Research Paper

GEOSPHERE Orogen proximal sedimentation in the foreland basin Graham M. Soto-Kerans1,2, Daniel F. Stockli1, Xavier Janson2, Timothy F. Lawton2, and Jacob A. Covault2 1Department of Geological Sciences, Jackson School of Geosciences, The University of at Austin, Austin, Texas 78712, USA 2Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, Austin, Texas 78758, USA GEOSPHERE, v. 16, no. 2

https://doi.org/10.1130/GES02108.1 ABSTRACT ■■ INTRODUCTION

15 figures; 1 table; 1 set of supplemental files The sedimentary fill of peripheral foreland basins has the potential to pre- Major tectonic and geologic events, such as ocean closure and continen- serve a record of the processes of ocean closure and continental collision, as tal collision, and the evolution of associated -routing systems, are CORRESPONDENCE: well as the long-term (i.e., 107–108 yr) sediment-routing evolution associated recorded within the sedimentary fill of foreland basins (Jordan et al., 1988; [email protected] with these processes; however, the detrital record of these deep-time tec- DeCelles and Giles, 1996; Ingersoll, 2012). Detrital zircon (D2) U-Pb geochronol- CITATION: Soto-Kerans, G.M., Stockli, D.F., Janson, tonic processes and the sedimentary response have rarely been documented ogy is a powerful provenance analysis tool for elucidating hinterland tectonic X., Lawton, T.F., and Covault, J.A., 2020, Orogen proxi- during the final stages of supercontinent assembly. The stratigraphy within and erosional processes and the associated depositional record (Dickinson, mal sedimentation in the Permian foreland basin: Geo- the southern margin of the Delaware Basin and Marathon fold and thrust 1988; Fedo et al., 2003; Gehrels, 2012; Gehrels, 2014). This methodology pro- sphere, v. 16, no. 2, p. 567–​593, https://doi​.org/10.1130​ ​ /GES02108.1. belt preserves a record of the –​Permian Pangean continental vides invaluable insights into the geological processes linked to continental assembly, culminating in the formation of the Delaware and Midland foreland assembly and can help track the complex temporal and spatial evolution of Science Editor: David E. Fastovsky basins of North America. Here, we use 1721 new detrital zircon (DZ) U-Pb these foreland basin systems (Haughton et al., 1991; Lawton et al., 2010). Associate Editor: Cathy Busby ages from 13 stratigraphic samples within the Marathon fold and thrust The Permian Basin is the result of the diachronous collision between the belt and Glass Mountains of in order to evaluate the prove- Gondwanan and Laurentian continental plates during the Late Mississippian Received 19 December 2018 nance and sediment-routing evolution of the southern, orogen-proximal to early Permian, completing the formation of the supercontinent Pangea. This Revision received 6 August 2019 Accepted 2 December 2019 region of this foreland basin system. Among these new DZ data, 85 core-rim collision led to the inversion of the Laurentian continental margin, over-thrust- age relationships record multi-stage crystallization related to magmatic or ing of Gondwanan crust, and the formation of several flexural foreland basins Published online 6 January 2020 metamorphic events in sediment source areas, further constraining source and uplifts (King, 1930; King, 1937; Ross, 1986; Yang and Dorobek, 1995; Poole terranes and sediment routing. Within samples, a lack of Neoproterozoic–​ et al., 2005). The Delaware Basin is the western of two major sub-basins that Cambrian zircon grains in the pre-orogenic Mississippian Tesnus Formation make up the Permian foreland basin. It contains extensive hydrocarbon reser- and subsequent appearance of this zircon age group in the syn-orogenic voirs in conventional deep-water as well as unconventional Haymond Formation point toward initial basin inversion and strata. These important resources have motivated numerous and the uplift and exhumation of volcanic units related to Rodinian rifting. studies of the depositional environments and reservoir quality across the Moreover, an upsection decrease in Grenvillian (ca. 1300–920 Ma) and an basin (Hills, 1984; Harms et al., 1988; Dutton et al., 2003; Gardner et al., 2003; increase in zircons denote a progressive provenance shift from Pyles et al., 2010). The Pennsylvanian to middle Permian stratigraphy exposed that of dominantly orogenic highland sources to that of sediment sources along the southern margin of the Delaware Basin (Figs. 1 and 2) offers the deeper in the Gondwanan hinterland during tectonic stabilization. Detrital unique opportunity to study geological processes related to late Paleozoic zircon core-rim age relationships of ca. 1770 Ma cores with ca. 600–300 Ma Laurentia- continental collision and foreland basin deposition near rims indicate Amazonian cores with peri-Gondwanan or Pan-African rims, the orogenic front. Grenvillian cores with ca. 580 Ma rims are correlative with Pan-African vol- Provenance studies have been employed to reconstruct the evolution of canism or the ca. 780–560 Ma volcanics along the rifted Laurentian margin, numerous orogenic systems, such as the Andes or the Himalayas, which still and Paleozoic core-rim age relationships are likely indicative of volcanic arc preserve much of both the orogenic hinterland and the adjacent basins. In activity within peri-Gondwana, Coahuila, or Oaxaquia. Our results suggest contrast, the Permian Basin represents a more significant challenge because dominant sediment delivery to the Marathon region from the nearby south- a significant portion of the Gondwanan hinterland subsequently rifted away ern orogenic highland; less sediment was delivered from the axial portion of during opening of the (Hatcher et al., 1989; Thomas, the Ouachita or Appalachian regions suggesting that this area of the basin 2006, 2011b), and what likely remains in northern and northeastern Mexico was not affected by a transcontinental drainage. The provenance evolution is covered and poorly understood. Hence, the sedimentary strata along the of sediment provides insights into how continental collision directs the southern margin of the Permian Basin offer the potential to help unravel the This paper is published under the terms of the dispersal and deposition of sediment in the Permian Basin and analogous long-term geological and tectonic evolution of this collisional episode despite CC‑BY-NC license. foreland basins. the missing or inaccessible hinterland. Several models for the sediment

