<<

Research Paper

GEOSPHERE A Late to Early record of orogenic wedge GEOSPHERE; v. 14, no. 3 evolution in the Western Interior basin, USA and https://doi.org/10.1130/GES01606.1 Garrett M. Quinn1, Stephen M. Hubbard1, Peter E. Putnam2, William A. Matthews1, Benjamin G. Daniels1, and Bernard Guest1 1Department of Geoscience, University of Calgary, Earth Science 118, 2500 University Dr. NW, Calgary, , T2N 1N4, Canada 12 figures; 1 table; 1 supplemental file 2Osum Oil Corp., Suite 1900, 255 5th Ave SW, Calgary, Alberta, T2P 3G6, Canada

CORRESPONDENCE: garrett​.quinn@​ucalgary.ca; gquinn7@​gmail​.com ABSTRACT INTRODUCTION

CITATION: Quinn, G.M., Hubbard, S.M., Putnam, P.E., Matthews, W.A., Daniels, B.G., and Guest, B., The Late Jurassic to fill of the Western Interior foreland Aggradation and denudation of proximal deposits in the 2018, A Late Jurassic to Early Cretaceous record basin is characterized using geochronological data in order to assess the wedge-top depozone are first-order controls on sediment flux to more-distal of orogenic wedge evolution in the Western Interior stratigraphic expression of wedge-top geomorphology, as controlled by sedi- parts of foreland basin systems (Ben-Avraham and Emery, 1973; DeCelles, basin, USA and Canada: Geosphere, v. 14, no. 3, ment cover and denudation. In northern Montana, USA, and Alberta, Canada, 1994; DeCelles and Giles, 1996; Roddaz et al., 2005; Ross et al., 2005; Horton, p. 1187–1206, https://​doi​.org​/10​.1130​/GES01606.1. wedge-top deposits are poorly preserved; however, their former presence may 2018). Burial of the frontal toe of the orogen can heal complex topography

Science Editor: Raymond M. Russo be inferred from the detrital record in the foreland basin. We present new U/Pb and enhance direct sediment transfer from the orogenic hinterland to a basin. detrital zircon data from nine samples collected near Great Falls, Montana, Conversely, denudation of these wedge-top sediments can lead to structural Received 24 August 2017 augmented with field data. The at Great Falls is characterized by control of river pathways; the process also can expose older stratigraphy in Revision received 21 February 2018 Late Jurassic marine and nonmarine deposits, which are truncated by a basin-­ the orogenic wedge, which can be reflected by provenance changes in the Accepted 16 April 2018 wide sub-Cretaceous . and lower strata overlying adjacent foreland basin (Ross et al., 2005; Lawton et al., 2010). Published online 7 May 2018 the unconformity are dominated by nonmarine deposits, which transition Sediments in the wedge-top depozone accumulate on top of orogenic up-section into a predominantly marine succession related to a major trans- structures and are part of the frontal toe of the orogenic wedge (DeCelles and gression of the Boreal Seaway in the Albian. Giles, 1996). These synorogenic strata accumulate near the erosional/deposi- Detrital zircon grains from Great Falls strata yield age spectra that can be tional surface and are characterized by low preservation potential such that the subdivided into three groups using multidimensional scaling. Group 1 is char- history of in this zone is difficult to discern (e.g., Coogan, 1992; acterized by diverse zircon populations, which are interpreted to record recy- Frisch et al., 2001; McMechan et al., 2018). More distal parts of the foreland cling of pre-Cordilleran sedimentary strata transported via foreland basin-axial basin contain well-established records of orogenic processes (e.g., Heller and river systems with headwaters in the southwestern United States. Group 2 is Paola, 1989; Ross et al., 2005; Quinn et al., 2016). Therefore, we hypothesize characterized by the dominance of Mesozoic detrital zircon grains, which are that the history of the wedge-top depozone can be elucidated by investigating interpreted to record sediment dispersal by fluvial systems with headwaters strata in more-distal parts of the foreland basin. This is particularly relevant to in the Cordillera. Group 3 is intermediate between groups 1 and 2, based on understanding topographic evolution in fold-thrust belts that have been deeply OLD G its proportion of Mesozoic zircon grains. This group records a diversification incised during subsequent orogenesis and glaciation (Osborn et al., 2006). of the provenance from one dominated by Cordilleran igneous rocks to include We present a detrital zircon data set from Western Interior basin strata ex- recycled sedimentary strata. posed near Great Falls, Montana, USA. The units provide a unique window New data are integrated with three other data sets from Montana and into the Late Jurassic–Early Cretaceous evolution of the foreland basin be- OPEN ACCESS Alberta such that stratal thicknesses (a proxy for accommodation devel- cause outcrops of these strata are rare except within the fold-thrust belt, which opment) and provenance evolution can be compared across the basin. The is ~100 km west of the study area. New data are integrated with previously pre- ­detrital record in each area, which transitions from diverse provenance to sented data sets from across the basin in order to consider the history of burial predominantly Cordilleran through the entire stratigraphic section, can be and exhumation of the frontal toe of the orogenic wedge (Fig. 1) (­Fuentes et al., linked to the burial of the pre-foreland strata in the wedge-top depozone. 2011; Leier and Gehrels, 2011; Raines et al., 2013; Benyon et al., 2014; Blum and This record elucidates a period of evolution of the western margin of North Pecha, 2014; Quinn et al., 2016). This study emphasizes the potential impact This paper is published under the terms of the America to a more Andean-type system with primary input to the basin from of wedge-top dynamics on sediment dispersal across foreland basins and the CC‑BY-NC license. an active magmatic arc. structural evolution of the orogen.

© 2018 The Authors

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1187 Research Paper

A 55°0′0″N W WHR TH ALBERTA GC CLSASKATCHEWAN AA′ AC G GR GF C Fold-Thrust Belt YM AFB Belt-Purcell JE Basin GRP A MONTANA N Great Falls

v v v v v v v 1000 km v v v YC v v 25°0′0″N vv 110°0′0″W v 80°0′0″W vv EM

Figure 1. (A) Modified geological map of Montana, Little Belt USA, which highlights the Great Falls outcrop belt (af- Mountains ter Garrity and Soller, 2009). The inset shows the set- ting of the Great Falls and other areas with detrital zir- con data sets relative to the major bedrock provinces of North America (modified after Dickinson and Gehrels, 2009b). Inset: A—Appalachian Orogen; AC—Archean Craton; AFB—Appalachian Foreland Basin; C—Cordi­ Idaho lleran magmatic arc and accreted terranes; CL—Cold Batholith 100 km Lake; EM—East Mexico Arc; G—Grenville Orogen; GC—; GF—Great Falls; GR—Gibson Res- ervoir; GRP—Granite–Rhyolite Province; JE—Jurassic eolianites; TH—Trans-Hudson Orogen; YC—Yucatan– SedimentaryVolcanic Intrusive Campeche Terrane; W—Wopmay Orogen; WHR—Wyo- ming–Hearne–Rae cratons. (B) Balanced cross section Late Cretaceous to Quaternary Late Cretaceous Late Cretaceous showing a reconstruction of the Early Cretaceous oro- Cretaceous Early to Middle Cretaceous Early to Middle gen–foreland basin system of the southern Canadian Cretaceous Cordillera (after Price, 1994). Line of section is shown in Great Falls outcrop belt Jurassic the inset of Figure 1A. (Jurassic to Early Cretaceous) Jurassic Paleozoic Proterozoic

A B A′ km 0

50 km 50 Foreland basin Late Proterozoic to Middle Jurassic sedimentary rocks Mesozoic intrusions Mid-Proterozoic sedimentary rocks Allochthonous terranes Basement

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1188 Research Paper

STUDY AREA AND STRATIGRAPHY Groff, 1966; Suttner, 1969; Brenner and Davies, 1974; Walker, 1974; Hayes, 1983). Jurassic strata are truncated by the basin-wide sub-Cretaceous uncon- The outcrops of interest to this study are within 30 km of Great Falls, Montana formity (Fig. 2). The overlying is dominated by fluvial (Fig. 1). This location is ideal for the study of basin evolution in the Late Jurassic deposits, with marine influence interpreted by some workers (Glaister, 1959; to Early Cretaceous because of the unique access to outcropping undeformed­ Oakes, 1966; Shelton, 1967; Walker, 1974; Mudge and Rice, 1982; Hopkins, 1985; strata of this age exposed in the distal foreland plains. Because the stratigraphy Hayes, 1986; Farshori and Hopkins, 1989; Hayes, 1990). Two major flooding can be correlated into western Montana and the Alberta foreland basin to the events are evident in Lower Cretaceous strata, recorded by (1) the lacustrine north, the Great Falls area is key for integrating data across the region. or restricted marine Ostracod Member of the Kootenai Formation, and (2) the The stratigraphic nomenclature for this study (Figs. 2, 3) is mostly based marine and shoreface deposits of the Flood Member of the Blackleaf For- on the convention defined by Walker (1974); however, we refer to the basal mation, which overlies the Kootenai Formation (Cannon, 1966; Fox and Groff, of the Kootenai Formation as the Cut Bank Member (Glaister, 1959; 1966; Finger, 1983). Hopkins, 1985) and the Quartzose Sandstone unit as the Sunburst Member (Glaister, 1959; Hopkins, 1985; Hayes, 1990). The Upper Jurassic–Lower Creta- ceous stratigraphy is more than 200 m thick in the study area (Figs. 2 and 3). METHODS Sedimentological characteristics and the interpreted depositional environ- ments for individual units are summarized in Table 1. Measured stratigraphic sections totaling ~290 m were used to compile Jurassic deposits transition from the marine Swift Formation to the non- sedimentological data. Paleocurrent measurements were acquired from dune marine (Table 1) (Cobban, 1945; Harris, 1966; Fox and foresets, ripple foresets, and bed scours (n = 630).

Age Period Epoch Age (Ma) Gibson Reservoir, Montana Great Falls, Montana Grande Cache, Alberta Cold Lake, Alberta Upper Belle Fourche 101 Blackleaf Blackleaf Shaftesbury Fish Scales Fish Scales Flood Member Flood Member Westgate

olorado Viking C Paddy Joli Fou Cadotte Albian Harmon Notikewin Grand Rapids

Upper Kootenai Upper Kootenai Ft. St. John Spirit A to r Undi erentiated Undi erentiated River Falher Clearwater

Upper Upper F ville Kootenai Kootenai Moulton Wilrich Upper

Cretaceous Ostracod Ostracod Bluesky Wabiskaw r r Lowe Mann 113 Red Sandstone r Gething Sunburst McMurray Lowe Lowe Lowe Aptian Kootenai Basal Kootenai Cut Bank

Bullhead 125 Hauterivian 133 Valanginian Monteith/ Berriasian 140 145 Nikanassin Tithonian 152 Morrison Morrison Upper Fernie Kimmeridgian 157 Swift Swift Upper Oxfordian Ellis Fernie Callovian 164 Elli s Jurassic Bathonian Rierdon Middle Bajogian Sawtooth Aalenian

Sandstone, commonly interbedded with shale, mudstone or Location of detrital zircon sample Shale, commonly interbedded with sandstone

Figure 2. Stratigraphic chart of areas within Montana, USA, and Alberta, Canada, that are discussed in this study (after Glaister, 1959; Walker, 1974; Hopkins, 1985; Hayes, 1986, 1990; Fuentes et al., 2011; and Alberta Geological Survey, 2015).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1189 Research Paper

240 Rock samples weighing 3–4 kg were collected in the field for detrital zircon measurements. Two samples were collected from Upper Jurassic Upper Sandstone strata of the Swift Formation (Ellis Group) and of the overlying Morrison 220 Formation (Fig. 2). Five samples were selected from the Early Cretaceous

rmation Kootenai Formation (two from the Cut Bank Member and one each from the Flood Mb. overlying Sunburst, Red Sandstone, and Upper Kootenai Members) (Fig. 2).

leaf Fo Two samples were collected from the Albian–Cenomanian Blackleaf For- 200 mation. Both these samples are from regionally mappable sandstone units

