, Jurassic magmatic arc terranes from the Cordilleran retroarc hinterland (Fig. 1B)

farther to the west and southwest constituted the sediment supply (Lawton and Bradford, 2011;

Hampson, 2014).

The Cordilleran orogenic system

An orogenic front-perpendicular transect of the Cordilleran orogenic system (e.g.

California to Utah) contains a forearc accretionary complex and basin, magmatic arc, retroarc

hinterland, fold-thrust belt, foreland basin, and Laramide foreland belt (Fig. 1B). The orogeny

evolved time-transgressively from west to east through protracted Middle Mesozoic to Early

Cenozoic paleo-Pacific plate collision with and subduction beneath the North American

continent (Yonkee and Weil, 2015).

The Cordilleran fold-thrust belt (Fig. 1B) formed during the Cretaceous Sevier orogeny.

Sevier thin-skinned deformation was caused by eastward high-angle (~30°) subduction of the

Farallon plate beneath the western margin of North America’s Cordilleran magmatic arc

(DeCelles and Coogan, 2006; Yonkee and Weil, 2015). The fold-thrust belt translated the

Paleozoic and Lower to Middle Mesozoic sedimentary cover eastward along basal décollement

surfaces (Yonkee and Weil, 2015). Western Sevier thrusts slipped above Neoproterozoic (~720-

660 Ma) syn-rift micaceous strata, whereas eastern thrusts translated across middle Cambrian

shale and limestone (DeCelles and Coogan, 2006; Yonkee and Weil, 2015). The eastward-

migration of the orogenic front stalled with a critical tapering of thrust sheets in central Utah at

the Wasatch Hinge (Figs. 1B and 2A) (DeCelles and Coogan, 2006; Yonkee and Weil, 2015).

Shortening in triangle zones formed backthrusts, duplexes, and thrust-tip anticlines at the

orogenic front (DeCelles et al., 1995). By Late Cretaceous, the Sevier structural front was in

Page 13: High-frequency tectonic sequences in the Campanian

6

central Utah (Fig. 2A), ~ 700 km from the Farallon trench (DeCelles and Coogan, 2006; Fig.

1B). Middle Jurassic carbonates and evaporites formed detachments in eastern Sevier thrusts

(DeCelles and Coogan, 2006; Yonkee and Weil, 2015). In Utah, there were four major detached

thrust fault systems (Fig. 2A), the Canyon Range, Pavant, Paxton and Gunnison thrusts, which

collectively contributed to ~220 km of crustal shortening (DeCelles and Coogan, 2006).

The early to middle Cretaceous Canyon Range thrust system was the first of the four to

develop. Activity of the Canyon Range thrust was ensued by shortening of the Pavant thrust

system (DeCelles and Coogan, 2006). These two thrusts accommodated most of the shortening

in the Sevier belt as they had the deepest detachment levels (Constenius, 1988). Motion in the

Charleston-Nebo thrust and triangle zone, which exposed Pennsylvanian Oquirrh Group

quartzite and supplied a large portion of sediments deposited as the Castlegate Sandstone, ceased

sometime during Campanian (Constenius, 1998; Horton et al., 2004). Eastward translation of the

Charleston-Nebo thrust stalled in a triangle zone at the Sevier front promoting crustal contraction

and generation of the Santaquin Culmination consisting of fault-bend-folds and east-dipping

backthrusts (Constenius, 1998). The Paxton thrust is the least described of the four major thrust

systems because it is not exposed at the surface such that known information regarding the

kinematic evolution and unroofing history stems entirely from subsurface data, which are mostly

proprietary (DeCelles and Coogan, 2006). The Paxton thrust likely activated during Santonian

and began shortening during Campanian (DeCelles and Coogan, 2006). These duplexes were

eroded, shedding mostly fine-grained detritus from Jurassic terrane (DeCelles and Coogan,

2006). The Gunnison thrust, the latest of the four, was significantly buried by sediments in the

wedge-top depozone and was predominantly a Maastrichtian event (DeCelles and Coogan,

2006). Sediments entered the Cordilleran foreland basin likely within fluvial entry points

Page 14: High-frequency tectonic sequences in the Campanian

7

adjacent to growth structures, such as between the Charleston-Nebo and Paxton thrusts

(Hampson et al., 2012).

The Cordilleran foreland basin formed coeval to and principally because of loading in the

Sevier thrust belt (DeCelles and Coogan, 2006). Foreland basins migrate orogen-perpendicularly

and are lithospheric depressions formed cratonward of fold-thrust belts (Jordan, 1981; Ettensohn,

2004). Models attempting to recreate the process of foreland basin formation were unsuccessful

until they assumed a lithosphere having viscoelastic instead of elastic rheology (Beaumont,

1981; Jordan, 1981). From proximal to distal, foreland basins have four major components—

wedge-top, foredeep, forebulge and backbulge (Jordan, 1981). The migration and loading of a

thrust-wedge cause cratonward lithospheric downbending creating a proximal foredeep, and

basinward upwarping creating a distal forebulge (Jordan, 1981). Foreland basin-fills in cross-

section therefore appear asymmetric, with thicker sedimentary accumulations towards the

orogenic load.

The Cordilleran Laramide foreland belt (Fig. 1B) formed during the Late Cretaceous to

Eocene Laramide orogeny, which partitioned the foreland basin into smaller sub-basins by

localized basement-involved uplifts (Cross, 1986). The uplifts occurred along reverse faults with

deeper detachment levels than in the Sevier fold-thrust belt. The Laramide orogeny is

hypothesized to record a transition to shallower and more-rapid Farallon plate subduction than

during the Sevier Orogeny (Coney and Reynolds, 1977; Cross, 1986; Liu et al., 2010; Painter

and Carrapa, 2013; Yonkee and Weil, 2015). The transition at ca. 81 Ma is also coincident with a

major shutdown of arc magmatism (DeCelles, 2015; Yonkee and Weil, 2015). The decreased

subduction angle has been attributed to increased buoyancy of the subducting slab (Liu et al.,

2010; Painter and Carrapa, 2013). This portion of the Farallon plate is often referred to as the

Page 15: High-frequency tectonic sequences in the Campanian

8

Shatsky Rise (Liu et al., 2010), which was likely an aseismic ridge like the Juan Fernandez Rise

of the South American Plate that triggered basement-cored uplifts of the Sierra Pampeanas far

inboard of the Andean subduction zone (Jordan and Allmendinger, 1986; Aschoff and Steel,

2011). Laramide deformation occurred with an early and late phase that varied in style (Cross,

1986), as much as ~700 km cratonward of the Franciscan subduction zone. During the early pre-

Paleocene phase, deformation was relatively widespread, defined by gentle folding and flexure

(Cross, 1986; Marshak et al., 2000). This embryonic phase of deformation began during

Campanian, arguably coeval to deposition of the Castlegate Formation (Van Wagoner, 1995;

Miall and Arush, 2001; Aschoff and Steel, 2011). During the late phase subsequent to the Early

Paleocene and to an apparent ~5 m.y. pause in deformation, structures that developed were more

localized and brittle, characterized by large-offset fault-bounded uplifts commonly in extant

weakness zones (Cross, 1986).

Campanian stratigraphy of central and eastern Utah

Strata have pristine and nearly continuous exposures in the Book Cliffs and eastern

Wasatch Plateau but are difficult to access as sandstones form prominent cliffs between steep

scree-covered slopes. Spieker (1949) noted, “…the rocks are easy to see but hard to reach”.

These rocks are the regressive Mesaverde Group, which, from central to eastern Utah, form a

series of basinward-tapering clastic wedges (Fig. 3; Seymour and Fielding, 2013). These strata

are renowned for their apparent interfingering relationships and the ease at which they can be

correlated down depositional-dip in the Book Cliffs and along depositional-strike in the Wasatch

Plateau (Spieker, 1949). At Price Canyon in the northwestern-most Book Cliffs near the town of

Helper (Fig. 2B), the Mesaverde Group consists of the intercalated offshore-marine Mancos

Shale and shallow-marine Star Point Sandstone which transition upwards to the coastal-plain,

Page 16: High-frequency tectonic sequences in the Campanian

9

coal-bearing and fluvial Blackhawk Formation, which is capped by the primarily fluvial

Castlegate Formation (Fig. 3).

At the type-section in Price Canyon, the middle to upper Campanian Castlegate

Formation is divided into three members on the basis of weathering profile (Olsen et al., 1995).

The lowest member is the middle Campanian, cliff-forming Lower Castlegate Sandstone, which

is characterized by ~60-80 m of amalgamated fine-to-medium grained sandstone, interpreted as

braided fluvial deposits (Fig. 3). The overlying unit is ledge-forming, middle to upper

Campanian, and known as the Middle Castlegate Sandstone. The ledges form at mudstone-

encased fine-grained sandbodies (i.e. lenticular) interpreted as meandering and tidally-influenced

fluvial, and bayhead delta deposits (Fig. 3). The upper unit which is cliff-forming, upper

Figure 3. Biostratigraphically-constrained chronologic and spatial stratigraphic diagram showing variation of the

Campanian Mesaverde Group from the eastern Wasatch Plateau, to the western and central Book Cliffs. The

diagram incorporates data from Miall and Arush (2001), Aschoff and Steel (2011), Seymour and Fielding

(2013), and this study. A—Aberdeen Member; AMT—Anchor Mine Tongue; BT—Buck Tongue; D—Desert

Member; G—Grassy Member; K—Kenilworth Member; S—Sunnyside Member; SC—Spring Canyon Member.

Page 17: High-frequency tectonic sequences in the Campanian

10

Campanian and referred to as the Bluecastle Tongue, consists of ~20 m of amalgamated

medium-grained sandstone locally with pebble conglomerate interpreted as braided fluvial

deposits (Fig. 3).

Along-strike, to the southwest, at Joe’s Valley the Castlegate Formation lacks the

alternated cliff and slope weathering profile of the type-section, and is entirely cliff-forming. The

Joe’s Valley section is also coarser grained and thinner (Fig. 3). Paleo-landward, to the west

from the type-section, sandstones interfinger with conglomerates, where boulder-sized clasts

locally exist in a thick coeval-succession of piedmont-type deposits at the Cedar Hills and

Gunnison Plateau (Spieker, 1949). Paleo-seaward, to the east of the type-section near the Green

River in the Book Cliffs, the Lower Castlegate Sandstone is finer-grained, reduced to ~20 m, and

is capped by the paleolandward-tapering, offshore-marine Buck Tongue Member of the Mancos

Shale (Fig. 3; Aschoff and Steel, 2011). The estuarine to shallow-marine Sego Sandstone

truncates the Buck Tongue and is conformable to the overlying coal-bearing Neslen Formation

(Fig. 3; Miall and Arush, 2001). The Neslen Formation intercalates basinward with the marine

Anchor Mine Tongue (Fig. 3) and is likely age-equivalent to the up-dip Middle Castlegate

Sandstone (Yoshida et al., 1996; Miall and Arush, 2001; Aschoff and Steel, 2011). The

Bluecastle Tongue also caps the succession in these more basinward locations (Fig. 3) but is

finer-grained (Aschoff and Steel, 2011).

