Faulty foundations: Early breakup of the southern Utah Cordilleran foreland basin

Gabriela A. Enriquez St. Pierre† and Cari L. Johnson University of Utah, Department of Geology and Geophysics, 115 S 1460 E, Salt Lake City, Utah 84112, USA

ABSTRACT cations for understanding variability and - during active thrust phases (Burbank and Beck, dynamics of foreland basin evolution. 1988; Garcia-Castellanos, 2002; Horton et al., Anomalous features of Upper Cretaceous 2004; Thomson et al., 2017). Although flexure strata in southern Utah challenge existing INTRODUCTION is considered the main subsidence control on tectonic and depositional models of the Cor- foreland basin evolution and forebulge migra- dilleran foreland basin. Extreme thickness Global archives suggest that retroarc foreland tion (Catuneanu et al., 1997; Currie, 1998; variations, net to gross changes, and facies basins are broadly defined by several common Houston et al., 2000), dynamic subsidence is distributions of nonmarine to marginal characteristics (Jordan, 1981; DeCelles and also an important potential driving mechanism marine strata of the Turonian–early Campan- Giles, 1996; Catuneanu, 2004; DeCelles, 2012). (Mitrovica et al., 1989; Liu and Nummedal, ian Straight Cliffs Formation are documented These fundamental attributes include (Fig. 1): a 2004; Romans et al., 2010; Painter and Carrapa, across the Southwestern High Plateaus. Con- basinward-propagating, thin-skinned (detach- 2013; Fosdick et al., 2014). Finally, onset of trary to most traditional models of foreland ment-style) fold-thrust belt; concomitant migra- Laramide-style (basement-involved) tectonism basin architecture, regional correlations dem- tion of four main depositional zones (wedge-top, in latest Cretaceous time (ca. 75–50 Ma; Dumi- onstrate abrupt stepwise thickening, with a foredeep, forebulge, and backbulge); a strongly tru et al., 1994; Crowley et al., 2002; Saleeby, punctuated increase in average of asymmetric, continuous foredeep with greatest 2003; Tindall et al., 2010; Peyton et al., 2012; key intervals from west to east, i.e., proximal subsidence most proximal to the flexural load; Weil and Yonkee, 2012; Yonkee and Weil, 2015) to distal relative to the fold-thrust belt. Except clastic material sourced largely from the fold- fundamentally changed Cordilleran foreland in the most proximal sections, fluvial drainage thrust belt (particularly in proximal sections); basin architecture, subdividing the foredeep into systems were oriented predominantly subpar- and general coarse to fine-grained facies tran- multiple isolated basins thus forming a “broken allel to the fold-thrust belt. Combined, these sitions across the depozones from proximal foreland” (Dickinson et al., 1988; Schmidt et al., results suggest that modern plateau-bounding to distal. 1993; Erslev, 1993; Strecker et al., 2011; Dávila faults may have had topographic expressions Many of these characteristics are well and Carter, 2013). as early as Cenomanian time, and influenced expressed in the North American Cordillera, Evolution of the Sevier fold-thrust belt the position of the main axial river system by particularly in the Cretaceous Sevier fold- and foreland basin system in southern Utah creating northeast-trending paleotopography thrust belt and foreland basin system of Utah (Fig. 1B) is poorly understood relative to other and sub-basins. Laramide-style tectonism (Armstrong, 1968; Jordan, 1981; Lawton et al., parts of the Cordillera. Major thrust structures (e.g., basement-involved faults) is already 1994; DeCelles, 1994; DeCelles and Giles, appear to have been emplaced by the Cenoma- cited as a driver for sub-basin development 1996; DeCelles and Currie, 1996; Currie, 2002; nian (DeCelles, 2004), with little evidence of in latest Cretaceous–Cenozoic time, but new DeCelles, 2004; Fig. 1). Variations on the broad major basinward propagation over time. Pre- data presented here suggest that this part of conceptual themes of foreland basin evolution liminary data from foreland basin strata suggest the foredeep was “broken” into distinct sub- are also widely acknowledged. For example, unusually high accommodation, sediment sup- basins from its earliest stages. We suggest that proximal to distal fining grain size trends can ply, and coarse-grained deposition in distal parts flexural foundering of the lithosphere may be disrupted by complex sediment routing sys- of the foredeep, with relatively stationary (non- have caused early stage normal faulting in the tems, such as large axial rivers and distributive migrating) depozones (Eaton and Nations, 1991; foredeep. Regional implications of these new fluvial systems (DeCelles and Cavazza, 1999; Goldstrand et al., 1993; Mulhern and Johnson, data indicate that both detachment-style and Hartley et al., 2010; Weissmann et al., 2010), 2016). Although flexure adjacent to the Sevier basement-involved structures were simulta- (Swift et al., 1987; Slingerland fold-thrust belt is thought to be a primary con- neously active in southern Utah earlier than and Keen, 1999), progradation of major delta trol on subsidence and sedimentation patterns in previously recognized. These structures were systems (Ahmed et al., 2014; Fielding, 2015), northern Utah and Wyoming before ca. 83 Ma likely influenced by inherited Proterozoic and tidal deposits (Chentnik et al., 2015; Van (DeCelles, 2004), previous studies of southern basement heterogeneities along the edge of Cappelle et al., 2018). Similarly, punctuated Utah foreland basin stratigraphy indicate a more the Colorado Plateau. This interpretation and widespread deposition of gravel and complicated history from at least Cenomanian suggests that tectonic models for the region sheets extending beyond the proximal fore- time, long before onset of Laramide-style tec- should be reevaluated and has broader impli- deep are documented (Weissmann et al., 2011; tonics (Gustason, 1989; Eaton and Nations, Lawton et al., 2014), and these deposits likely 1991; Laurin and Sageman, 2001). reflect redistribution of sediment during iso- Cenozoic normal faults bound a series of †[email protected]. static rebound, and/or syntectonic deposition plateaus in the study area (Southwestern High

GSA Bulletin; Month/Month 2021; 0; p. 1–20; https://doi.org/10.1130/B35872.1; 10 figures; 2 tables. published online 28 May 2021

© 2021 The Authors. Gold Open Access: 1 This paper is published under the terms of the CC-BY license.

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

Figure 1. Simplified maps of A B key Precambrian to Late Cre- taceous structures across the Utah (UT) Cordilleran oro- genic system (with parts of Ne- vada (NV) and Arizona (AZ)). The field area for this study is marked with a red box outlin- ing the Southwestern High Pla- teaus, including the Gunlock and Pine Valley areas (G and PV), the Markagunt Plateau (MP), the Paunsaugunt Pla- teau (PP), and the Kaiparow- its Plateau (KP). (A) Relevant basement structures include the Wasatch Hinge Line (Picha and Gibson, 1985); the suture zone between the Mojave and Yavapai provinces (Karlstrom and Shaw, 1999); and Neopro- C terozoic normal faults and lin- eaments (Picha, 1986; Currie, 2002). Black line A–A′ repre- sents cross section locations for Figures 7 and 8. (B) Relevant structures formed during Late Cretaceous time include the Sevier fold-thrust belt (SFTB) and Laramide monoclines. Key thrust sheets of the Sevier fold- thrust belt in central Utah in- clude the Canyon Range (CR), Pavant (PV), Paxton (PX), and Gunnison (GU) thrusts, and in southern Utah the Blue Moun- tain (BM), Wah Wah (WW), and Iron Springs (IS) thrusts. Laramide monoclines (orange; arrows showing dip direction) include the Monument Uplift (MU), Circle Cliffs Uplift (CC); East Kaibab mono- cline (KM), and Keystone Uplift/ Fault (KF). (C) Schematic cross section of the Sevier foreland basin system in southern Utah highlighting the location of the field area for this study, and relative positions of the foreland basin depocenters during the Late Cretaceous. WY—Wyoming.

Plateaus, SWHP; Fig. 2), exposing relatively We completed a regional correlation across and Mesozoic strata which provided sediment to undeformed Upper Cretaceous strata ∼200 km the southern margin of the Sevier foreland basin the adjacent foreland basin to the east (DeCelles west to east (proximal to distal relative to the to investigate controls on foreland basin devel- and Coogan, 2006). At its southwestern extent, fold-thrust belt), and ∼50 km of outcrop from opment, and in particular their effect on fluvial the Sevier fold-thrust belt formed a high-angle south to north. Whereas fluvial architecture systems during Late Cretaceous time. Targeted junction with the Mogollon Highlands (Fig. 1) along the western Kaiparowits Plateau has been measured sections document stratigraphic archi- in northwestern Arizona. The geometry of these studied in some detail (Gooley et al., 2016; Ben- tecture, proximal-to-distal changes in thickness features allowed for incursion of the Utah , hallam et al., 2016; Primm et al., 2018; Koch and average grain size, and paleocurrent indi- an embayment filled by the Cretaceous Western et al., 2019), much less is known about fluvial cators. Ultimately, these data allow us to infer Interior Seaway (Fig. 2 inset; Zapp and Cobban, systems to the west in the Markagunt and Paun- controls on sedimentation and to test prevailing 1960; McGookey et al., 1972). saugunt Plateaus, which preserve more proxi- models for foreland basin development. The Sevier fold-thrust belt in northern and mal deposits of the Cretaceous foreland basin. central Utah shows evidence for eastward Regional correlations to date focus mainly on GEOLOGIC BACKGROUND propagation of thin-skinned thrust sheets, biostratigraphy and marine deposits, but do not punctuated periods of out-of-sequence thrust- detail regional fluvial interactions (Eaton and Structural Evolution of Southern Utah ing, and growth of structural culminations, all Nations, 1991; Eaton et al., 1993; Goldstrand, of which influenced subsidence and sediment 1992, 1994; Moore and Straub, 2001; Biek The Sevier fold-thrust belt trends northeast supply (Fig. 1B; DeCelles et al., 1995). The et al., 2015). across Utah, deforming and exhuming Paleozoic Canyon Range thrust system represents the first

2 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

Figure 2. Inset map shows present day location of the Southwestern High Plateaus (SWHP) in Utah including relevant structures from Late Cretaceous to Cenozoic time. WIS—Western Interior Seaway; AZ—Arizona; CA—California; CO—Colorado; NM—New Mexico; NV— Nevada; UT—Utah; WY—Wyoming. Main map of the study area shows the locations of measured sections and relevant stratigraphy across the SWHP. Sections of the Straight Cliffs and Iron Springs Formations completed for this study are shown in orange. Generic lithologies are represented by brick (mixed limestone and sandstone) and stippled (dominantly sandstone) patterns; non-stippled is a mixture of mud- stone and sandstone lithologies. Sections from this study (orange) and previous studies include: (1) Pine Valley Mountains (Cook, 1957); (2) Parowan Gap west of the SWHP; (3) Parowan Canyon; (4) Cedar Canyon (Eaton et al., 2001); (5) Orderville Gulch in the Markagunt Plateau; (6) Glendale and (7) Tenny Canyon (Tilton, 1991); (8, 9) Willis Creek and Heward Creek from the Paunsaugunt Plateau; and (10) Shakespeare Mine from the northern Kaiparowits Plateau. Previously measured sections in the Kaiparowits Plateau include: (11) Buck Hollow (Mulhern and Johnson, 2016); (12) Main Canyon (Chentnik et al., 2015); (13) Left Hand Collet (Dooling, 2013); (14) Rogers Can- yon (Allen and Johnson, 2010); (15) Kelly Grade (Gallin et al., 2010); (16) Tibbet Canyon (Peterson, 1969); (17) Bull Canyon (Gooley et al., 2016); and (18) Rock House (Gooley et al., 2016).

major thrusting episode in central Utah, initiat- vation and incorporation of the Canyon Range sional overprinting. The Wah Wah thrust sheet ing during Barremian–early Aptian time (ca. and Pavant thrust sheets, which created a topo- was active from ca. 120 to 97 Ma and is primar- 130–120 Ma; DeCelles and Coogan, 2006). The graphic high, providing sediment to the foreland ily composed of back-breaking imbricate thrusts Pavant thrust system propagated east during Late basin through Paleocene time (DeCelles et al., (Fig. 1B; Miller, 1963; Fillmore and Middleton, Aptian–Cenomanian time. By the Cenomanian 1995; Stockli et al., 2001). 1989; Friedrich and Bartley, 2003; DeCelles, (ca. 100 Ma), the Paxton thrust sheet initiated The structure and timing of major thrust faults 2004). The nearby Blue Mountain thrust system growth of the Sevier culmination (DeCelles in southwestern Utah are poorly known relative ∼10 km to the southeast was emplaced by lat- et al., 1995). This culmination included reacti- to central Utah, partly due to Cenozoic exten- est Cenomanian–Turonian time (ca. 95–85 Ma;