Lower Sandstone Black within the Flood Member, the lowermost unit within the (­Cannon, 1966). Sandstone samples were pulverized, and initial separation of heavy miner- 180 als was performed on an MD Gemini Goldharvester™ shaking table (a water­ table). Zircon was further concentrated using heavy liquids and magnetic separation. A representative fraction of the zircon-rich separate was mounted into a 25.44-mm form and cast in epoxy. Mounts were then ground to expose 160 Upper Kootenai Mb. the cores of the zircons and polished. U-Pb isotopic data were collected using laser ablation–inductively coupled plasma mass spectrometry (LA-ICPMS) at the University of Calgary by the methods of Matthews and Guest (2017). Zir- con grains were ablated using a 33-µm beam diameter, which was chosen Figure 3. Composite stratigraphic section 140 of the Great Falls, Montana, USA, area. to minimize grain-size bias. In each session, four reference materials were The section was produced with field ablated: (1) Temora-2 (Black et al., 2004), (2) 91500 (Wiedenbeck et al., 1995), measurements and published measured (3) FC-1 (Paces and Miller, 1993), and (4) 1242 (Mortensen and Card, 1993). sections from Ballard (1966), Cannon Moulton Mb. 120 This was done to correct for laser-induced elemental fractionation, to account rmation (1966), Fox and Groff (1966), Harris (1966), ness (m) Ostracod Mb. and Walker (1974). c—coarse-grained; for instrumental mass fractionation and drift, to calibrate the measured iso- cl/coal—/coal; f—fine-grained; Fm.— Thick topic ratios, and to validate the measurement method. For data reduction, University of Calgary, Centre for Pure and Applied Tectonics and Thermochronology Formation; m—medium-grained; Mb. — Red Sandstone Mb. 2 Upper Sandstone, Flood Mb ( 47.386600°, -111.388053°) 207 206 1 206 204 Data for Tera-Wasserburg plot 208 206 f206c Pb CPS Pb CPS U (ppm) U/Th Pb/ Pb 1σ % 238 206 207 206 Pb/ Pb 2σx (%) Sample Spot U/ Pb 2σx (%) Pb/ Pb 2σx (%) we used Iolite™ v. 2.5 (Paton et al., 2010) and the custom Microsoft Excel™ UK058UK058_101 NA 65914420 34 0.0NANA62.6174 2.00.04967.8 NA NA UK058UK058_174 NA 297650080.0NANA62.4610 2.10.04979.1 NA NA Member; sl—silt; vf—very fine-grained. UK058UK058_204 NA 108219350.1NANA61.5006 2.20.053512.3NANA 100 UK058UK058_272 NA 3898090230.0 NA NA 60.6428 2.20.05499.8 NA NA macro ARS4.0 (Matthews and Guest, 2017). We performed data visualization UK058UK058_257 NA 50811900 21 0.0NANA60.2047 2.40.04689.5 NA NA Yellow—sandstone; gray—shale, silt- Kootanai Fo UK058UK058_131 NA 123266040.0NANA59.9880 2.10.052511.3NANA UK058UK058_213 NA 212472090.0NANA59.8802 2.20.04919.9 NA NA UK058UK058_172 NA 1994660120.0 NA NA 59.7372 2.30.050610.9NANA UK058UK058_191 NA 58513170 70.1 NA NA 59.6659 2.00.05018.3 NA NA stone; blue—limestone. using the Excel plugin Isoplot (Ludwig, 2012) and plotted normalized proba- UK058UK058_138 NA 1704000100.1 NA NA 59.5238 2.20.046110.6NANA UK058UK058_203 NA 245546070.1NANA59.4530 2.10.04869.7 NA NA UK058UK058_215 NA 220496040.1NANA58.9623 2.20.048410.2NANA UK058UK058_94NA320 7560 13 0.0NANA58.8582 2.00.04759.0 NA NA Sunburst Mb. UK058UK058_260 NA 231451090.0NANA58.6166 2.10.05429.7 NA NA bility density data using an Excel macro from the Arizona LaserChron Center. UK058UK058_270 NA 2746310120.0 NA NA 58.3431 2.10.04829.7 NA NA UK058UK058_244 NA 83 2220 20.1 NA NA 58.1395 2.30.042312.9NANA UK058UK058_51NA88209030.0NANA58.1058 2.40.048712.9NANA UK058UK058_275 NA 1037 22430130.0 NA NA 57.7701 2.00.05037.3 NA NA UK058UK058_102 NA 149318040.1NANA57.6369 2.30.050511.4NANA Full isotopic data and sample locations can be found in the Supplemental UK058UK058_21NA121826300 70.1 NA NA 56.6893 2.00.05067.3 NA NA UK058UK058_70NA136 2920 50.1 NA NA 55.1876 2.50.049411.1NANA 80 UK058UK058_78NA148 3340 10 0.1NANA54.7645 2.50.048012.3NANA 1 UK058UK058_141 NA 312728060.1NANA54.1126 2.10.04779.2 NA NA Data Table . UK058UK058_253 NA 84 1840 20.1 NA NA 54.0833 2.30.051112.3NANA The maximum depositional age (MDA) of the samples was calculated as 1 Supplemental Data Table. U-Pb isotopic data and the weighted average of the youngest dates (minimum of n = 3) that overlap ages of detrital zircon grains. Please visit https://doi​ ​ Cut Bank Mb. .org/10​ ​.1130/GES01606​ ​.S1 or the full-text article on 60 within uncertainty at the 2-sigma level (Dickinson and Gehrels, 2009a). To stan- www​.gsapubs.org​ to view Supplemental Data Table. dardize error propagation between the Great Falls and other data sets used in this study, analytical and systematic errors were added in quadrature to the errors reported in previous studies following recommendations by Horstwood 40 et al. (2016). Morrison Fm. To compare the detrital zircon populations objectively, multidimensional scaling (MDS) plots were constructed using a MATLAB script (MuDisc) pro- vided by Vermeesch (2013). MDS analysis is based on D-values, which are pro- 20 duced as part of the Kolmogorov–Smirnov test and represent “distance” be- tween two samples (Vermeesch, 2013; Spencer and Kirkland, 2015). D-values­ are arranged in a matrix and plotted on a Euclidian plane while attempting to Swift Fm. honor the “distances” in the matrix. Normal-distribution, synthetic age popu- 0 lations were created in Excel and added to the MDS plot to show important age cl/ sl vf fm c coal inputs (Spencer and Kirkland, 2015).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1190 Research Paper

TABLE 1. SUMMARY OF STRATIGRAPHIC UNITS IN THE GREAT FALLS, MONTANA, USA, AREA Lithology and Sedimentary Architectural Additional Interpreted depositional Unit grain-size trends structures elements notes Processes environment References Flood Coarsening upward, shale to Cross-stratification Sheet Bioturbation common in Shoreline ShorefaceCannon (1966); Fox and Member medium-grained sandstone the upper sandstone progradation Groff (1966) of the Flood Member Upper Fining upward, medium-grained Cross-stratification Lenticular sandstone bodies Unidirectional Fluvial Glaister (1959); Walker Kootenai sandstone to mudstone. Lignite (up to 10 m thick) currents, (1974) Member at base of succession channel cut and fill Moulton Coarsening upward, shale and Ripples, horizontal Sheet sandstone (about 2 m Sandstone is often Shoreline Shallow marine or Oakes (1966); Walker Member siltstone to fine-grained sandstone lamination thick) bioturbated progradation lacustrine (1974) Ostracod Limestone and mudstone Ostracod and gastropod High biological Restricted marine Walker (1974); Finger Member body . Regional stress, calcite (1983) stratigraphic marker precipitation in restricted basin Red Fine- to medium-grained sandstone Cross-stratification, Lenticular sandstone Planolites burrows Unidirectional Fluvial, shallow marine Glaister (1959); Walker Sandstone with rare pebbles and current ripples bodies (up to 35 m thick), observed in some currents, (1974); Hopkins (1985) Member mudstone breccia. Red and gray inclined heterolithic sheet sandstones channel cut mudstone stratification, off-channel and fill, point- sheet sandstones (up to bar accretion 2 m thick) Sunburst Fining upward, coarse-grained Cross-stratification Lenticular sandstone bodies Simple horizontal Unidirectional Fluvial, estuarine, and Glaister (1959); Walker Member sandstone to mudstone. Fine- (up to 15 m thick), inclined burrows in siltstone currents, shallow marine (1974); Hopkins (1985); grained interbedded with heterolithic stratification, and sandstone of channel cut Farshori and Hopkins siltstone. Molluscan biomicrite and inclined point-bar inclined heterolithic and fill, point- (1989); Hayes (1990) in mudstone-filled channel accretion sets stratification bar accretion Cut Bank Fining upward, coarse-grained Cross-stratification Lenticular sandstone bodies Mudstone interbedded Unidirectional Fluvial, pedogenic- Glaister (1959); Oakes Member sandstone with rare boulder-size (up to 25 m thick), inclined with calcite-concretion currents, concretion formation (1966); Shelton (1967); chert clasts. Red mudstone accretion surfaces horizons and micritic channel cut may indicate semiarid Walker (1974); Mudge limestone and fill climate, limestone may and Rice (1982); Hayes indicate lacustrine (1986) environments Morrison Fine-grained sandstone, siltstone, Current ripples, root Lenticular sandstone bodies Unidirectional Fluvial with coal-forming Ballard (1966); Fox and Formation mudstone, coal imprints (up to 3 m thick) currents, floodplain Groff (1966); Harris channel cut (1966); Suttner (1969); and fill Walker (1974) Swift Dark-gray shale and fine-grained Cross-stratification, Sheet sandstone bodies Accessory glaucontie Waning storm Shallowing upward marine Cobban (1945); Ballard Formation sandstone interbedded with hummocky/swaley (24 m thick at type section and coaly fragments. energy, wave (1966); Harris (1966); mudstone. Isolated chert-pebble/ cross-stratification, in northwest Montana) Chondrites burrows currents, Brenner and Davies belemnite conglomerate flaser bedding, unidirectional (1974); Hayes (1983) ripples, wavy currents bedding

RESULTS and 1200–950 Ma), (2) Neoproterozoic populations (at 950 Ma and 615– 590 Ma), and (3) a prominent population (at 310 Ma). The zircon Detrital zircon results are presented as probability density plots and maxi- populations of the Swift and Morrison Formations are composed of 20% Me- mum depositional ages are reported in Figure 4. sozoic detrital zircons with age modes in the (235 Ma) and Jurassic The oldest zircon populations in the Swift and Morrison Formations are (160–150 Ma). In the Swift Formation, the Jurassic mode dominates, whereas Archean in age (2700–2640 Ma) (Fig. 4). Both samples contain (1) prominent the Triassic and Jurassic modes are almost equal in the Morrison Formation. Mesoproterozoic to earliest Neoproterozoic populations (at 1470–1400 Ma Both these units exhibit average paleocurrent directions to the north (Fig. 4B).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1191 Research Paper

Upper Sandstone; Flood Mb. n=271 Lower Sandstone; A MDA: 103±1.7 Ma B Flood Mb. n = 30 Mean = 351° Upper Kootenai Mb. n = 42 Mean = 102° Lower Sandstone; Flood Mb. n=283 MDA: 109±1.7 Ma Red Sandstone Mb. n = 105 Mean = 20°

Sunburst Mb. n = 206 Upper Kootenai Mb. n=283 Mean = 256° MDA: 104±1.8 Ma

Cut Bank Mb. n = 158 Red Sandstone Mb. n=249 Mean = 353° MDA: 108±1.6 Ma Sunburst Mb. n=244 MDA: 154±3.2 Ma Cut Bank Mb. n=256 MDA: 158±2.6 Ma Morrison Fm. Cut Bank Mb. n=266 MDA: 155±2.9 Ma n = 12 Mean = 7˚ Morrison Fm. n=241 MDA: 151±2.2 Ma Swift Fm. n=247 Swift Fm. Figure 4. (A) Normalized probability density plots MDA: 152±2.4 Ma n = 78 (errors incorporated at the 1-sigma level) and 0500 1000 1500 2000 2500 3000 3500 Mean = 359° maxi­mum depositional ages (MDA) for nine de- Age (Ma) trital zircon samples. (B) Paleoflow rose diagrams for seven stratigraphic units. North is to the top of each rose diagram. (C) Probability density plots C Upper Sandstone; Flood Mb. n=271 (errors incorporated at the 2-sigma level) for each of the stratigraphic units sampled with a scale change at 350 Ma to highlight young detrital zircon populations. The two Cut Bank Member samples Lower Sandstone; Flood Mb. n=283 are grouped because of their similarity. Colors indicate the major magmatic provinces of North America. Fm—Formation; Mb—Member; MDA— maximum depositional age. N—number of sam- Upper Kootenai Mb. n=283 ples; n—number of individual measurements (ages and flow directions).

Red Sandstone Mb. n=249

Sunburst Mb. n=244

Cut Bank Mb. N=2; n=522

Morrison Fm. n=241

Swift Fm. n=247

50 100 150 200 250 300 350 600 1100 1600 2100 2600 3100 Age (Ma) Cordilleran Orogen (<245 Ma) Appalachian Orogen (850–285 Ma) Granite-Rhyolite Province/A-Type Plutons (1500–1300 Ma) East Mexico Arc (284–232 Ma) Grenville Orogen (1300–900 Ma) Yavapai-Mazatzal (1800–1600 Ma) Trans-Hudson Orogen (1900–1800 Ma)

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1192 Research Paper

The Cut Bank and Sunburst Member samples share many populations Windley, 1988; Anderson and Bender, 1989; Whitmeyer and Karlstrom, 2007; with the underlying strata. The Cut Bank and Sunburst Members have more- Dickinson, 2008). Neoproterozoic detrital zircon grains (850–550 Ma) are de- prominent­ populations in the Archean to Paleoproterozoic (2800–2600 Ma and rived from peri-Gondwanan terranes that collided with North America during 2150–2000 Ma) and the (430–410 Ma) than do the Swift and the Appalachian Orogen (Eriksson et al., 2004; Becker et al., 2005; Park et al., Morrison strata. The Cut Bank Member samples also have a prominent mode 2010). Plutons associated with the Appalachian Orogen and precursor rift as- in the Paleoproterozoic at 1840–1780 Ma, and the Sunburst Member has a semblages span the Neoproterozoic to the Paleozoic and supply a portion of prominent mode at 1620 Ma. Both have populations in the –Triassic the 760–285 Ma grains (Dickinson and Gehrels, 2009b; Park et al., 2010). (260–230 Ma). The zircon populations of the Cut Bank and Sunburst Members Magmatic assemblages younger than the middle Permian are associated are composed of 10%–11% and 8% Mesozoic detrital zircons, respectively, in- with the East Mexico Arc (284–232 Ma) (Torres et al., 1999), the Cordilleran mag- cluding a Jurassic population (160–155 Ma). Like the Jurassic strata, measure- matic arc (<245 Ma), and accreted terranes of the Cordilleran Orogen (Fig. 5). ments from the Cut Bank Member indicate an average paleoflow to the north. Triassic and later Mesozoic igneous rocks and detrital zircon are present on the Measurements from the Sunburst Member display a large spread but average Wallowa and Olds Ferry Terranes of Washington, Idaho, and Oregon (LaMaskin paleocurrent direction is to the west (Fig. 4B). et al., 2011; LaMaskin, 2012; LaMaskin et al., 2015; Gaschnig et al., 2017a). Triassic The zircon populations of the Red Sandstone Member, Upper Kootenai to Early Jurassic magmatic assemblages are reported from the Mojave Desert Member and the lower sandstone of the Flood Member are distinct from the and the Eastern Coast Plutonic Complex (Barth and Wooden, 2006; Gehrels et al., older samples in that they are composed mainly of Mesozoic detrital zircon 2009). Jurassic plutons are present in the Western Coast Plutonic Complex, Si- grains (84%–95%) (Fig. 4C). There are three major Late Triassic to Cretaceous erra Nevada Batholith, and the Omineca Belt of interior (Fig. 5) populations in the probability density plots. The most prominent is Jurassic (Archibald et al., 1983; Hyndman, 1983; Armstrong, 1988; Ducea, 2001; Gaschnig in age (160 Ma), and the two subordinate populations are Late Triassic–Early et al., 2009; Gehrels et al., 2009; Paterson et al., 2011). Similar segments of the Jurassic (220–190 Ma) and Cretaceous (115–110 Ma) in age. The Red Sand- Cordillera were active in the Early Cretaceous. An additional episode of Early stone Member is characterized by north-directed paleocurrent measurements. Cretaceous magmatism is evident in the Suture Zone Suite of the western mar- Paleocurrent indicators in the Upper Kootenai Member are characterized by gin of the Idaho Batholith (Manduca et al., 1993; Gaschnig et al., 2017b). mean paleocurrent direction to the east. Paleocurrent measurements collected from the lower sandstone of the Flood Member at one outcrop location are characterized by paleoflow to the north (Fig. 4B). Provenance Interpretation and Evolution The zircon population of the upper sandstone of the Flood Member is com- posed of 51% Mesozoic detrital zircon grains (Fig. 4). The Mesozoic zircon pop- Samples are subdivided into two endmember groups and one intermedi- ulations are the same as underlying strata (160 and 110 Ma). There are Protero- ate group based on the MDS plot and visual inspection of their detrital zircon zoic populations, most notably at 1830 and 1030 Ma, and minor populations at populations (Fig. 6). The first endmember is characterized by diverse detrital 650, 450, and 360 Ma. zircon spectra with populations spanning the Mesozoic to Archean. The other endmember is characterized by detrital zircon spectra dominated by Mesozoic populations derived from Cordilleran magmatic rocks. The intermediate group ANALYSIS AND INTERPRETATION appears to be transitional with both Mesozoic populations and Paleo­zoic-to- zircon populations. The ultimate sources of detrital zircon grains are interpreted on the ba- sis of correlation with published ages of North American magmatic assem- blages (Figs. 1 and 4C). Archean detrital zircons (>2500 Ma) ultimately derive Group 1—Diverse Detrital Zircon Spectra from the cratonic cores of North America (Hoffman, 1988; Card, 1990). Paleo­ proterozoic­ detrital zircon grains are associated with the Trans-Hudson Orogen Group 1 consists of the basal units in this study (i.e., Swift, Morrison, Cut (1900–1800 Ma), similar orogens that stitch together the Archean cores, and Bank and Sunburst strata) and is characterized by detrital zircon populations the Yavapai–Mazatzal Province (1790–1610 Ma) (Hoffman, 1988; Van Schmus derived from all the major magmatic provinces of North America (Fig. 6). The et al., 1993; Van Kranendonk et al., 1993; Zhao et al., 2002; Whitmeyer and distribution of zircon populations is consistent with recycling of sedimentary Karlstrom, 2007). Mesoproterozoic detrital zircon sources include the Granite– strata uplifted along the Mogollon highlands and incipient Cordilleran fold- Rhyolite Province in the southern and eastern United States (1550–1300 Ma) thrust belt in the southwestern United States (Fig. 7; Dickinson and Gehrels, (Whitmeyer and Karlstrom, 2007), A-Type plutons found throughout the 2008a, 2009b; Leier and Gehrels, 2011; Laskowski et al., 2013; May et al., 2013). United States (1480–1340 Ma) and the Grenville Orogen of the eastern and Group 1 samples are also characterized by a Jurassic detrital zircon population southern United States and eastern Canada (1300–900 Ma) (Easton, 1986; from primary igneous sources in the Cordillera (Fig. 4).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1193 Research Paper