The Mesaverde Group is progressively, east to west tilted (Spieker, 1949) because of

deformation in the Sevier thrust belt (Horton et al., 2004). In the Wasatch Plateau, the Star Point

Sandstone, and Blackhawk and Castlegate Formations in combination form two and arguably

three 3rd-order, ~3-5 m.y. sequences (Miall and Arush, 2001; Seymour and Fielding, 2013). The

lowest Mesaverde Group sequence extends from the forced-regressive Panther Tongue (of the

Page 18: High-frequency tectonic sequences in the Campanian

11

Star Point Sandstone) up to the base of the Lower Castlegate Sandstone (Seymour and Fielding,

2013). The overlying Mesaverde Group strata may actually represent two distinct 3rd-order

sequences (Miall and Arush, 2001), but were originally defined (by Olsen et al., 1995) as a single

sequence from the base of the Lower Castlegate Sandstone to the base of the Bluecastle Tongue.

Farther basinward in the Book Cliffs, these 3rd-order sequences split into numerous (10-15)

higher frequency sequences (Seymour and Fielding, 2013).

In the Wasatch Plateau, most of the Star Point Sandstone and all of the Blackhawk

Formation has been interpreted to represent the highstand systems tract (HST) of the lowest 3rd-

order Mesaverde Group sequence (Seymour and Fielding, 2013). Within this HST, nested

higher-frequency sequences likely occur (Gani et al., 2015). Widespread sheet-like foreland

basin deposits like the Lower Castlegate Sandstone have basal, sequence-bounding

unconformities (Schwans, 1995). This basal Castlegate unconformity marks the base of the

second 3rd-order sequence. This surface is angular close to the Sevier thrust belt, and forms a

“smooth” denudational erosion surface near the Price Canyon type-section and narrow paleo-

valleys further basinward near the Green River (Spieker, 1949). Farther down-dip, east of the

Green River, the surface is likely a correlative conformity (Spieker, 1949; and Bhattacharya and

Holbrook, 2011).

The base of the Sego Sandstone is likely correlative up-dip to surface “D” defined by

Miall and Arush (2001) at the Castlegate type-section. They postulate that their “D” surface is a

‘cryptic’ sequence boundary forming the base of the potential third, uppermost (3rd-order)

Mesaverde Group sequence, with the uppermost sequence boundary as the base of the Bluecastle

Tongue. Miall and Arush (2001) identify surface “D” as a cryptic sequence boundary based on

sandstone petrography, diagenesis and paleocurrent data because regional erosion surfaces are

Page 19: High-frequency tectonic sequences in the Campanian

12

commonly obfuscated by the immensity of localized channel erosions. Below surface “D”, Miall

and Arush (2001) identified evidence for early diagenesis and a relative abundance of lithic and

feldspar grains. They found sandstones above surface “D” to contain higher modal-percentages

of quartz and porosities (i.e. less cement), indicating minimal early diagenesis (i.e. prolonged

subaerial exposure) and possibly source area (thrust) rejuvenation of the quartzitic Oquirrh

Group source.

Methods

In order to address the questions of the representation of transitioning deformation style

in stratigraphic surfaces, we developed a dataset that includes: (1) a 70-km-long transect (Fig.

2B) that correlates the Castlegate Formation along-strike from Joe’s Valley in the central

Wasatch Plateau to the Price Canyon area in the northwestern-most Book Cliffs, tied to (2) a

rendering of Aschoff and Steel’s (2011) approximately 400-km-long along-dip transect

correlating the Castlegate Formation from source-to-sink, perpendicular to the Wasatch Plateau

and through the Book Cliffs, (3) sediment-dispersal patterns obtained from 257 paleocurrent

measurements of dune trough cross-stratifications , (4) a QFL (quartz, feldspar, and lithics)

diagram for samples collected in the central Wasatch Plateau and its comparison to type-area

samples collected by Miall and Arush (2001), (5) thin-section photomicrographs, (6) and high-

resolution outcrop photo-mosaics. The new data were obtained from 14 Castlegate Formation

measured sections (Fig. 2B). Sections were analyzed for grain-sizes, sedimentary structures,

ichnology, facies, channel and facies architecture, and the nature of bounding discontinuities.

Seven of the 14 sections had covered portions or were incomplete (eroded at the top). Some

sections have vertical-cliff faces that were measured on rope. Others were traversed via physical

climbing without rope through scree-covered slopes. With the exception of the type-area and

Page 20: High-frequency tectonic sequences in the Campanian

13

Joe’s Valley, each section measured contains ‘frontier’ data on the Castlegate Formation as

accessing outcrops required daring and strenuous efforts. 138 hand samples were collected, and

18 were examined by petrographic microscope. Thin-sections were analyzed for QFL ratios,

nature of grain contacts, and types and percentages of cement.

The correlation of measured sections forms a two-dimensional cross-section parallel to

depositional-strike (the Wasatch Plateau) that has been tied to Aschoff and Steel’s (2011)

depositional-dip correlation. The tying of strike and dip transects provides a regional three-

dimensional perspective of the studied successions. Detailed interpretations (e.g. surface-tracing)

of high-resolution photomosaics, hand sample (including thin-section) analysis and comparison,

and field mapping aided key surface correlation between sections. Uncertainty is common to

nonmarine bounding-surface correlation. To counter this uncertainty, marine flooding surfaces or

condensed sections of the Western Interior Seaway in which high-resolution ammonite biozones

were established (Cobban et al., 2006), have been correlated up-dip to the type-section by

Aschoff and Steel (2011) and to the Wasatch Plateau herein. Notably, the ammonite zones have

~200 k.y. of time-constraint (Cobban et al., 2006). Marine flooding surfaces in predominantly

nonmarine successions are manifested by tidally-influenced facies. Sequence boundaries were

correlated based on mineralogy, diagenesis, paleocurrent patterns, and prominent erosional

surfaces, as demonstrated by Miall and Arush (2001). This method assumes that sandstones

above similarly-aged (more or less) erosion surfaces will have similar mineralogical

compositions and sediment dispersal patterns, and that strata below a sequence boundary can

show evidence for early diagenesis.

Page 21: High-frequency tectonic sequences in the Campanian

14

RESULTS

Facies

At the type-locality, McLaurin and Steel (2007) characterized three Castlegate Formation

depositional environments. In order of decreasing prevalence, they identify thalweg-fill, barform,

and floodplain deposits. Thalweg-fill facies occupied the basal portion of a channelized

succession, and were thus least susceptible to post-depositional erosion (McLaurin and Steel,

2007). Most barforms were classified as downstream accreting with rare lateral and upstream

accretion. Floodplain deposits, which are least preserved are present up-section.

Of the nine facies established in this study, six are deemed in some cases, sequence

stratigraphically significant. They are bioturbated sandstone (1), intraformational lag (2),

extraformational lag (3), sandstone with soft-sediment deformation (4), sandstone with rhythmic

mudstone (5), and mudstone with lenticular sandstone (6). The bioturbated sandstone facies

represents floodplain deposits subject to considerable primary diagenesis as evidenced by

ichnology and concretion. The prevalence of Teredolites longissimus and clavatus ichnofossils in

this facies indicate brackish-water deposition, particularly during marine transgression (Shanley

and McCabe, 1992). Also present are siderite nodules, indicative of burial and humid, oxygen-

reducing environment. The intraformational and extraformational lag facies coincide with

thalweg-fills but in some cases major erosional surfaces. These thalweg-fill, matrix-supported

conglomerate facies suggest deposition from dense-inertia flows or “traction carpets” likely

triggered during high-energy, flood-stages. Linking a depositional environment to the sandstone

with soft-sediment deformation facies is difficult as preexistent sedimentary structures were

destroyed. Beds or sandbodies with pervasive (km’s continuity) convolutions can indicate

seismicity (Kundu et al., 2011; Balsamo et al., 2013). Both the sandstone with rhythmic

Page 22: High-frequency tectonic sequences in the Campanian

15

mudstone and mudstone with subordinate sandstone facies record tidal influence indicated by

sedimentary structures and ichnology.

Strike-variability of stratigraphy

The datum for this new, ~70-km-long stratigraphic correlation along depositional-strike

(Fig. 4) is a maximum flooding surface (TS4) correlative to the Didymoceras nebrascense

ammonite biozone from the Anchor Mine Tongue condensed section. This surface was traced

up-dip through the Book Cliffs to the type-section (Price Canyon) by Aschoff and Steel (2011)

and herein was correlated in the Wasatch Plateau (Fig. 4). Surface TS4 at the type-section is

capped by the mudstone with lenticular sandstone facies, and by a sandstone with rhythmic

mudstone facies near Joe’s Valley. Transgressive surfaces in the proximal setting are linked to

the fine-grained heterolithic deposits (associated with tidal-influence), which have

southwestward-tapering thicknesses (Fig. 4). In fluvial-dominated successions, transgressive and

flooding surfaces are commonly underlain by ellipsoidal diagenetic concretions or thin

calcareous-cemented horizons (Taylor et al., 2000; Al-Ramadan et al., 2013).

The correlation reveals significant spatio-temporal facies and architectural variability

induced by protracted tectonism within the Cordilleran foreland. There is an abrupt southward

increase in sandbody amalgamation and grain sizes along-strike from the type-section onto an

interpreted structural paleo-high that began forming during early Castlegate Sandstone

deposition in the Wasatch Plateau (Fig. 4). Mudstone-rich intervals at the type-locality are the

age-equivalents of sandstone-dominated intervals along-strike (Fig. 4). As the Castlegate

Formation becomes increasingly amalgamated, it also thins (Fig. 4). The Lower Castlegate

Sandstone between the basal sequence boundary (CGSB) and Miall and Arush’s (2001) surface

“D”, which is ~20 m-thick at the type section, is as thin as 5 m on the paleo-high (Fig. 4).