Geological Society of America Bulletin, v. 130, no. XX/XX 3

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

Miller, 1966; DeCelles, 2004; Fig. 1B). The Iron Present-day structure and topography in section and an overlying estuarine-influenced Springs thrust system is the youngest and east- the study area is reflected in the Markagunt, section (Primm et al., 2018). ernmost expression of the Sevier fold-thrust belt Paunsaugunt, and Kaiparowits Plateaus, which The Coniacian–Santonian John Henry Mem- in southern Utah, and it is characterized by three together comprise the Southwestern High Pla- ber comprises the thickest part of the Straight east-vergent, forward-breaking thrust sheets. It is teaus. These modern features are bounded by Cliffs Formation and preserves fluvial-alluvial estimated to have been active in Santonian-Cam- Cenozoic normal faults (from west to east: the deposits in the west, with marginal marine and panian time (ca. 84–70 Ma; Goldstrand, 1994), Hurricane, Sevier, and Paunsaugunt faults), deposits in the eastern Kaiparowits but drag folds in Paleocene-Eocene conglomer- which mark the transition between the Basin and Plateau (Peterson and Waldrop, 1965; Eaton ates indicate minor late stage displacement as Range and Colorado Plateau provinces (Fig. 2). and Nations, 1991; Shanley and McCabe, 1995; well (Anderson and Dinter, 2010). These faults have offsets of up to ∼1000 m, and Hettinger et al., 1996). Marine deposits record In contrast with the central Utah fold-thrust are thought to have initiated during the early shoreface and barrier systems belt (e.g., Canyon Range, Pavant, Paxton-Gun- Miocene, with the Hurricane fault showing mod- with initially mainly transgressive then regressive nison thrusts), the southern thrust systems did ern active seismicity (Maldonado et al., 1997; stacking patterns throughout the Santonian (Het- not consistently propagate eastward, and surface Davis, 1999; Lund et al., 2008; Biek et al., 2015). tinger et al., 1993; Shanley and McCabe, 1995; expressions stalled ∼100 km farther west of the Allen and Johnson, 2010; Mulhern and Johnson, easternmost central Utah thrust sheet (Fig. 1B). Stratigraphic Evolution of Southern Utah 2016). Extensive coal zones occur within the Younger faults in the area include the Ruby’s Inn John Henry Member of the central Kaiparowits and Paunsaugunt thrust faults, but these struc- The SWHP preserve relatively undeformed Plateau, which are correlated to fluvial sections tures are attributed to the emplacement of the Upper Cretaceous strata of the Cordilleran fore- in the west that vary in terms of fluvial style and Oligocene-Miocene Marysvale Volcanic Field land basin system. These deposits include marine tidal influence (Shanley and McCabe, 1993; (Merle et al., 1993; Davis, 1999; Biek et al., to fluvial-alluvial strata of the Naturita (formerly Hettinger et al., 2009). The upper boundary of 2019). Proprietary industry data and a regional Dakota), Tropic, Straight Cliffs, Wahweap, and the John Henry Member records a transition gravity survey (e.g., Upper Valley Oil field near Kaiparowits Formations (Fig. 3; Peterson, 1969; to higher-energy, coarser grained fluvial sheet- Cedar City; Cook and Hardman, 1967; Van Gustason, 1989; Eaton and Nations, 1991). The deposits of the basal Drip Tank Member (Lawton Kooten, 1988) indicate no evidence for sub- Straight Cliffs Formation, of primary interest et al., 2003, 2014; Gooley et al., 2016). surface thrust faults east of the Iron Springs to this study, overlies the marine Tropic Shale Unconformities throughout the Cretaceous thrust system. and records the initial regression of the Western sections are important considerations for inter- Basement-involved features in southern Utah Interior Seaway (Fig. 3, Peterson, 1969). Depo- preting thickness variations. Unconformities began to form during the Late Cretaceous as sitional environments of the Straight Cliffs For- within the Straight Cliffs Formation are con- asymmetric folds with varying orientations mation transition from proximal fluvial-alluvial sidered relatively minor sequence boundaries, (Fig. 1B; Davis and Bump, 2009). The Kanarra, systems west of the Markagunt Plateau to coal- with minimal erosion into underlying sections East Kaibab, and Circle Cliffs monoclines are bearing marginal marine strata in the eastern (Peterson, 1969; Biek et al., 2015; Primm the most prominent Laramide-style features in Kaiparowits Plateau (Eaton and Nations, 1991; et al., 2018). More significant unconformities this region (Fig. 1B). Movement along the East Little, 1997). Changes in fluvial and marginal across the SWHP occur in latest Campanian- Kaibab monocline during deposition of the marine stacking patterns within the Straight Maastrichtian successions, between deposition Campanian Wahweap Formation is the earliest Cliffs Formation have been attributed to sea- of the Kaiparowits, Canaan Peak, and Claron evidence for Laramide-style deformation in this level fluctuations (Shanley and McCabe, 1993) Formations, and are principally located in the area (Fig. 1B; Tindall and Davis, 1999; Hilbert- and/or interactions between axial and transverse Paunsaugunt and Kaiparowits Plateaus (Fig. 3; Wolf et al., 2009). drainages, which suggest tectonic and possibly Goldstrand, 1994; Biek et al., 2015; Beveridge The nature and orientations of Cretaceous climatic influences on deposition (Lawton et al., et al., 2020). These unconformities are largely structures in southern Utah are influenced by 2014; Szwarc et al., 2015; Gooley et al., 2016; attributed to Laramide-style deformation and basement features underlying the Colorado Pla- Primm et al., 2018). Recent geochronologic data formation of structures like the East Kaibab teau (Picha and Gibson, 1985; Schwans, 1995; from the Straight Cliffs Formation indicate that and Dutton monoclines (Roberts, 2007; Hilbert- Neely and Erslev, 2009). Beneath Phanerozoic deposition spans Turonian–early Campanian Wolf et al., 2009; Lawton and Bradford, 2011). strata is a northeast-trending shear zone of the time (ca. 94–81 Ma; Primm et al., 2018; Fig. 3). To date, there is little evidence that sub-Claron– Paleoproterozoic Yavapai-Mojave basement, The Tibbet Canyon Member of the Straight aged erosion significantly affected the Straight which overlaps with the southwestern edge Cliffs Formation is a thick shoreface sandstone Cliffs Formation in this area (Lawton et al., of the Colorado Plateau (Fig. 1A; Shaw and deposited during shoreline regression overly- 2003; Biek et al., 2015). Karlstrom, 1999; Davis and Bump, 2009). Pro- ing the marine Tropic Shale (Peterson, 1969; terozoic normal faults overprinted the Yavapai- Shanley and McCabe, 1991, 1995; Fig. 3). It is METHODS Mojave shear zones during breakup and rifting overlain by the Turonian-Coniacian fluvial and of Rodinia (Dickinson, 2004). Basement-cored paralic (marginal marine), coal-bearing Smoky Measured sections across the SWHP docu- faults of the Laramide orogeny align with these Hollow Member (Bobb, 1991; Hettinger et al., ment stratigraphic architecture (Figs. 4, 5, and Proterozoic normal faults (e.g., the East Kaibab 1993; Hettinger, 2000). This unit is capped by 6) as well as proximal-to-distal changes in monocline; Fig. 1), suggesting that basement the Calico Bed, a distinctive regional marker thickness, grain size, and paleocurrent indica- heterogeneities influenced the location of Cre- unit composed of gravelly sheet- (Peter- tors (Fig. 7). Six new measured sections of the taceous structures (Picha, 1986; Huntoon, 1993; son, 1969; Bobb, 1991; Little, 1997). In the John Henry Member are combined with previ- Stone, 1993; Tindall and Davis, 1999; Marshak eastern Kaiparowits Plateau, a minor unconfor- ous work from Tilton (1991) in the Paunsau- et al., 2000; Bump and Davis, 2003; Neely and mity (ca. 87–89 Ma) divides the Calico Bed into gunt Plateau to understand basin-wide strati- Erslev, 2009). two informal units: a lower amalgamated fluvial graphic and thickness trends. Net to gross

4 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

Figure 3. Chronostratigraphy diagram across the Southwest- ern High Plateaus (SWHP), with relevant formation names, thicknesses, and ages, compiled from biostratigraphic studies, maximum depositional ages of detrital zircons, and ashes where present. Maximum depositional ages calculated using the weighted average of the youngest three ages that overlap at 2σ; Youngest single grains (YSG) of two members of the Straight Cliffs Forma- tion included. Error bars are 2σ uncertainty. References are as follows: a—Roberts (2007); b—Jinnah et al. (2009); c—Sz- warc et al. (2015); d—Primm et al. (2018); e—Barclay et al. (2015); and f—Dyman et al. (2002). ID-TIMS—Isotope di- lution–thermal ionization mass spectrometry.

(NTG) is calculated by measuring sandstone these widespread deposits are likely to be at elements include bedforms and individual bed thickness (net) relative to total section thick- least partly time-transgressive, previous studies boundary relationships that fill a single ness (gross) and reported as a decimal num- and available age-control suggests they are rea- (e.g., scour-and-fill structures). Channel stories ber (0.0–1.0, net thickness/gross thickness). sonable datums relative to the age of the entire are defined as a single channel unit with related Paleocurrent data were collected from trough section and are suitable to guide regional cor- channel-fill elements, separated by conspicuous axes, ripples, and barforms where possible, and relations (Peterson, 1969; Lawton et al., 2014; basal scours that are laterally continuous. Chan- combined with previous work of Tilton (1991) Biek et al., 2015). nel belts are defined as laterally and/or vertically to create a more holistic picture of paleoflow coalescing channel stories that form a single directions (Fig. 2). RESULTS elongate sandstone body (Gibling, 2006; Ford Correlation across the SWHP from proximal and Pyles, 2014). to distal (west to east) is based on the formation Lithofacies and Facies Associations descriptions of Peterson (1969) from the Kai- Floodplain Facies Associations (FA-1 and parowits Plateau, and extrapolated to the west Thirteen lithofacies were identified through FA-2) with the aid of publications, theses, and regional field observations (Table 1), and these were Description. Facies association 1 (FA-1) reports (Tilton, 1991; Eaton and Nations, 1991; grouped into four facies associations (Table 2: consists of fine-grained lithofacies, including Moore and Straub, 2001; Lawton et al., 2014; FA-1 to FA-4) that are distinguished by distinct massive (Fm) and laminated mudstones (Fl) Biek et al., 2015), as well as new measured sec- channel architectural patterns and recurring and siltstones with minor lenticular to tabular tions presented herein. Regional marker beds associations of lithofacies. Channel facies asso- sandstones (Table 2; Fig. 4). Mudstones are used for correlation within the Straight Cliffs ciations were distinguished based on channel expressed as structureless beds ranging from Formation include the Tibbet Canyon Member architecture, bed boundary relationships, and 0.1 to 30 m thick and are commonly light gray, shoreface and the Calico Bed and Drip Tank grain-size distributions (e.g., Miall, 1996; Sling- but can be yellow, red, and tan in color (Fig. 4B, Member gravelly sheet sands (Fig. 3). Although erland and Smith, 2004). Specific channel-fill 4C). Locally, mudstones have blocky structures,

Geological Society of America Bulletin, v. 130, no. XX/XX 5

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

A B C

D

E

6 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

Figure 4. Representative stratigraphic column of the John Henry Member from Orderville mottled coloring (Fig. 4B), and may contain Gulch (Markagunt Plateau, Fig. 2) showing the relative abundance of Facies Association 1 plant rootlets. (FA-1) and FA-3 within the section (see Fig. 7 for key to stratigraphic columns). Outcrop Facies association 2 (FA-2; Fig. 4) consists photographs illustrating the differences between well-drained and poorly drained flood- of organic-rich fine-grained lithofacies (Fc) plain lithofacies of FA-1 and FA-2, and small, laterally restricted fluvial channels found and coal (C) with minor components of iso- within these facies. (A) Carbonaceous mudstone overlain by siltstone lithofacies from poorly lated sandstones (Table 1: Fc and C; Fig. 4A, drained floodplain deposits. (B) Variegated mudstone with blocky texture, (C) Siltstone to 4E). Fine-grained lithofacies of FA-2 are very fine-grained sandstone lithofacies. (D) Isolated asymmetric sandstone body with basal expressed as structureless beds ranging from cross-bedding. (E) Example of poorly drained floodplain carbonaceous mudstones and coal, red, yellow, light gray, and black to charcoal showing light red and yellow paleosol development. gray in color (Fm, Fl, Fc, C). Coal seams (sub- bituminous grade) are common and are up to

A

Figure 5. Representative strati- graphic column featuring facies association 3 (FA-3) and FA-1 within the John Henry Mem- B ber at Heward Creek (see Fig- ure 7 for key to stratigraphic columns). Outcrop photographs show: (A) Channel thalweg with erosive scour surfaces outlined in black. Note heterolithic nature of channel margins; (B) Distinct ribbon “steer’s head” channel C belt architecture in cross section with upper and lower boundar- ies outlined from Heward Creek section; (C) Well-developed gray mudstone lithofacies over- lain by tan siltstone facies com- monly found above and below isolated sandstone channels; (D) Soft ­sediment deformation typi- cal within thick sandstone units; (E) Horizontal unlined bur- D E row clusters found within thick sandstone bodies.

Geological Society of America Bulletin, v. 130, no. XX/XX 7

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

A

B

C

DE

8 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

Figure 6. Representative stratigraphic column of facies association 4 (FA-4) within the Iron and asymmetric) but can also be tabular, and are Springs Formation at Parowan Gap (see Fig. 7 for key to stratigraphic columns). Note the no more than 2 stories (Fig. 4D). Individual beds prevalence of internal scouring, trough cross-bedding, and woody fragments. Outcrop pho- are composed of very fine- to medium-grained tographs show: (A) sheet-like channel sandstones with erosive surfaces marked in black sand. Bedsets thin upwards from 0.1 to 0.5 m with scour surfaces (s); (B) soft sediment deformation features; (C) lags at the base thick. Sedimentary structures within FA-1 typi- of channels; (D) a representative photo of FA-4 within the Calico Bed of the Smoky Hollow cally include soft sediment deformation (Sc), Member at Orderville Gulch, showing large trough cross-bedding, and internal scouring low-angle cross-lamination (Sl), and asymmet- of sandstone units; and (E) preserved mudstone and siltstone fine-grained layers between ric ripples, particularly at the tops of sandstone sandstone channels. bedsets (Fig. 4D). Sedimentary structures within FA-2 include asymmetric ripples, planar to low- angle cross-bedding (Sl), with no visible struc- 1 m thick in the northern Kaiparowits Plateau Both FA-1 and FA-2 contain distinctive fine- tures in the thinnest beds (Sm). Bioturbation is (Fig. 4A). Darker carbonaceous shales are pres- to medium-grained, isolated sandstone units/ variable and infrequent for both FA-1 and FA-2, ent, with laterally variable oxidized lamina, and bedsets (Fig. 4D). These sandstone bedsets are and where present, individual Skolithos traces siderite concretions (Fig. 4A). Plant, coal, and up to 4 m thick and less than ∼10 m wide lat- are observed. Vertical successions of sandstones woody fragments exist throughout the carbona- erally. FA-1 and FA-2 sandstones commonly within FA-1 and FA-2 include basal beds with ceous mudstones. exhibit a lenticular geometry (both symmetric erosive bases, and trough cross-stratification

A A¢

Figure 7. Correlation of John Henry Member strata across the Sevier and Paunsaugunt faults, showing the relative sandstone to mudstone rations (net to gross, NTG), grain size, and thickness across the Southwestern High Plateaus (SWHP). Inset map shows cross section line across SWHP; see Figure 2 for detailed section locations. Upper datum of the correlation is the base of the Drip Tank Member, the top Calico Bed (upper Smoky Hollow Member) is the lower correlation marker. KP—Kaiparowits Plateau; MP—Markagunt Plateau; PP—Paunsaugunt Plateau.