115 Ma 160 Ma 200 Ma 235 Ma 260 Ma Figure 5. Ages of magmatic events in North America between 280 and 80 Ma (Folinsbee Omineca Belt et al., 1957; Pearce, 1970; Archi­bald et al., Wallowa/Olds Ferry Terrane Sierra Nevada Batholith 1983; Hyndman, 1983; Torres et al., 1999; Idaho Batholith Mojave Desert Barth and Wooden, 2006; Gehrels­ et al., Western Coast Plutonic Complex 2009; Gaschnig et al., 2009; LaMaskin et al., 2011; Gaschnig et al., 2017b). Ages with Crowsnest Volcanics Eastern Coast Plutonic Complex East Mexico Arc arrows at the top denote significant zircon populations in the Great Falls, Montana, 80 100 120 140 160 180 200 220 240 260 280 USA, data set. Age (Ma)

Mesozoic eolianites in the southwestern United States (Fig. 7) are char- tions similar to the southwestern eolianites (LaMaskin et al., 2011; LaMaskin, acterized by broad detrital zircon populations derived from the Appalachian 2012; Gehrels and Pecha, 2014). The Belt Supergroup, which outcrops in orogeny (populations at 615 and 420 Ma) and the Grenville orogeny (popula- northwestern Montana, contains abundant Proterozoic and Archean detrital tions at 1160 and 1055 Ma) (Dickinson and Gehrels, 2008a, 2009b; Laskowski zircons (Ross and Villeneuve, 2003). The Yavapai–Mazatzal-age Big Sky Oro- et al., 2013). Paleozoic and Proterozoic detrital zircon populations in Group 1 gen in southwestern Montana is an alternative source for Mesoproterozoic samples are similar to the Mesozoic eolianites (Fig. 7). Additionally, the eoli- detrital zircon (Harms et al., 2004). Three distinct chert varieties—black, red, anite strata contain Late Permian–Triassic populations (260–235 Ma) (Fig. 4), and phosphatic—and sponge spicules are reported from petrographic study which constitute a major population in each Group 1 sample and are also of the Morrison and Kootenai Formations and give the sandstones a “salt- common in Triassic strata (e.g., Chinle Group, Moenkopi Formation) of the and-pepper” appearance in the field (Suttner, 1969). Suttner (1969) ascribes southwestern United States (Dickinson and Gehrels, 2008b). Group 1 samples provenance of these components to the Permian–Pennsylvanian Phosphoria, have Jurassic populations (165–150 Ma) that are too young to be a significant Wood River, and Quadrant Formations of western North America and notes component of the eolianite strata and must be attributed to the Cordilleran that because the Permian–Pennsylvanian section is relatively intact in Mon- magmatic arc. Plutonic and volcanic strata of this age are widespread along tana, these diagnostic components are derived from west or south of the the strike of the western continental margin, limiting their usefulness in prov- Montana–Idaho border. Texturally, framework grains of the Swift, Morrison, enance interpretation (Fig. 5). and Kootenai Formations have been reported as round to subround, which Local sources of detrital zircon grains could also account for specific supports the hypothesis of second-cycle deposition of these units (Ballard, modes in these samples. Sedimentary strata of Cordilleran terranes and North 1966). Given the limitations of this data set, it is difficult to rule out these local American passive-margin units west of the study area have zircon popula- sources; however, north-directed paleoflow for the Swift, Morrison, and Cut Bank units, and the continuity of similar Morrison Formation (Dickinson and Gehrels, 2008a; May et al., 2013) and Early Cretaceous (Leier and Gehrels, Group 1. Diverse Group 2. Mesozoic- 2011) detrital zircon spectra with samples south of the Great Falls area support Group 3. Intermediate Detrital Zircon Spectra dominated Detrital southern provenance as most likely. Zircon Spectra

0.2

Sunburst Mb. 0.1 Upper Kootenai Mb. Group 1; Great Falls; N=5, n=1253 Cut Bank Mb. 1 Upper Flood Mb. 0 Cut Bank Mb. 2 Red Swift Fm. Sandstone Mb. Lower Mesozoic Eolianites, SW U.S., –0.1 Morrison Fm. Flood Mb. (Dickinson and Gehrels, 2008a);N=10, n=942 Dimension 2 (unitless) –0.2 0500 1000 1500 2000 2500 3000 3500 Age (Ma) –0.4 –0.3 –0.2 –0.1 00.1 0.20.3 0.40.5 0.6 Dimension 1 (unitless) Figure 7. Composite probability density plot of all the detrital zircon ages from Group 1 samples from Great Falls, Montana, USA, compared Figure 6. Multidimensional scaling plot of the nine detrital zircon samples from to a composite probability density plot of detrital zircon ages from the Great Falls, Montana, USA, showing two endmember groups of zircon spectra southwestern United States eolianites (Dickinson and Gehrels, 2008a). and one intermediate group. Solid lines indicate nearest neighbors, and dashed Error incorporated into both probability density functions are at the lines indicate second nearest neighbors. Mb.—Member; Fm.—Formation. 2-sigma level.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1194 Research Paper

The Sunburst Member is unique in that it is characterized by an average INTEGRATION OF STRATIGRAPHY AND GEOCHRONOLOGY paleoflow direction to the west, which is difficult to reconcile with detrital zir- OF THE WESTERN INTERIOR BASIN con spectra that suggest a source area to the south (Fig. 4). The population at 260 Ma in the Sunburst Member overlaps in age with plutonism in the East Expansive detrital zircon data sets from Upper Jurassic to Lower Creta- Mexico Arc, reinforcing a southern component of the provenance. Of note, ceous strata across the Western Interior basin were integrated with the data the Sunburst Member is also distinct petrographically, with higher proportions collected from Great Falls (Fig. 8). These areas include Gibson Reservoir of of quartz and less chert than the underlying units (Walker, 1974; Hayes, 1990). western Montana (Fuentes et al., 2011), the Grande Cache area of west-central We speculate that the drainage system deviated east of Great Falls, perhaps Alberta (Raines et al., 2013; Quinn et al., 2016), and Cold Lake in east-central deflecting away from uplifted highlands (cf. Christopher, 2003) and eventually Alberta (Blum and Pecha, 2014). The lithostratigraphic nomenclature for each diverting westward through the study area. As such, the ultimate sources of area is presented in Figure 2. zircon remain unchanged from underlying units; the deviation of the sediment A generalized stratigraphic cross section shows variability in the thickness routing system away from the fold-thrust belt led to reduced enrichment of and stratigraphic characteristics across the four regions (Fig. 9). This variability chert and lithic grains in the unit. reflects differences in generation and destruction of accommodation from the proximal foredeep to the distal foreland basin. An MDS plot was generated using all samples from the four areas Group 2—Cordilleran-dominated Detrital Zircon Spectra (Fig. 10). Samples fall into similar groupings as those described from the Great Falls data set, with one endmember population dominated by Cor- Group 2 includes samples from the Red Sandstone Member, the Upper dilleran magmatic source rocks and the other endmember comprised of Kootenai Member, and the lower sandstone of the Flood Member, which are diverse detrital zircon populations. Samples yielding diverse zircon spec- characterized by 84–95% Mesozoic detrital zircon ages (Fig. 6). Because the tra fall on a continuum between synthetic age populations representing timing of magmatic episodes is common across many parts of the Cordillera, the Trans-Hudson (1850 ± 50 Ma), Grenville (1050 ± 50 Ma), and Gond- attributing the populations in this group (160 Ma, 110–115 Ma, and 190–220 Ma) wanan orogens associated with the Appalachians (600 ± 50 Ma) (Fig. to individual segments of the Cordillera is difficult (Fig. 5). Paleoflow mea- 10). Samples that plot closer to the Grenville and Gondwanan synthetic surements from the Red Sandstone Member and the lower sandstone of the populations, including those of Group 1 from Great Falls, are interpreted Flood Member are consistent with a southern source and basin-axial sediment to record recycling of sedimentary strata from the southwestern United routing (Fig. 4B). East-directed paleoflow measurements support transverse States (Dickinson and Gehrels, 2009b; Leier and Gehrels, 2011; Gehrels sediment dispersal from the orogen during deposition of the Upper Kootenai and Pecha, 2014), or deposition from continental-scale river systems such Member (Fig. 4B). as interpreted for Cold Lake (Blum and Pecha, 2014). Samples that plot closer to the Trans-Hudson synthetic population are interpreted to de- rive from uplifted passive margin strata in the northern Cordillera (Leier Group 3—Intermediate Detrital Zircon Spectra and Gehrels, 2011; Laskowski et al., 2013; Raines et al., 2013; Gehrels and ­Pecha, 2014; Quinn et al., 2016). Samples that cluster near the Jurassic The upper sandstone of the Flood Member, the stratigraphically youngest synthetic population include samples that are similar to Group 2 from sample in this study, is composed of 51% Mesozoic detrital zircon grains, Great Falls and record provenance dominated by Cordilleran magmatic making it intermediate between the two endmembers, Group 1 and Group rocks (Fig. 10). 2 (Fig. 6). This suggests that the sediment source areas diversified upwards The samples exhibit some evidence of geographic clustering (Fig. 10). through the stratigraphic section. The Mesozoic populations in Group 3 over- Samples from Grande Cache are associated with sources in the Canadian lap with Group 2 samples, indicating that similar Cordilleran arc sediment passive margin (e.g., Monteith A member, Cadomin and Bluesky Forma- sources contributed detritus to this unit. Proterozoic detrital zircon popula- tions), and Great Falls samples are associated with sources in the southwest- tions are dominated by a Grenville-age population (1030 Ma) and a subor- ern United States (e.g., Swift and Morrison Formations, and Cut Bank and dinate Trans-Hudson-age population (1830 Ma). Neoproterozoic to Paleozoic Sunburst Members). The samples from the Cold Lake area are consistent (670–370 Ma) zircon ages also occur in composite signature of the Mesozoic with either a southwestern United States source or the continental-river hy- eolianites of the southwestern United States and in populations of Group 1 potheses of Blum and Pecha (2014) and Benyon et al. (2014, 2016). Jurassic (Fig. 4) (Dickinson and Gehrels, 2009b). The upper sandstone of the Flood samples from Gibson Reservoir likely also derive from the southwestern Member could record either the recycling of Jurassic–Lower Cretaceous fore- United States, whereas Early Cretaceous samples (e.g., basal conglomer- land basin strata or the reincorporation of pre-foreland sedimentary strata ate of the Kootenai Formation) reflect derivation from uplifted pre-foreland into the drainage system. ­basin strata.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1195 Research Paper

A Grande Cache, Alberta B Cold Lake, Alberta SPATIAL–TEMPORAL BASIN EVOLUTION Upper Grand Rapids (AOS9) n=99 By comparing new data from Great Falls with data from Gibson Reservoir, Middle Grand Rapids (AOS12) n=103 Grande Cache, and Cold Lake, the evolution of sediment sources, sediment dispersal, and basin partitioning can be assessed over a period of orogen evo- Chinook Fm n=86 lution and basin filling that lasted for ~50 m.y. (Fig. 11). The interpretation of Lower Grand Rapids (AOS20) n=102 sediment routing systems from the Cordillera and other areas of North Amer- Cardium Fm n=140 ica to the foreland basin illustrate large-scale trends in the evolution of drain- age in the region. Dunvegan Fm n=92 Upper Clearwater (AOS18) n=103 Cadotte Mb n=134 Late Jurassic