Page 23: High-frequency tectonic sequences in the Campanian

16

Figure 4. An approximately 70-km-long, along-strike stratigraphic transect of middle Campanian in the eastern

Wasatch Plateau and northwestern Book Cliffs of Utah. Note distinct southwestward thinning, coarsening and

amalgamation of strata towards an uplifted structure. Refer to Fig. 2B for location abbreviations and location of

cross-section. CGSB—Castlegate sequence boundary; “D”—Miall and Arush’s (2001) surface D (up-dip

equivalent to base of Sego Sandstone); “E”—Miall and Arush’s surface E; TS—transgressive surface; AU—

angular unconformity; MCGU—Middle Castlegate unconformity.

Page 24: High-frequency tectonic sequences in the Campanian

17

Extraformational lag deposits approximately age-equivalent to strata resting on Miall’s surface

“E” are prevalent on this paleo-high, and taper northeastward along-strike into a topographic

paleo-depression flanking the paleo-high (Fig. 4). The Lower Castlegate Sandstone is

progressively northeast-tilted away from the paleo-high that formed its margin between Wattis

Road and Price Canyon, beneath a subtle, low-angle angular unconformity (surface AU1) (Fig.

4). The structural-tilt abruptly steepens southwestward between Wilberg Mine and Joe’s Valley,

as the Middle Castlegate Sandstone’s ledge-forming weathering profile disappears (Fig. 4). Near

Joe’s Valley, the lowest three Castlegate Formation transgressive surfaces (TS1, TS2, and TS3)

are truncated by surface AU1 (Fig. 4).

In the northwestern Book Cliffs, strata are northeast-tilted below angular unconformity 2

(AU2), away from the younger of the two identified structural paleo-highs (Fig. 4). Surface AU2,

which forms the base of the Bluecastle Tongue, truncates strata resting on a Middle Castlegate

Sandstone unconformity (MCGU) (Fig. 4). In places such as at the type-section, strata between

surfaces AU2 and MCGU are completely absent (Fig. 4). Below surfaces AU1 and AU2, strata

commonly display Liesegang banding and laterally-extensive soft sediment deformation (Fig. 4).

Paleocurrent analysis

Measurements (𝑛 = 111) of fluvial paleocurrent direction gathered from dune trough

cross-stratifications were analyzed temporally between significant bounding surfaces and

spatially between the northwestern Book Cliffs and the Wasatch Plateau (Fig. 5).

In the northwestern Book Cliffs, Castlegate Formation fluvial paleocurrent directions

gradually shift anticlockwise from SE (131°) near the base to NE (53°) in the Bluecastle Tongue,

maintaining an overall basinward-dispersal pattern (Fig. 5). In the Wasatch Plateau, there are two

Page 25: High-frequency tectonic sequences in the Campanian

18

distinct temporal perturbations in Castlegate Formation fluvial paleocurrent direction (Fig. 5).

Below the lowermost perturbation, within the basal 5-20 m of the Lower Castlegate Sandstone,

fluvial transport is basinward (SE) as is age-equivalent northwestern Book Cliff strata. The

perturbations are abrupt, 136-171° shifts in fluvial transport direction (Fig. 5). The older of the

two is defined across surface “E” above the basal 5-20 m of the Lower Castlegate Sandstone,

Figure 5. Rose diagram analysis of paleocurrent measurements. Notice how drainage patterns were basinward

prior to 78 Ma. Around 78 Ma, fluvial systems developed an axial trend and seemingly converged in the

northern Wasatch Plateau. Subsequent to 78 Ma, drainage direction gradually shifted clockwise in the Wasatch

Plateau and anticlockwise in the northwestern Book Cliffs. Ensuing the development of surface AU2, drainages

became axial again but fluvial systems between the Wasatch Plateau and northwestern Book Cliffs were

divergent to each other.

Page 26: High-frequency tectonic sequences in the Campanian

19

where an anticlockwise (SE to NW), 136° shift from basinward to basin-axial fluvial transport

direction is observed (Fig. 5). Between surface “E” and the base of the Bluecastle Tongue,

fluvial paleocurrent direction in the Wasatch Plateau gradually rotates clockwise (from NW to

NE) and basinward by ~40° (Fig. 5). Across the Bluecastle sequence boundary, the younger of

the two perturbations is expressed as an abrupt, 171° clockwise (NE to SW) change in fluvial

transport direction to a basin-axial pattern (Fig. 5).

Sandstone petrography

Thin

sections (𝑛 = 18)

are from samples

gathered at the

Wattis Road, Trail

Canyon Road, Joe’s

Valley, Price

Canyon, and

Straight Canyon

measured section

localities (Fig. 2B

and 6). QFL

estimates from

these samples were compared with those gathered from the type-area (Fig. 6) by Miall and Arush

(2001). Photomicrographs (𝑛 = 4) of samples at Joe’s Valley and the type-area were compared

(Fig. 7).

Figure 6. Quartz-feldspar-lithics (QFL) diagram of middle to late Campanian

Sandstones of this study in the Wasatch Plateau, and compiled by Miall and Arush

(2001) in the northwestern Book Cliffs (B). The provenance divisions in the ternary

diagrams are taken from Lawton et al. (2014).

Page 27: High-frequency tectonic sequences in the Campanian

20

Sandstone mineralogy and provenance

Based on analysis of sandstone mineralogy, Miall and Arush (2001) hypothesized that the

Castlegate Formation contains three, ‘shingled’ tectonic sequences with the bases of the Lower

Castlegate Sandstone, Sego Sandstone, and Bluecastle Tongue forming the boundaries. They

discovered a mineralogical similarity between the Joe’s Valley section and the Bluecastle

Tongue at Price Canyon. They therefore suggest that the amalgamated sandstones resting on the

Blackhawk Formation at Joe’s Valley are younger than and not age-equivalent to those at the

type-section. Compare Fig. 7A to Fig. 7C to more thoroughly understand their reasoning.

Figure 7. Thin section photomicrographs of comparing samples from Joe’s Valley and the type-area, at the base

of the Castlegate Formation (B and D) and the base of the Bluecastle Tongue (A and C). Notice the coarser

grains and relative abundance of lithics and feldspars in the Joe’s Valley samples.

Page 28: High-frequency tectonic sequences in the Campanian

21

The proximal 2-D seismic dip-transect of Horton et al. (2004) reveals that the Castlegate

Sandstone at the type-section was unroofed from Pennsylvanian Oquirrh Group quartzite in the

Santaquin culmination. Miall and Arush’s (2001) type-area samples are also quartz-rich

sandstones (Fig. 6), suggesting this quartzitic orogenic-source. Samples presented here from the

Wasatch Plateau still contain mostly quartz but cluster separately from the type-section sample

suite, and contain higher percentages of carbonate, lithic, and feldspars grains (Fig. 6). Quartz

grains in the Wasatch Plateau are typically coarser and more rounded than those at the type-

section (Fig. 7). The main source areas in the Wasatch Plateau are likely the Canyon Range and

Pavant thrust systems. The prevalence of carbonate grains strongly suggests that the Wasatch

Plateau between 75-80 Ma was fed sediments from the Canyon Range and Pavant thrust systems,

which were the only of the Sevier thrusts in Utah to expose limestones (DeCelles and Coogan,

2006). Some of the feldspars appear to be of Cordilleran hinterland, arc-terrane origin, as

evidenced by QFL-plotting (Fig. 6) and mineralogical properties. For example, some feldspar

gravels at the Wattis Road location display exsolutional lamellae and still have well-defined

cleavages. The presence of mineral-cleavage in fluvially-transported grains strongly might

suggest they are not orogenically-recycled.

Sandstone diagenesis

Sandstone samples contain variable modal-percentages and types of cement. The cements

are mostly interparticle with calcareous, ferrous and siliceous compositions. Samples contain as

much as ~28% cement. Macroscopically, carbonate-cementation was documented occurring in

localized, as much as 4-m thick concretionary-bodies or in thinner more laterally-extensive

horizons. Both features likely reflect early meteoric diagenesis (Taylor et al., 2000). The

localized concretionary diagenetic features are truncated by sequence boundaries and thus likely

Page 29: High-frequency tectonic sequences in the Campanian

22

indicate diagenesis during subaerial exposure (Taylor et al., 2000). The thin laterally-extensive

horizons exist below surfaces that correlate to marine flooding events farther in the basin, likely

suggesting diagenesis when there was a high water-table, and low-sediment supply (Taylor et al.,

2000).

Page 30: High-frequency tectonic sequences in the Campanian

23

DISCUSSION

Manifestation of deformation-style in stratigraphic surfaces

In the Cordilleran foreland basin, several studies (Cross, 1986; Pang and Nummedal,

1995; Aschoff

and Steel, 2011;

Liu et al., 2011)

linked

subsidence and

uplift patterns to

spatio-temporal

variability of: (1)

supralithospheric

flexural loading

by thrust sheets,

(2)

sublithospheric

dynamic loading

by a shallowly-

subducted

Farallon plate,

and/or (3)

differential sediment loading. Using the technique of along-strike stratigraphic correlation, two

distinct intra-foreland basin uplifts were documented here. The earliest of the two events is

Figure 8. Photograph (upper diagram) and stratigraphic architectural analysis (lower diagram)

of the Straight Canyon measured section (Figs. 2B and 4). Notice the distinct structural tilt of

strata (~8° NE) beneath surface AU1 and relatively flat nature of surfaces above AU1.

Page 31: High-frequency tectonic sequences in the Campanian

24

apparent towards the southwest in the Wasatch Plateau as the Lower Castlegate Sandstone

abruptly thinned, coarsened, and northeast-tilted onto a structural paleo-high at Joe’s Valley

(Figs. 4 and 8). The later of the two deformation phases also caused strata to develop a northeast

tectonic-tilt but below the Bluecastle Tongue (Figs. 4, 9 and 10) onto a paleo-high in the

northwestern Book Cliffs. Each phase was ensued by erosional beveling and development of a

low-angle angular unconformity, beneath which as much as 30-m of strata were removed.

In the Alberta foreland basin, flexure-induced erosional beveling has been interpreted to

reflect the location and transgressive truncation of the forebulge (Plint et al., 1993). Van

Wagoner (1995) hypothesized that the forebulge coeval to the Castlegate Sandstone is overlain

by oolitic ironstones near the Utah-Colorado border. The erosional beveling documented in this

Figure 9. Photographs of the uppermost Castlegate Formation at Deadman’s Creek Road (Fig. 2B) illustrating

the angular unconformity, where surface AU2 truncates underlying tilted (~10° N) and soft-sediment deformed

strata (orange). In both A (distant view) and B (close-up view), upper diagram is un-interpreted and lower

diagram is interpreted.