Geological Society of America Bulletin, v. 130, no. XX/XX 9

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

TABLE 1. LITHOFACIES OF THE SMOKY HOLLOW AND JOHN HENRY MEMBERS Code Lithofacies Description Interpretation Gt Trough cross-stratified gravel Matrix-supported conglomerate with trough cross-stratification; clasts range from Migration of 3D unidirectional and sandstone <1 mm to 5 cm and are subangular to subrounded; matrix grain sizes range from fine- to very-coarse sand; cross-set height ranges from 10 cm to 0.5 m Gm Massive gravel conglomerate Matrix-supported conglomerate; clasts range from 2 mm to 5 cm and are High strength flow (debris flows) subangular to subrounded; matrix grain sizes range from very fine-very coarse sand St Trough cross-stratified Fine- to coarse-grained sandstone with cross-stratification; occasional pebble- Migration of 3D unidirectional dunes sandstone sized clasts; cross-set height ranges from 3 cm to 1 m Sr Ripple-laminated sandstone Very fine- to medium-grained sandstone with ripples; ripple heights are Current ripples under lower flow regime <5 cm currents Sc Convolute-bedded sandstone Fine- to coarse-grained sandstone with convolute and deformed bedding; Dewatering structures or post-depositional dewatering structures visible; height up to 1 m slumping Sp Planar cross-bedded Very fine- to medium-grained sandstone with planar cross-stratification; cross set Upper flow regime sandstone height up to 0.5 m Se Erosional sandstone Fine- to very coarse-grained sandstone with cross-stratification; cross- Rapid deposition of coarse bed-load stratification are up to 0.5 m thick and <15° angle Sl Low-angle cross-bedded Very fine- to medium-grained sandstone with planar to very low angle cross- Upper flow regime sandstone stratification Sm Massive sandstone Very fine- to medium-grained sandstone with no visible features Rapid sedimentation in channel-fill and splay deposits Fl Laminated sandstone, Interlamination of mudstone, siltstone and mudstones; may have small-scale Suspension settling during weak traction siltstone, and mudstone ripples currents Fm Massive mudstone/siltstone Structureless gray to variegated mudstone and siltstone; may contain fragments Waning-flow sedimentation of plant and coal material Fc Carbonaceous mudstone/ Structureless dark gray to black mudstone and siltstone with red and yellow Suspension settling with high occurrence siltstone nodules; abundant plant material and coal fragments of organic matter C Coal Black coal beds up to 1 m thick Floodplain proximal to peat swamps

that transitions upwards to planar or low-angle ceous material within mudstone and siltstone sandstone body, specifically narrow and discrete cross stratified beds, capped by ripple-laminated lithofacies of FA-2 suggests there was abundant channel belts versus broader sheet-like units. sand beds. organic matter present during floodplain depo- Description FA-3. Facies association 3 Fine-grained Facies Associations (FA-1 and sition. Coal lithofacies indicate the substantial (FA-3) consists of sandstone-rich lithofacies FA-2) Interpretation. FA-1 represents well- growth of vegetation and an elevated water table, (Fig. 5), including very fine- to coarse-grained drained floodplain deposition between active interpreted as deposition in peat swamps during sandstones (St, Sr, Sc, Sp; Table 1) with infre- ­fluvial channels. Siltstone and mudstone units times of limited clastic input (Fielding, 1987; quent gravels (Gt). Sandstone beds are 0.5–3 m represent overbank fines deposited during flood- Shanley et al., 1992; Hettinger et al., 1996). This thick, and form bedsets that are 5–7 m thick. ing events when bankfull discharge is exceeded rapid plant accumulation occurred under humid Observed sedimentary structures include trough and channel levees are breached (Slingerland and tropical conditions, as was common during depo- cross-beds that have a cross-bed height of Smith, 2004). The blocky mudstone structures, sition of the John Henry Member (Wolfe and 5–25 cm (St) and frequent soft sediment defor- mottled coloring, and plant rootlets represent Upchurch, 1987; McCabe and Shanley, 1992). mation, more commonly present in the thicker weakly developed paleosols and indicate brief Sandstone lithofacies are interpreted as overbank sandstone beds, including convolute bedding periods of subaerial exposure between active splay deposits in which channels were breached and dewatering structures (Sc; Figs. 5D). The channels and well drained floodplains (Kraus during times of high discharge or flooding events margins of the larger channel belts (>3 m) pre- and Wells, 1999; Kraus, 2002). Sandstone litho- (Slingerland and Smith, 2004; Donselaar et al., serve lateral accretion elements. Bioturbation facies represent crevasse splay deposits with cre- 2012). Lenticular and tabular sandstone lithofa- is scarce and varies by location, with the most vasse-channel development in some areas; these cies are interpreted as minor channel-fill succes- common identifiable traces being Skolithos and formed as the suspended and bedload material sions, abandoned during channel migration or Asthenopodichnium (Moran et al., 2010). Areas within a channel spills into the floodplain during (Slingerland and Smith, 2004). with higher bioturbation intensity are typically levee-breaching flood events (Smith et al., 1989; unlined burrow clusters within sandstone chan- Perez-Arlucea and Smith, 1999). Channel-Fill Facies Associations (FA-3 and nels (Fig. 5E). FA-2 represents poorly drained floodplain FA-4) Overall, channels within FA-3 exhibit lens- development between active fluvial channels Two channel facies (FA-3 and FA-4) were shaped architecture and are composed of a (Davies-Vollum and Smith, 2008). Carbona- distinguished based on the geometry of the sandstone-rich core with wings that thin laterally

TABLE 2. FACIES ASSOCIATIONS OF THE SMOKY HOLLOW AND JOHN HENRY MEMBERS Facies associations Lithofacies Architectural Elements Description FA 1: Well drained Fl, Fm, St, Asymmetric–symmetric Massive mudstones and siltstones dominate; mudstones variegated in places with pedogenic floodplain Sm, Sl, lenticular to tabular sandy structures; Fine- to medium- grained 0.5–3 m thick tabular to lenticular sandstones; Sr, Sh bedforms; accreting sandstones up to 7 m thick; beds thin and fine upward to fine-grained 0.1–0.5 m thick beds; barforms convolute bedding present; individual Skolithos burrows and clusters present. FA 2: Poorly drained C, Fc, Fl, Tabular to lenticular isolated Massive mudstone and siltstones; carbonaceous shale and 0.1–1 m thick coal beds; very floodplain Fm, Sm, sandy bedforms fine- to medium-grained sandstones beds< 3 m thick; less common than in well-drained Sr, Sh floodplain. FA 3: Major channel St, Sr, Sc, Complex major sandy Very fine- to medium-grained sandstone; tabular-lenticular beds range from 1–3 m thick; 5–7 m within well-developed Se, Fl ribbon channels and thick bed sets; trough cross-beds and soft sediment deformation common; mud clasts up to floodplain channel fill; often 4 cm wide at base of erosive scours; channel fill includes barforms and channel lags. multistory FA 4: Gravelly to sandy Gt, Gm, St, Tabular to lenticular laterally Medium- to very coarse-grained, fining-up sandstone and sandy conglomerate; 0.1–2 m thick fluvial sheet deposit Sc, Sr, Sl, extensive gravelly and bed sets; mud clasts and common at base of erosive scours; fluvial channel fill Fl, Fm sandy sheets includes channel lag and barform deposits.

10 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

into adjacent overbank and floodplain deposits Sandstone architecture is dominated by later- sandy to gravelly bedload. The multi-story char- (Kjemperud et al., 2008). Channels extend lat- ally extensive tabular sandstones, with erosive acter of these deposits records episodes of cut- erally from 40 to 100 m (Figs. 5A, 5B). Chan- scours between stories (Figs. 6A, 6E). Tabular ting and filling, likely a result of fluctuations in nel-fill style varies from simple single channels sandstones are often vertically and laterally discharge causing incision during high discharge to complex channel belts that comprise 6–10 amalgamated forming laterally extensive sheet events, and subsequent deposition during waning stories separated by erosive scours (Ford and complexes, but lateral and vertical channel flow (Holbrook, 2001). A lack of distinct channel Pyles, 2014). Larger channel-fill elements can amalgamation varies by location. Some sheet boundaries and laterally extensive tabular geom- be strongly vertically amalgamated, incising into deposits form channel belts that are laterally etries indicate lateral instability and a dominance lower stories and preserving incomplete chan- extensive and can be mapped for 100s of meters of unconfined flow (Smith et al., 1989; Owen nel fill and barforms (cf. Chamberlin and Hajek, with >5 stories identifiable, e.g., the Calico Bed et al., 2017). The coarse-grained nature of the 2019). Mud rip-up clasts up to 4 cm in diameter (Fig. 6E). In contrast, some channel belts have channel fills, numerous mud-rip-ups, and grav- are common at the base of these scour surfaces. a reduced lateral extent (10s of meters), but elly lags suggests high discharge with frequent In general, there is a decrease in grain size, bed instead have high vertical amalgamation (typi- bankfull flows (Gibling, 2006). The cut-and-fill thickness, and bedform height, with an increase cally between 1 and 3 stories) with rare fine- patterns combined with coarser grained chan- in mudstone content moving from the channel grained deposits preserved (Fig. 6A). Tabular nel fill may also indicate some level of inter- axis toward the channel wings (Fig. 5B). Lateral sandy channel belts are commonly 1–3 m thick nal reworking, suggesting a fluvial system that accretion sets occur along the margins of chan- but can be up to ∼5 m thick. migrated or avulsed more frequently (Straub and nels (Hassanpour et al., 2013). Esposito, 2013) or during relatively slow accom- Description FA-4. Facies association 4 (FA- Channel-Fill Facies Associations (FA-3 and modation creation (Miall and Arush, 2001). 4; Fig. 6) typically consists of coarser-grained FA-4) Interpretation bedforms with fine to gravelly sandstone litho- FA-3 Interpretation. Channel-fill successions Regional Stratigraphic Patterns facies (Table 2). Locally, gravel lithofacies are of FA-3 represent deposition of discrete channel more common (Gt, Gm; Fig. 6D). Gravel clast belts within well-developed floodplain, similar Figure 7 shows a regional correlation com- sizes range from 2 mm to 4 cm, and commonly to the “steers-head” channels of Kjemperud et al. posed of six new measured sections of the John consist of subangular to angular gray to black (2008). Channel deposits of FA-3 represent flu- Henry Member combined with two sections chert, quartzite and carbonate lithic clasts, and vial systems with a higher sediment suspended from Tilton (1991; Glendale and Tenny Can- mud rip-ups (Fig. 6D). These beds are often nor- load and lateral bank stability, interpreted as yon). Sections range from the northern Kaip- mally graded with larger grains and clasts at the distributary channels draining into floodplain arowits Plateau (Table Cliffs) region, to the east base of trough cross-beds and erosional scours deposits (Figs. 5A, 5B). Distinct channel belt and southern flanks of the Paunsaugunt Plateau (Se, St). Sandstone beds are 0.1–2.5 m thick forms suggest bank stability, allowing for lateral (PP; Fig. 2), and into the central and western and bedsets (greater than 2 m) typically exhibit and vertical aggradational deposition within the Markagunt Plateau (MP; Fig. 2). a fining upwards trend. Sedimentary structures channel margins (Gibling, 2006). observed within the sandstone and gravel-prone Channel-fill facies are locally adjacent to Thickness Trends bedsets include trough cross-beds and convolute lateral-accretion elements, although complete bedding (Se, St; Fig. 6B). Many bedsets contain barform preservation is rare. The lateral wings of Markagunt Plateau thicker beds at the base that have trough-cross channels are typically thinner-bedded and finer Measured sections from the northwestern and convolute bedding, which are overlain by grained with a higher preservation of interbedded edge of the Markagunt Plateau (Parowan Gap thinner, or “flaggy” bedding that is commonly mudstone and siltstones. These are interpreted and Parowan Canyon; Fig. 7) represent the ripple-laminated. Mud rip-up clasts, pebbles, as channel levees, where sediment spilled onto most proximal strata in the study area. Near and coarse-grained sandstones are common at overbank regions (Allen, 1978; Friend et al., the Wah Wah and Blue Mountain thrust faults the base of erosive scours and trough cross-beds, 1979). Channel levee development and preser- (Fig. 8), the Iron Springs Formation thickens particularly at the base of channel bedsets. - vation indicates a high level of bank stability that from at least 350 m as measured in this study stones and mudstones are infrequently preserved is concurrent with the higher frequency of veg- (Parowan Canyon), and up to ∼900 m thick to between channel stories (Fig. 6E). etated floodplain deposition seen in FA-3 (Sling- the southwest (Fig. 8, Gunlock and Pine Valley Bioturbation is variable, ranging from scarce erland and Smith, 2004). Additionally, preserva- Mountains; Cook, 1957; Hintze, 1986; Fillmore, to common, with occurrences of Skolithos, tion of overbank and inter-channel fine-grained 1991). The base of the Straight Cliffs Formation Asthenopodichnium (Moran et al., 2010) and material suggests there was a higher volume of (Tibbet Canyon and Smoky Hollow Members) escape traces. Bioturbation intensity is variable suspended sediment load relative to the sand- is not exposed in Parowan Gap. There, the par- across the plateaus and can be relatively high rich sheet deposits that comprise FA-4. Lateral tially exposed upper John Henry Member is locally (∼15% bioturbated). Wood casts (cm- accretion bedsets and barforms are infrequently ∼80 m thick (Fig. 7), and this section thickens scale to >1 m in length) are commonly preserved preserved and vary by location, but where pres- to ∼360 m in Parowan Canyon (∼15 km south- at the base of erosive scours. Asthenopodichnium­ ent indicate some level of reworking within east; Fig. 7). In the central part of the Marka- traces are visible within wood casts. Casts of channel margins (Mohrig et al., 2000; Hassan- gunt Plateau, the Smoky Hollow Member ranges therapod and ornithopod tracks are preserved pour et al., 2013). These fixed-location channels from ∼110–117 m thick (Orderville Gulch, on the bases of some channel deposits (e.g., could indicate a less mobile river system, with Fig. 7). The Calico Bed has a well-developed Parowan Canyon; Milner et al., 2006). Croco- less frequent avulsion and stable, well-developed ∼17-m-thick lower bed but pinches out along dilian, dinosaur, and turtle bone fragments erode river channels (Straub and Esposito, 2013). the western margin of the Markagunt Plateau. out of sandstone and siltstones of FA-4 (Eaton, FA-4 Interpretation. Sheet-like sandstone The John Henry Member has an average thick- 1999). Concretions 0.3–1.0 m thick are com- deposits of FA-4 represent unconfined flow within ness of ∼200 m along the south-central margin monly found in the larger sandstone beds. poorly developed fluvial channels dominated by of the Markagunt Plateau.

Geological Society of America Bulletin, v. 130, no. XX/XX 11

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

A A¢

Figure 8. Regional cross section showing Cretaceous strata across southwest Utah from Gunlock to the northern Kaiparowits Plateau (see inset map for cross section line, and Fig. 2 for detailed section locations). Datum is top Kaiparowits Formation.

Paunsaugunt Plateau ber in the southern portion of the Kaiparowits Paunsaugunt Plateau Thinning of the John Henry and Smoky Hol- Plateau is only ∼200 m thick (Gallin et al., 2010; Along the western and southern margins of low Members occurs across the Sevier and Paun- Allen and Johnson, 2010; Dooling, 2013; Pet- the Paunsaugunt plateau (Glendale and Tenny saugunt faults (Fig. 7). Within the Paunsaugunt tinga, 2013; Gooley et al., 2016). Canyon; Fig. 7), the John Henry Member con- Plateau, the Smoky Hollow Member is roughly sists of laterally and vertically amalgamated half as thick as it is to the west (∼30 m). John Architectural Trends sheet-like deposits (Fig. 6; FA-4) with minor Henry Member strata thin across the Sevier fault components of isolated channel belts within from ∼200 m to ∼160 m (Glendale, Fig. 7; Til- Markagunt Plateau well-developed floodplain deposits (FA-1 and ton, 1991). This thickness is maintained along Fluvial architecture in the most proximal FA-3). This architecture changes dramatically the western and southern flanks of the Paun- John Henry Member sections at Parowan Gap in the northeast margin near the Paunsaugunt saugunt Plateau (Fig. 7; Tilton, 1991). Local- and Parowan Canyon is characterized by sheet fault (Fig. 7; Heward Creek), where the John ized thickening occurs along the northeastern deposits of FA-4 intercalated with minor com- Henry Member consists primarily of FA-3 with edge of the Paunsaugunt Plateau, in both the ponents of FA-1 and some FA-3 (Fig. 7). These minor components of FA-1. Here, channel belts John Henry Member (∼160 to ∼270 m) and strata are typically composed of medium- to of FA-3 are typically ∼8 stories high and situ- the Smoky Hollow Member (∼30 m to ∼60 m; coarse-grained sandstones with pebble lags at ated within well-developed gray to red mud- Heward Creek, Fig. 7). the base of scours. Channel belts are vertically stone deposits (Fig. 5A, B). Typical bedforms amalgamated with minor floodplain develop- are difficult to identify in many of the thickest Kaiparowits Plateau ment; NTG in these proximal sections is ∼0.64 channels due to the abundance of convolute and Across the Paunsaugunt fault, strata from and ∼0.48 (Fig. 7). slumped bedding disrupting bedding structures the Smoky Hollow Member and John Henry In contrast, across the rest of the Markagunt (Fig. 5D). Overall, the average NTG of the John Member strata thicken across the Paunsaugunt Plateau, the John Henry Member consists of Henry Member in the south and southwest of fault, in the northern Kaiparowits Plateau–Table well-drained floodplain material (FA-1) with the Paunsaugunt Plateau is ∼0.64 (average Cliffs region (Fig. 7; Shakespeare Mine). Here, isolated sandstone channels of FA-3 (Fig. 7; of Tenny Canyon and Glendale), but drops to the Smoky Hollow Member is ∼70 m thick, a Orderville Gulch). Instances of FA-4 within the ∼0.32 at Heward Creek. The average paleocur- slight thickness increase relative to the west. The middle John Henry Member can be traced later- rent direction is to the northeast, with no signifi- John Henry Member, however, is ∼456 m thick, ally for 100s of meters. Sheet deposits of FA-4 cant changes in trends across the plateau (Tilton, a 69% increase in thickness compared to Heward within the John Henry Member are composed 1991; Lawton et al., 2014; Fig. 7). Creek, across the Paunsaugunt fault (Shake- of fine- to medium-grained sandstone and lack speare Mine; Fig. 7). These thickness changes pebble lags. NTG estimates of the John Henry Kaiparowits Plateau continue into the paralic and shallow-marine Member decrease from ∼0.56 along the west- At Shakespeare Mine in the northern Kai- John Henry Member strata to the east, which are ern margin of the plateau to ∼0.37 in the central parowits Plateau (Table Cliffs) region, the ∼460 m thick along the northeastern margin of plateau. Overall paleocurrent indicators are ori- lower John Henry Member is composed of the Kaiparowits Plateau in Buck Hollow (Mul- ented north/northeast (average paleoflow∼ 055; the poorly drained floodplain and coals of hern and Johnson, 2016). The John Henry Mem- n = 120; Fig. 7). FA-2 and defined channels of FA-3. Channel