Jurassic deposits record marine to nonmarine basin filling, which is typical Middle Clearwater (AOS16) n=103 Notikewin Mb n=137 of foreland basin clastic wedges (Table 1) (Cant and Stockmal, 1989; Fuentes et al., 2011; Miles et al., 2012; Kukulski et al., 2013; Raines et al., 2013). Overall, the Lower Clearwater (AOS14) n=93 thickness of these units is controlled by variations in accommodation and dif- ferential erosion of pre-Cretaceous units during incision of the sub-Cretaceous Falher A Mb n=137 Lower Clearwater (AOS6) n=102 unconformity (Fig. 9) (Leckie and Smith, 1992; Hayes et al., 1994; Gillespie­ and Falher D Mb n=127 Heller, 1995; Miles et al., 2012). That continental deposits at Grande Cache, Al- McMurray Fm (AOS17) n=89 berta (e.g., Monteith Formation), are stratigraphically younger than continental Bluesky Fm n=86 McMurray Fm (AOS7) n=83 deposits in Montana (e.g., Morrison Formation) is interpreted to record south- Gething Fm n=114 0500 1000 1500 2000 2500 3000 3500 4000 to-north, basin-axial filling of foreland accommodation space (Fig. 2) (Poulton Age (Ma) Cadomin Fm n=134 et al., 1990; Kukulski et al., 2013). Monteith A Mb n=100 MDAs for the Swift and Morrison Formations from Great Falls (152 ± 2.4 Ma Monteith A Mb n=152 and 151 ± 2.2 Ma respectively) and Gibson Reservoir (168 ± 5.5 Ma and 162 ± Monteith B Mb n=139 4.5 Ma support Late Jurassic deposition of these units; these ages indicate that Monteith C Mb n=130 deposition may have been diachronous, or that different sources of Jurassic 0500 10001500 20002500 3000 3500 4000 zircon were available to the different areas. Age (Ma) Detrital zircon data, including the close association of Jurassic units at Gibson Reservoir, Montana CA Great Falls and Grande Cache (Fig. 10), support the interpretation that Late Jurassic sediment dispersal was characterized by basin-axial fluvial systems Figure 8. Normalized probability density plots of detrital zir- with headwaters in the southwestern United States (Fig. 11) (Hamblin and con data from three other areas in the Western Interior basin. Walker, 1979; Miles et al., 2012; Raines et al., 2013). While this interpretation Errors are incorporated at the 1-sigma level. (A) Data from Blackleaf Fm (1SR80) n=95 Grande Cache, Alberta, Canada (Raines et al., 20136; Quinn apparently best fits this data set, transverse sediment routing during this time et al., 2016; items in parentheses correspond to the sample period has also been reported. For example, the upper part of the Monteith designations in these papers). (B) Data from Cold Lake, Al- Formation (Monteith A member) records local recycling of pre-foreland sedi- berta, Canada (Blum and Pecha, 2014). (C) Data from Gibson mentary strata of the Canadian passive margin (Figs. 10 and 11). Raines et al. Kootenai Fm (1FG70) n=99 Reservoir, Montana, USA (Fuentes et al., 2011). Fm—Forma- tion, Mb—Member. (2013) attributed this to the propagation of transverse drainage systems across the foredeep. Additionally, the Morrison Formation in Utah has been shown Kootenai Fm (1SFSR1) n=97 to include the record of transverse sediment dispersal by distributive fluvial Kootenai Fm (1GR100) n=95 systems (Hartley et al., 2015; Owen et al., 2015). Previously, Upper Jurassic strata in Montana have been assigned to a backbulge setting with passage of Morrison Fm (1GRX) n=91 the forebulge interpreted to have created the sub-Cretaceous uncon­ formity­ Morrison Fm (1GRZ) n=87 (DeCelles, 2004; Fuentes et al., 2009; 2011). The detrital zircon spectra are Swift Fm (1GR14) n=98 Sawtooth Fm (EB) n=97 statistically similar at Great Falls, Gibson Reservoir, and Grande Cache, and 0 500 1000 1500 2000 2500 3000 3500 4000 all the samples contain Jurassic populations, which derive from the west Age (Ma) though they may have been delivered to the basin aerially as ash fall (Fig. 8).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1196 Research Paper m—medium sandstone; Ss.—sandstone; yellow—sandstone; gray—shale; blue—limestone. the subsurface at Cold Lake, Alberta, Canada (Hutcheon et al., 1989; Feldman et al., 2008). Fm.—Formation; Mb.—Member; cl/coal—clay/coal, vf—very fine sandstone; section from the northwest Alberta foothills, Grande Cache, Alberta, Canada (Miles et al., 2012; Kukulski et al., 2013; and Quinn et al., 2016). (D) Stratigraphic section from from Great Falls, Montana, USA, from field measurements and Ballard (1966), Cannon (1966), Fox and Groff (1966), Harris (1966), and Walker (1974). (C) Stratigraphic each unit. Location map is inset. (A) Stratigraphic section from the foothills of western Montana, USA, at Gibson Reservoir (Fuentes et al., 2011). (B) Stratigraphic section Vertical scale is in meters. Units with detrital zircon data are indicated by the black diamonds with numbers in parentheses indicating the number of zircon samples for Figure 9. Cross section correlating composite stratigraphic sections from each area using the top of the Kootenai Formation (and equivalent surfaces) as the datum. 10 0 20 0 30 0 Gibson Reser 0 coal cl/ A vf Upper Kootanai Fm Co Kootenai Morrison voir Swif m nglomerat C Basal t Fm Gr (2) ande Cache (2) Fm . Fm Co e . . Gibson Rese rdilleran Idaho Sub . -C retaceous Unconformit Alber fo ld-thrust belt 10 0 rv ta 0 Co oi r A ld La coal Great F cl/ ke D Montana B Gr B Morrison vf eat 100 km Upper Kootanai Fm alls N m Fl C Fa Sunburst Mb Swif Red Ss Ostracod Mb y ut Bank ood Mb lls Mb T t Fm Fm op Kootenai . . Mb . . (2) . . . . . Fm . 800 600 700 400 500 200 300 100 0 Grande Cach coal cl/ C vf Wi m Gethin Fe Pa Bluesk lrich Mb Fm Shaf rnie Cadotte mb e Mont ddy/Harmon Fm Cadomin Spirit River Fm. Notikewin Mb . Fa g . Fm tesbur Fm Fm y lher Mb eith Fm . . . (3) . (4)

y . . . / . Su b- Cretaceous Unconformi 100 0 Co coal cl/ ld Lake D vf McMurr m Clear Upper Grand Lowe Fm R R Ca De apids Fm apids Fm ty Fm rbonates . (4) water vonian r Grand ay . (2)

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1197 Research Paper

A Increasing Canadian 1850±50 Ma Passive Margin Signal Sample Key: Great Falls, Montana (this study) Gibson Reservoir , Montana (Fuentes et al., 2011) 0.6 Cold Lake, Alberta (Blum and Pecha, 2014) Grande Cache, Alberta (Raines et al., 2013; Quinn et al., 2016) Monteith A Synthetic age population

0.4 Increasing Cordilleran Magmatic Arc Signal

Bluesky

Dunvegan Basal Kootenai (1SFSR1) Blackleaf (1SR80) Cadomin 0.2 Sawtooth Basal Kootenai (1GR100) (Eb) Cardium Gething 160±50 Ma Falher D Chinook Cadotte L. Grand Rapids (AOS20) Lower Flood 0 McMurray Monteith C Red Sandstone Upper Kootenai (AOS7) L. Clearwater M. Clearwater (AOS16) Dimension 2 (unitless) L. Clearwater Upper (AOS6) (AOS14) Flood U. Clearwater Notikewin (AOS18) Upper Kootenai (1FG70) McMurray M. Grand (AOS17) –0.2 Rapids (AOS12) U. Grand Rapids Monteith B (AOS9) Part B Cut Sunburst B Bank Falher A 1 Morrison (GRX) –0.4 Cut Bank 2 Mon- Morrison teith A Swift (GRZ) 1050±50 Ma (1GR14) Swift Note: Box edge lengths –0.6 Increasing Southwestern span a “distance” of 0.2 in each dimension Morrison US Signal 600±50 Ma

–0.4 –0.2 0 0.2 0.4 0.6 0.8 Dimension 1 (unitless)

Figure 10. Multidimensional scaling (MDS) plot of data from areas in Montana, USA, and Alberta, Canada. Solid lines indicate nearest neighbors, and dashed lines indicate second nearest neighbors. (A) MDS plot of all the detrital zircon data from each of the four study areas. Synthetic age popu- lations are shown with stars to highlight important inputs to the samples. Annotations indicate the interpretation of the provenance signals. The dashed box indicates the expanded area in Figure 10B. (B) Scale expansion to show the statistical differences between samples that are too similar to observe at the scale of Figure 10A. L—Lower; M—Middle; U—Upper.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1198 Research Paper

Legend Upper Jurassic strata of the Monteith Formation in the Grande Cache area A GC Sedimentary have been assigned to the foredeep on the basis of thickness, sandstone CL Great Falls provenance, and composition (Price and Mountjoy, 1970; Monger et al., 1982; Outcrop Belt Cant and Stockmal, 1989; Leckie and Smith; 1992, Miles et al., 2012). The Paleozoic similarity between the detrital zircon spectra in the Late Jurassic system and Proterozoic the occurrence of Cordilleran magmatic-arc grains in these deposits do not support the hypothesis that a forebulge significantly influenced sediment GR GF Volcanic dispersal in the basin during the Late Jurassic. The basin configuration is po- CM E.-M. Cretaceous Jurassic tentially analogous to the modern foreland east of the central Andes, where PF the forebulge is buried and does not influence sediment routing patterns Granitic (Chase et al., 2009). E.-M. Cretaceous Jurassic Early Cretaceous–Aptian JrE General sediment Overall, the thickness of Early Cretaceous strata (e.g., Lower Mannville dispersal Group in Alberta) is fairly consistent across the region (Hayes et al., 1994), B GC CL Mapped indicating that foreland subsidence associated with a crustal load to the west highlands had limited impact on accommodation development (Gillespie and Heller, 1995). The thickness variations that do exist in Lower Cretaceous deposits are significantly influenced by differential erosion on the angular sub-Cretaceous unconformity, with thicker sediment accumulations occupying paleovalley GR GF systems incised into Jurassic and pre-foreland sedimentary rocks (Jackson, CM 1984; Hayes, 1986; Ranger and Pemberton, 1988; Wightman and Pemberton, 1997; Ardies et al., 2002; Zaitlin et al., 2002). PF Figure 11. Maps showing generalized sedi­ Fluvial sandstones of the Cut Bank Member and the McMurray Formation ment dispersal into the Western Interior basin, modified after Jackson (1984), overlie the sub-Cretaceous unconformity in the Great Falls and Cold Lake Leckie and Smith (1992), Dickinson and areas, respectively (Table 1; Figs. 2 and 9). At Gibson Reservoir and Grande Gehrels (2009b), and Garrity and Soller Cache, the sub-Cretaceous unconformity is overlain by fluvial conglomerate (2009). CL—Cold Lake; CM—area domi­ (the basal conglomerate of the Kootenai Formation and the Cadomin Forma- JrE nated by Cordilleran magmatic rocks (green shading); GC—Grande Cache; GF— tion, respectively (Figs. 2 and 9) (McLean, 1977; Heller and Paola, 1989; Leier Great Falls; GR—Gibson Reservoir; JrE: and Gehrels, 2011). approximate extent of Jurassic eolianites MDAs for the McMurray Formation (117 ± 4.7 Ma) (Benyon et al., 2014, 2016) C GC (yellow shading); PF—area dominated CL by pre-foreland sedimentary strata (blue and the (117 ± 2.4 Ma) are consistent with Aptian deposi- shading). E.-M.—Early-Middle.­ (A) Late tion and validate the stratigraphic correlation of these units. Syndepositionally Jurassic sediment dispersal (i.e., Morrison formed zircon grains are not abundant in basal Cretaceous strata of the south- Formation, Monteith Member). (B) Aptian sediment dispersal (i.e., Cut Bank Mem- ern areas, so robust MDAs could not be calculated; however, biostratigraphic ber, basal conglomerate of the Kootenai and lithostratigraphic correlation indicate the lower Kootenai Formation strata GR GF Formation, Cadomin Formation, McMurray are likely time equivalent to the northern units (Burden, 1984) (Fig. 2). Formation). (C) Albian sediment dispersal CM (i.e., Upper Kootenai Member of Great After the incision of the sub-Cretaceous unconformity, basin-axial river sys- Falls and Grande Cache, Notikewin Mem- tems routed sediment northwards from the United States to Alberta and Sas- PF ber, Grand Rapids Formation). katchewan (Leier and Gehrels, 2011; Benyon et al., 2014, 2016; Blum and Pecha, 2014). Diverse detrital zircon spectra for eastern-area samples from Great Falls (Cut Bank Member) and Cold Lake (McMurray Formation) are consistent with interpretations of broad drainage basins in the southwest and/or southeast JrE United States (Figs. 10 and 11). Detrital zircon populations in the western areas­ (Cadomin Formation and basal conglomerate of the Kootenai Formation) con-

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1199 Research Paper

tain abundant detrital zircons that ultimately derive from the Trans-Hudson dominated by Mesozoic detrital zircon grains, sharing the two populations Orogen but are attributed to recycling of uplifted pre-foreland basin strata in at 160 and 110 Ma with the Red Sandstone Member (Fig. 8B; samples AOS6, the Canadian segment of the fold-thrust belt (Leier and Gehrels, 2011). The dif- AOS16, AOS18). The zircon data are consistent with petrographic studies that ferences in the detrital zircon spectra between the western areas and the east- reported large proportions of volcanic and plutonic detritus preserved by early ern areas confirm that the basin was segregated into multiple paleodrainages, hydrocarbon charge at Cold Lake (Putnam and Pedskalny, 1983; Hutcheon presumably controlled by topography on the unconformity surface (Jackson, et al., 1989). The sandstone in the Cold Lake area con- 1984; Leckie and Smith; 1992; Leckie and Cheel, 1997; Wightman and Pember- tains angular to subangular grains, including abundant intact feldspar grains, ton, 1997; Benyon et al., 2014, 2016; Blum and Pecha, 2014). and volcanic rock fragments, which support first-cycle deposition (Putnam and Pedskalny, 1983). The detrital zircon spectra from the at Grande Cache Early Cretaceous–Aptian-Albian have a significant proportion of Jurassic detrital zircon grains; however, the population at ca. 1800 Ma and the diversity of other ages in the sample indicate Strata deposited near the Aptian–Albian boundary are less strongly con- the continued importance of uplifted pre-foreland basin sedimentary sources trolled by accommodation development in major valley systems as topogra- (Fig. 10). Distinction between zircon spectra from the west (i.e., Bluesky For- phy across much of the basin was infilled by this time (Jackson, 1984; Leckie mation) and east (i.e., Clearwater Formation) could be attributed to basin par- and Smith, 1992). A transgression of the basin at this time is consistent with titioning associated with the initial incursion of the Boreal Sea into the basin, the onset of increased tectonic subsidence, possibly associated with dynamic which may have segregated sediment sinks adjacent to the fold-thrust belt mantle subsidence and eustatic sea-level rise (Fig. 9) (Jackson, 1984; Mitrovica from those hundreds of kilometers to the east (cf. Somme et al., 2009; Blum et al., 1989; Leckie and Smith, 1992; Hayes et al., 1994), which is reflected in et al., 2013; Bhattacharya et al., 2016). This mode of basin partitioning could be subsidence models from western Montana (Fuentes et al., 2009). analogous to the documented effect of post-glacial sea-level rise on drainage Flooding of the foreland basin by the Boreal Sea is variably expressed in basins in the southern United States and Europe (Blum and Womack, 2009; the stratigraphic record (McLean and Wall, 1981; Jackson, 1984; Leckie and Maselli et al., 2011). Smith, 1992). At Great Falls, fluvial deposits with possible marine influence in the Red Sandstone Member underlie the lacustrine to restricted marine lime- stone of the Ostracod Member (Table 1). In Alberta, the Bluesky Formation Early Cretaceous–Middle Albian records deposition in a marine to estuarine setting (Jackson 1984; Smith et al., 1984; Hubbard et al., 2004; MacKay and Dalrymple, 2011). The Cold Lake area Middle Albian strata thicken westward owing to asymmetric accommoda- is characterized by a thick succession of deltaic and estuarine valley-fill facies tion development associated with flexural subsidence of the foreland basin in the Clearwater Formation (Putnam, 1982; Putnam and Pedskalny, 1983; (Fig. 9) (Hayes et al., 1994). The Upper Kootenai Formation is nearly twice as ­Hutcheon et al., 1989, Feldman et al., 2008; Maynard et al., 2010). The high thick at Gibson Reservoir relative to Great Falls. The Spirit River Formation at proportion of sandstone at Cold Lake was linked to the persistence of a major Grande Cache was deposited in a foredeep setting and is hundreds of meters continental river system by Blum and Pecha (2014) on the basis of pre-Meso- thick (Smith et al., 1984), whereas the equivalent units at Cold Lake are less zoic detrital zircon populations. than 100 m thick. The Red Sandstone is stratigraphically older than the other units consid- The Upper Kootenai Member is largely nonmarine at Great Falls and Gibson ered for this time period because of its position below the Ostracod Member, Reservoir (Table 1) (Fuentes et al., 2011). The Spirit River Formation at Grande which is generally considered to be time-equivalent to the Wabiskaw Member Cache consists of marine shoreline units at its base, which transition upwards and Bluesky Formation (Figs. 2 and 9). However, the MDA calculated for the into nonmarine deposits of the Falher and Notikewin Members (Quinn et al., Red Sandstone Member (108 ± 1.6 Ma) is younger than the Clearwater For- 2016). The stratigraphy at Cold Lake is characterized by shoreface and fluvial mation (115 ± 3.4 Ma) and the Bluesky Formation (114 ± 2.2 Ma). While these valley deposits of the Grand Rapids Formation (Maynard et al., 2010). data are conflicting, they generally support deposition of these units near the MDAs calculated for strata are consistent with deposition from the Aptian Aptian–Albian boundary. This analysis suggests that either the propagation boundary to the middle Albian (104 ± 1.8 Ma, Upper Kootenai Member, Great of uncertainty in these data sets or fine-scale correlations of this time period Falls; 110 ± 2.0 Ma, Notikewin Member, Grande Cache; 113 ± 2.4 Ma, Upper should be revisited (i.e., correlation of the Ostracod Member from Montana to ­Kootenai Member, Gibson Reservoir; 115 ± 1.0 Ma, Middle Grand Rapids,­ Alberta could be problematic). AOS12, Cold Lake). Within the uncertainty of the method, the diachro­neity According to stratigraphic position, the Red Sandstone Member is the old- of the strata cannot be ascertained. However, physical correlation in the sub­ est unit in this study that is dominated by detrital zircon grains from Mesozoic surface of Alberta has demonstrated an overall northward progradation of sources in the Cordillera (Figs. 9 and 10). The Clearwater Formation is also fluvial-deltaic­ depositional systems during this time interval (Smith et al., 1984).