Page 32: High-frequency tectonic sequences in the Campanian

25

study is ~60 km from the Sevier thrust front, whereas the interpreted forebulge locations of Van

Wagoner (1995) and Plint et al. (1993) are several hundred kilometers basinward from the source

area and in a marine-setting. As evidenced by down depositional-dip middle Campanian facies

changes, the observed beveling in this study is not related to truncation of the forebulge.

Another hypothesis to be considered is that the structural-tilts beneath surfaces AU1 and

AU2 (Figs. 4, 8, 9 and 10) developed during and in response to Sevier deformation, implying a

‘growth strata’ origin. Bounded by progressive unconformities, growth strata develop

Figure 10. Along-strike (Wasatch Plateau) and down-dip (Book Cliffs) sequence stratigraphic correlation of the

middle to upper Campanian strata. This 3-D perspective accentuates tilting of strata below AU1 and doming of

strata below the Bluecastle Tongue (AU2). NMC—Nine Mile Canyon; HoC—Horse Canyon; TrC—Trail

Canyon; GrR—Green River; TCb—East Tuscher Canyon (Fig. 2A/B). The down-dip cross-section incorporates

data from Yoshida et al. (2006) and Aschoff and Steel (2011).

Page 33: High-frequency tectonic sequences in the Campanian

26

synchronously to growth folding, whereby erosion records thrusting and sedimentation reflects

tectonic quiescence (Vergés et al., 2002). Growth strata typically develop only within a few

kilometers of the deformation front (Aschoff, 2008). They inherit wedge shapes that

progressively taper and tilt towards the thrust front (Aschoff, 2008). However, wedge-tapering

and structural-tilts documented here progress in a basin-parallel direction. Another hypothesis

that could explain intra-foreland uplift is post-thrusting crustal rebound (Heller et al., 1988;

Flemings and Jordan, 1990; Ettensohn, 2004). This is unlikely however as Horton et al. (2004)

showed that the deposition of the Castlegate Sandstone was coeval to Sevier thrusting.

The structural-tilts beneath surfaces AU1 and AU2 (Figs. 7 and 8) most likely formed

during incipient episodes of Laramide deformation linked to sub-lithospheric loading by a

shallowly-subducted flat slab (e.g. an aseismic ridge) in the Farallon plate (Yonkee and Weil,

2015). Deep-seated subsurface faults (e.g. blind thrusts) were probably (re)activated by stress

transfer between the flat slab and the North American lithosphere. Aschoff and Steel (2011)

calculated that the Middle Castlegate Sandstone and down-dip lithostratigraphic equivalents

(Neslen Formation and Sego Sandstone) prograded at an anomalous rate and extent. They

contributed this unusual long-transit, rapid progradation, hence reduced subsidence, to

embryonic development of the Laramide San Rafael Swell basement-cored uplift. The evidence

presented here suggests that incipient growth of the San Rafael Swell was episodic (not a gradual

transition from Sevier to Laramide-style deformation as they describe), and that the initial phase

occurred even earlier, during deposition of the Lower Castlegate Sandstone spanning the

biozones Baculites asperiformis and Baculites perplexus (late), between ca. 79 and 77 Ma. The

anomalous Middle Castlegate Sandstone progradation commenced with the development of

Page 34: High-frequency tectonic sequences in the Campanian

27

surface AU1, and was subsequent, not coeval, to a preliminary growth-episode of the San Rafael

Swell.

Manifestation of deformation-style in drainage patterns and provenance

Basin-margin structures strongly dictate the volume and transport-destination of a

sediment supply (Heller et al, 1993; Ettensohn, 2004). In addition to surfaces AU1 and AU2,

evidence for middle to late Campanian growth of the San Rafael Swell is recorded in paleo-

drainage patterns (Fig. 5). In the lowermost Castlegate Sandstone (between CGSB and AU1),

fluvial channel networks were likely distributive with a fairly direct cratonward path. Up-section

(between surfaces “D” and AU2), paleocurrents have orogen-parallel orientations, where fluvial

systems flowed southward in the northwestern Book Cliffs and northward in the Wasatch

Plateau. Thus, they seemingly converged into an area in the northernmost Wasatch Plateau.

Farther up-section (between surfaces “D” and AU2), drainage patterns are characterized even

more by channel confluence rather than bifurcation (Fig. 5) because accommodation space was

increasingly confined between the fold-thrust belt and growing San Rafael Swell. Just below the

Bluecastle Tongue (between MCGU and surface AU2), drainage patterns were again distributive

and basinward, reflecting erosional-flattening of positive surface-topography generated by the

early San Rafael Swell folding-episode. Within the Bluecastle Tongue (above surface AU2),

transport directions are parallel (NE) or oblique (SW) to the fold-thrust belt and divergent around

the San Rafael Swell, which underwent rejuvenated uplift around 75 Ma.

Evidence for middle to late Campanian Laramide-modification of the Cordilleran

foreland also comes from spatio-temporally variable sedimentary provenance (Figs. 6 and 7).

The Lower Castlegate Sandstone (from CGSB to surface TS1) in the northwestern Book Cliffs

appears enriched with eroded Late Paleozoic quartzite (Oquirrh Group) from the Charleston-

Page 35: High-frequency tectonic sequences in the Campanian

28

Nebo thrust and triangle zone (Miall and Arush, 2001). Mineralogy and associated thrust-

orthogonal (south-southeast) paleocurrent directions in the northwestern Book Cliffs confirm the

source of these middle Campanian sandstones. The Castlegate Formation’s muddy middle

member (between surfaces TS1 to MCGU), stems from a Jurassic clay-rich lithology unroofed

from the Paxton thrust sheet, ~100 km south of the northwestern Book Cliffs. Grain mineralogy,

textures and sizes coupled to coeval along-strike Wasatch Plateau paleocurrent-directions helped

pinpoint this Paxton thrust source, located in south-central Utah. Though the mudstones in the

muddy, middle member of the Castlegate Formation likely originated from the Paxton thrust, it is

conceivable that the sandstones maintained an Oquirrh Group derivation in the northwestern

Book Cliffs.

Quartz grains have a different origin, and feldspar and calcite grains are more abundant in

the Wasatch Plateau compared to coeval deposits in the northwestern Book Cliffs. Related-

paleocurrents are directed away from the Canyon Range, Pavant and Paxton thrust systems (Fig.

2) in south-central Utah. The coarse-grained material was derived mostly from the oldest and

westernmost mechanically-rigid Canyon Range and Pavant thrusts (DeCelles and Coogan, 2006).

Though the Canyon Range and Pavant thrusts had early Cretaceous and middle Cretaceous

emplacements, respectively, they must have constituted the Campanian coarse-grained sediment

supply as the Paxton and Gunnison thrusts exposed mainly fine-grained rocks and their frontal

structures became buried by wedge-top Indianola conglomerate and sandstone (DeCelles et al.,

1995; DeCelles and Coogan, 2006). Provenance data from Indianola wedge-top sediments

indicate that most quartz-grains were derived from Neoproterozoic quartzite, which was only

exposed at the Canyon Range thrust hanging wall (DeCelles and Coogan, 2006). Majority of

calcite (or dolomite) grains were shed from the Canyon Range or Pavant thrusts as they exposed

Page 36: High-frequency tectonic sequences in the Campanian

29

thick Cambro-Ordovician successions of passive-margin carbonate rocks (DeCelles and Coogan,

2006).

Though clays shed from the Paxton thrust prior to its late Campanian burial likely passed

through fluvial systems in the Wasatch Plateau, they were not significantly preserved there

because of the presence of a paleo-high generated by preliminary uplift of the San Rafael Swell.

Upon this early growth of San Rafael Swell, fluvial systems in the Wasatch Plateau were drained

towards the northwestern Book Cliffs rather than in an immediate basinward direction. Of the

sediments passing through middle to late Campanian axial, northeast-flowing fluvial systems in

the Wasatch Plateau, conceivably only the fine-grained fraction reached downstream areas in the

northwestern Book Cliffs.

Arkosic gravels were observed on surface “E” in the Wasatch Plateau, displaying

exsolutional-lamellae and well-intact cleavage planes, which conceivably been destroyed during

fold-thrust related metamorphism. They most likely originated west of the Sevier thrust belt in

Jurassic back-arc plutons of eastern Nevada and western Utah (Fig. 1B). Their presence within

the basin-fill is indicative of catchment-enlargement (Hampson et al., 2014), which was likely

facilitated by a reduction of subsidence.

Origin of orogeny

With the waning of Sevier deformation and onset of Laramide deformation, around 81

Ma, the process of foreland accommodation creation or destruction shifted from flexural by

supralithospheric loading to dynamical by sublithospheric loading (Figs 11 and 12; Painter and

Carrapa, 2013). This cratonward and detachment-level shift in deformation was accompanied by

a major shutdown of arc magmatism (Yonkee and Weil, 2015; DeCelles, 2015) and a reduction

Page 37: High-frequency tectonic sequences in the Campanian

30

of Farallon plate subduction angle (Figs. 11 and 12; Coney and Reynolds, 1977; Cross, 1986; Liu

et al., 2010; Painter and Carrapa, 2013; Yonkee and Weil, 2015). Steep subduction promotes

near-trench arc-magmatism, slab-dehydration, and the basalt-eclogite transition as the subducting

oceanic crust is relatively dense. Reduction of oceanic plate density occurs at lesser distances

Figure 11. Paleogeographic

reconstructions of the Cordilleran

orogenic system illustrating the

evolution of surface topography,

depositional environments, thrust faults,

and subduction of an oceanic plateau

during the Late Cretaceous. A.

Paleogeographic reconstruction prior to

a major Late Cretaceous (ca 81 Ma)

shutdown of arc magmatism. During this

time, the foreland basin in Utah was

dominated by thrust-related flexural

subsidence, where coarse grained

sediments were trapped adjacent to the

orogenic front. Notice that the subducted

oceanic plateau was far southward from

the study area in Utah. B.

Paleogeographic reconstruction

subsequent to the major arc magmatism

shutdown. Note how the foreland

depocenter migrated significantly

cratonward as basin accommodation

began reacting to dynamic effects in

response to NE propagation of the

subducted oceanic plateau. Maps

incorporate data from Yonkee and Weil

(2015). For location, compare state

boundaries with Fig. 1B.

Page 38: High-frequency tectonic sequences in the Campanian

31

inboard of the plate-boundary (Liu et al., 2010). With the reduction of subduction angle hence

plate density, the basalt-eclogite transition is reached farther inboard of the subduction zone,

allowing for surface topography to adjust dynamically well-within cratonal regions (Fig. 11; Liu

et al., 2010; Painter and Carrapa, 2013; Yonkee and Weil, 2015).