12 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

­amalgamation increases both laterally and ver- As mentioned previously, intraformational recorded in extensive “sheet-sands” that punctu- tically up-section, transitioning from small, iso- unconformities and inferred sequence boundar- ate the Cretaceous stratigraphy in the area (e.g., lated channels of FA-2, to larger defined channels ies in Upper Cretaceous sections are thought to Calico Bed, Drip Tank Member, and Wahweap of FA-3, and to sheet-like sandstones of FA-4. be relatively minor in terms of duration, potential Formation; Lawton et al., 2003, 2014; Lawton Fluvial John Henry Member NTG in the northern eroded section, and angular stratigraphic rela- and Bradford, 2011; Szwarc et al., 2015; Primm Kaiparowits region is ∼0.87. Overall paleocur- tionships (Peterson, 1969; Gustason, 1989; Til- et al., 2018). These distinctive and widespread rent direction is to the east-northeast (Fig. 7). ton, 1991; Shanley and McCabe, 1995; Lawton units typically have east-directed paleocurrent et al., 2003, 2014; Szwarc et al., 2015; Primm indicators and provenance signatures that indi- DISCUSSION et al., 2018). While the sub-Claron unconformity cate a Sevier fold-thrust belt source. is regionally significant, available data suggest Thus, evidence for orogen-transverse drain- Stratigraphic data presented here show that the Kaiparowits Formation was mainly age systems appears to be localized and episodic increasing thickness and NTG to the east in deposited in an isolated depocenter located in relative to most of the Cretaceous foreland basin distal parts of the foredeep, which raises ques- the northern Kaiparowits Plateau (Fig. 9; Eaton succession in southern Utah. Previous interpre- tions regarding controls on basin architecture et al., 1993; Roberts, 2007; Sampson et al., 2013; tations of major axial fluvial systems are partly and evolution of the Late Cretaceous southern Biek et al., 2015) and that variations in thickness based on facies distributions combined with Utah foreland. This discussion focuses pri- are not entirely a function of subsequent erosion. paleocurrent analyses, which show dominantly marily on thickness and facies variations of Thus, observed stratigraphic thicknesses of Cre- northeast transport in Upper Cretaceous fluvial the fluvial John Henry Member of the Straight taceous sections in the SWHPs likely represent sections of the Kaiparowits Plateau (Szwarc Cliffs Formation, and factors controlling sedi- paleo-accommodation variations. et al., 2015; Gooley et al., 2016). Our data ment supply and accommodation within the compilation confirms overall northeast-directed foreland basin. Facies Trends paleocurrent directions for the John Henry Member across the Markagunt and Paunsaugunt Thickness Trends Using fluvial architecture and NTG as prox- Plateaus (Fig. 7). Similarly, U-Pb detrital zircon ies for grain size and facies trends across the geochronology of the Straight Cliffs Formation Regional stratigraphic correlations of Upper foreland basin (Fig. 9A), results show consistent in the Kaiparowits Plateau (Szwarc et al., 2015; Cretaceous strata across the SWHP of southern “distal fining” only from the most proximal Iron Primm et al., 2018) indicates relatively minor Utah indicate abrupt, stepwise thickness varia- Springs Formation (NTG 0.64) to the Straight detrital zircon input from the Sevier fold-thrust tions in key parts of the section that coincide Cliffs Formation in the Markagunt Plateau, belt. Rather, major zircon source areas lie to the with modern plateau-bounding faults (Figs. 8, with its well-developed floodplain deposits south (e.g., 1.4 and 1.7 Ga zircon populations 9). The most proximal sections of the Iron and isolated fluvial channel belts (NTG 0.37). sourced from the Mogollon Highlands) and Springs Formation (e.g., Parowan Canyon) are In the Paunsaugunt Plateau, distal fining trends southwest (e.g., ca. 147 Ma zircon populations at least 350 m thick, whereas the age-equivalent become more complex. Across the Sevier fault likely sourced from the Mojave Dike Swarm in Straight Cliffs Formation thins by 50% to the in the western and southern Paunsaugunt Pla- southeastern California; Fig. 1). east across the Markagunt and Paunsaugunt Pla- teau, channel belt amalgamation increases, In sum, fluvial sections in the Straight Cliffs teaus (average thickness ∼175 m). This proxi- becoming more sheet-like with less floodplain Formation show distinct facies and thickness mal eastward-thinning trend is reversed across deposition (NTG 0.64). In contrast, to the north- variations by location; these features are oriented the Paunsaugunt Plateau where John Henry east along the Paunsaugunt fault (e.g., Heward sub-parallel to the fold-thrust belt and coincide Member sections thicken from ∼270 m at Creek; Fig. 7), the John Henry Member section with modern plateau-bounding faults. The inter- Heward Creek to over ∼450 m at Shakespeare becomes thicker with much higher floodplain pretation of these basin-axial corridors suggests Mine over a distance of ∼25 km (Fig. 7). This deposition and well-developed isolated chan- some allogenic control on local accommodation thickening into the distal foredeep also occurs nels (NTG 0.32). Coincident with dramatic rather than a continuous foredeep throughout the from south to north across the Kaiparowits Pla- thickening across the Paunsaugunt fault, the Late Cretaceous. teau in the John Henry Member, as well as the John Henry Member sections in the northern overlying Wahweap and Kaiparowits Forma- Kaiparowits Plateau (Shakespeare Mine; Fig. 7) Accommodation Controls tions (Fig. 9B; Lawton et al., 2003; Gallin et al., have a notably high NTG (∼0.87) with highly 2010; Gooley et al., 2016). amalgamated fluvial channel belts and a few Flexural Loading Present-day stratigraphic thicknesses are well-developed coal zones primarily in the lower Previous studies have interpreted the Upper reported here, and thus are minimum sediment parts of the section (Fig. 4; Hettinger, 2000; Het- Jurassic–Lower Cretaceous Morrison, Carmel, accumulation estimates. The observed east- tinger et al., 2009). and Cedar Mountain Formations to record fore- ward- and northward-thickening trends in the These Late Cretaceous facies trends can be bulge migration and deposition, with the foreb- Kaiparowits Plateau (Fig. 9) would be exagger- partly explained by fluvial drainage networks ulge lying just east of the Kaiparowits Plateau ated by decompaction, given the concentration that were aligned subparallel to the trend of the by Cenomanian or Turonian time (Royse, 1993; of mudstone and coal deposits in paralic and fold-thrust belt in southern Utah, thus implying Currie, 1998, 2002; DeCelles, 2004). An impor- marine sections within the John Henry Member an axial fluvial system Fig. 10( ). Direct evidence tant implication of this interpretation is that high (Hettinger, 2000; Hettinger et al., 2009). Mul- for orogen-transverse drainage patterns is essen- Late Cretaceous sediment accumulation rates of hern and Johnson (2016) estimated the entire tially limited to the most proximal deposits of the Kaiparowits Plateau appear to have formed Straight Cliffs Formation could be expanded the Iron Springs Formation (Gunlock and Pine in the distal foredeep to forebulge depozones, from 456 m to over 800 m at Buck Hollow Valleys; Fig. 8; Fillmore, 1991; Goldstrand, 150–200 km east of the Wah Wah and Blue (Fig. 2), even with conservative decompaction 1991). In addition, episodic progradation of Mountain thrust systems. This pattern is par- estimates. transverse fluvial systems across the foreland is ticularly odd given thinning of time-equivalent

Geological Society of America Bulletin, v. 130, no. XX/XX 13

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

A

B

Figure 9. Map view showing depocenters of the fluvial John Henry Member (A) and the total thickness distribution of the Upper Cretaceous section (B). Thicknesses are averaged from measured sections shown with circles (see Fig. 2 for section references of Straight Cliffs Forma- tion. Additional references include: Eaton et al. (2001); Moore and Straub (2001); Lawton et al. (2003); Roberts (2007); Lawton et al. (2014).

strata to the west in more proximal parts of the its Plateau due to movement along the Pavant/ Santonian, simultaneous with deposition of the foredeep, where thicker sediment accumulation Paxton thrust sheets of central Utah, ∼75 km to John Henry Member (DeCelles et al., 1995; would be expected (Fig. 9). the northwest (Fig. 1B). The thrusting style of Coogan and DeCelles, 1996). During this time, Differential loading could have increased flex- the Pavant/Paxton system consisted of duplexing sediment supply into the central Utah foreland ural accommodation in the northern Kaiparow- and culmination growth during the Coniacian- basin was likely high, but provenance data and

14 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

Figure 10. Revised broken foreland basin model which includes active basement-in- volved structures partitioning the foredeep prior to Laramide deformation (schematic faults in red; MP—Markagunt Plateau; PP—Paunsaugunt Plateau; KP—Kaiparowits Plateau). The altered paleoto- pography influenced foreland basin sub-basin formation, stratal stacking patterns, and the position and architecture of the axial fluvial system (brown- white color gradient). No verti- cal scale implied.

evidence for dominant paleoflow to the east and foreland is thought to transition from a narrow place onset of Laramide-style deformation up to northeast suggest the majority of sediment was flexural profile to a longer-wavelength foreland ∼20 m.y. later than the suspected formation of being transported away from the Kaiparowits basin (Roberts and Kirschbaum, 1995). Dynamic sub-basins within the Straight Cliffs Formation Plateau (Lawton, 1983; DeCelles et al., 1995; subsidence likely contributed 10s of meters of and thickness changes in the Naturita Formation. Fielding et al., 2010). Furthermore, a 47% subsidence across the entire southern Utah fore- This timing is similar to the emergence of base- expansion of the John Henry Member occurs land basin, a wavelength of 100s of kilometers ment cored structures of the northern foreland in over just 15 km between Main Canyon and (Painter and Carrapa, 2013; Heller and Liu, southwest Montana, where thermochronologic Buck Hollow, too short of a distance to relate to 2016). This effect is on a much broader scale evidence shows exhumation of Laramide struc- regional flexural patterns Fig. 2( ; Chentnik et al., than the evidence for localized variable accom- tures occurred prior to ca. 80 Ma, and as early 2015; Mulhern and Johnson, 2016). modation presented in this study (10s–100s of m as ca. 100–120 Ma (Schwartz and DeCelles, variable thickness over <25 km distance). These 1988; Carrapa et al., 2019; Garber et al., 2020; Blind Thrust Faults abrupt thickness changes suggest that dynamic Orme, 2020). The early (Aptian-Albian) onset Given evidence for early emplacement and a subsidence alone cannot explain evidence for of Laramide-style deformation is similar to our relatively stationary fold-thrust belt in the Wah discrete sub-basins in the distal foredeep. earliest estimates of foreland partitioning from Wah and Blue Mountain thrust systems, the pos- the Cenomanian Naturita Formation (Gusta- sibility of buried Sevier-related thrust faults may Laramide-style Tectonics son, 1989), and argues against migration of the be considered to explain observed foredeep sedi- The idea of sub-basin formation within the Shatsky rise as a driver of Laramide deformation mentation patterns. No subsurface thrusts have Cordilleran foreland basin is commonly cast in (Carrapa et al., 2019; Garber et al., 2020). Fur- been reported across the SWHP during Late the context of the Laramide orogeny, or “broken thermore, classic Laramide-style “broken fore- Cretaceous time (Cook and Hardman, 1967; Van foreland” of Dickinson et al. (1988). Laramide- land” basins (e.g., the Uinta basin) are generally Kooten, 1988), nor is there outcropping evidence style, basement-involved uplifts exist on the much larger than the sub-basins now exposed of significant detachment horizons. Late-break- margins of the SWHP, including the Kanarra along the SWHP (Picard and High, 1968; Dick- ing thrust faults at Parowan Gap and Ruby’s Inn Fold to the west, and the East Kaibab monocline inson et al., 1988). likely occurred after deposition of the Straight and Circle Cliffs uplifts to the east (Fig. 1). The Cliffs Formation (Merle et al., 1993). Further- earliest evidence of this foreland break-up in Structural Inheritance and Normal Faults more, we note that studies of wedge-top basins southern Utah includes growth faults in the Cam- The location of modern plateau-bounding indicate very different stratigraphic patterns than panian Wahweap Formation in the northern Kai- faults was likely influenced by preexisting those documented here, including common pro- parowits Plateau (e.g., Hilbert-Wolf et al., 2009; crustal weaknesses. Inherited rift-related base- gressive unconformities and coarse-grained allu- Simpson et al., 2014). One of the main drivers ment faults across southern Utah (Fig. 1A) are vial deposits with local provenance (Lawton and for Laramide tectonism is flat-slab subduction, known to have been reactivated during mul- Trexler, 1991; Horton, 1998; Quinn et al., 2018). which is thought to have initiated at the latitude tiple tectonic events throughout the Phanerozoic of southern Utah during the earliest Campanian (Schwartz, 1982; Picha and Gibson, 1985). Evi- Dynamic Load (Simpson et al., 2014). Inverse mantle convec- dence for Late Cretaceous–early Cenozoic activ- Subsidence mechanisms across the southern tion models show the flat-slab segment, likely ity along the Hurricane fault (Cook and Hard- Utah foreland can be attributed to flexural load- the conjugate of the Shatsky Rise, entering man, 1967) suggests that other SWHP-bounding ing of the Sevier fold-thrust belt and dynamic the subduction zone along the Mojave area of faults may have been prone to structural inheri- processes (Jordan, 1981; Mitrovica et al., 1989; California at ca. 90 Ma, followed by northeast tance controls and reactivation as well. We sug- Pang and Nummedal, 1995; Liu and Nummedal, underthrusting beneath Arizona to Wyoming gest that reactivation of Proterozoic basement 2004; Liu and Gurnis, 2010). During the Conia- during the latest Cretaceous (Liu et al., 2010; features, possibly as normal faults, was a factor cian to Santonian (i.e., during deposition of the Liu and Currie, 2016). Although this is a plausi- in early partitioning of the Cretaceous southern John Henry Member), subsidence within the ble driving mechanism, timing estimates would Utah foreland basin.