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1200 Research Paper

Mesozoic zircons derived from the Cordillera were the dominant compo- fore, to preserve Mesozoic-dominated detrital zircon spectra in the foreland, nent of age spectra from Great Falls (Upper Kootenai Formation), Gibson Res- rivers that supplied sediment to the Western Interior basin from the magmatic ervoir (Upper Kootenai, 1FG70), and Grande Cache (Notikewin Member) areas hinterland would need to have traversed the fold-thrust belt without incor- by this time (Figs. 10 and 11). In southwestern Alberta, there are orogen-derived porating detrital zircon grains from pre-foreland sedimentary strata, which in gold-bearing igneous-clast conglomerates, consistent with the dominance aggregate yield diverse detrital zircon spectra (Fig. 7) (Ross and Villeneuve, of Cordillera-derived detritus (Leckie and Craw, 1995). In the Cold Lake area, 2003; Dickinson and Gehrels, 2009b; May et al., 2013; Gehrels and Pecha, 2014; the detrital zircon record in the Grand Rapids Formation exhibits an upward Golding et al., 2015). Aerial delivery of Early Cretaceous zircon to the basin is transition from dominance by Cordilleran magmatic arc zircon (Lower Grand possible or even probable; however, most of the zircon in these samples is Ju- Rapids, AOS20) to more-diverse age spectra (Upper Grand Rapids, AOS9). It is rassic in age, supporting an interpretation that fluvial sediment routing was an plausible that for these units the Cordilleran signature was diluted by an influx important process (Figs. 4 and 8). We suggest that these fluvial systems and/or of sediment from the south and east at this time (cf. Blum and Pecha, 2014). volcanic rocks aggraded a significant thickness of sediment on top of the oro- genic wedge, thereby shielding pre-foreland rocks from erosion (Fig. 12). DISCUSSION

The detrital zircon data from each of the four areas featured in this study A 5 km show a transition from diverse spectra (i.e., Group 1) to spectra dominated by Cordilleran magmatic rocks (i.e., Group 2). When the data are considered as a whole, Cordilleran magmatic sources of sediment dominate sediment supply for a period of several million years from the Aptian boundary to the middle 10 km Albian (Figs. 4, 8, and 10). Because changes in the detrital zircon record on the order of 10–100 Ma logically are linked to changes in the tectonic setting of

the basin (LaMaskin, 2012), we suggest that the change in zircon spectra ob- 5 km served in the data from the Western Interior foreland basin reflects the change B from a margin dominated by uplifted pre-foreland sedimentary strata and basin-axial­ sediment routing to a margin with extensive sediment aggradation on top of the orogenic wedge, similar to what is observed in the modern fore- 10 km land basin of the central Andes (Horton, 1998). The wedge-top depozone overlies the front of the orogenic wedge and is

both an important sink for sediments derived from the adjacent orogen and Foreland basin Volcanogenic wedge-top sediments and volcanic carapace a potential source of sediments as it undergoes deformation and denudation Mesozoic intrusions Late Proterozoic to Middle Jurassic sedimentary rocks Basement Mid-Proterozoic sedimentary rocks (DeCelles and Giles, 1996). Deformation in the underlying orogenic wedge controls the pattern of sedimentation in this zone. It has been suggested that C shortening in the wedge is associated with sediment bypass of the wedge- 1830 top zone because thickening the wedge leads to uplift and the destruction of accommodation (DeCelles, 1994; DeCelles and Giles, 1996; Horton, 1998). In 280 1050 n=222 active orogenic belts, the frontal thrusts of the orogen are blind and wedge-top 430 560 2700 sediments commonly blanket the deforming orogenic wedge (Ben Avraham and Emery, 1973; Vann et al., 1986; Horton and DeCelles; 1997; Horton, 1998). 200600 1000 1400 1800 2200 2600 3000 3400 These deposits can be kilometers thick and extend from tens to hundreds of Age (Ma) kilometers from the orogen (Burbank et al., 1997; Horton, 1998). Wedge-top de- posits have rarely been described from the northern Cordilleran foreland basin Figure 12. (A) Diagram showing fluvial systems incising into pre-foreland sedimentary strata that contain diverse detrital zircon spectra (modified after Price, 1994). (B) Diagram showing aggra- (e.g., McMechan et al., 2018). Presumably, the paucity of these strata has to do dation of volcanic rocks and volcanogenic sedimentary strata on top of the orogen, shielding with their preservation potential. This naturally raises the question of whether pre-foreland sedimentary strata from erosion (modified after Price, 1994). (C) Probability den- their former presence can be inferred from the detrital record in the basin. sity plot of pre-Mesozoic detrital zircon grains from Cordilleran magmatic arc-dominated sam- In North America, magmatic rocks generated during formation of the ples (Group 2) from Great Falls, Montana; Gibson Reservoir, Montana (Upper Kootenai, 1FG70) (Fuentes­ et al., 2011); Grande Cache, Alberta (Notikewin Member) (Quinn et al., 2016); and Cold Cordillera are generally preserved to the west of the fold-thrust belt, which Lake, Alberta (lower Clearwater, AOS6; middle Clearwater, AOS16; upper Clearwater, AOS18; is largely composed of pre-foreland sedimentary strata (Figs. 1 and 11). There- Lower Grand Rapids, AOS20) (Blum and Pecha, 2014). Error is incorporated at the 2-sigma level.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1201 Research Paper

The deposition and removal of wedge-top strata have been shown to sig- The orogenic influence on foreland basin strata is well established (e.g., nificantly impact landscape evolution and sediment dispersal. Osborn et al. Heller and Paola, 1989; Ross et al., 2005; DeCelles et al., 2009; Horton, 2018). Re- (2006) hypothesized that the Canadian Rocky Mountains as a landscape in its cent research has hypothesized that the rheology of foreland basin deposits can present form did not emerge until ~30 m.y. after the end of orogenesis because control the structural evolution of the fold-thrust belt (Chapman and DeCelles, a thick sequence (~2–20 km) of Mesozoic wedge-top strata covered pre-Meso- 2015) and that changes in sediment yield may be linked to deformation in the zoic sedimentary rocks. Ross et al. (2005) presented detrital zircon data from interior of the orogen (Wipple, 2009). Therefore, the recognition of sediment that first demonstrated the transition from Mesozoic-domi- aggradation on top of orogenic structures is essential to testing hypotheses of nated detrital zircon spectra to more-diverse spectra in the Paleocene, attrib- orogen–foreland basin interaction. In the case explored here, an episode of sub- uting this pattern to uncovering of the orogenic wedge. The new data, pre- stantial sedimentation on the orogenic wedge during the Albian hypothetically sented herein, establish that this transition was synchronous across the basin would correspond to thrust movements in the orogen. In combination with to within a few millions of years and likely tectonically controlled. late Albian–early Cenomanian cooling ages from thrust faults in the Canadian Mesozoic-dominated zircon spectra from the Western Interior basin do Rockies reported by Pana and Van der Pluijm (2015), the connection between contain pre-Mesozoic zircon populations, which are characteristic of pre-Cor- wedge-top dynamics and foreland basin strata warrants further investigation. In dilleran foreland strata of North America (Fig. 12). Blum and Pecha (2014) this study, we provide key indirect evidence for significant Jurassic–Early Cre- use the pre-Mesozoic zircon populations of the Clearwater and Grand Rapids taceous wedge-top deposits in the Northern Cordillera foreland basin system. Formations to argue for the longevity of a proposed continental river sys- tem. However, it is unclear whether across the Western Interior basin these pre-Mesozoic grains represent input via recycling from early foreland basin CONCLUSIONS strata (Late Jurassic–Aptian) in the wedge-top; pre-foreland basin strata up- lifted in the orogenic wedge; sedimentary strata from Appalachia, the south- Detrital zircon data from Upper Jurassic–Lower Cretaceous strata of Great western United States, or the accreted terranes; or some combination of Falls, Montana, can be segregated into three groups using multidimensional these sources. scaling. Upper Jurassic–Aptian strata exhibit diverse detrital zircon spectra Detrital zircon spectra from Great Falls, Cold Lake, and Grande Cache diver- (Group 1) interpreted to record sediment recycling from pre-foreland strata sify up stratigraphic section after a long period of being dominated by Meso­ of the southwest United States and the Cordilleran margin of North Amer- zoic Cordilleran sources (Fig. 8). The reincorporation of many detrital com- ica. Near the Aptian–Albian boundary, the detrital zircon provenance shifts to ponents that were characteristic of earlier Mesozoic strata suggests that the spectra dominated by Mesozoic grains originating from magmatic rocks of the wedge-top may have become an important sediment source in the middle to Cordillera (Group 2). In the Albian, the uppermost unit of this study is char- late Albian (cf. Ross et al., 2005); however, contributions from the other poten- acterized by zircon spectra less dominated by Cordilleran magmatic sources tial sources previously discussed cannot be ruled out by these data. Neverthe- (Group 3), recording a diversification of the provenance area to include recy- less, we speculate that the record of burial and exhumation of the wedge-top cling of sedimentary strata. depozone should be more widespread in the detrital record of foreland basins Data from Great Falls are compared to data from Gibson Reservoir in west- than has been reported to date. ern Montana, Grande Cache in west-, and Cold Lake in east-cen- Numerical and physical modeling studies have demonstrated that sedi- tral Alberta in order to evaluate the evolution of sediment routing systems to mentation is a key variable in the structural evolution of a mountain belt be- the foreland basin. This comparison shows that Late Jurassic sediment rout- cause surficial processes are an important control on the taper angle of the ing was dominated by basin-axial transport systems. Detrital zircon data con- orogen (Storti and McClay, 1995; Lageson et al., 2001; Simpson, 2010; Buiter, firm that in the Aptian, sediment-dispersal patterns were strongly controlled 2012). The rate of sediment generation in an orogen is controlled by elevation by basin partitioning linked to topography created during formation of the and climate, which are to some degree tectonically controlled and linked (e.g., sub-Cretaceous unconformity. Near the Aptian–Albian boundary, Boreal Sea rain-shadow effects, glaciation). It also has been hypothesized that the addition transgression potentially led to the east-west segregation of the basin. of volcanic rocks to the wedge can alter the taper state of the orogenic wedge The data from each of the areas show an evolution from diverse spectra (Lageson et al., 2001). Therefore, deciphering the record of aggradation in the to Cordilleran magmatic-dominated spectra. The dominance of Cordilleran wedge-top in ancient orogen-basin systems allows for conjecture about tec- magmatic sources plausibly records covering of the orogenic wedge by the tonic and structural evolution of the orogen as well as inferences concerning aggradation of sediments and volcanic strata on top of the orogenic wedge, episodes of increased tectonic activity or tectonic quiescence. Some attempts insulating pre-foreland strata with diverse detrital zircon spectra from being re- to decipher this record exist for the Cordilleran Orogen of North America; how- worked into the basin. The recognition of this episode of Cordilleran–foreland ever, these studies focus on time periods after the Early Cretaceous (DeCelles, basin evolution has implications for the application of orogenic models to the 1994; DeCelles and Mitra, 1995; Lageson et al., 2001). geological record.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1202 Research Paper