The interacting effects between an oceanic flat slab and overriding continent on dynamic

topography are heavily debated (Yonkee and Weil, 2015). It is broadly assumed that

sublithospheric loading by a flat slab generates dynamic subsidence and widespread marine

incursion (Cross and Pilger, 1978; Liu et al., 2011; Painter and Carrapa, 2013). Two hypotheses

have emerged (Painter and Carrapa, 2013) for the creation of dynamic subsidence. The earlier

one favors near-horizontal subduction whereby coupling of oceanic and continental plates

Figure 12. Cross-sections illustrating Late Cretaceous evolution of the North American Cordilleran orogenic

system prior to (A) and during (B) the deposition of the Castlegate Formation Cross-sections incorporate

interpretation of DeCelles (2015).

Page 39: High-frequency tectonic sequences in the Campanian

32

generate subsidence by replacement of the asthenosphere with colder, denser oceanic lithosphere

(Cross and Pilger, 1978). The newer model favors shallow (less horizontal) subduction such that

dynamic subsidence results from viscous flow within the asthenosphere above the plunging

oceanic plate (Liu et al., 2010; Painter and Carrapa, 2013). Dávila and Lithgow-Bertelloni (2015)

argue however that positive dynamic topography (uplift) develops above anomalously buoyant

flat slabs (Fig. 11), and that negative dynamic topography (subsidence) develops at the slab’s

leading-edge.

It is useful to draw clues from a modern analog where the effects of subduction zone

plate-interactions on dynamic topography can be modeled inversely. The most widely accepted

modern analog to Laramide-style deformation are the uplifts and piggyback basins inboard of the

Andean subduction zone such as the Sierra Pampeanas in Argentina and Fitzcarrald Arch in

Brazil (Dávila and Lithgow-Bertelloni, 2015). These uplifts partitioned cratonic regions along

localized basement-bounded reverse faults beginning around 6 Ma. As evidenced by geophysical

data, these South American intra-cratonic uplifts are directly linked to flat-slab subduction of

aseismic ridges embedded within the Nazca plate (Gutscher et al., 1999). Dávila and Lithgow-

Bertelloni (2015) found that even though the subducted intra-plate ridges are relatively cold, they

represent anomalously buoyant segments of lithosphere because of prior and substantial partial-

melting, which made them less dense and drier than typical mid-ocean ridge basalt (MORB). The

low water content and cold temperature of the flat slabs inhibit the basalt to eclogite transition,

thus making them buoyant (Dávila and Lithgow-Bertelloni, 2015). The effects of flat-slab

subduction on dynamic topography as explained by Dávila and Lithgow-Bertelloni (2015) can be

applied to uplift of the San Rafael Swell. Dynamic uplift above a flat-slab segment could

correspond to basal traction between the flat-slab and overriding continental plate or flat-slab

Page 40: High-frequency tectonic sequences in the Campanian

33

dehydration (Yonkee and Weil, 2015), rather than to crustal rebound upon foundering of the

sublithospheric load (Liu et al., 2010). Regardless, preexisting weakness zones are likely

required to generate the Laramide, multi-km reverse-fault offsets in mechanically-rigid

continental crust (Marshak et al., 2000).

Evidence in the form of angular unconformities and drainage deflection have been

presented herein for middle to late Campanian Laramide-style deformation in Utah. The angular

unconformities are consistently underlain by laterally-extensive (multi-km) horizons of

sandstone with soft sediment deformation. The lateral continuity of this facies is suggested to

record seismicity (Kundu et al., 2011), which may provide a record of the stress transfer

mechanism from the subducted flat slab to overlying lithosphere during the early Laramide

orogeny. If the angular unconformities are linked to Laramide-style deformation, the seismicity-

origin that generated the sandstone with soft-sediment deformation is probably sublithospheric.

Sublithospheric seismicity is mostly known to occur within the Wadati-Benioff zone, either via

internal slab deformation (e.g. the compositional transformation from basalt to eclogite) or by

basal traction at the top of the subducting slab (Coney and Reynolds, 1977). Flat subduction in

the study region implies that the basalt to eclogite transition had yet to be reached (Dávila and

Lithgow-Bertelloni, 2015). Therefore, these middle to late Campanian probable seismic events

were focused likely at the top of the aseismic ridge, induced by basal traction with the overriding

plate.

The NE-SW axial trend of the San Rafael Swell parallels the compressive-stress that

would have been applied by the subducting aseismic ridge, assumed to be a conjugate of the

Shatsky-Hess Rise (Liu et al., 2010) (Figs. 11 and 12). Therefore, the San Rafael Swell should

display some tensional or trans-tensional geometry. Like many Laramide structures in the

Page 41: High-frequency tectonic sequences in the Campanian

34

Colorado plateau, the San Rafael Swell is a fault-propagation fold (Marshak et al., 2000) that

likely formed by inversion of a Proterozoic extensional fault. Incipient slip of the fault

underlying the swell was likely oblique given the NE-propagation of the subducted Shatsky

plateau. However, detailed mapping of the structure would be required to validate this claim.

Paradoxically, formation of the San Rafael Swell required significant compressive forces

oriented orthogonally to the NE-propagation direction of the Shatsky plateau. We therefore

predict that the stress transfer mechanism to generate Laramide uplifts likely evolved.

Embryonic stages (e.g. incipient growth of the San Rafael Swell) were likely induced by

sublithospheric basal traction between the North American plate and NE-propagating Shatsky

plateau often involving transpressional deformation. The later, better-known Laramide phase of

deformation that brought about uplift of the Colorado Plateau and reactivation of Ancestral

Rocky Mountain faults resulted from the Cenozoic foundering of the Shatsky-Hess Rise

conjugate (Liu et al., 2010).

Page 42: High-frequency tectonic sequences in the Campanian

35

CONCLUSIONS

Two incipient growth-episodes of the San Rafael Swell recorded in middle to late

Campanian stratigraphy played a significant role in reshaping the Cordilleran foreland in Utah.

Documented low-angle (<10°) angular unconformities are the removal-sites of thick (as much as

30-m) sediment-accumulations, which were redistributed in long, rapid, basinward-transits. The

earliest episode, around 78 Ma that records growth in the southern San Rafael Swell caused

uplift in the Joe’s Valley area of the Wasatch Plateau. The later phase began around 75 Ma, prior

to deposition of the Bluecastle Tongue in the northern San Rafael Swell, causing uplift in the

northwestern Book Cliffs.

Fluvial systems developed longer, winding transits as the San Rafael Swell grew as an

intra-foreland basin-boundary. One of the effects was that clay-rich sediments were deposited

northeast of their Jurassic Paxton thrust source area by basin-axial trending rivers. The incipience

of Laramide deformation and waning of Sevier deformation also allowed for sediments shed

from back-arc plutons to appear within the sediment supply, as catchments enlarged due to

thrust-inactivity-and-burial and as thrust-induced subsidence was reduced.

The NE-propagation of deformation in the San Rafael Swell can be attributed to NE-

propagation of an underlying, subducted oceanic plateau conjugate of the Shatsky-Hess Rise.

The documented angular unconformities are consistently underlain by laterally-extensive (multi-

km) and horizon-like sandstone with soft-sediment deformation, interpreted to record Wadati-

Benioff Zone seismicity. Early (ca. 80-75 Ma) Laramide deformation associated with flat-slab

plateau subduction likely caused uplift prior to Cenozoic slab steepening as Laramide-

deformation became isostatic in nature.

Page 43: High-frequency tectonic sequences in the Campanian

36

REFERENCES

Adams, M.M., and Bhattacharya, J.P., 2005, No change in fluvial-style across a sequence

boundary, Cretaceous Blackhawk and Castlegate Formations of Central Utah, U.S.A.:

Journal of Sedimentary Research, v. 75, p. 1038-1051.

Al-Ramadan, K., Morad, S., and Plink-Bjorklund, P., 2013, Distribution of diagenetic alterations

in relationship to depositional facies and sequence stratigraphy of a wave- and t-de-

dominated siliciclastic shoreline complex: Upper Cretaceous Chimney Rock sandstones,

Wyoming and Utah, USA, in Linking Diagenesis to Sequence Stratigraphy (eds S.

Morad, J.M. Ketzer, and L.F. De Ros), John Wiley & Sons, Inc., West Sussex, UK.

Aschoff, J.L., 2008, Controls on the development of growth strata and clastic wedges in foreland

basins: Examples from the Cordilleran foreland basin [Ph.D. thesis]: Austin, Texas,

University of Texas at Austin, 198 p.

Aschoff, J.L., and Steel, R., 2011, Anomalous clastic wedge development during the Sevier-

Laramide transition, North American Cordilleran foreland basin, USA: Geological

Society of America Bulletin, v. 123, p. 1822-1835.

Balsamo, F., Bezerra, F.H.R., Vieira, M.M., and Storti, F., 2013, Structural control on the

formation of iron-oxide concretions and Liesegang bands in faulted, poorly lithified

Cenozoic sandstones of the Paraíba Basin, Brazil: Geological Society of America

Bulletin, v. 125, p. 913-931.

Beaumont, C., 1981, Foreland basins: Geophysical Journal of the Royal Astronomical Society, v.

65, p. 291-329.

Holbrook, J.M., and Bhattacharya, J.P., 2012, Reappraisal of the sequence boundary in time and

space: Case and considerations for an SU (subaerial unconformity) that is not a sediment

bypass surface, a time barrier, or an unconformity: Earth-Science Reviews, v. 113, p.

271-302.

Blair, T.C., and Bilodeau, W.L., 1988, Development of tectonic cyclothems in rift, pull-apart,

and foreland basins: Sedimentary response to episodic tectonism: Geology, v. 16, p. 517-

520.

Bradley, M.D., and Bruhn, R.L., 1988, Structural interactions between the Uinta arch and the

overthrust belt, north-central Utah; Implications of strain trajectories and displacement

modeling: Geological Society of America Memoirs, v. 171, p. 431-446.

Chan, M.A., and Pfaff, B.J., 1991, Fluvial sedimentology of the Upper Cretaceous Castlegate

Sandstone, Book Cliffs, Utah, in Chidsey, T.C., Jr., Geology of east-central Utah;

Frontmatter: Utah Geological Association Publication 19, p. 95-110.

Cobban, W.A., Obradovich, J.D., Walaszcyk, I., and McKinney, K.C., 2006, A USGS Zonal

Table for the Upper Cretaceous Middle Cenomanian-Maastrichtian of the Western

Interior of the United States Based on Ammonites, Inoceramids, and Radiometric Ages:

U.S. Geological Survey Open-File Report 2006-1250, 46 p.