© 2020 The Authors

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 567 by guest on 30 September 2021 Research Paper

Southern Permian Basin and sourcing and routing have been proposed (Thomson and McBride, 1964; Marathon Fold and Thrust Belt Regional Map: Kocurek and Kirkland, 1998; Soreghan and Soreghan, 2013). The recent advent of isotopic provenance analysis through DZ U-Pb dating has provided critical Delaware Basin Central Basin Platform process-oriented insights into the linkages between basin formation, hinter- Hovey and Ozona Arch land tectonics and/or unroofing, sediment routing, drainage reorganization, Diablo Channel Platform Val Verde Basin and basin subsidence as read in the basin fill history (Lawton et al., 2010, GLASS MTNS 2015a, 2015b; Gehrels, 2012; Sharman et al., 2015; Thomson et al., 2017). Orogenic Recent provenance studies have employed DZ U-Pb geochronology in order Belt to interpret sediment sources for the voluminous clastic reservoirs within Devils the Permian Basin (Soreghan and Soreghan, 2013; Anthony, 2015; Xie et al., River 2018). Specifically, these studies identified a previously unexpected signal Uplift Marfa of southern sediment provenance (i.e., peri-Gondwana). However, inter- Marathon Basin pretations of the sediment-routing scenario continue to be developed. For example, Soreghan and Soreghan (2013) interpret southerly provenance for the section in the northwestern Delaware Basin, but not by a N 0 50 km direct feeder system. Rather, sediment was contributed via transcontinental