ACKNOWLEDGMENTS Controls on Deep Water Depositional Systems: Climate, Sea-Level, and Sediment Flux: Funding for this research was generously provided by Osum Corp (master’s thesis SEPM Special Publication, 92, p. 15–39, https://​doi​.org​/10​.2110​/sepmsp​.092​.015​. of G. Quinn), the University of Calgary, and an NSERC Discovery Grant (RG-PIN/341715-201) to Brenner, R.L., and Davies, D.K., 1974, Oxfordian sedimentation in western interior United States: S. Hubbard. Reviews by Marc Hendrix and an anonymous scientist improved the content and The American Association of Geologists Bulletin, v. 58, no. 3, p. 407–428. clarity of the manuscript and are greatly appreciated. Buiter, S.J.H., 2012, A review of brittle compressional wedge models: Tectonophysics, v. 530– 531, p. 1–17, https://​doi​.org​/10​.1016​/j​.tecto​.2011​.12​.018​. Burbank, D.W., Beck, R.A., and Mulder, T., 1997, The Himalayan foreland basin, in Yin, A., and Harrison, M., eds., The tectonic evolution of Asia: Cambridge, UK, Cambridge University REFERENCES CITED Press, p. 149–188. Alberta Geological Survey, 2015, Alberta Table of Formations: Alberta Energy Regulator: https://​ Burden, E.T., 1984, Terrestrial palynomorph biostratigraphy of the lower part of the Mannville www​.aer​.ca​/documents​/catalog​/TOF​.pdf (accessed December 2015). Group (Lower Cretaceous), Alberta and Montana, in Stott, D.F., and Glass, D.J., eds., The Anderson, J.L., and Bender, E.E., 1989, Nature and origin of Proterozoic A-type granitic magma- Mesozoic of Middle North America: Canadian Society of Petroleum Geologists Memoir 9, tism in the southwestern United States of America: Lithos, v. 23, p. 19–52, https://​doi​.org​/10​ p. 249–269. .1016​/0024​-4937​(89)90021​-2​. Cannon, J.L., 1966, Outcrop examination and interpretation of paleocurrent pattern of the Black- Archibald, D.A., Glover, J.K., Price, R.A., Farrarr, E., and Carmichael, D.M., 1983, Geochronology leaf Formation near Great Falls, Montana, in Cox, J.E., ed., 17th Annual Field Conference: and tectonic implications of magmatism and metamorphism, southern Kootenay Arc and Billings, Montana, Billings Geological Society, p. 71–111. neighboring regions, southeastern British Columbia, Part I: Jurassic to mid-Cretaceous: Ca- Cant, D.J., and Stockmal, G.S., 1989, The Alberta foreland basin: Relationship between stratig- nadian Journal of Earth Sciences, v. 20, p. 1891–1913, https://​doi​.org​/10​.1139​/e83​-178​. raphy and Cordilleran terrane-accretion events: Canadian Journal of Earth Sciences, v. 26, Ardies, G.W., Dalrymple, R.W., and Zaitlin, B.A., 2002, Controls on the geometry of incised val- p. 1964–1975, https://​doi​.org​/10​.1139​/e89​-166​. leys in the Basal Quartz unit (Lower Cretaceous), Western Canada Sedimentary Basin: Jour- Card, K.D., 1990, A review of the Superior Province of the , a product of nal of Sedimentary Research, v. 72, p. 602–618, https://​doi​.org​/10​.1306​/032101720602​. ­Archean accretion: Precambrian Research, v. 48, p. 99–156, https://​doi​.org​/10​.1016​/0301​-9268​ Armstrong, R.L., 1988, Mesozoic and early Cenozoic magmatic evolution of the Canadian Cor­ (90)90059​-Y​. dillera, in Clark, S.P., Burchfiel, B.C., and Suppe, J., eds., Processes in Continental Litho- Chapman, J.B., and DeCelles, P.G., 2015, Foreland basin stratigraphic control on thrust belt evo- spheric Deformation: Geological Society of America Special Paper 218, p. 55–91, https://​doi​ lution: Geology, v. 43, p. 579–582, https://​doi​.org​/10​.1130​/G36597​.1​. .org​/10​.1130​/SPE218​-p55​. Chase, C.G., Sussman, A.J., and Coblentz, D.D., 2009, Curved Andes: Geoid, forebulge, and flex- Ballard, W.W., 1966, Petrography of the Jurassic and Cretaceous sandstone, north flank Little ure: Lithosphere, v. 1, p. 358–363, https://​doi​.org​/10​.1130​/L67​.1​. Belt Mountains, Montana, in Cox, J.E., ed., 17th Annual Field Conference: Billings, Montana, Christopher, J.E., 2003, Jura–Cretaceous Success Formation and Lower Cretaceous Mannville Billings Geological Society, p. 56–70. Group of : Saskatchewan Industry and Resources, Saskatchewan Geological Barth, A.P., and Wooden, J.L., 2006, Timing of magmatism following initial convergence at a Survey, Report 223, 352 p. passive margin, southwestern U.S. Cordillera, and ages of lower crustal magma sources: Cobban, W.A., 1945, Marine Jurassic formations of the Sweetgrass Arch, Montana: Bulletin of The Journal of Geology, v. 114, p. 231–245, https://​doi​.org​/10​.1086​/499573​. the American Association of Petroleum Geologists, v. 29, no. 9, p. 1262–1303. Becker, T.P., Thomas, W.A., Samson, S.D., and Gehrels, G.E., 2005, Detrital zircon evidence of Coogan, J.C., 1992, Structural evolution of piggyback basins in the Wyoming–Idaho–Utah thrust Laurentian crustal dominance in the lower Pennsylvanian deposits of the Alleghanian clastic belt, in Link, P.K., Kuntz, M.A. and Piatt, L.B., eds., Regional Geology of Eastern Idaho and wedge in eastern North America: Sedimentary Geology, v. 182, p. 59–86, https://doi​ ​.org​/10​ Western Wyoming: Geological Society of America Memoir 179, p. 55–81. .1016​/j​.sedgeo​.2005​.07​.014​. DeCelles, P.G., 1994, Late Cretaceous–Pleistocene synorogenic sedimentation and kinematic Ben-Avraham, Z., and Emery, K.O., 1973, Structural framework of Sunda Shelf: The American history of the Sevier thrust belt, northeast Utah and southwest Wyoming: Geological So- Association of Petroleum Geologists Bulletin, v. 57, no. 12, p. 2323–2366. ciety of America Bulletin, v. 106, p. 32–56, https://​doi​.org​/10​.1130​/0016​-7606​(1994)106​<0032:​ Benyon, C., Leier, A., Leckie, D.A., Webb, A., Hubbard, S.M., and Gehrels, G., 2014, Provenance LCPSSA>2​.3​.CO;2​. of the Cretaceous , Canada: Implications for Continental-Scale Sedi- DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland ment Transport: Journal of Sedimentary Research, v. 84, p. 136–143, https://​doi​.org​/10​.2110​ basin system, Western U.S.A: American Journal of Science, v. 304, p. 105–168, https://doi​ ​ /jsr​.2014​.16​. .org​/10​.2475​/ajs​.304​.2​.105​. Benyon, C., Leier, A.L., Leckie, D.A., Hubbard, S.M., and Gehrels, G.E., 2016, Sandstone prove- DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems: Basin Research, v. 8, p. 105–123, nance and insights into the paleogeography of the McMurray Formation from detrital zir- https://​doi​.org​/10​.1046​/j​.1365​-2117​.1996​.01491​.x​. con geochronology, Athabasca Oil Sands, Canada: The American Association of Petroleum DeCelles, P.G., and Mitra, G., 1995, History of the Sevier orogenic wedge in terms of critical taper Geol­ogists Bulletin, v. 100, no. 2, p. 269–287, https://​doi​.org​/10​.1306​/10191515029​. models, northeast Utah and southwest Wyoming: Geological Society of America Bulletin, Bhattacharya, J.P., Copeland, P., Lawton, T.F., and Holbrook, J., 2016, Estimation of source area, v. 107, no. 4, p. 454–462, https://​doi​.org​/10​.1130​/0016​-7606​(1995)107%3C0454:​HOTSOW%3E2​ river paleodischarge, paleoslope, and sediment budgets of linked deep-time depositional .3​.CO;2​. systems and implications for hydrocarbon potential: Earth-Science Reviews, v. 153, p. 77– DeCelles, P.G., Ducea, M.N., Kapp, P., and Zandt, G., 2009, Cyclicity in Cordilleran orogenic sys- 110, https://​doi​.org​/10​.1016​/j​.earscirev​.2015​.10​.013​. tems: Nature Geoscience, v. 2, p. 251–257, https://​doi​.org​/10​.1038​/NGEO469​. Black, L.P., Kamo, S.L., Allen, C.M., Davis, D.W., Aleinikoff, J.N., Valley, J.W., Mundil, R., Camp- Dickinson, W.R., 2008, Impact of differential zircon fertility of granitoid basement rocks in North bell, I.H., Korsch, R.J., Williams, I.S., and Foudoulis, C., 2004, Improved 206Pb/238U micro- America on age populations of detrital zircons and implications for granite petrogenesis: Earth probe geochronology by the monitoring of trace-element-related matrix effects; SHRIMP, and Planetary Science Letters, v. 275, p. 80–92, https://doi​ .org​ /10​ .1016​ /j​ .epsl​ .2008​ .08​ .003​ .​ ID-TIMS, ELA-ICP-MS and oxygen isotope documentation for a series of zircon standards: Dickinson, W.R., and Gehrels, G.E., 2008a, Sediment delivery to the Cordilleran foreland basin: Chemical Geology, v. 205, p. 155–170, https://​doi​.org​/10​.1016​/j​.chemgeo​.2004​.01​.003​. Insights from U-Pb ages of detrital zircons in Upper Jurassic and Cretaceous strata of the Blum, M., and Pecha, M., 2014, Mid-Cretaceous to Paleocene North American drainage reorgani- Colorado Plateau: American Journal of Science, v. 308, p. 1041–1082. zation from detrital zircons: Geology, v. 42, no. 7, p. 607–610, https://doi​ .org​ /10​ .1130​ /G35513​ .1​ ​. Dickinson, W.R., and Gehrels, G.E., 2008b, U-Pb ages of detrital zircons in relation to paleogeog- Blum, M., Martin, J., Milliken, K., and Garvin, M., 2013, Paleovalley systems: Insights from Qua- raphy: Triassic paleodrainage networks and sediment dispersal across southwest Laurentia: ternary analogs and experiments: Earth-Science Reviews, v. 116, p. 128–169, https://​doi​.org​ Journal of Sedimentary Research, v. 78, p. 745–764, https://​doi​.org​/10​.2110​/jsr​.2008​.088​. /10​.1016​/j​.earscirev​.2012​.09​.003​. Dickinson, W.R., and Gehrels, G.E., 2009a, Use of U–Pb ages of detrital zircons to infer maximum Blum, M.D., and Womack, J.H., 2009, Climate change, sea-level change, and fluvial sediment depositional ages of strata: A test against a Colorado Plateau Mesozoic database: Earth and supply to deepwater systems, in Kneller, B., Martinsen, O.J., McCaffery, B., eds., External Planetary Science Letters, v. 228, p. 115–125, https://​doi​.org​/10​.1016​/j​.epsl​.2009​.09​.013​.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1203 Research Paper