Coney, P., and Reynolds, S.J., 1977, Cordilleran Benioff zones: Nature, v. 270, p. 403-406.

Page 44: High-frequency tectonic sequences in the Campanian

37

Constenius, K.N., 1998, Extensional tectonics of the Cordilleran fold-thrust belt and the Jurassic-

Cretaceous Great Valley forearc basin [Ph.D. thesis]: Tucson, Arizona, University of

Arizona, 232 p.

Cross, T.A., 1986, Tectonic controls of foreland basin subsidence and Laramide style

deformation, western United States, in Allen, P.A., and Homewood, P., eds., Foreland

basins: International Association of Sedimentologists Special Publication 8, p. 15-40.

Cross, T.A., and Pilger, R.H., 1978, Tectonic controls of late Cretaceous sedimentation, western

interior, USA: Nature, v. 274, p. 653-657.

Dávila, F.M., and Lithgow-Bertelloni, C., 2015, Dynamic uplift during slab flattening: Earth and

Planetary Science Letters, v. 425, p. 34-43.

DeCelles, P.G., Lawton, T.F., and Mitra, G., 1995, Thrust timing, growth of structural

culminations, and synorogenic sedimentation in the type Sevier orogenic belt, western

United States: Geology, v. 23, p. 699-702.

DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and

foreland basin system, western U.S.A.: American Journal of Science, v. 304, p. 105-168.

DeCelles, P.G., and Coogan, J.C., 2006, Regional structure and kinematic history of the Sevier

fold-and-thrust belt, central Utah: GSA Bulletin, v. 118, p. 841-864.

DeCelles, P.G., and Graham, S.A., 2015, Cyclical processes in the North American Cordilleran

orogenic system: Geology, v. 43, p. 499-502.

Erskine, M.C., 1997, The Oquirrh Basin Revisited: American Association of Petroleum

Geologists Bulletin, v. 81, p. 624-636.

Ettensohn, F.R., 2004, Modeling the nature and development of major Paleozoic clastic wedges

in the Appalachian Basin, USA: Journal of Geodynamics, v. 37, p. 657-681.

Flemings, P.B., and Jordan, T.E., 1989, A synthetic stratigraphic model of foreland basin

development: Journal of Geophysical Research, v. 94, p. 3851-3866.

Jordan, T.E., and Flemings, B., 1991, Large-scale stratigraphic architecture, eustatic variation,

and unsteady tectonism: A theoretical evaluation: Journal of Geophysical Research, v. 96,

p. 6681-6699.

Franczyk, K.J., and Pitman, J.K., 1991, Latest Cretaceous nonmarine depositional systems in the

Wasatch Plateau area: Reflections of foreland to intermontane basin transition in

Chidsey, T.C., Jr., Geology of east-central Utah; Frontmatter: Utah Geological

Association Publication 19, p. 77-94.

Fricke, H.C., Foreman, B.Z., and Sewall, J.O., 2010, Integrated climate model-oxygen isotope

evidence for a North American monsoon during the Late Cretaceous: Earth and Planetary

Science Letters, v. 289, p. 11-21.

Gani, M.R., Ranson, A., Cross, D.B., Hampson, G.J., Gani, N.D., and Sahoo, H., 2015, Along-

strike sequence stratigraphy across the Cretaceous shallow marine to coastal-plain

transition, Wasatch Plateau, Utah, U.S.A.: Sedimentary Geology, v. 325, p. 59-70.

Page 45: High-frequency tectonic sequences in the Campanian

38

Garcia-Castellanos, D., 2002, Interplay between lithospheric flexure and river transport in

foreland basins: Basin Research, v. 14, p. 89-104.

Gutscher, M.A., Malavieille, J., Lallemand, S., and J.Y., Collot, 1999, Tectonic segmentation of

the North Andean margin: impact of the Carnegie Ridge collision: Earth and Planetary

Science Letters, v. 168, p. 255-270.

Hampson, G.J., 2010, Sediment dispersal and quantitative stratigraphic architecture across an

ancient shelf: Sedimentology, v. 57, p. 96-141.

Hampson, G.J., Gani, M.R., Sahoo, H., Rittersbacher, A., Irfan, N., Ranson, A., Jewell, T.O.,

Gani, N.D.S., Howell, J.A., Buckley, S.J., and Bracken, B., 2012, Controls on large-scale

patterns of fluvial sandbody distribution in alluvial to coastal plain strata: Upper

Cretaceous Blackhawk Formation, Wasatch Plateau, Central Utah, USA: Sedimentology,

v. 59, p. 2226-2258.

Hampson, G.J., Duller, R.A., Petter, A.L., Robinson, R.A.J., and Allen, P.A., 2014, Mass-

balance constraints on stratigraphic interpretation of linked alluvial-coastal-shelfal

deposits form source to sink: Example from Cretaceous Western Interior Basin, Utah and

Colorado, U.S.A.: Journal of Sedimentary Research, v. 84, p. 935-960.

Heller, P.L., Angevine, C.L., and Winslow, N.S., 1988, Two-phase stratigraphic model of

foreland-basin sequences: Geology, v. 16, p. 501-504.

Heller, P.L., Beekman, F., Angevine, C.L., and Cloetingh, S.A.P.L., 1993, Cause of tectonic

reactivation and subtle uplifts in the Rocky Mountain region and its effect on the

stratigraphic record: Geology, v. 21, p. 1003-1006.

Horton, B.K., Constenius, K.N., and DeCelles, P.G., 2004, Tectonic control on coarse-grained

foreland-basin sequences: An example from the Cordilleran foreland basin, Utah:

Geology, v. 32, p. 637-640.

Jordan, T.E., 1981, Thrust loads and foreland basin evolution, Cretaceous, western United States:

American Association of Petroleum Geologists Bulletin, v. 65, p. 2506-2520.

Kauffman, E.G., 1977, Geological and biological overview: Western Interior Cretaceous basin:

The Mountain Geologist, v. 14, p. 75-99.

Krystinik, L.F., and DeJarnett, B.B., 1995, Lateral variability of sequence stratigraphic

framework in the Campanian and lower Maastrichtian of the Western Interior Seaway, in

Van Wagoner, J.C., and Bertram, G.T., eds., Sequence Stratigraphy of Foreland Basin

Deposits: American Association of Petroleum Geologists, Memoir 64, p. 137-223.

Lawton, T. F., 1986, Fluvial Systems of the Upper Cretaceous Mesaverde Group and Paleocene

North Horn Formation, central Utah: A record of transition from thin-skinned to thick-

skinned deformation in the foreland region, in Peterson, J.A., ed., Paleotectonics and

sedimentation in the Rocky Mountain Region, United States: American Association of

Petroleum Geologists Memoir 41, p. 423-442.

Lawton, T.F., and Bradford, B.W., 2011, Correlation and provenance of Upper Cretaceous

(Campanian) fluvial strata, Utah, U.S.A., from zircon U-Pb geochronology and

petrography: Journal of Sedimentary Research, v. 81, p. 495-512.

Page 46: High-frequency tectonic sequences in the Campanian

39

LeClair, S.F., and Bridge, J.S., 2001, Quantitative interpretation of sedimentary structures

formed by river dunes: Journal of Sedimentary Research, v. 71, p. 713-716.

Liu, L., Gurnis, M., Seton, M., Saleeby, J., Muller, R. D., and Jackson, J.M., 2010, The role of

oceanic plateau subduction in the Laramide orogeny: Nature Geoscience, v. 3, p. 353-

357.

Liu, S., Nummedal, D., and Liu, L., 2011, Migration of dynamic subsidence across the Late

Cretaceous United States Western Interior Basin in response to Farallon plate subduction:

Geology, v. 39, p. 555-558.

Marshak, S., Karlstrom, K., and Timmons, J.M., 2000, Inversion of Proterozoic extensional

faults: An explanation for the pattern of Laramide and Ancestral Rockies intracratonic

deformation, United States: Geology, v. 28, p. 735-738.

Martinsen, O.J., and Helland-Hansen, W., 1995, Strike variability of clastic depositional

systems: Does it matter for sequence-stratigraphic analysis?: Geology, v. 23, p. 439-442.

McLaurin, B.T., and Steel, R.J., 2000, Fourth-order nonmarine to marine sequences, middle

Castlegate Formation, Book Cliffs, Utah: Geology, v. 28, p. 359-362.

Miall, A.D., and Arush, M., 2001, The Castlegate Sandstone of the Book Cliffs, Utah: Sequence

stratigraphy, paleogeography, and tectonic controls: Journal of Sedimentary Research, v.

71, p. 537-548.

Miall, A.D., The emptiness of the stratigraphic record: A preliminary evaluation of missing time

in the Mesaverde Group, Book Cliffs, Utah, U.S.A.: Journal of Sedimentary Research, v.

84, p. 457-469.

McLaurin, B.T., and Steel, R.J., 2007, Architecture and origin of an amalgamated fluvial sheet

sand, lower Castlegate Formation, Book Cliffs, Utah: Sedimentary Geology, v. 197, p.

291-311.

Olsen, T., Steel, R., Hogseth, K., Skar, T., Roe, S.L., 1995, Sequential architecture in a fluvial

succession: sequence stratigraphy in the Upper Cretaceous Mesaverde Group, Price

Canyon, Utah: Journal of Sedimentary Research, v. 65, p. 265-280.

Painter, C.S., and Carrapa, B., 2013, Flexural versus dynamic processes of subsidence in the

North American Cordillera foreland basin: Geophysical Research Letters, v. 50, p. 4249-

4253.

Posamentier, H.W., and Allen, G.P., 1993, Variability of the sequence stratigraphic model:

effects of local basin factors: Sedimentary Geology, v.86, p. 91-109.

Robinson, R.A.J., and Slingerland, 1998, Grain-size trends, basin subsidence and sediment

supply in the Campanian Castlegate Sandstone and equivalent conglomerates of central

Utah: Basin Research, v. 10, p. 109-127.

Schwans, P., 1995, Controls on sequence stacking and fluvial to shallow-marine architecture in a

foreland basin, in Van Wagoner, J.C., and Bertram, G.T., eds., Sequence Stratigraphy of

Foreland Basin Deposits: American Association of Petroleum Geologists, Memoir 64, p.

137-223.

Page 47: High-frequency tectonic sequences in the Campanian

40

Seymour, D.L., and Fielding, C.R., 2013, High-resolution correlation of the Upper Cretaceous

stratigraphy between the Book Cliffs and the Western Henry Mountains Syncline, Utah,

U.S.A.: Journal of Sedimentary Research, v. 83, p. 475-494.