Geological Society of America Bulletin, v. 130, no. XX/XX 15

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

Structural variations within the crust of fore- typical foredeep migration and may have favored Eaton. This research was supported by the National land basins, like those in southern Utah, have the formation of normal faults on the bending Science Foundation Graduate Research Fellowship Program, AAPG student grants, and industry support- been shown to influence crustal deflection and plate (Fig. 10), which ultimately triggered for- ers of the Rocks2Models consortium, in particular evolution of the foredeep (e.g., Romans et al., mation of isolated sub-basins within the south- ConocoPhillips. Thank you to the Bureau of Land 2010). In some cases, they may cause segmenta- ern Utah foreland (Fig. 9). Restricted foreland Management, the Grand Staircase–Escalante National tion of the foreland (Waschbusch and Royden, basin migration could have increased local stress Monument, and the Dixie National Forest for permits 1992) or rapid migration of the foredeep dep- within the plate allowing for faulting to occur to conduct research in the Kaiparowits, Paunsaugunt, and Markagunt Plateau region. We would like to ac- ocenter (DeCelles et al., 1995). For example, along preexisting heterogeneities (e.g., Bradley knowledge that although we are using state and local eastward propagation of the central Utah thrust and Kidd, 1991; Hudson, 2000). names for location markers, it is important to remem- sheets may have been aided by the Ancient Even given uncertainty regarding their orien- ber that these places are known in many ways, and this Ephraim fault and other basement structures tation and kinematics, we infer that movement research was done on the land of the Southern Paiute, Pueblo, Dine, Western Shoshone, and Ute peoples, (Fig. 1). The northeast-trending Paragonah lin- along inherited structures during Late Creta- and others that we may be unaware of. We want to eament separates the southern and central Utah ceous time influenced sediment transport and recognize that in western and especially academic and fold-thrust belt and is speculated to have offset accumulation within the foreland basin (Fig. 10). STEM spaces, indigenous people’s ability to describe the southern Utah thrust sheets, inhibiting their In some cases, changes in thickness and fluvial and name places have been diminished if not fully further propagation to the east (Picha, 1986). architecture are abrupt on either side of modern taken. Last, we are also grateful to the University of Utah Basins Research Group, Magdalena Curry, and In the Magallanes foreland basin in Argentina, plateau-bounding structures such as the Sevier an anonymous reviewer, as well as John Byrd for their pre-foreland attenuated lithosphere led to atypi- and Paunsaugunt normal faults (Eaton et al., thoughtful feedback and suggestions. cal foredeep subsidence, amplified by sediment 1993). Sub-basin formation partitioned the axial loading (Romans et al., 2010; Fosdick et al., fluvial system into distinct northeast-trending REFERENCES CITED 2014; Gianni et al., 2015). fluvial facies belts (Fig. 10) with variable NTG Ahmed, S., Bhattacharya, J.P., Garza, D.E., and Li, Y., 2014, Multiple mechanisms have been suggested for and channel belt architectures preserved. Thin- Facies architecture and stratigraphic evolution of a riv- normal faulting associated with flexure. Exten- ning of foreland basin strata across the Marka- er-dominated delta front, Turonian Ferron Sandstone, Utah, U.S.A: Journal of Sedimentary Research, v. 84, sion observed along the outer rise in subduction gunt and Paunsaugunt Plateaus is reversed in the p. 97–121, https://doi.org/10.2110/jsr.2014.6. zones may provide a simple conceptual analog most distal parts of the foredeep, particularly in Allen, J.R.L., 1978, Studies in fluviatile sedimentation: an (Ludwig et al., 1966). The Taconic peripheral the northern Kaiparowits Plateau. Similar evi- exploratory quantitative model for the architecture of avulsion-controlled alluvial suites: Sedimentary Geol- foreland basin was accompanied by motion dence for foredeep partitioning is reported as ogy, v. 21, p. 129–147, https://doi.org/10.1016/0037- along normal faults due to inelastic extensional early as Cenomanian time (Gustason, 1989; 0738(78)90002-7. deformation on the convex side of the flexed Eaton and Nations, 1991), and we interpret these Allen, J.L., and Johnson, C.L., 2010, Sedimentary facies, paleoenvironments, and relative sea level changes in plate, which suggests bending of the lithosphere signals to have intensified by Santonian time the John Henry Member, Cretaceous Straight Cliffs beyond its elastic limit (Bradley and Kidd, 1991). during deposition of the John Henry Member. Formation, southern Utah, USA, in Carney, S.M., Ta- bet, D.E., and Johnson, C.L., eds., Geology of South- A possible expression of flexural foundering has Central Utah: Utah Geological Society Publication, also been documented in the Apulian foreland CONCLUSIONS v. 39, p. 225–247. basin in Italy, where systematic joints formed Anderson, L.P., and Dinter, D.A., 2010, Deformation and sedimentation in the southern Sevier Foreland, Red along the forebulge due to flexure-related fiber New stratigraphic data from the Straight Cliffs Hills, southwestern Utah, in Carney, S.M., Tabet, D.E., stresses (Billi and Salvini, 2003). This flexural Formation of the SWHP, including sediment and Johnson, C.L., eds., Geology of South-Central foundering within the forebulge can alter typical thickness, facies architecture, and net to gross Utah: Utah Geological Association Publication 39, p. 338–366. basin depocenters by stopping forebulge migra- estimates, indicate that the Cretaceous southern Armstrong, R.L., 1968, Sevier Orogenic Belt in Nevada and tion (Waschbusch and Royden, 1992). Normal Utah foredeep was divided into discrete sub- Utah: Geological Society of America Bulletin, v. 79, https://doi.org/10.1130/0016-7606(1968)79[429:SOBI faulting due to extensional stresses within flex- basins at least 10–20 million years prior to onset NA]2.0.CO;2. ural foreland basins is likely more common than of known local Laramide-style deformation. Barclay R.S., Rioux, M., Meyer, L.B., Bowring, S.A., John- is recognized because it may present as subtle Flexural foundering at the edge of the Colorado son, K.R., Miller, I.M., 2015, High precision U–Pb zir- con geochronology for Cenomanian Dakota Formation shifts in stratigraphic architecture that cannot be Plateau may have favored sub-basin formation floras in Utah: Cretaceous Research, v. 52, part A, p. explained by other allogenic or autogenic signals within the foreland, thinning strata across the 213–237, https://doi.org/10.1016/j.cretres.2014.08.006. (Bradley and Kidd, 1991; Waschbusch and Roy- Paunsaugunt Plateau, and dramatically thicken- Benhallam, W., Turner, A., Stright, L., and Johnson, C.L., 2016, Spatial analysis of channel-belt stacking patterns: den, 1992; Londoño and Lorenzo, 2004; Gianni ing strata in the northern Kaiparowits Plateau. Metrics to discriminate between local and regional con- et al., 2015). We suggest that the Paunsaugunt, Sevier, and trols on deposition in the fluvial John Henry Member of the Straight Cliffs Formation, southern Utah, U.S.A: Hurricane faults are the modern expressions Journal of Sedimentary Research, v. 86, p. 1210–1227, Syndepositional Tectonic History—An of these inherited features, which ultimately https://doi.org/10.2110/jsr.2016.77. Integrated Model reflect lithospheric heterogeneities formed in Beveridge, T.L., Roberts, E.M., and Titus, A.L., 2020, Volca- niclastic member of the richly fossiliferous Kaiparowits the Proterozoic. These findings demonstrate Formation reveals new insights for regional correlation Traditional models of foreland basin evolu- that “broken forelands” evolve different basin and tectonics in southern Utah during the latest Campa- tion conflict with aspects of stratigraphic archi- geometries, sedimentary facies, and stratigraphic nian: Cretaceous Research, v. 114, 104527, https://doi​ .org/10.1016/j.cretres.2020.104527. tecture and depocenters across the southern stacking patterns than those predicted by tradi- Biek, R.F., Rowley, P.D., Anderson, J.J., Maldonado, F., Utah Cretaceous foreland basin. Specifically, tional models of foreland basin systems. Moore, D.W., Hacker, D.B., Eaton, J.G., Hereford, R., Sable, E.G., Filkorn, H.F., and Matyjasik, B., 2015, thickening and increases in NTG in the distal Geologic Map of the Panguitch 30′ × 60′ Quadrangle, foredeep or forebulge area (the northern Kaip- ACKNOWLEDGMENTS Garfield, Iron, and Kane Counties, Utah: Utah Geologi- arowits Plateau) is unusual for classic foreland cal Survey Map 270DM, scale 1:62,500. This work would not be possible without the tech- Biek, R.F., Rowley, P.D., and Hacker, D.B., 2019, The Gi- basin models. We propose that inherited struc- nical guidance, logistical support, and immense in- gantic Markagunt and Sevier Gravity Slides Resulting tural heterogeneities within the crust restricted sight into southern Utah stratigraphy provided by Jeff from Mid-Cenozoic Catastrophic Mega-scale Failure of