Dickinson, W.R., and Gehrels, G.E., 2009b, U-Pb ages of detrital zircons in Jurassic eolian and Gillespie, J.M., and Heller, P.L., 1995, Beginning of foreland subsidence in the Columbian–Sevier associated sandstones of the Colorado Plateau: Evidence for transcontinental dispersal and belts, southern Canada and northwest Montana: Geology, v. 23, p. 723–726, https://doi​ ​.org​ intraregional recycling of sediment: Geological Society of America Bulletin, v. 121, p. 408– /10​.1130​/0091​-7613​(1995)023​<0723:​BOFSIT>2​.3​.CO;2​. 433, https://​doi​.org​/10​.1130​/B26406​.1​. Glaister, R.P., 1959, Lower Cretaceous of southern Alberta and adjoining areas: Bulletin of the Ducea, M., 2001, The California Arc: Thick granitic batholiths, eclogitic residues, lithospheric-­ American Association of Petroleum Geologists, v. 43, no. 3, p. 590–640. scale thrusting, and magmatic flare-ups: GSA Today, v. 11, no. 11, p. 4–10. Golding, M.L., Mortensen, J.K., Ferri, F., Zonneveld, J.-P., and Orchard, M.J., 2015, Determining Easton, R.M., 1986, Geochronology of the Grenville Province, in Moore, J.M., Davidson, A., and the provenance of Triassic sedimentary rocks in northeastern British Columbia and western Baer, A.J., eds., The Grenville Province: Ottawa, Ontario, Geological Association of Canada, Alberta using detrital zircon geochronology, with implications for regional tectonics: Ca- Special Paper 31, p. 127–173. nadian Journal of Earth Sciences, v. 53, p. 140–155, https://​doi​.org​/10​.1139​/cjes​-2015​-0082​. Eriksson, K.A., Campbell, I.H., Palin, J.M., Allen, C.M., and Bock, B., 2004, Evidence for multiple Hamblin, A.P., and Walker, R.G., 1979, Storm-dominated shallow marine deposits: The Fernie– recycling in Neoproterozoic through Pennsylvanian sedimentary rocks of the Central Appa­ Kootenay (Jurassic) transition, southern Rocky Mountains: Canadian Journal of Earth Sci- lachian Basin: The Journal of Geology, v. 112, p. 261–276, https://​doi​.org​/10​.1086​/382758​. ences, v. 16, p. 1673–1690. Farshori, M.Z., and Hopkins, J.C., 1989, Sedimentology and petroleum geology of fluvial and Harms, T.A., Brady, J.B., Burger, H.R., and Cheney, J.T., 2004, Advances in the geology of the shoreline deposits of the Lower Cretaceous Sunburst Sandstone Member, , Tobacco Root Mountains, Montana, and their implications for the history of the northern southern Alberta: Bulletin of Canadian Petroleum Geology, v. 37, no. 4, p. 371–388. Wyoming province, in Brady, J.B., Burger, H.R., Cheney, J.T., and Harms, T.A., eds., Precam- Feldman, H.R., McCrimmon, G.G., and De Freitas, T.A., 2008, Fluvial to estuarine valley-fill mod- brian Geology of the Tobacco Root Mountains, Montana: Geological Society of America els without age-equivalent sandy shoreline deposits, based on the Clearwater Formation Special Paper 377, p. 227–243. (Cretaceous) at Cold Lake, Alberta, Canada, in Hampson, G.J., Steel, R.J., Burgess, P.M., and Harris, W.L., 1966, The stratigraphy of the Upper Jurassic–Lower Cretaceous rocks in the Great Dalrymple, R.W., eds., Recent Advances in Models of Shallow Marine Stratigraphy: SEPM Falls–Lewistown coal field, central Montana,in Cox, J.E., ed., 17th Annual Field Conference: Special Publication 90, p. 443–472. Billings, Montana, Billings Geological Society, p. 164–177. Finger, K.L., 1983, Observations of the Lower Cretaceous Ostracode Zone of Alberta: Bulletin of Hartley, A.J., Owen, A., Swan, A., Weissmann, G.S., Holzweber, B.I., Howell, J., Nichols, G., and Canadian Petroleum Geology, v. 31, no. 4, p. 326–337. Scuderi, L., 2015, Recognition and importance of amalgamated sandy meander belts in the Folinsbee, R.E., Ritchie, W.D., and Stansberry, G.F., 1957, The Crowsnest Volcanics and Creta- continental rock record: Geology, v. 43, p. 679–682, https://​doi​.org​/10​.1130​/G36743​.1​. ceous Geochronology, in Taylor, N.W., ed., Guide Book, Seventh Annual Field Conference Hayes, B.J.R., 1983, Stratigraphy and petroleum potential of the Swift Formation (Upper Juras- Waterton: Alberta Society of Petroleum Geologists, p. 20–26. sic), southern Alberta and north-central Montana: Bulletin of Canadian Petroleum Geology, Fox, R.D., and Groff, S.L., 1966, Stratigraphic and structural investigations of the Cascade–Ulm v. 31, p. 37–52. area, in Cox, J.E., ed., 17th Annual Field Conference: Billings, Montana, Billings Geological Hayes, B.J.R., 1986, Stratigraphy of the basal Cretaceous Lower Mannville Formation, Southern Society, p. 36–55. Alberta and north-central Montana: Bulletin of Canadian Petroleum Geology, v. 34, p. 30–46. Frisch, W., Kuhlemann, J., Dunkl, I., and Szekely, B., 2001, The Dachstein paleosurface and the Hayes, B.J.R., 1990, A perspective on the Sunburst Member of southern Alberta: Bulletin of Augenstein Formation in the northern calcareous Alps: A mosaic stone in the geomorpho- Canadian Petroleum Geology, v. 38, p. 483–484. logical evolution of the eastern Alps: International Journal of Earth Sciences, v. 90, p. 500– Hayes, B.J.R., Christopher, J.E., Rosenthal, L., Los, G., McKercher, B., Minken, D., Tremblay, Y.M., 518, https://​doi​.org​/10​.1007​/s005310000189​. and Fennell, J., 1994, Cretaceous Mannville Group of the Western Canada Sedimentary Ba- Fuentes, F., DeCelles, P.G., and Gehrels, G.E., 2009, Jurassic onset of foreland basin deposition in sin, in Mossop, G.D., and Shetsen, I., compilers, Geological Atlas of the Western Canada northwestern Montana, USA: Implications for along-strike synchronicity of the Cordilleran Sedimentary Basin: , Alberta, Canadian Society of Petroleum Geologists and Al- orogenic activity: Geology, v. 37, p. 379–382, https://​doi​.org​/10​.1130​/G25557A​. berta Research Council, p. 317–334. Fuentes, F., DeCelles, P.G., Constenius, K.N., and Gehrels, G.E., 2011, Evolution of the foreland ba- Heller, P.L., and Paola, C., 1989, The paradox of Lower Cretaceous gravels and the initiation sin system in northwestern Montana, U.S.A: Geological Society of America Bulletin, v. 123, of thrusting in the Sevier orogenic belt, United Sates Western Interior: Geological Society p. 507–533, https://​doi​.org​/10​.1130​/B30204​.1​. of America Bulletin, v. 101, p. 864–875, https://​doi​.org​/10​.1130​/0016​-7606​(1989)101​<0864:​ Garrity, C.P., and Soller, D.R., 2009, Database of the Geologic Map of North America: Adapted TPOLCG>2​.3​.CO;2​. from the map by J.C. Reed, Jr., and others (2005): U.S. Geological Survey Data Series 424, Hoffman, P.F., 1988, United plates of America, the birth of a craton: Early Proterozoic assembly https://​pubs​.usgs​.gov​/ds​/424/ (May 2017). and growth of Laurentia: Annual Review of Earth and Planetary Sciences, v. 16, p. 543–603, Gaschnig, R.M., Vervoort, J.D., Lewis, R.S., and McClelland, W.C., 2009, Migrating magmatism https://​doi​.org​/10​.1146​/annurev​.ea​.16​.050188​.002551​. in the northern US Cordillera: In situ U-Pb geochronology of the Idaho Batholith: Contri- Hopkins, J.C., 1985, Channel-fill deposits formed by aggradation in deeply scoured, superim- butions to Mineralogy and Petrology, v. 159, p. 863–883, https://​doi​.org​/10​.1007​/s00410​-009​ posed distributaries of the Lower Kootenai Formation (Cretaceous): Journal of Sedimentary -0459​-5​. Petrology, v. 55, p. 42–52. Gaschnig, R.M., Macho, A.S., Fayon, A., Schmitz, M., Ware, B.D., Vervoort, J.D., Kelso, P., Horstwood, M.S.A., Kosler, J., Gehrels, G., Jackson, S.E., McLean, N.M., Paton, C., Pearson, N.J., ­LaMaskin, T.A., Kahn, M.J., and Tikoff, B., 2017a, Intrusive and depositional constraints on Sircombe, K., Sylvester, P., Vermeesch, P., Bowring, J.F., Conden, D.J., and Schoene, B., 2016, the Cretaceous tectonic history of the southern Blue Mountains, eastern Oregon: Litho- Community-derived standards for LA-ICP-MS U-(Th-)Pb geochronology: Uncertainty prop- sphere, v. 9, no. 2, p. 265–282, https://​doi​.org​/10​.1130​/L554​.1​. agation, age interpretation and data reporting: Geostandards and Geoanalytical Research, Gaschnig, R.M., Vervoort, J.D., Tikoff, B., and Lewis, R.S., 2017b, Construction and preservation v. 40, p. 311–332, https://​doi​.org​/10​.1111​/j​.1751​-908X​.2016​.00379​.x​. of batholiths in the northern U.S. Cordillera: Lithosphere, v. 9, no. 2, p. 315–324, https://doi​ ​ Horton, B.K., 1998, Sediment accumulation on top of the Andean orogenic wedge: Oligocene to late .org​/10​.1130​/L497​.1​. Miocene basins of the Eastern Cordillera, southern Bolivia: Geological Society of America Bul- Gehrels, G., and Pecha, M., 2014, Detrital zircon U-Pb geochronology and Hf isotope geochem- letin, v. 110, p. 1174–1192, https://doi​ .org​ /10​ .1130​ /0016​ -7606​ (1998)110​ <1174:​ SAOTOT>2​ .3​ .CO;2​ .​ istry of Paleozoic and Triassic passive margin strata of western North America: Geosphere, Horton, B.K., 2018, Sedimentary record of Andean mountain building: Earth-Science Reviews, v. 10, p. 49–65, https://​doi​.org​/10​.1130​/GES00889​.1​. v. 178, p. 270–309, https://​doi​.org​/10​.1016​/j​.earscirev​.2017​.11​.025​. Gehrels, G., Rushmore, M., Woodsworth, G., Crawford, M., Andronicos, C., Hollister, L., Patchett,­ Horton, B.K., and DeCelles, P.G., 1997, The modern foreland basin system adjacent to the Cen- J., Ducea, M., Butler, R., Klepeis, K., Davidson, C., Friedman, R., Haggart, J., Mahoney, B., tral Andes: Geology, v. 25, p. 895–898, https://​doi​.org​/10​.1130​/0091​-7613​(1997)025​<0895:​ Crawford, W., Pearson, D., and Girardi, J., 2009, U-Th-Pb geochronology of the Coast Moun- TMFBSA>2​.3​.CO;2​. tains batholith in north-coastal British Columbia: Constraints on age and tectonic evolu- Hubbard, S.M., Gingras, M.K., and Pemberton, G., 2004, Paleoenvironmental implications of tion: Geological Society of America Bulletin, v. 121, p. 1341–1361, https://doi​ ​.org​/10​.1130​ trace fossils in estuarine deposits of the Cretaceous Bluesky Formation, Cadotte region, /B26404​.1​. Alberta, Canada: Fossils and Strata, v. 51, p. 68–87.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1204 Research Paper

Hutcheon, I., Abercrombi, H.J., Putnam, P., Gardner, R., and Krouse, H.R., 1989, Diagenesis and stratigraphic evolution and paleogeography: Geological Society of America Bulletin, v. 125, sedimentology of the Clearwater Formation at Tucker Lake: Bulletin of Canadian Petroleum p. 1403–1422, https://​doi​.org​/10​.1130​/B30824​.1​. Geology, v. 37, p. 83–97. Maynard, J.R., Feldman, H.R., and Alway, R., 2010, From bars to valleys: The sedimentology Hyndman, D.W., 1983, The Idaho batholith and associated plutons, Idaho and Western Montana, and seismic geomorphology of fluvial to estuarine incised-valley fills of the Grand Rapids in Roddick, J.A., ed., Circum-Pacific Plutonic Terranes: Geological Society of America Mem- Formation (Lower Cretaceous), Iron River field, Alberta, Canada: Journal of Sedimentary oir 159, p. 213–240. Research, v. 80, p. 611–638, https://​doi​.org​/10​.2110​/jsr​.2010​.060​. Jackson, P.C., 1984, Paleogeography of the Lower Cretaceous Mannville Group of Western Can- McLean, J.R., 1977, The Cadomin Formation: Stratigraphy, sedimentology, and tectonic implica- ada, in Masters, J.A., ed., Elmworth: Case Study of a Deep Basin Gas Field: American Asso- tions: Bulletin of Canadian Petroleum Geology, v. 25, p. 792–827. ciation of Petroleum Geologists Memoir 38, p. 49–77. McLean, J.R., and Wall, J.H., 1981, The Early Cretaceous Moosebar Sea in Alberta: Bulletin of Kukulski, R.B., Hubbard, S.M., Moslow, T.F., and Raines, M.K., 2013, Basin-scale stratigraphic archi- Canadian Petroleum Geology, v. 29, p. 334–337. tecture of upstream fluvial deposits: Jurassic–Cretaceous foredeep, Alberta Basin, Canada: McMechan, M.E., Currie, L.D., Matthews, W.A., Sweet, A.R., and Reyes, J., 2018, Cretaceous Journal of Sedimentary Research, v. 83, no. 8, p. 704–722, https://doi​ .org​ /10​ .2110​ /jsr​ .2013​ .53​ .​ strata at the west edge of the Canadian Rocky Mountains: A piggyback basin remnant of Lageson, D.R., Schmitt, J.G., Horton, B.K., Kalakay, T.J., and Burton, B.R., 2001, Influence of Late the western Canada foreland basin: Geological Society of America Bulletin, https://​doi​.org​ Cretaceous magmatism on the Sevier orogenic wedge, western Montana: Geology, v. 29, /10​.1130​/B31821​.1 ​(in press)​. p. 723–726, https://​doi​.org​/10​.1130​/0091​-7613​(2001)029​<0723:​IOLCMO>2​.0​.CO;2​. Miles, B.D., Kukulski, R.B., Raines, M.K., Zonneveld, J.-P., Leier, A.L., and Hubbard, S.M., 2012, LaMaskin, T.A., 2012, Detrital zircon facies of Cordilleran terranes in western North America: GSA A stratigraphic framework for Late Jurassic–Early Cretaceous gas-bearing strata (Monteith Today, v. 22, p. 4–11, https://​doi​.org​/10​.1130​/GSATG142A​.1​. Formation) in the subsurface of northwestern Alberta: Bulletin of Canadian Petroleum Geol- LaMaskin, T.A., Vervoort, J.D., Dorsey, R.J., and Wright, J.E., 2011, Early Mesozoic paleogeog- ogy, v. 60, p. 3–36, https://​doi​.org​/10​.2113​/gscpgbull​.60​.1​.3​. raphy and tectonic evolution of the western United States: Insights from detrital zircon Mitrovica, J.X., Beaumont, C., and Jarvis, G.T., 1989, Tilting of continental interiors by the dy- U-Pb geochronology, Blue Mountains Province, northeastern Oregon: Geological Society of namical effects of subduction: Tectonics, v. 8, no. 5, p. 1079–1094, https://doi​ ​.org​/10​.1029​ America Bulletin, v. 123, p. 1939–1965, https://​doi​.org​/10​.1130​/B30260​.1​. /TC008i005p01079​. LaMaskin, T.A., Dorsey, R.J., Vervoort, J.D., Schmitz, M.D., Tumpane, K.P., and Moore, N.O., 2015, Monger, J.W.H., Price, R.A., and Tempelman-Kluit, D.J., 1982, Tectonic accretion and the origin Westward growth of Laurentia by pre-Late Jurassic terrane accretion, eastern Oregon and of the two major metamorphic and plutonic welts in the Canadian Cordillera: Geology, v. 10, western Idaho, United States: The Journal of Geology, v. 123, p. 233–267, https://doi​ ​.org​/10​ p. 70–75, https://​doi​.org​/10​.1130​/0091​-7613​(1982)10​<70:​TAATOO>2​.0​.CO;2​. .1086​/681724​. Mortensen, J.K., and Card, K.D., 1993, U-Pb age constraints for the magmatic and tectonic Laskowski, A.K., DeCelles, P.G., and Gehrels, G.E., 2013, Detrital zircon geochronology of Cor­di­ evolution of the Pontiac Subprovince, Quebec: Canadian Journal of Earth Sciences, v. 30, lleran retroarc foreland basin strata, western North America: Tectonics, v. 32, p. 1027–1048, p. 1970–1980, https://​doi​.org​/10​.1139​/e93​-173​. https://​doi​.org​/10​.1002​/tect​.20065​. Mudge, M.R., and Rice, D.D., 1982, Lower Cretaceous Mount Pablo Formation, northwestern Lawton, T.F., Hunt, G.J., and Gehrels, G.E., 2010, Detrital zircon record of thrust belt unroofing in Montana: U.S. Geological Survey Bulletin 1502-D, 19 p., http://pubs​ ​.usgs​.gov​/bul​/1502d​ Lower Cretaceous synorogenic conglomerates, central Utah: Geology, v. 38, no. 5, p. 463– /report​.pdf. 466, https://​doi​.org​/10​.1130​/G30684​.1​. Oakes, M.H., 1966, North Cut Bank field and the Moulton Sandstone,in Cox, J.E., ed., 17th An- Leckie, D.A., and Cheel, R.J., 1997, Sedimentology and depositional history of Lower Cretaceous nual Field Conference: Billings, Montana, Billings Geological Society, p. 191–200. coarse-grained clastics, southwest Alberta and southeast British Columbia: Canadian Bulle- Osborn, G., Stockmal, G., and Hapsel, R., 2006, Emergence of the and adjacent tin of Petroleum Geology, v. 45, p. 1–24. plains: A comparison of physiography between end-of-Laramide time and the present day: Leckie, D.A., and Craw, D., 1995, Westerly derived Early Cretaceous gold paleoplacers in the Geomorphology, v. 75, p. 450–477, https://​doi​.org​/10​.1016​/j​.geomorph​.2005​.07​.032​. western Canada foreland basin, southwestern Alberta: Tectonic and economic implications: Owen, A., Nichols, G.J., Hartley, A.J., Weismann, G.S., and Senderi, L.A., 2015, Quantification Canadian Journal of Earth Sciences, v. 32, p. 1079–1092, https://​doi​.org​/10​.1139​/e95​-090​. of a distributive fluvial system: The Salt Wash DFS of the Morrison Formation, SW USA: Leckie, D.A., and Smith, D.G., 1992, Regional setting, evolution, and depositional cycles of the Journal of Sedimentary Research, v. 85, p. 544–561. western Canada foreland basin, in Macqueen, R.W., and Leckie, D.A., eds., Foreland Basins Paces, J.B., and Miller, J.D., 1993, Precise U-Pb ages of Duluth Complex and related mafic intru- and Fold Belts: American Association of Petroleum Geologists Memoir 55, p. 9–46. sions, northeastern Minnesota: Geochronological insights to physical, petrogenetic, paleo­ Leier, A.L., and Gehrels, G.E., 2011, Continental-scale detrital zircon provenance signatures in magnetic, and tectonomagmatic processes associated with the 1.1 Ga Midcontinent Rift Lower Cretaceous strata, western North America: Geology, v. 39, p. 399–402, https://doi​ ​.org​ System: Journal of Geophysical Research Solid Earth, v. 98, no. B8, p. 13997–14013, https://​ /10​.1130​/G31762​.1​. doi​.org​/10​.1029​/93JB01159​. Ludwig, K.R., 2012, Isoplot v. 3.75: A geochronological toolkit for Microsoft Excel: Berkeley Geo- Pana, D.I., and van der Pluijm, B.A., 2015, Orogenic pulses in the Alberta Rocky Mountains: chronology Center, Special Publication, no. 5, p. 1–75. of major faults and comparison with the regional tectono-stratigraphic MacKay, D.A., and Dalrymple, R.W., 2011, Dynamic deposition in a tidal environment: The record: Geological Society of America Bulletin, v. 127, p. 480–502, https://doi​ ​.org​/10​.1130​ record of fluid mud deposition in the Cretaceous Bluesky Formation, Alberta, Canada: Jour- /B31069​.1​. nal of Sedimentary Research, v. 81, p. 901–920, https://​doi​.org​/10​.2110​/jsr​.2011​.74​. Park, H., Barbeau, D.L., Rickenbaker, A., Bachmann-Krug, D., and Gehrels, G.E., 2010, Application Manduca, C.A., Kuntz, M.A., and Silver, L.T., 1993, Emplacement and deformation history of of foreland basin detrital-zircon geochronology to the reconstruction of the southern and the western margin of the Idaho batholith near McCall, Idaho: Influence of a major terrane central Appalachian Orogen: The Journal of Geology, v. 118, p. 23–44, https://​doi​.org​/10​.1086​ boundary: Geological Society of America Bulletin, v. 105, p. 749–765, https://​doi​.org​/10​.1130​ /648400​. /0016​-7606​(1993)105​<0749:​EADHOT>2​.3​.CO;2​. Paterson, S.R., Okaya, D., Memeti, V., Economos, R., and Miller, R.B., 2011, Magma addition Maselli, V., Hutton, E.W., Kettner, A.J., Syvitski, J.P.M., and Trincardi, F., 2011, High-frequency sea and flux calculations of incrementally constructed magma chambers in continental mar- level and sediment supply fluctuation during Termination 1: An integrated sequence-stratig- gin arcs: Combined field, geochronologic, and thermal modeling studies: Geosphere, v. 7, raphy and modeling approach from the Adriatic Sea (Central Mediterranean): Marine Geol- p. 1439–1468, https://​doi​.org​/10​.1130​/GES00696​.1​. ogy, v. 287, p. 54–70, https://​doi​.org​/10​.1016​/j​.margeo​.2011​.06​.012​. Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., and Maas, R., 2010, Im- Matthews, W.A., and Guest, B., 2017, A practical approach for collecting large-n detrital zircon proved laser ablation U-Pb zircon geochronology through robust downhole fractionation U-Pb data sets by quadrupole LA-ICP-MS: Geostandards and Geoanalytical Research, v. 41, correction: Geochemistry Geophysics Geosystems, v. 11, p. 1–36, https://doi​ ​.org​/10​.1029​ p. 161–180, https://​doi​.org​/10​.1111​/ggr​.12146​. /2009GC002618​. May, S.R., Gray, G.G., Summa, L.L., Stewart, N.R., Gehrels, G.E., and Pecha, M.E., 2013, Detrital Pearce, T.H., 1970, The analcite-bearing volcanic rocks of the Crowsnest Formation, Alberta: Cana- zircon geochronology from the Bighorn Basin, Wyoming, USA: Implications for tectono­ dian Journal of Earth Sciences, v. 7, no. 46, p. 46–66, https://​doi​.org​/10​.1139​/e70​-004​.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1205 Research Paper