Shanley, K.W., McCabe, P.J., and Hettinger, R.D., 1992, Tidal influence in Cretaceous fluvial

strata from Utah, USA: a key to sequence stratigraphic interpretation: Sedimentology, v.

39, p. 905-930.

Sohn, Y.K., Rhee, C.W., and Kim, B.C., 1999, Debris flow and hyperconcentrated flood-flow

deposits in an alluvial fan, northwestern part of the Cretaceous Yongdong Basin, central

Korea: The Journal of Geology, v. 107, p. 111-132.

Spieker, E.M., 1949, Sedimentary facies and associated diastrophism in the Upper Cretaceous of

central and eastern Utah: Geological Society of America Memoir 39, p. 55-82.

Taylor, K.G., Gawthorpe, R.L., Curtis, C.D., Marshall, J.D., and Awwiller, D.N., 2000,

Carbonate cementation in a sequence-stratigraphic framework: Upper Cretaceous

sandstones, Book Cliffs, Utah-Colorado: Journal of Sedimentary Research, v. 70, p. 360-

372.

Vakarelov, B.K., Bhattacharya, J.P., and Nebrigic, D.D., 2006, Importance of high-frequency

tectonic sequences during greenhouse times of Earth history: Geology, v. 34, p. 797-800.

Vakarelov, B.K., and Bhattacharya, J.P., 2009, Local tectonic control on parasequence

architecture: Second Frontier sandstone, Powder River Basin, Wyoming: American

Association of Petroleum Geologists Bulletin, v. 93, p. 295-327.

Van De Graaff, F.R., 1972, Fluvial-deltaic facies of the Castlegate Sandstone (Cretaceous), east-

central Utah: Journal of Sedimentary Petrology, v. 42, p. 558-571.

Van Wagoner, 1995, Sequence stratigraphy and marine to nonmarine facies architecture of

foreland basin strata, Book Cliffs, Utah, U.S.A., in Van Wagoner, J.C., and Bertram,

G.T., eds., Sequence Stratigraphy of Foreland Basin Deposits: American Association of

Petroleum Geologists, Memoir 64, p. 137-223.

Vergés, J., Marzo, M., and Muñoz, J.A., 2002, Growth strata in foreland settings: Sedimentary

Geology, v. 146, p. 1-9.

Willis, A., 2000, Tectonic control of nested sequence architecture in the Sego Sandstone, Neslen

Formation and Upper Castlegate Sandstone (Upper Cretaceous), Sevier Foreland Basin,

Utah, USA: Sedimentary Geology, v. 136, p. 277-317

Yonkee, W.A., and Weil, A.B., 2015, Tectonic evolution of the Sevier and Laramide belts within

the North American Cordillera orogenic system: Earth-Science Reviews, v. 150, p. 531-

593.

Yoshida, S., Willis, A., and Miall, A.D., 1996, Tectonic control of nested sequence architecture

in the Castlegate Sandstone (Upper Cretaceous), Book Cliffs, Utah: Journal of

Sedimentary Research, v. 66, p. 737-748.

Young, R.G., 1955, Sedimentary facies and intertonguing in the Upper Cretaceous of the Book

Cliffs, Utah-Colorado, GSA Bulletin, v. 66, p. 177-202.

Page 48: High-frequency tectonic sequences in the Campanian

41

Zaleha, M.J., Way, J.N., and Suttner, L.J., 2001, Effects of syndepositional faulting and folding

on Early Cretaceous rivers and alluvial architecture (Lakota and Cloverly Formations,

Wyoming, U.S.A.): Journal of Sedimentary Research, v. 71, p. 880-894.

Page 49: High-frequency tectonic sequences in the Campanian

42

APPENDIX

Regional Geology (Extended)

Pre-Cordillera sedimentary cover and tectonics

The oldest sedimentary rocks of western North America record localized intracratonic

deposition on the Mesoproterozoic to Early Neoproterozoic supercontinent Rodinia. These strata

are highly localized but thick (~5 km) basin-accumulations, one of which is an aulacogen-fill

now inverted in the E-W trending Uinta-Cottonwood Arch of northeast Utah (Bradley and

Bruhn, 1988). With the ensuing breakup of Rodinia, volcanic and siliciclastic rocks accumulated

in the Neoproterozoic to Early Cambrian rift basin bounding western Laurentia (Yonkee and

Weil, 2015). The Wasatch hinge line centered in the Sevier belt, represents the eastern limit of

Proterozoic rifting (DeCelles, 2004). Pre-Cordillera sedimentary cover west of the Wasatch

hinge-line is on average 12-km-thick whereas in regions east, cover is on average ~1.5-km-thick

(DeCelles, 2004). Proterozoic extensional faults likely played a significant role during

Phanerozoic, thick-skinned (Ancestral Rockies and Laramide) orogenies as they acted as

preexisting weakness planes necessary for uplift (Marshak et al., 2000). The thick-skinned

orogenies were likely confined to the east of the Wasatch hinge-line because of the relatively

thin pre-Cordillera sedimentary cover (DeCelles, 2015). Western Laurentia from the Middle

Cambrian to Middle Devonian records post-rifting, passive-margin, and predominately carbonate

deposition during the Sauk and Tippecanoe transgressions (Yonkee and Weil, 2015).

The western (present-day coordinates) Laurentian passive-margin trend that was

subsequently modified by Late Jurassic sinistral offset (induced by oblique Farallon Plate

convergence) strongly dictated the future curvature of the Sevier fold-and-thrust belt (Yonkee

Page 50: High-frequency tectonic sequences in the Campanian

43

and Weil, 2015). The western Laurentian margin became convergent during the Late Devonian-

Mississippian Antler Orogeny. The long-standing yet enigmatic hypothesis is that the Antler

Orogeny involved a continent-arc collision wherein an oceanic basin was closed and the Lower

Paleozoic deep-marine sediments deposited within were emplaced onto Laurentia (Speed and

Sleep, 1982). The allocthonous blocks were translated to present day Nevada, ~140 km from the

subsiding western-edge of the continent and contemporaneously eroded into the load-induced

foreland basin continentward of the orogeny (Speed and Sleep, 1982). During Late

Mississippian-Pennsylvanian, Laurentia collided with Gondwana, the Ancestral Rocky

Mountains uplifted, and ~6 km of extensional growth-faulted strata were deposited in the N-S

Oquirrh Basin of northern Utah (Erskine, 1997). The Oquirrh Basin was later inverted along the

Charleston-Nebo thrust during the Sevier orogeny (Erskine, 1997) and reworked into the

Cordilleran foreland basin-fill (Horton et al., 2004). Additional terranes were welded to the

Antler terranes during the Early Mesozoic Sonoman orogeny. From west to east in the Lower

Mesozoic, marine strata transition to continental deposits. Middle Jurassic marine carbonates and

evaporites were deposited on the subsiding Utah-Idaho trough of the Sonoman retroarc foreland,

and subsequently formed detachments in the eastern Sevier belt (DeCelles and Coogan, 2006;

Yonkee and Weil, 2015).

The Cordilleran orogenic system (Extended)

The westernmost and oldest portion of the Cordilleran is the forearc region. The forearc

developed with the accretion of Middle Jurassic (~165 Ma) volcanic arcs and interarc sediments

to form California’s Sierra Foothills and Coast Range (DeCelles, 2004; Yonkee and Weil, 2015).

Though speculative, their accretion onto the North American continent was manifested by

foundering of the “Mezcalera plate” between two Late Triassic to Early Jurassic volcanic arcs

Page 51: High-frequency tectonic sequences in the Campanian

44

during the Nevadan Orogeny (DeCelles, 2004; Yonkee and Weil, 2015). Coeval subduction

(Franciscan) generated metamorphism of the accreted terranes, intraplate shortening and

extensional arc-plutonism. These early Cordilleran plutons were partially eroded and deposited

in the Franciscan subduction trench and then underplated, offscraped, and accreted as the

Franciscan complex in California’s Coast Range by Early Cretaceous (Yonkee and Weil, 2015).

The Nevadan Orogeny culminated with the accretion of an ophiolite terrane in California’s Coast

Range. Upon accretion of the arcs, interarc-basin and trailing-ophiolite (complete foundering of

the “Mezcalera plate”), the Farallon plate began subducting beneath the western North American

margin. The Early Cretaceous involved the initial major flare-up of the Cordilleran magmatic arc

(e.g. Sierra Nevada’s calc-alkaline plutons) (DeCelles and Coogan, 2006). The second and final

major Cordilleran arc flare-up occurred during the Late Cretaceous, at ca. 90 Ma (DeCelles,

2004; DeCelles, 2015; Yonkee and Weil, 2015). The forearc basin consists mostly of clastic

sediments shed westward from the magmatic arc deposited as the ~2 to 10 km-thick Great Valley

Group, draping the forearc accretionary complex (Yonkee and Weil, 2015).

The retroarc hinterland which was dramatically subject to extension during the

Oligocene-Miocene Basin and Range event, encompasses the area between the magmatic arc and

fold-thrust belt (Fig. 1B), and mostly developed during Late Jurassic (Nevadan) crustal

thickening, metamorphism, and igneous intrusion events (Yonkee and Weil, 2015). From

Nevada to western Utah, the hinterland’s major components (Fig. 1B) include the Luning-

Fencemaker fold-thrust belt, the central Nevada fold-thrust belt, mid-crustal metamorphic rocks,

and gently deformed Paleozoic strata (DeCelles and Coogan, 2006; Yonkee and Weil, 2015).

The Luning-Fencemaker fold-thrust belt (Fig. 1B) which is the oldest of the retroarc elements

Page 52: High-frequency tectonic sequences in the Campanian

45

formed with incipient Franciscan subduction, coeval to the obduction of the Coast Range

ophiolite (Yonkee and Weil, 2105).

Campanian stratigraphy of central and eastern Utah (extended)

The Lower Campanian Star Point Sandstone contains several shallow-marine

parasequences that pass basinward into the Mancos Shale (Fig. 3). The Panther Tongue of the

lower Star Point Sandstone has been interpreted as a detached fluvial delta from which some of

the shallow-marine parasequences received sediments (Posamentier and Allen, 1993). The upper

parts of the Star Point Sandstone such as the Spring Canyon Member pass landward into the

Blackhawk Formation (Fig. 3). The lower to middle Blackhawk Formation is mostly mudstone

but contains isolated fluvial sandbodies and numerous coal seams that decrease in prevalence

upwards (Fig. 3). The proportion of channelized sandbodies increases upwards, from ca. 10% to

ca. 30% (Hampson et al., 2012). The Blackhawk Formation transitions basinward from coastal-

plain and fluvial strata to shoreface sandbodies, which pass further basinward into the Mancos

Shale (Fig. 3). The Blackhawk Formation has a rising, concave-down shoreline trajectory

(Hampson, 2010). The total Blackhawk Formation thickness (~250 m in the Wasatch Plateau)

tapers down-dip and pinches-out within the Mancos Shale in the Book Cliffs near the Green

River area (Fig. 3). Stratigraphic architecture changes abruptly across the unconformable contact

between the Blackhawk and Castlegate Formations.