16 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

the Marysvale Volcanic Field, Utah, USA: Geological a modern avulsive floodplain: Examples from the Dumitru, T., Duddy, I., and Green, P., 1994, Mesozoic- Society of America Field Guide 56, 121 p., https://doi​ Cumberland Marshes, Saskatchewan, Canada: Journal Cenozoic burial, uplift, and erosion of the west-central .org/10.1130/2019.0056(01). of Sedimentary Research, v. 78, no. 10, p. 683–692, Colorado Plateau: Geology, v. 22, p. 499–502, https:// Billi, A., and Salvini, F., 2003, Development of systematic https://doi.org/10.2110/jsr.2008.077. doi.org/10.1130/0091-7613(1994)022<0499:MCBUA joints in response to flexure-related fibre stress in flexed Dávila, F.M., and Carter, A., 2013, Exhumation history of the E>2.3.CO;2. foreland plates: The Apulian forebulge case history, Ita- Andean broken foreland revisited: Geology, v. 41, no. 4, Dyman, T.S., Cobban, W.A., Titus, A.L., Obradovich, J.D., ly: Journal of Geodynamics, v. 36, p. 523–536, https:// p. 443–446, https://doi.org/10.1130/G33960.1. Davis, L.E., Eves, R.L., Pollock, G.L., Takahashi, K.I., doi.org/10.1016/S0264-3707(03)00086-3. Davis, G.H., 1999, Structural Geology of the Colorado Pla- and Hester, T.C., 2002, New biostratigraphic and ra- Bobb, M.C., 1991, The Calico bed, Upper Cretaceous, south- teau Region of Southern Utah, with Special Emphasis diometric ages for Albian–Turonian Dakota Formation ern Utah: A fluvial sheet deposit in the Western Interior on Deformation Bands: Geological Society of America and Tropic Shale at Grand Staircase–Escalante National foreland basin and its relationship to eustasy and tecton- Special Paper 342, 157 p., https://doi.org/10.1130/ Monument and Iron Springs Formation near Cedar ics [M.S. thesis]: University of Colorado. SPE342. City, Parowan, and Gunlock, Utah: Geological Society Bradley, D.C., and Kidd, W.S.F., 1991, Flexural exten- Davis, G.H., and Bump, A.P., 2009, Structural geologic of America Abstracts with Programs, v. 34, A13. sion of the upper continental crust in collisional evolution of the Colorado Plateau, in Kay, S.M., Ra- Eaton, J.G., 1999, Vertebrate paleontology of the Iron foredeeps: Geological Society of America Bulletin, mos, V.A., and Dickinson, W.R., eds., Backbone of Springs Formation, Upper Cretaceous, southwestern v. 103, p. 1416–1438, https://doi.org/10.1130/0016- the Americas: Shallow Subduction, Plateau Uplift, Utah, in Gillette, D.D., ed., Vertebrate Paleontology 7606(1991)103<1416:FEOTUC>2.3.CO;2. and Ridge and Terrane Collision, Geological Soci- in Utah: Utah Geological Survey Miscellaneous Pub- Bump, A.P., and Davis, G.H., 2003, Late Cretaceous–early ety of America Memoir 204, p. 99–124, https://doi​ lication 99-1, p. 339–343, https://ugspub.nr.utah.gov/ Tertiary Laramide deformation of the northern Colo- .org/10.1130/2009.1204(05). publications/misc_pubs/MP-99-1.pdf (accessed March rado Plateau, Utah and Colorado: Journal of Structural DeCelles, P.G., 1994, Late Cretaceous–Paleocene syn- 2021). Geology, v. 25, p. 421–440, https://doi.org/10.1016/ orogenic sedimentation and kinematic history of Eaton, J.G., and Nations, J.D., 1991, Introduction; Tectonic S0191-8141(02)00033-0. the Sevier thrust belt, northeast Utah and southwest setting along the margin of the Cretaceous Western In- Burbank, D.W., and Beck, R.A., 1988, Thrusting and gravel Wyoming: Geological Society of America Bulle- terior Seaway, southwestern Utah and northern Arizona, progradation in foreland basins : A test of post-thrusting tin, v. 106, p. 32–56, https://doi.org/10.1130/0016- in Eaton, J.G., and Nations, J.D., eds., Stratigraphy, gravel dispersal: Geology, v. 16, p. 1143–1146, https:// 7606(1994)106<0032:LCPSSA>2.3.CO;2. depositional environments; and sedimentary tectonics doi.org/10.1130/0091-7613(1988)016<1143:TAGPIF DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of of the western margin, Cretaceous Western Interior Sea- >2.3.CO;2. the Cordilleran thrust belt and foreland basin system, way: Geological Society of America Special Paper 260, Carrapa, B., DeCelles, P.G., and Romero, M., 2019, Early Western U.S.A: American Journal of Science, v. 304, p. p. 1–8, https://doi.org/10.1130/SPE260-p1. inception of the Laramide Orogeny in southwestern 105–168, https://doi.org/10.2475/ajs.304.2.105. Eaton, J.G., Goldstrand, P.M., and Morrow, J., 1993, Com- Montana and northern Wyoming: Implications for DeCelles, P.G., 2012, Foreland basin systems revisited: position and stratigraphic interpretation of Cretaceous models of flat-slab subduction: Journal of Geophysical Variations in response to tectonic settings, in Busby, strata of the Paunsaugunt Plateau, Utah: Museum of Research. Solid Earth, v. 124, p. 2102–2123, https://doi​ C., and Perez Azor, A., eds., Tectonics of Sedimentary Northern Arizona Bulletin, v. 59, p. 153–162. .org/10.1029/2018JB016888. Basins: Recent Advances: Blackwell Publishing Ltd, Eaton, J.G., Kirkland, J.I., Tjjert, N.E., Leckie, R.M., Brad- Catuneanu, O., 2004, Retroarc foreland systems-evo- p. 405–426, https://doi.org/10.1002/9781444347166​ ley, B., Goldstrand, P.M., Moore, D.W., Straub, A.W., lution through time: Journal of African Earth Sci- .ch20. Cobban, W.A., and Dalebout, J.D., 2001, Cretaceous ences, v. 38, p. 225–242, https://doi.org/10.1016/ DeCelles, P.G., and Cavazza, W., 1999, A compari- and Early Tertiary geology of Cedar and Parowan Can- j.jafrearsci.2004.01.004. son of fluvial megafans in the Cordilleran (Up- yons, western Markagunt Plateau, in Erskine, M.C., Catuneanu, O., Beaumont, C., and Waschbusch, P., 1997, In- per Cretaceous) and modern Himalayan foreland Faulds, J.E., Bartley, J.M., Rowley, P.D., eds., The terplay of static loads and subduction dynamics in fore- basin systems: Geological Society of America Bulle- Geologic Transition, High Plateaus to Great Basin: A land basins: Reciprocal stratigraphies and the “missing” tin, v. 111, p. 1315–1334, https://doi.org/10.1130/0016- Symposium and Field Guide: The Mackin Volume: peripheral bulge: Geology, v. 25, p. 1087–1090, https:// 7606(1999)111<1315:ACOFMI>2.3.CO;2. Utah Geological Association Publication 30 and Pacific doi.org/10.1130/0091-7613(1997)025<1087:IOSLAS DeCelles, P.G., and Coogan, J.C., 2006, Regional structure Section American Association of Petroleum Geologists >2.3.CO;2. and kinematic history of the Sevier fold-and-thrust Publication GB78, p. 337–363. Chamberlin, E.P., and Hajek, E.A., 2019, Using bar pres- belt, central Utah: Bulletin of the Geological Society Erslev, E.A., 1993, Thrusts, back-thrusts, and detachment ervation to constrain reworking in channel-dominated of America, v. 118, p. 841–864, https://doi.org/10.1130/ of Rocky Mountain foreland arches, in Schmidt, C.J., fluvial stratigraphy: Geology, v. 47, p. 531–534, https:// B25759.1. Chase; R.B., and Erslev, E.A., eds., Laramide Basement doi.org/10.1130/G46046.1. DeCelles, P.G., and Currie, B.S., 1996, Long-term sediment Deformation in the Rocky Mountain Foreland of the Chentnik, B.M., Johnson, C.L., Mulhern, J.S., and Stright, accumulation in the Middle Jurassic–early Eocene Western United States: Geological Society of America L., 2015, Valleys, , and : Paleoenvi- Cordilleran retroarc foreland-basin system: Geol- Special Paper 280, p. 339–358, https://doi.org/10.1130/ ronments and regressive-transgressive architecture ogy, v. 24, p. 591–594, https://doi.org/10.1130/0091- SPE280-p339. of the Upper Cretaceous Straight Cliffs Formation, 7613(1996)024<0591:LTSAIT>2.3.CO;2. Fielding, C.R., 1987, Coal depositional models for deltaic and Utah, U.S.A: Journal of Sedimentary Research, v. 85, DeCelles, P.G., and Giles, K.A., 1996, Foreland basin sys- alluvial plain sequences: Geology, v. 15, p. 661–664, p. 1166–1196, https://doi.org/10.2110/jsr.2015.70. tems: Basin Research, v. 8, p. 105–123, https://doi​ https://doi.org/10.1130/0091-7613(1987)15<661:CD Coogan, J.C., and DeCelles, P.G., 1996, Extensional col- .org/10.1046/j.1365-2117.1996.01491.x. MFDA>2.0.CO;2. lapse along the Sevier Desert reflection, northern DeCelles, P.G., Lawton, T.F., and Mitra, G., 1995, Thrust tim- Fielding, C.R., 2015, Anatomy of falling-stage deltas in Sevier Desert basin, western United States: Geol- ing, growth of structural culminations, and synorogenic the Turonian Ferron Sandstone of the western Henry ogy, v. 24, p. 933–936, https://doi.org/10.1130/0091- sedimentation in the type Sevier orogenic belt, western Mountains Syncline, Utah: Growth faults, slope failures 7613(1996)024<0933:ECATSD>2.3.CO;2. United States: Geology, v. 23, p. 699–702, https://doi. and mass transport complexes: Sedimentology, v. 62, p. Cook, E.F., 1957, Geology of the Pine Valley Mountains org/10.1130/0091-7613(1995)023<0699:TTGOSC>2 1–26, https://doi.org/10.1111/sed.12136. [Ph.D. thesis]: Salt Lake City, University of Utah. .3.CO;2. Fielding, C.R., Antia, J., Birgenheier, L.P., and Corbett, M.J., Cook, K.L., and Hardman, E., 1967, Regional gravity Dickinson, W.R., 2004, Evolution of the North American 2010, A Field Guide to the Cretaceous Succession of survey of the Hurricane Fault Area and Iron Springs Cordillera: Annual Review of Earth and Planetary Sci- the Western Henry Mountains Syncline, South-Central District, Utah: Geological Society of America Bulle- ences, v. 32, p. 13–45, https://doi.org/10.1146/annurev. Utah, in Carney, S.M., Tabet, D.E., and Johnson, C.L., tin, v. 78, p. 1063–1076, https://doi.org/10.1130/0016- earth.32.101802.120257. eds., Geology of South-Central Utah: Utah Geological 7606(1967)78[1063:RGSOTH]2.0.CO;2. Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Association, Publication 39, p. 276–316. Crowley, D., Reiners, W., Reuter, J.M., and Kaye, G.D., 2002, Lundin, E.R., McKittrick, M.A., and Olivares, M.D., Fillmore, R., 1991, Tectonic influence on sedimentation in Laramide exhumation of the Bighorn Mountains, Wyo- 1988, Paleogeographic and paleotectonic setting of southern Sevier foreland Iron Springs in Schmidt, C.J., ming: An apatite (U-Th)/He thermochronology study: Laramide sedimentary basins in the central Rocky Chase; R.B., and Erslev, E.A., eds., Laramide Basement Geology, v. 30, p. 27–30, https://doi.org/10.1130/0091- Mountain region: Geological Society of America Bulle- Deformation in the Rocky Mountain Foreland of the 7613(2002)030<0027:LEOTBM>2.0.CO;2. tin, v. 100, p. 1023–1039, https://doi.org/10.1130/0016- Western United States: Geological Society of America Currie, B.S., 1998, Upper Jurassic–Lower Cretaceous Mor- 7606(1988)100<1023:PAPSOL>2.3.CO;2. Special Paper 260, p. 9–26, https://doi.org/10.1130/ rison and Cedar Mountain formations, NE Utah–NW Donselaar, M.E., Cuevas Gozalo, M.C., and Moyano, S., SPE260-p9. Colorado; relationships between nonmarine deposition 2012, Avulsion processes at the terminus of low-gra- Fillmore, R.P., and Middleton, L.T., 1989, Tectonic and and early Cordilleran foreland-basin development: dient semi-arid fluvial systems: Lessons from the Rio transport controls on conglomerate composition, Upper Journal of Sedimentary Research, v. 68, p. 632–652, Colorado, Altiplano endorheic basin, Bolivia: Sedimen- Cretaceous of Southwest Utah, in Colburn, I.P., Abbott, https://doi.org/10.2110/jsr.68.632. tary Geology, v. 55, p. 1–14, https://doi.org/10.1016/ P.L., and Minch, J., eds., Conglomerates on Basin Anal- Currie, B.S., 2002, Structural configuration of the Early Cre- j.sedgeo.2012.10.007. ysis: A Symposium Dedicated to A.O. Woodford: Pa- taceous Cordilleran foreland basin system and Sevier Dooling, P.R., 2013, Tidal Facies, Stratigraphic Architec- cific Section Society of Economic Paleontologists and thrust belt, Utah and Colorado: The Journal of Geology, ture, and Along-Strike Variability of a High Energy, Mineralologists, A.O. Woodford Volume, p. 113–122. v. 110, p. 697–718, https://doi.org/10.1086/342626. Transgressive Shoreline, Late Cretaceous, Kaiparowits Ford, G.L., and Pyles, D.R., 2014, A hierarchical ap- Davies-Vollum, K.S., and Smith, N.D., 2008, Factors af- Plateau, Southern Utah [M.S. thesis]: Salt Lake City, proach for evaluating fluvial systems: Architectural fecting the accumulation of organic-rich deposits in University of Utah, 135 p. analysis and sequential evolution of the high net-sand

Geological Society of America Bulletin, v. 130, no. XX/XX 17

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

­content, middle Wasatch Formation, Uinta Basin, Heller, P.L., and Liu, L., 2016, Dynamic topography and ver- relation, provenance, and biostratigraphy: Cretaceous Utah: The American Association of Petroleum Geolo- tical motion of the U.S. Rocky Mountain region prior Research, v. 30, p. 287–299, https://doi.org/10.1016/ gists Bulletin, v. 98, no. 7, p. 1273–1304, https://doi​ to and during the Laramide orogeny: Geological Soci- j.cretres.2008.07.012. .org/10.1306/12171313052. ety of America Bulletin, v. 128, p. 973–988, https://doi​ Jordan, T.E., 1981, Thrust loads and foreland basin evolution, Fosdick, J.C., Graham, S.A., and Hilley, G.E., 2014, Influ- .org/10.1130/B31431.1. Cretaceous, western United States: AAPG Bulletin, v. ence of attenuated lithosphere and sediment loading on Hettinger, R.D., 2000, A Summary of Coal Distribution and 65, p. 2506–2520, https://doi.org/10.1306/03B599F4- flexure of the deep-water Magallanes retroarc foreland Geology in the Kaiparowits Plateau, Utah: Geologic as- 16D1-11D7-8645000102C1865D. basin, Southern Andes: Tectonics, v. 33, p. 2505–2525, sessment of coal in the Colorado Plateau: National Coal Kjemperud, A.V., Schomacker, E.R., and Cross, T.A., https://doi.org/10.1002/2014TC003684. Resource Assessment in the Colorado Plateau: Arizona, 2008, Architecture and stratigraphy of alluvial de- Friedrich, A.M., and Bartley, J.M., 2003, Three-dimensional Colorado, New Mexico, and Utah: U.S. Geological Sur- posits, Morrison Formation (Upper Jurassic), Utah: structural reconstruction of a thrust system overprint- vey Professional Paper 1625-B, Chapter J, p. 1–64. AAPG Bulletin, v. 92, p. 1055–1076, https://doi​ ed by postorogenic extension, Wah Wah thrust zone, Hettinger, R.D., McCabe, P.J., and Shanley, K.W., 1993, .org/10.1306/03250807115. southwestern Utah: Geological Society of America Detailed facies anatomy of transgressive and highstand Koch, A.R., Johnson, C.L., and Stright, L., 2019, Does fluvial Bulletin, v. 115, p. 1473–1491, https://doi.org/10.1130/ systems tracts from the Upper Cretaceous of southern channel-belt clustering predict net sand to gross rock B25148.1. Utah, U.S.A., in Weimer, P., and Posamentier, H.W., volume? Architectural metrics and point-pattern analy- Friend, P.F., Slater, M.J., and Williams, R.C., 1979, Vertical eds., Siliciclastic Sequence Stratigraphy: American sis of a digital outcrop model: Journal of Sedimentary and lateral building of river sandstone bodies, Ebro Ba- Association of Petroleum Geologists Memoir 58, Research, v. 89, p. 1109–1126, https://doi.org/10.2110/ sin, Spain: Journal of the Geological Society, v. 136, p. p. 235–257. jsr.2019.60. 39–46, https://doi.org/10.1144/gsjgs.136.1.0039. Hettinger, R.D., Roberts, L.N.R., Biewick, L.R.H., and Kraus, M.J., 2002, Basin-scale changes in floodplain paleo- Gallin, W., Johnson, C., and Allen, J., 2010, Fluvial and Mar- Kirschbaum, M.A., 1996, Preliminary investigations of sols: Implications for interpreting alluvial architecture: ginal Marine Architecture of the John Henry Member, the distribution and resources of coal in the Kaiparowits Journal of Sedimentary Research, v. 72, p. 500–509, Straight Cliffs Formation, Kelly Grade of the Kaip- Plateau, southern Utah: U.S. Geological Survey Open- https://doi.org/10.1306/121701720500. arowits Plateau, South-Central Utah in Carney, S.M., File Report 96-539, p. 1–64. Kraus, M.J., and Wells, T.M., 1999, Facies and facies archi- Tabet, D.E., and Johnson, C.L., eds., Geology of South- Hettinger, R.D., Roberts, L.N.R., Biewick, L.R.H., and tecture of Paleocene floodplain deposits, Fort Union Central Utah: Utah Geological Association Guidebook, Kirschbaum, M.A., 2009, Geologic Overview and Re- Formation, Bighorn Basin, Wyoming: The Mountain v. 39, p. 248–275. source Assessment of Coal in the Kaiparowits Plateau, Geologist, v. 36, p. 57–70. Garber, K.L., Finzel, E.S., and Pearson, D.M., 2020, Prov- Southern Utah: National Coal Resource Assessment in Laurin, J., and Sageman, B.B., 2001, Tectono-sedimentary enance of Synorogenic Foreland Basin Strata in the Colorado Plateau: Arizona, Colorado, New Mexico, evolution of the western margin of the Colorado Plateau Southwestern Montana Requires Revision of Exist- and Utah: U.S. Geological Survey Professional Paper during the latest Cenomanian and Early Turonian, in ing Models for Laramide Tectonism: North Ameri- 1625-B, Chapter T, p. 1–64. Erskine, M.C., Faulds, J.E., Bartley, J.M., Rowley, P.D., can Cordillera: Tectonics, v. 39, p. 1–26, https://doi​ Hilbert-Wolf, H.L., Simpson, E.L., Simpson, W.S., Tindall, eds., The Geologic Transition, High Plateaus to Great .org/10.1029/2019TC005944. E., and Wizevich, M.C., 2009, Insights into syndeposi- Basin: A Symposium and Field Guide: The Mackin Garcia-Castellanos, D., 2002, Interplay between lithospheric tional fault movement in a foreland basin; trends in seis- Volume: Utah Geological Association Publication 30, flexure and river transport in foreland basins: Basin Re- mites of the Upper Cretaceous, Wahweap Formation, p. 57–74. search, v. 14, p. 89–104, https://doi.org/10.1046/j.1365- Kaiparowits Basin: Basin Research, v. 21, p. 856–871, Lawton, T.F., 1983, Late Cretaceous fluvial systems and the 2117.2002.00174.x. https://doi.org/10.1111/j.1365-2117.2009.00398.x. age of foreland uplifts in central Utah, in Lowell, J.D., Gianni, G., Navarrete, C., Orts, D., Tobal, J., Folguera, A., Hintze, L.F., 1986, Stratigraphy and Structure of the Beaver ed., Rocky Mountain Foreland Basins and Uplifts: Den- and Giménez, M., 2015, Patagonian broken foreland Dam Mountains, Southwestern Utah: Thrusting and Ex- ver, Colorado, Rocky Mountain Association of Geolo- and related synorogenic rifting: The origin of the tensional Structures and Mineralization in the Beaver gists, p. 181–199. Chubut Group Basin: Tectonophysics, v. 649, p. 81–99, Dam Mountains, Southwestern Utah: Utah Geological Lawton, T.F., and Bradford, B.A., 2011, Correlation and https://doi.org/10.1016/j.tecto.2015.03.006. Association Publication, v. 15, p. 1–36. Provenance of Upper Cretaceous (Campanian) Fluvial Gibling, M.R., 2006, Width and thickness of fluvial chan- Holbrook, J., 2001, Origin, genetic interrelationships, and Strata, Utah, U.S.A., from Zircon U-Pb Geochronol- nel bodies and valley fills in the geological record: A stratigraphy over the continuum of fluvial channel- ogy and Petrography: Journal of Sedimentary Research, literature compilation and classification: Journal of form bounding surfaces: an illustration from middle v. 81, p. 495–512, https://doi.org/10.2110/jsr.2011.45. Sedimentary Research, v. 76, p. 731–770, https://doi​ Cretaceous strata, southeastern Colorado: Sedimentary Lawton, T.F., and Trexler, J.H., 1991, Piggyback basin in the .org/10.2110/jsr.2006.060. Geology, v. 144, p. 179–222, https://doi.org/10.1016/ Sevier orogenic belt, Utah: implications for development Goldstrand, P.M., 1992, Evolution of Late Cretaceous and S0037-0738(01)00118-X. of the thrust wedge: Geology, v. 19, p. 827–830, https:// Early Tertiary basins of southwest Utah based on clas- Horton, B.K., 1998, Sediment accumulation on top of the doi.org/10.1130/0091-7613(1991)019<0827:PBITSO tic petrology: Journal of Sedimentary Research, v. 62, Andean orogenic wedge: Oligocene to late Mio- >2.3.CO;2. p. 495–507, https://doi.org/10.1306/D4267933-2B26- cene basins of the Eastern Cordillera, southern Bo- Lawton, T.F., Boyer, S.E., and Schmitt, J.G., 1994, Influ- 11D7-8648000102C1865D. livia: Geological Society of America Bulletin, v. ence of inherited taper on structural variability and Goldstrand, P.M., 1994, Tectonic development of Upper Creta- 110, p. 1174–1192, https://doi.org/10.1130/0016- conglomerate distribution, Cordilleran fold and thrust ceous to Eocene strata of southwestern Utah: Geological 7606(1998)110<1174:SAOTOT>2.3.CO;2. belt, western United States: Geology, v. 22, p. 339–342, Society of America Bulletin, v. 106, p. 145–154, https:// Horton, B.K., Constenius, K.N., and DeCelles, P.G., 2004, https://doi.org/10.1130/0091-7613(1994)022<0339:IO doi.org/10.1130/0016-7606(1994)106<0145:TDOUC Tectonic control on coarse-grained foreland-basin se- ITOS>2.3.CO;2. T>2.3.CO;2. quences: An example from the Cordilleran foreland Lawton, T.F., Pollock, S.L., and Robinson, R.J., 2003, Inte- Goldstrand, P.M., Trexler, J.H., Kowallis, B.J., and Eaton, basin, Utah: Geology, v. 32, p. 637–640, https://doi​ grating sandstone petrology and nonmarine sequence J.G., 1993, Late Cretaceous to Early Tertiary tec- .org/10.1130/G20407.1. stratigraphy: Application to the Late Cretaceous flu- tonostratigraphy of southwestern Utah: Aspects of Houston, W.S., Huntoon, J.E., and Kamola, D.L., 2000, Mod- vial systems of southwestern Utah, U.S.A: Journal of Mesozoic geology and paleontology of the Colorado eling of cretaceous foreland-basin parasequences, Utah, Sedimentary Research, v. 73, p. 389–406, https://doi​ Plateau: Museum of Northern Arizona Bulletin, v. 59, with implications for timing of Sevier thrusting: Geol- .org/10.1306/100702730389. p. 181–188. ogy, v. 28, p. 267–270, https://doi.org/10.1130/0091- Lawton, T.F., Schellenbach, W.L., and Nugent, A.E., 2014, Gooley, J.T., Johnson, C.L., and Pettinga, L., 2016, Spatial 7613(2000)28<267:MOCFPU>2.0.CO;2. Megafan and axial-river systems in the southern Cor- and temporal variation of fluvial architecture within a Hudson, M.R., 2000, Coordinated strike-slip and nor- dilleran foreland basin: Drip tank member of Straight prograding clastic wedge of the Late Cretaceous West- mal faulting in the southern Ozark dome of Cliffs Formation and adjacent strata, Southern Utah, ern Interior Basin (Kaiparowits Plateau), U.S.A: Jour- northern Arkansas: Deformation in a late Paleo- USA: Journal of Sedimentary Research, v. 84, p. 407– nal of Sedimentary Research, v. 86, p. 125–147, https:// zoic foreland: Geology, v. 28, p. 511–514, https://doi​ 434, https://doi.org/10.2110/jsr.2014.33. doi.org/10.2110/jsr.2016.11. .org/10.1130/0091-7613(2000)28<511:CSANFI>2.0 Little, W.W., 1997, Tectonic and eustatic controls on cyclical Gustason, E.R., 1989, Stratigraphy and Sedimentology of the .CO;2. fluvial patterns, Upper Cretaceous strata of the Kaip- Middle Cretaceous (Albian-Cenomanian) Dakota For- Huntoon, P.W., 1993, Influence of inherited Precambrian arowits Basin, Utah, in Hill, L.M., ed., Learning from mation, Southwestern Utah [Ph.D. thesis]: University basement structure on the localization and form of the Land: Grand Staircase–Escalante National Monu- of Colorado, Boulder. Laramide monoclines, Grand Canyon, Arizona, in ment Science Symposium Proceedings: Bureau of Land Hartley, A.J., Weissmann, G.S., Nichols, G.J., and Warwick, Schmidt, C.J., et al., eds., Laramide Basement Defor- Management, Salt Lake City, p. 489–504. G.L., 2010, Large distributive fluvial systems: Char- mation in the Rocky Mountain Foreland of the Western Liu, S., and Currie, C.A., 2016, Farallon plate dynamics prior acteristics, distribution, and controls on development: United States: Geological Society of America Spe- to the Laramide orogeny: Numerical models of flat sub- Journal of Sedimentary Research, v. 80, p. 167–183, cial Paper 280, p. 243–256, https://doi.org/10.1130/ duction: Tectonophysics, v. 666, p. 33–47, https://doi​ https://doi.org/10.2110/jsr.2010.016. SPE280-p243. .org/10.1016/j.tecto.2015.10.010. Hassanpour, M.M., Pyrcz, M.J., and Deutsch, C.V., 2013, Jinnah, Z.A., Roberts, E.M., Deino, A.L., Larsen, J.S., Link, Liu, L., and Gurnis, M., 2010, Dynamic subsidence and uplift Improved geostatistical models of inclined hetero- P.K., and Fanning, C.M., 2009, New 40Ar-39Ar and of the Colorado Plateau: Geology, v. 38, p. 663–666, lithic strata for McMurray Formation, Alberta, Can- detrital zircon U-Pb ages for the Upper Cretaceous https://doi.org/10.1130/G30624.1. ada: AAPG Bulletin, v. 97, p. 1209–1224, https://doi​ Wahweap and Kaiparowits formations on the Kaip- Liu, L., Gurnis, M., Seton, M., Saleeby, J., Müller, R.D., .org/10.1306/01021312054. arowits Plateau, Utah: Implications for regional cor- and Jackson, J.M., 2010, The role of oceanic plateau