Poulton, T.P., Tittemore, J., and Dolby, G., 1990, Jurassic strata of northwestern (and west-cen- Spencer, C.J., and Kirkland, C.L., 2015, Visualizing the sedimentary response through the oro- tral) Alberta and northeastern British Columbia: Bulletin of Canadian Petroleum Geology, genic cycle: A multidimensional scaling approach: Lithosphere, v. 8, p. 29–37, https://doi​ ​.org​ v. 38A, p. 159–175. /10​.1130​/L479​.1​. Price, R.A., 1994, Cordilleran tectonics and the evolution of the Western Canada Sedimentary Storti, F., and McClay, K., 1995, Influence of syntectonic sedimentation on thrust wedges in ana- Basin, in Mossop, G.D., and Shetsen, I., comp., Geological Atlas of the Western Canada Sedi­ logue models: Geology, v. 23, p. 999–1002, https://​doi​.org​/10​.1130​/0091​-7613​(1995)023​<0999:​ mentary Basin: Edmonton, Alberta, Canadian Society of Petroleum Geologists and Alberta IOSSOT>2​.3​.CO;2​. Research Council, p. 13–25. Suttner, L.J., 1969, Stratigraphic and petrographic analysis of Upper Jurassic–Lower Cretaceous Price, R.A., and Mountjoy, E.W., 1970. Geologic structure of the Canadian Rocky Mountains be- Morrison and Kootenai Formations, southwest Montana: The American Association of tween Bow and Athabasca Rivers: A progress report, in Wheeler, J.O., ed., Structure of Petroleum Geologists Bulletin, v. 53, no. 7, p. 1391–1410, https://​doi​.org​/10​.1306​/5D25C859​ the Southern Canadian Cordillera: Geological Association of Canada, Special Publication -16C1​-11D7​-8645000102C1865D​. 6, p. 7–25. Torres, R., Ruiz, J., Patchett, P.J., and Grajales, J.M., 1999, Permo–Triassic continental arc in east- Putnam, P.E., 1982, Fluvial channel sandstones within upper Mannville (Albian) of Lloydminis- ern Mexico: Tectonic implications for reconstruction of southern North America, in Barto- ter area, Canada: Geometry, petrography, and paleogeographic implications: The American lini, C., Wilson, J.L., and Lawton, T.F., eds., Mesozoic Sedimentary and Tectonic History of Asso­ciation of Petroleum Geologists Bulletin, v. 66, p. 436–459. North-Central Mexico: Geological Society of America Special Paper 340, p. 191–196. Putnam, P.E., and Pedskalny, M.A., 1983, Provenance of Clearwater Formation reservoir sand- Van Kranendonk, M.J., St-Onge, M.R., and Henderson, J.R., 1993, Paleoproterozoic tectonic as- stones, Cold Lake, Alberta, with comments on feldspar composition: Bulletin of Canadian sembly of northeast Laurentia through multiple indentations: Precambrian Research, v. 63, Petroleum Geology, v. 31, no. 3, p. 148–160. p. 325–347, https://​doi​.org​/10​.1016​/0301​-9268​(93)90039​-5​. Quinn, G.M., Hubbard, S.M., van Drecht, R., Guest, B., Matthews, W.A., and Hadlari, T., 2016, Van Schmus, W.R., Bickford, M.E., Anderson, J.L., Bender, E.E., Anderson, R.R., Bauer, P.W., Record of orogenic cyclicity in the Alberta foreland basin, Canadian Cordillera: Lithosphere, Robertson, J.M., Bowring, S.A., Condie, K.C., Denison, R.E., Gilbert, M.C., Grambling, J.A., v. 8, p. 317–332, https://​doi​.org​/10​.1130​/L531​.1​. Mawer, C.K., Shearer, C.K., Hinje, W.J., Karlstrom, K.E., Kisvarsanyi, E.B., Lidiak, E.G., Reed, Raines, M.K., Hubbard, S.M., Kukulski, R.B., Leier, A.L., and Gehrels, G.E., 2013, Sediment dis- J.C., Jr., Sims, P.K., Tweto, O., Silver, L.T., Treves, S.B., Williams, M.L., and Wooden, J.L., persal in an evolving foreland: Detrital zircon geochronology from Upper Jurassic and low- 1993, Transcontinental Proterozoic provinces, in Reed, J.C., Jr., Bickford, M.E., Houston, R.S., ermost Cretaceous strata, Alberta Basin, Canada: Geological Society of America Bulletin, Link, P.K., Rankin, D.W., Sims, P.K., and Van Schmus, W.R., eds., Precambrian Conterminous v. 125, p. 741–755, https://​doi​.org​/10​.1130​/B30671​.1​. U.S.: Boulder, Colorado, Geological Society of America, Geology of North America, v. C-2, Ranger, M.J., and Pemberton, G.S., 1988, Marine influence on the McMurray Formation in the p. 171–334, https://​doi​.org​/10​.1130​/DNAG​-GNA​-C2​.171​. Primrose Area, Alberta, in James, D.P., and Leckie, D.A., eds., Sequences, Stratigraphy, Sedi­ Vann, I.R., Graham, R.H., and Hayward, A.B., 1986, The structure of mountain fronts: Journal of mentology: Surface and Subsurface: Canadian Society of Petroleum Geologists Memoir 15, Structural Geology, v. 8, no. 3–4, p. 215–227, https://​doi​.org​/10​.1016​/0191​-8141​(86)90044​-1​. p. 439–450. Vermeesch, P., 2013, Multi-sample comparison of detrital age distributions: Chemical Geology, Roddaz, M., Viers, J., Brusset, S., Baby, P., and Herail, G., 2005, Sediment provenances and drain- v. 341, p. 140–146, https://​doi​.org​/10​.1016​/j​.chemgeo​.2013​.01​.010​. age evolution of the Neogene Amazonian foreland basin: Earth and Planetary Science Let- Walker, T.F., 1974, Stratigraphy and depositional environments of the Morrison and Kootenai ters, v. 239, p. 57–78, https://​doi​.org​/10​.1016​/j​.epsl​.2005​.08​.007​. Formations in the Great Falls area, central Montana [Ph.D. thesis]: University of Montana, Ross, G.M., and Villeneuve, M., 2003, Provenance of the Mesoproterozoic (1.45 Ga) Belt basin 195 p. (western North America): Another piece in the pre-Rodina paleogeographic puzzle: Geologi- Whitmeyer, S.J., and Karlstrom, K.E., 2007, Tectonic model for the Proterozoic growth of North cal Society of America Bulletin, v. 115, p. 1191–1217, https://​doi​.org​/10​.1130​/B25209​.1​. America: Geosphere, v. 3, no. 4, p. 220–259, https://​doi​.org​/10​.1130​/GES00055​.1​. Ross, G.M., Patchett, P.J., Hamilton, M., Heaman, L., DeCelles, P.G., Rosenberg, E., and Giovanni, Wiedenbeck, M., Allen, P., Corfu, F., Griffin, W.L., Meier, M., Oberli, F., Von Quadt, A., Roddick, M.K., 2005, Evolution of the Cordilleran orogeny (southwestern Alberta, Canada) inferred J.C., and Spiegel, W., 1995, Three natural zircon standards for U-Th-Pb, Lu-Hf, trace element from detrital mineral geochronology, geochemistry, and Nd isotopes in the foreland basin: and REE analysis: Geostandards and Geoanalytical Research, v. 19, p. 1–23, https://​doi​.org​ Geological Society of America Bulletin, v. 117, p. 747–763, https://​doi​.org​/10​.1130​/B25564​.1​. /10​.1111​/j​.1751​-908X​.1995​.tb00147​.x​. Shelton, J.W., 1967, Stratigraphic models and general criteria for recognition of alluvial, barrier-­ Wightman, D.M., and Pemberton, S.G., 1997, The Lower Cretaceous (Aptian) McMurray Forma- bar, and turbidity-current sand deposits: The American Association of Petroleum Geologists tion: An overview of the Fort McMurray Area, northeastern Alberta, in Pemberton, S.G., and Bulletin, v. 51, no. 12, p. 2441–2461. James, D.P., eds., Petroleum Geology of the Cretaceous Mannville Group, western Canada: Simpson, G.D.H., 2010, Formation of accretionary prisms influenced by sediment subduction Canadian Society of Petroleum Geologists Memoir 18, p. 312–344. and supplied by sediments from adjacent continents: Geology, v. 38, p. 131–134, https://​doi​ Wipple, K.X., 2009, The influence of climate on the tectonic evolution of mountain belts: Nature .org​/10​.1130​/G30461​.1​. Geoscience, v. 2, https://​doi​.org​/10​.1038​/ngeo413​. Smith, D.G., Zorn, C.E., and Sneider, R.M., 1984, The Paleogeography of the Lower Cretaceous Windley, B.G., 1988, Anorogenic magmatism and the Grenville Orogeny: Canadian Journal of of western Alberta and northeastern British Columbia in and adjacent to the Deep Basin of Earth Sciences, v. 26, p. 479–489. the Elmworth area, in Masters, J.A., ed., Elmworth: Case Study of a Deep Basin Gas Field: Zaitlin, B.A., Warren, M.J., Potocki, D., Rosenthal, L., and Boyd, R., 2002, Depositional styles in a American Association of Petroleum Geologists Memoir 38, p. 79–114. low accommodation foreland basin setting: An example from the Basal Quartz (Lower Cre- Somme, T.O., Helland-Hansen, W., Martinsen, O.J., and Thurmond, J.B., 2009, Relationships taceous), southern Alberta: Bulletin of Canadian Petroleum Geology, v. 50, no. 1, p. 31–72. between morphological and sedimentological parameters in source-to-sink systems: A ba- Zhao, G., Cawood, P., Wilde, S.A., and Sun, M., 2002, Review of global 2.1–1.8 Ga orogens: Impli- sis for predicting semi-quantitative characteristics in subsurface systems: Basin Research, cations for a pre-Rodinian supercontinent: Earth-Science Reviews, v. 59, p. 125–162, https://​ v. 21, p. 361–387, https://​doi​.org​/10​.1111​/j​.1365​-2117​.2009​.00397​.x​. doi​.org​/10​.1016​/S0012​-8252​(02)00073​-9​.

GEOSPHERE | Volume 14 | Number 3 Quinn et al. | A record of orogenic wedge evolution 1206