In the lower Blackhawk Formation of the Wasatch Plateau, multi-story fluvial sandbodies

confined within sequence-bounding incised valleys transition laterally to well-developed

interfluvial calcareous paleosols (Gani et al., 2015). Gani et al. (2015) also identified coal seams

to correlate with marine flooding surfaces in these 4th order sequences. This presence of coeval

marine and nonmarine strata along basin-strike serves as one line of evidence for a marine-

Page 53: High-frequency tectonic sequences in the Campanian

46

embayment that encompassed the northern Wasatch Plateau and northwestern Book Cliffs.

Known as the Utah Bight (Hampson, 2010), this marine embayment was a long-lived (>5 m.y.)

depocenter and may have influenced Castlegate Formation sequence architecture as discussed

later herein.

Page 54: High-frequency tectonic sequences in the Campanian

47

RESULTS (EXTENDED)

Facies

1. Organic-rich mudstone

Dark gray mudstone containing a greenish-gray weathering-tint has massive or nodular

bedding. These deposits are moderately pedogenic (e.g. gleysol) with root tubes that harbor

diagenetic concretions. Woody debris and charcoal also occur.

This facies is interpreted to have accumulated in a vegetated floodplain. The dark color of

this facies is related to the organic content.

2. Bioturbated sandstone

Fine to medium grained sandstone displays bioturbation. Bedding is thin to massive.

Some thinly bedded sandstone contains siderite nodules. Coal chips occur locally. Ichnofossils

include Teredolites longissimus, Teredolites clavatus and Fugichnia. The bioturbation index (BI)

of this facies does not exceed 2.

This facies is interpreted as coastal plain deposits. The presence of wood-boring

Teredolites trace fossils indicates brackish-water setting (Shanley and McCabe, 1992). Siderite

nodules are diagenetic and formed during burial in a reducing environment such as a sediment-

starved swamp or floodplain (Taylor et al., 2000).

3. Intraformational lag

Conglomerate with fine to medium grained sandy matrix. The deposits are poorly sorted

and infill scour pits or blanket erosion. Fills comprise pebble and cobble-sized clay rip-up clasts

and plant materials such as large tree stumps, small twigs and leaves. Some plant materials

display Teredolites longissimus and Teredolites clavatus ichnofossils.

Page 55: High-frequency tectonic sequences in the Campanian

48

This facies represents sand-dominated fluvial deposition. Intraformational lags are

interpreted as channel-thalweg deposits (McLaurin and Steel, 2007). They represent high-energy,

probable debris flows (Sohn et al., 1999).

4. Extraformational lag

Quartz, chert, or arkosic pebble-cobble conglomerate with a medium to coarse-grained

sandy matrix. Most commonly, the facies contains sub-rounded gravels. Sub-angular gravels are

typically arkosic or cherty and less common. Grains are poorly sorted and randomly-oriented.

This facies is interpreted as alluvial fan and gravel-dominated fluvial deposits. The lack

of plant material and mud rip-up clasts, and presence of rounded pebbles suggest that these

deposits are extraformational, potentially transported by antecedent rivers that dissected the

Sevier thrust belt. The round pebbles and cobbles indicate a long fluvial transit. The matrix-

support of this facies suggests deposition from dense inertia flows or “traction carpets” (Sohn et

al., 1999). Poor sorting and absence of grading supports that this facies records river flood-stage

deposition (Sohn et al., 1999).

5. Sandstone with soft-sediment deformation

Sandstone with convolute laminae, slump folds, recumbent folds, flame structures, or

dish-and-pillar structures.

Soft-sediment deformation doesn’t necessarily indicate an environment of deposition as it

is the product of sudden incompetency or instability (i.e. loss of strength) in semi-liquefied

sediments (Kundu et al., 2011). These instabilities are commonly local in origin, controlled by

slope-failure, rapid sedimentation or differential loading. Seismicity can also produce soft-

sediment deformation (Kundu et al., 2011; Balsamo et al., 2013). Seismic events are evident if

Page 56: High-frequency tectonic sequences in the Campanian

49

the sedimentary structures are spatially extensive, forming a horizon bounded by undeformed

strata (Kundu et al., 2011). McLaurin and Steel (2007) argue though that at the type-locality,

soft-sediment deformations in sandstone lack continuity, and thus represent bank slumping into

adjacent channels. Further south, we found however rather continuous (multi-km) soft-sediment

deformation horizons in the Wasatch Plateau, likely suggesting their seismic origin. Some

occurrences of soft-sediment deformation appear more localized as McLaurin and Steel (2007)

suggested.

6. Ripple cross-laminated sandstone

Fine to medium-grained sandstone displays asymmetric ripple cross-lamination. Rare

climbing ripples occur.

Ripple cross-laminated sandstone reflects either bar-top, levee, or floodplain deposition

(Reference). Vertical grain size trends within rippled intervals assist in specifying whether these

are bar or over bank deposits.

7. Dune cross-stratified sandstone

Fine to medium grained sandstone displays dune-scale trough cross-stratification. Cross-

set thicknesses vary between 5 and 50 cm.

This facies is interpreted to record deposition in either the middle-portion of channels

bars or in overbank splays (Reference). Dune set-thickness is proportional to formative water

depth whereby the smaller the set-thickness, hence dune, the shallower the water (LeClair and

Bridge, 2001).

8. Parallel-laminated sandstone

Page 57: High-frequency tectonic sequences in the Campanian

50

Fine to medium grained sandstone displays planar cross-lamination. The planar cross-

laminations are typically inclined-gently relative to horizontal.

This facies is interpreted to reflect deposition on the lower to middle portions of braided-

channel bars, adjacent to the channel thalweg (Reference).

9. Sandstone with rhythmic mudstone

Sandstone with subordinate rhythmic mudstone occurs as thin to thick beds. Sedimentary

structures include herringbone cross-stratification, flaser bedding, sigmoidal bedding, inclined

heterolithic strata, and dune or ripple cross-stratification with single or double mud drapes.

Cross-stratifications commonly have bi-directional paleocurrents. Ichnofossils include Fugichnia

and Ophiomorpha.

This facies is interpreted to reflect tidally-influenced fluvial and estuarine deposits

(Yoshida et al., 2006). The heterolithic yet sandstone-dominated nature of this facies is due to

higher depositional energy with a relative abundance of coarse-grained compared to fine-grained

detritus.

10. Mudstone with lenticular sandstone

This facies consists of mudstone with lensoidal fine-grained sandstone. These deposits

show coarsening-upward trend. Sedimentary structures include wavy-lenticular bedding.

Locally, there is comminute-distribution of organic detritus. Ichnofossils include Fugichnia and

Ophiomorpha.

This facies is interpreted as part of bay-head or tidal-deltaic deposits. Despite having a

coarsening-upward grain size trend, a fluvial overbank splay interpretation for these deposits is

Page 58: High-frequency tectonic sequences in the Campanian

51

discarded because of the trace fossil assemblage, wavy bedding, and absence of pedogenic

modifications.

Page 59: High-frequency tectonic sequences in the Campanian

52

DISCUSSION (EXTENDED)

Manifestation of deformation-style in stratigraphic surfaces (Extended)

Surface-loading by retroarc fold-thrust belts cause lithospheric depression in the direction

of thrust-advancement, strongly dictating sediment-accumulation across broad areas (Ettensohn,

2004). Therefore, if fold-thrust belt loads decrease along-strike, the cratonward accommodation

does too. Observable stratigraphic and facies architecture in the underlying Blackhawk

Formation and Star Point Sandstone form complex relationships along the margin of a long-lived

depocenter, where there was a large marine embayment known as the Utah Bight (Hampson,

2010; Gani et al., 2015) that spanned a region cratonward of the Charleston-Nebo thrust system.

The inter-depocenter regions therefore were hypothetically cratonward from inactive thrust

sheets. Although accommodation was relatively higher within the depocenter, it was still

relatively low when compared to the middle and upper Campanian. Accommodation was

reduced but active thrusting should promote the opposite, trapping of coarse-grained sediments

within the wedge-top depocenter, and fine-grained organic rich deposition across wide areas

further basinward. It is possible that variable loading within the fold-thrust belt caused the

lithosphere to rise or subside and had some contribution to the observed erosional beveling, but

given the long-transit progradation of coarse-grained material, there was likely an additional

force acting to reduce accommodation.

In Wyoming, Cenomanian Second Frontier marine strata record a similar tectonic

influence characterized by episodic, localized uplift ensued by development of angular

unconformities (Vakarelov et al., 2006). The localized uplift was attributed to nascent growth of

the Laramide Tisdale Anticline (Vakarelov and Bhattacharya, 2009). These Second Frontier

angular truncation surfaces commonly contain Glossifungites and overlie a pebble lag facies.

Page 60: High-frequency tectonic sequences in the Campanian

53

Surfaces AU1 and AU2 which formed in terrestrial setting, contain Teredolites rather than

Glossifungites but the bioturbation associated with angular unconformities from both the Second

Frontier and Castlegate Formation is strongly indicative of subaerial exposure.

Page 61: High-frequency tectonic sequences in the Campanian

54

VITA

David Cross was born in Lexington, KY on March 21, 1990 to Mark and Cindy Cross.

He graduated from Lexington’s Henry Clay High School in 2008. He later received a Bachelor

of Science degree in Geology from the University of Kentucky in 2012. Also in 2012, David was

awarded a prestigious internship with the United States Geological Survey in Menlo Park, CA as

part of the cooperative USGS-NAGT Field Training Program. He began pursuit of a Master of

Science degree focused on Geology at the University of New Orleans in 2013. While at the

University of New Orleans and under the supervision of Dr. M. Royhan Gani, he coauthored a

paper which is now published in Sedimentary Geology. In 2013, he participated in the Imperial

Barrel Award Program in which his team earned 2nd place honors in the eastern Gulf Coast

Region. In 2015, he was awarded to participate in the ExxonMobil-GSA Bighorn Basin Field

Program.

Permanent address: 2033 Hillgate Dr. Lexington, KY 40515

This thesis was typed by the author.