18 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 Faulty foundations: Early breakup of the southern Utah Cordilleran foreland

­subduction in the Laramide orogeny: Nature Geosci- Mohrig, D., Heller, P.L., Paola, C., and Lyons, W.J., 2000, Foreland in Utah, in Peterson, J.A., ed., Paleotecton- ence, v. 3, p. 353–357, https://doi.org/10.1038/ngeo829. Interpreting avulsion process from ancient alluvial ics and Sedimentation in the Rocky Mountain Region, Liu, S., and Nummedal, D., 2004, Late Cretaceous sub- sequences: Guadalope-Matarranya system (north- United States: Tulsa, Oklahoma, American Associa- sidence in Wyoming: Quantifying the dynamic com- ern Spain) and Wasatch Formation (western Colo- tion of Petroleum Geologists Memoir 41, p. 309–320, ponent: Geology, v. 32, no. 5, p. 397–400, https://doi​ rado): Geological Society of America Bulletin, v. https://doi.org/10.1306/M41456C14. .org/10.1130/G20318.1. 112, p. 1787–1803, https://doi.org/10.1130/0016- Picha, F., and Gibson, R.I., 1985, Cordilleran hingeline: Late Londoño, J., and Lorenzo, J.M., 2004, Geodynamics of conti- 7606(2000)112<1787:IAPFAA>2.0.CO;2. Precambrian rifted margin of the North American cra- nental plate collision during late tertiary foreland basin Moore, D.W., and Straub, A.W., 2001, Correlation of Upper ton and its impact on the depositional and structural evolution in the Timor Sea: constraints from foreland Cretaceous and Paleogene(?) Rocks Beneath the Claron history, Utah and Nevada: Geology, v. 13, p. 465–468, sequences, elastic flexure and normal faulting: Tecto- Formation, Crow Creek, Western Markagunt Plateau: https://doi.org/10.1130/0091-7613(1985)13<465:CHL nophysics, v. 392, p. 37–54, https://doi.org/10.1016/ Southwest Utah in Erskine, M.C., Faulds, J.E., Bart- PRM>2.0.CO;2. j.tecto.2004.04.007. ley, J.M., Rowley, P.D., eds., The Geologic Transition, Primm, J.W., Johnson, C.L., and Stearns, M., 2018, Basin- Ludwig, W.J., Ewing, J.I., Ewing, M., Murauchi, S., Den, High Plateaus to Great Basin: A Symposium and Field axial progradation of a sediment supply driven dis- N., Asano, S., Hotta, H., Hayakawa, M., Asanuma, T., Guide: The Mackin Volume: Utah Geological Associa- tributive fluvial system in the Late Cretaceous southern Ichikawa, K., and Noguchi, I., 1966, Sediments and tion Publication 30, p. 75–95. Utah foreland: Basin Research, v. 30, no. 2, p. 249–278, structure of the Japan Trench: Journal of Geophysical Moran, K., Hilbert-Wolf, H.L., Golder, K., Malenda, H.F., https://doi.org/10.1111/bre.12252. Research, v. 71, p. 2121–2137, https://doi.org/10.1029/ Smith, C.J., Storm, L.P., Simpson, E.L., Wizevich, M.C., Quinn, G.M., Hubbard, S.M., Putnam, P.E., Matthews, W.A., JZ071i008p02121. and Tindall, S.E., 2010, Attributes of the wood-boring Daniels, B.G., and Guest, B., 2018, A Late Jurassic to Lund, W.R., Knudsen, T.R., and Vice, G.S., 2008, Paleoseis- trace fossil Asthenopodichnium in the Late Cretaceous Early Cretaceous record of orogenic wedge evolution mic reconnaissance of the Sevier fault: Kane and Gar- Wahweap Formation, Utah, USA: Palaeogeography, in the Western Interior basin, USA and Canada: Geo- field Counties: Utah Geological Survey Special Study, Palaeoclimatology, Palaeoecology, v. 297, p. 662–669, sphere, v. 14, p. 1187–1206, https://doi.org/10.1130/ v. 122, p. 31. https://doi.org/10.1016/j.palaeo.2010.09.015. GES01606.1. Maldonado, F., Sable, E.G., and Nealey, D.L., 1997, Ce- Mulhern, J.S., and Johnson, C.L., 2016, Time–space vari- Roberts, E.M., 2007, Facies architecture and deposi- nozoic low-angle faults, thrust faults, and anastamos- ability of paralic strata deposited in a high accommo- tional environments of the Upper Cretaceous Kai- ing high-angle faults, western Markagunt Plateau dation, high sediment supply setting: example from the parowits Formation, southern Utah: Sedimentary southwestern Utah: Geologic studies in the Basin Cretaceous of Utah: Sedimentology of Paralic Reser- Geology, v. 197, p. 207–233, https://doi.org/10.1016/ and Range-Colorado Plateau transition zone in south- voirs: Recent Advances, Geological Society of Lon- j.sedgeo.2006.10.001. eastern Nevada, southwestern Utah and northwestern don, Special Publication 444, p. 349–392, https://doi​ Roberts, L.N., and Kirschbaum, M.A., 1995, Paleogeogra- Arizona, 1995: U.S. Geological Survey Bulletin 2153, .org/10.1144/SP444.7. phy of the Late Cretaceous of the Western Interior of p. 151–176. Neely, T.G., and Erslev, E.A., 2009, The interplay of fold Middle North America—Coal Distribution and Sedi- Marshak, S., Karlstom, K., and Timmons, J.M., 2000, mechanisms and basement weaknesses at the transi- ment Accumulation: U.S. Geological Survey Profes- Inversion of Proterozoic extensional faults: An ex- tion between Laramide basement-involved arches, sional Paper 1561, p. 1–120, https://doi.org/10.3133/ planation for the pattern of Laramide and Ancestral north-central Wyoming, USA: Journal of Structural pp1561. Rockies intracratonic deformation, United States: Ge- Geology, v. 31, p. 1012–1027, https://doi.org/10.1016/ Romans, B.W., Fildani, A., Graham, S.A., Hubbard, S.M., ology, v. 28, p. 735–738, https://doi.org/10.1130/0091- j.jsg.2009.03.008. and Covault, J.A., 2010, Importance of predecessor 7613(2000)28<735:IOPEFA>2.0.CO;2. Orme, D.A., 2020, New timing constraints for the onset of basin history on sedimentary fill of a retroarc fore- McCabe, P.J., and Shanley, K.W., 1992, Organic con- Laramide deformation in southwest Montana challenge land basin: provenance analysis of the Cretaceous trol on shoreface stacking patterns: bogged down our understanding of the development of a thick-skinned Magallanes basin, Chile (50–52°S): Basin Research, in the mire: Geology, v. 20, p. 741–744, https://doi​ structural style during flat-slab subduction: Tectonics, v. 22, p. 640–658, https://doi.org/10.1111/j.1365- .org/10.1130/0091-7613(1992)020<0741:OCOSSP> v. 39, p. 1–8, https://doi.org/10.1029/2020TC006193. 2117.2009.00443.x. 2.3.CO;2. Owen, A., Ebinghaus, A., Hartley, A.J., Santos, M.G.M., and Royse, F., 1993, Case of the phantom foredeep: McGookey, D.P., Haun, J.P., Heele, L.A., Goodell, H.G., Weissmann, G.S., 2017, Multi-scale classification of Early Cretaceous in west-central Utah: Geol- McCubbin, D.G., Wiemer, R.J., and Wulf, G.R., 1972, fluvial architecture: An example from the Palaeocene– ogy, v. 21, p. 133–136, https://doi.org/10.1130/0091- Cretaceous system, in Mallory, W.W., ed., Geologic Eocene Bighorn Basin, Wyoming: Sedimentology, v. 7613(1993)021<0133:COTPFE>2.3.CO;2. Map of the Rocky Mountain Region: Rocky Mountain 64, p. 1572–1596, https://doi.org/10.1111/sed.12364. Saleeby, J., 2003, Segmentation of the Laramide Slab— Association of Geologists, p. 190–233. Painter, C.S., and Carrapa, B., 2013, Flexural versus dynamic Evidence from the southern Sierra Nevada re- Merle, O.R., Davis, G.H., Nickelsen, R.P., and Gour- processes of subsidence in the North American Cordil- gion: Geological Society of America Bulletin, v. lay, P.A., 1993, Relation of thin-skinned thrusting lera foreland basin: Geophysical Research Letters, v. 115, p. 655–668, https://doi.org/10.1130/0016- of Colorado Plateau strata in southwestern Utah 40, p. 4249–4253, https://doi.org/10.1002/grl.50831. 7606(2003)115<0655:SOTLSF>2.0.CO;2. to Cenozoic magmatism: Geological Society of Pang, M., and Nummedal, D., 1995, Flexural subsidence Sampson, S.D., Loewen, M.A., Roberts, E.M., and Getty, America Bulletin, v. 105, p. 387–398, https://doi​ and basement tectonics of the Cretaceous Western In- M.A., 2013, A new macrovertebrate assemblage from .org/10.1130/0016-7606(1993)105<0387:ROTSTO> terior Basin, United States: Geology, v. 23, p. 173–176, the Late Cretaceous (Campanian) of southern Utah, in 2.3.CO;2. https://doi.org/10.1130/0091-7613(1995)023<0173:FS Titus, A.L., and Loewen, M.A., eds., At the Top of the Miall, A.D., 1996, The Geology of Fluvial Deposits: Sedi- ABTO>2.3.CO;2. Grand Staircase—The Late Cretaceous of Southern mentary Facies, Basin Analysis, and Petroleum Geol- Perez-Arlucea, M., and Smith, N.D., 1999, Depositional pat- Utah: Bloomington and Indianapolis, Indiana Univer- ogy: Berlin, Springer. terns following the 1870s avulsion of the Saskatchewan sity Press, p. 599–620. Miall, A.D., and Arush, M., 2001, the Castlegate Sand- River (Cumberland Marshes, Saskatchewan, Canada): Schmidt, C.J., Genovese, P.W., and Chase, R.B., 1993, Role stone of the Book Cliffs, Utah: Sequence stratigraphy, Journal of Sedimentary Research, v. 69, p. 62–73, of basement fabric and cover-rock lithology on the ge- paleogeography, and tectonic controls: Journal of https://doi.org/10.2110/jsr.69.62. ometry and kinematics of twelve folds in the Rocky Sedimentary Research, v. 71, p. 537–548, https://doi​ Peterson, F., 1969, Cretaceous sedimentation and tectonism Mountain foreland:, in Schmidt, C.J., Chase, R.B., and .org/10.1306/103000710537. in the southern Kaiparowits Region, Utah: U.S. Geo- Erslev, E.A., eds., Laramide Basement Deformation in Miller, G.M., 1963, Outline of structural-stratigraphic units logical Survey Open-File Report 69-202, 272 p., https:// the Rocky Mountain Foreland of the Western United of the Wah Wah Mountains, southwest Utah, in Hey- doi.org/10.3133/ofr69202. States: Geological Society of America Special Paper lmun, E.B., ed., Guidebook to the Geology of South- Peterson, F., and Waldrop, H.A., 1965, Jurassic and Creta- 280, p. 1–44, https://doi.org/10.1130/SPE280-p1. western Utah: Transition between Basin-Range and ceous stratigraphy of south-central Kaiparowits Pla- Schwans, P., 1995, Controls on sequence stacking and fluvial Colorado Plateau Provinces: Intermountain Associa- teau, Utah: Guidebook to the Geology of Utah, v. 19, to shallow-marine architecture in a foreland basin, in tion of Petroleum Geologists, 12th Annual Field Con- p. 47–69. Van Wagoner, J.C., and Bertram, G.T., eds., Sequence ference, p. 96–102. Pettinga, L.A., 2013, Tectonic controls on alluvial archi- Stratigraphy of Foreland Basin Deposits—Outcrop and Miller, G.M., 1966, Structure and stratigraphy of tecture in the Upper Cretaceous John Henry Member, Subsurface Examples from the Cretaceous of North southern part of Wah Wah Mountains, south- Straight Cliffs Formation, Southern Utah [Ph.D. thesis]: America: American Association of Petroleum Geolo- west Utah: AAPG Bulletin, v. 50, no. 5, Salt Lake City, University of Utah, 278 p. gists Memoir 64, p. 55–102. p. 858–900, https://doi.org/10.1306/5D25B601-16C1- Peyton, S.L., Reiners, P.W., Carrapa, B., and DeCelles, Schwartz, R.K., 1982, Broken Early Cretaceous foreland ba- 11D7-8645000102C1865D. P.G., 2012, Low-temperature thermochronology of the sin in southwestern Montana: sedimentation related to Milner, A.R.C., Vice, G.S., Harris, J.D., and Lockley, M.G., northern Rocky Mountains, Western U.S.A: American tectonism: Geologic studies of the Cordilleran Thrust 2006, Dinosaur tracks from the Upper Cretaceous Iron Journal of Science, v. 312, no. 2, p. 145–212, https:// Belt: Rocky Mountain Association of Geologists, v. 1, Springs Formation, Iron County, Utah: Bulletin of the doi.org/10.2475/02.2012.04. p. 159–183. New Mexico Museum of Natural History and Science, Picard, M.D., and High, L.R., 1968, Sedimentary cycles Schwartz, R.K., and DeCelles, P.G., 1988, Cordilleran fore- v. 35, p. 105–113. in the Green River Formation (Eocene), Uinta Ba- land basin evolution in response to interactive Creta- Mitrovica, J.X., Beaumont, C., and Jarvis, G.T., 1989, Tilt- sin, Utah: Journal of Sedimentary Petrology, v. 38, p. ceous thrusting and foreland partitioning, southwestern ing of continental interiors by the dynamical effects 378–383. Montana, in Schmidt, C.J., and Perry, W.J., Jr., eds., of subduction: Tectonics, v. 8, p. 1079, https://doi​ Picha, F., 1986, The Influence of Preexisting Tectonic Trends Interaction of the Rocky Mountain Foreland and the .org/10.1029/TC008i005p01079. On Geometries of the Sevier Orogenic Belt and Its Cordilleran Thrust Belt: Geological Society of ­America

Geological Society of America Bulletin, v. 130, no. XX/XX 19

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021 St. Pierre and Johnson

Memoir 171, p. 489–513, https://doi.org/10.1130/ extension along the Sevier Desert Detachment, west Van Cappelle, M., Hampson, G.J., and Johnson, H.D., 2018, MEM171-p489. central Utah: Tectonics, v. 20, p. 289–307, https://doi​ Spatial and temporal evolution of coastal depositional Shanley, K.W., and McCabe, P.J., 1991, Predicting fa- .org/10.1029/2000TC001237. systems and regional depositional process regimes: cies architecture through sequence stratigraphy—An Stone, D.S., 1993, Basement-involved thrust generated folds Campanian Western Interior Seaway, USA: Journal of example from the Kaiparowits Plateau, Utah: Geol- as seismically imaged in the subsurface of the central Sedimentary Research, v. 88, no. 8, p. 873–897, https:// ogy, v. 19, p. 742–745, https://doi.org/10.1130/0091- Rocky Mountain foreland: Laramide basement defor- doi.org/10.2110/jsr.2018.49. 7613(1991)019<0742:PFATSS>2.3.CO;2. mation in the Rocky Mountain foreland of the western Van Kooten, G.K., 1988, Structure and hydrocarbon po- Shanley, K.W., and McCabe, P.J., 1993, Alluvial architecture United States: in Schmidt, C.J., et al., eds., Geologi- tential beneath the Iron Springs laccolith, Southwest- in a sequence stratigraphic framework: A case history cal Society of America Special Paper 280, p. 271–318, ern Utah: Geological Society of America Bulletin, from the Upper Cretaceous of southern Utah, USA, https://doi.org/10.1130/SPE280-p271. v. 100, p. 1533–1540, https://doi.org/10.1130/0016- in Bryant, I.D., and Flint, S.S., eds., The Geological Straub, K.M., and Esposito, C.R., 2013, Influence of water 7606(1988)100<1533:SAHPBT>2.3.CO;2. Modelling of Hydrocarbon Reservoirs and Outcrop and sediment supply on the stratigraphic record of al- Waschbusch, P.J., and Royden, L.H., 1992, Spatial and tem- Analogues: International Association of Sedimentolo- luvial fans and deltas: Process controls on stratigraphic poral evolution of foredeep basins : lateral strength gists Special Publication 15, p. 21–56. completeness: Journal of Geophysical Research. Earth variations and inelastic yielding in continental litho- Shanley, K.W., and McCabe, P.J., 1995, Sequence stratigra- Surface, v. 118, p. 625–637, https://doi.org/10.1002/ sphere: Basin Research, v. 4, p. 179–196, https://doi​ phy of Turonian–Santonian strata, Kaiparowits Plateau, jgrf.20061. .org/10.1111/j.1365-2117.1992.tb00044.x. southern Utah, U.S.A.: Implications for regional cor- Strecker, M.R., Hilley, G.E., Bookhagen, B., and Sobel, E.R., Weil, A.B., and Yonkee, W.A., 2012, Layer-parallel short- relation and foreland basin evolution, in Van Wagoner, 2011, Structural, geomorphic, and depositional char- ening across the Sevier fold-thrust belt and Laramide J.C., and Bertram, G.T., eds., Sequence Stratigraphy of acteristics of contiguous and broken foreland basins: foreland of Wyoming: Spatial and temporal evolution Foreland Basin Deposits—Outcrop and Subsurface Ex- examples from the eastern flanks of the central Andes of a complex geodynamic system: Earth and Planetary amples from the Cretaceous of North America: Ameri- in Bolivia and NW Argentina, in Busby, C., and Azor, Science Letters, v. 357–358, p. 405–420, https://doi​ can Association of Petroleum Geologists Memoir 64, A., eds., Tectonics of Sedimentary Basins: Recent Ad- .org/10.1016/j.epsl.2012.09.021. p. 103–136. vances: Blackwell Publishing Ltd., p. 508–521, https:// Weissmann, G.S., Hartley, A.J., Nichols, G.J., Scuderi, L.A., Shanley, K.W., McCabe, P.J., and Hettinger, R.D., 1992, doi.org/10.1002/9781444347166.ch25. Olson, M., Buehler, H., and Banteah, R., 2010, Flu- Tidal influence in Cretaceous fluvial strata from Utah, Swift, D.J.P., Hudelson, P.M., Brenner, R.L., and Thomp- vial form in modern continental sedimentary basins: USA: A key to sequence stratigraphic interpreta- son, P., 1987, Shelf construction in a foreland basin: Distributive fluvial systems: Geology, v. 38, no. 1, p. tion: Sedimentology, v. 39, p. 905–930, https://doi​ storm beds, shelf sandbodies and shelf–slope deposi- 39–42, https://doi.org/10.1130/G30242.1. .org/10.1111/j.1365-3091.1992.tb02159.x. tional sequences in the Upper Cretaceous Mesaverde Weissmann, G.S., Hartley, A.J., Nichols, G.J., Sciences, E., Shaw, C.A., and Karlstrom, K.E., 1999, The Yavapai-Mazat- Group, Book Cliffs, Utah: Sedimentology, v. 34, p. Holloway, R., Scuderi, L.A., Olson, M.E., Buehler, zal crustal boundary in the Southern Rocky Moun- 423–457, https://doi.org/10.1111/j.1365-3091.1987​ H.A., and Massengill, L.C., 2011, Alluvial facies distri- tains: Rocky Mountain Geology, v. 34, no. 1, p. 37–52, .tb00578.x. butions in continental sedimentary basins—Distributive https://doi.org/10.2113/34.1.37. Szwarc, T.S., Johnson, C.L., Stright, L.E., and McFarlane, fluvial systems,in Davidson, S.K., Leleu, S., and North, Simpson, E.L., Koch, R., Heness, E.A., Wizevich, M.C., Tin- C.M., 2015, Interactions between axial and transverse C.P., eds., From River to Rock Record: The Preserva- dall, S.E., Hilbert-Wolf, H.L., Golder, K., and Steul- drainage systems in the Late Cretaceous Cordilleran tion of Fluvial Sediments and Their Subsequent Inter- let, A.K., 2014, Sedimentology and paleontology of foreland basin: Evidence from detrital zircons in the pretation: SEPM Special Publication 97, p. 327–355, the Upper Cretaceous Wahweap Formation sag ponds Straight Cliffs Formation, southern Utah, USA: Geo- https://doi.org/10.2110/sepmsp.097.327. adjacent to syndepositional normal faults, Grand Stair- logical Society of America Bulletin, v. 127, p. 372–392, Wolfe, J.A., and Upchurch, G.R., 1987, North case–Escalante National Monument, Utah: Cretaceous https://doi.org/10.1130/B31039.1. American nonmarine climates and vegetation Research, v. 50, p. 332–343, https://doi.org/10.1016/ Thomson, K.D., Stockli, D.F., Clark, J.D., Puigdefàbregas, during the Late Cretaceous: Palaeogeography, Palaeo- j.cretres.2014.05.001. C., and Fildani, A., 2017, Detrital zircon (U-Th)/(He- climatology, Palaeoecology, v. 61, p. 33–77, https://doi​ Slingerland, R., and Keen, T.R., 1999, Sediment transport in Pb) double-dating constraints on provenance and fore- .org/10.1016/0031-0182(87)90040-X. the Western Interior Seaway of North America: Predic- land basin evolution of the Ainsa Basin, south-central Yonkee, W.A., and Weil, A.B., 2015, Tectonic evolution tions from a climate- sediment model, in Berg- Pyrenees, Spain: Tectonics, v. 36, p. 1352–1375, https:// of the Sevier and Laramide belts within the North man, K.M., and Snedden, J.W., eds., Isolated Shallow doi.org/10.1002/2017TC004504. American Cordillera orogenic system: Earth-Science Marine Sand Bodies: Sequence Stratigraphic Analysis Tilton, T.L., 1991, Upper Cretaceous stratigraphy of the Reviews, v. 150, p. 531–593, https://doi.org/10.1016/ and Sedimentologic Interpretation: SEPM Society for southern Paunsaugunt Plateau, Kane County, Utah j.earscirev.2015.08.001. Sedimentary Geology Special Publication 64, p. 179– [Ph.D. thesis]: University of Utah, Salt Lake City, Zapp, A.D., and Cobban, W.A., 1960, Some Late Cretaceous 190, https://doi.org/10.2110/pec.99.64.0179. 162 p. strand lines in northwestern Colorado and northeastern Slingerland, R., and Smith, N.D., 2004, River avulsions and Tindall, S.E., and Davis, G.H., 1999, Monocline develop- Utah: U.S. Geological Survey Professional Paper 400- their deposits: Annual Review of Earth and Planetary ment by oblique-slip fault-propagation folding: The B, p. 246–249. Sciences, v. 32, p. 257–285, https://doi.org/10.1146/ East Kaibab monocline, Colorado Plateau, Utah: Jour- annurev.earth.32.101802.120201. nal of Structural Geology, v. 21, p. 1303–1320, https:// Science Editor: Rob Strachan Smith, N.D., Cross, T.A., Dufficy, J.P., and Clough, S.R., doi.org/10.1016/S0191-8141(99)00089-9. Associate Editor: Emily Finzel 1989, Anatomy of an avulsion: Sedimentology, v. 36, Tindall, S.E., Storm, L.P., Jenesky, T.A., and Simpson, E.L., Manuscript Received 20 August 2020 p. 1–23, https://doi.org/10.1111/j.1365-3091.1989​ 2010, Growth faults in the Kaiparowits Basin, Utah, Revised Manuscript Received 26 February 2021 .tb00817.x. pinpoint initial Laramide deformation in the western Manuscript Accepted 29 March 2021 Stockli, D.F., Linn, J.K., Walker, J.D., and Dumitru, T.A., Colorado Plateau: Lithosphere, v. 1, no. 4, p. 221–231, 2001, Miocene unroofing of the Canyon Range during https://doi.org/10.1130/L79.1. Printed in the USA

20 Geological Society of America Bulletin, v. 130, no. XX/XX

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/doi/10.1130/B35872.1/5400009/b35872.pdf by guest on 02 October 2021