W E river systems stemming from the Ouachita and Appalachian regions and 0 50 mi deflated into eolian erg systems ultimately transporting sediment to the S Delaware Basin. We used DZ U-Pb geochronology to define the signal of southern prov- Glass Mountains and enance and direct sediment supply from the Marathon fold and thrust belt Marathon Fold and Thrust Belt Geologic Map: and Gondwanan hinterland terranes. Moreover, the southern margin of the N Delaware Basin comprises strata that document the collisional phases of Pc Pc Pw 7 8 K IPh W E Pangea; the long-term (10 –10 yr) tectonic processes associated with this Pl Pwf S supercontinent assembly likely significantly influenced the provenance and Pw IPg sediment-routing evolution of the Delaware Basin (Graham et al., 1976). Pl IPd IPd Ci Duggout Creek Thrust IPh Pl Pwf ■■ GEOLOGIC SETTING AND STRATIGRAPHY Pc IPM Iapetan Rifted Margin—Rodinian Breakup K IPg Marathon IPd Prior to the assembly of Pangea, Neoproterozoic–Cambrian rifting of the lPz Rodinian continent formed the inherited promontories and embayments of IPM IPM K IPh lPz southern Laurentia (Fig. 3) (Arbenz, 1989a; Cawood et al., 2001; Thomas, 2011b). lPz IPM 0 10 km The traces of these large-scale features consist of the northwest-trending trans- IPh 0 10 mi form faults (Alabama-Oklahoma and Texas transforms) and northeast-trending rift segments (Blue Ridge, Ouachita, and Marathon) (Arbenz, 1989a; Thomas, 2006; Thomas, 2011b). Rifting of the Rodinian supercontinent occurred between Intrusive Pl Leonardian IPd Lower Pennsylvanian ca. 825 Ma and 740 Ma with notable volcanic pulses from 780 to 540 Ma (Li et al., K Cretaceous Pwf Wolfcampian IPM Upper Mississippian 2008; Hanson et al., 2016). Stabilization of the new Laurentian continent Pc Upper-Middle Guadalupian IPg Upper Pennsylvanian lPz Lower Paleozoic resulted in passive margin deposition of ~975–1100 m of thin-bedded lime- Pw Wordian IPh Middle Pennsylvanian DZ Sample Location stone, shale, chert, and shaly during the Late Cambrian–​Devonian or early Carboniferous (~170 m.y.) (King, 1978; McBride, 1989; Hickman et al., Figure 1. Regional map of the Glass Mountains and Marathon fold and thrust belt in re- 2009). The interior southern margin of Laurentia contained the Tobosa Basin, lation to Guadalupian paleogeography of the Permian Basin (from Hickman et al., 2009) and associated modern geologic map of the Marathon Region with sample locations. a remnant feature of the Rodinian breakup during Neoproterozoic–​Cambrian DZ—detrital zircon. time (Fig. 3) (Walper, 1982; Keller et al., 1983; Arbenz., 1989a; Li et al., 2008).

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 568 by guest on 30 September 2021 Research Paper

This region was characterized by weak crustal extension and a low subsidence 1989b; Muehlberger and Tauvers, 1989; Viele and Thomas, 1989; Reed and rate during the early Paleozoic (Galley, 1958; Horak, 1985). Stricker, 1990; Cawood et al., 2001; Thomas, 2006, 2011a, 2011b, 2014). During the Mississippian–Permian​ collision of Laurentia and Gondwana, mountain ranges formed across the axial continental divide through growth of oro- Convergence and Collision of Laurentia and Gondwana genic wedges and their thrusting onto the southern margin of the Laurentian continent (Fig. 4) (Ross, 1986; Poole et al., 2005; Keppie et al., 2008). The com- The Pangean accretionary margin preserves in outcrop the entire complex bined influences of tectonic stress related to the advancing Marathon fold and tectonic history of diachronous continental collision (Cawood and Buchan, thrust belt and preexisting lithospheric weakness from the antecedent Tobosa 2007), making it of unique structural interest and the focus of palinspastic Basin segmented and isolated the Permian Basin along renewed northwest- to reconstruction of numerous studies (Fig. 2) (Tauvers, 1988; Arbenz, 1989a, north-trending fault zones such as the Puckett-Gray Ranch fault, an intrabasinal

Glass Mountains and Marathon Fold Regional Cross-Section of the Glass Mountains, West Texas and Thrust Belt Geologic Map: ft m N NW SE 2000 W E S Gilliam Fm. 500

Capitan Fm. Guadalupian Vidrio Fm. Duggout Creek Thrust Altuda Fm. 1000

Hess Leonardian Word Fm. Marathon Road Canyon Fm. Cathedral Mountain Fm. Lenox Hills Fm. Wolfcampian Skinner Ranch IPh 0 0 0 10 km

0 1 2 3 4 5 6 mi Shelf 0 10 mi Shelf edge and ramp 0 2 4 6 8 10 km Vertical scale x3 Cenozoic Intrusive Pl Leonardian IPd Early Pennsylvanian Adapted from Janson and Hairabian, RCRL. K Cretaceous Pwf Wolfcampian IPM Late Mississippian Pc Late-Middle Guadalupian IPg Late Pennsylvanian lPz Lower Paleozoic Pw Wordian IPh Middle Pennsylvanian DZ Sample Location ft NW SE 6,000 4,000 SL 0 -4,000 -8,000 -12,000 -16,000 -20,000 -24,000 -28,000 -32,000

ft m Adapted from Reed and Stricker, 1990 12,000 Cenozoic Intrusive Middle Permian (Word) Pennsylvanian (Haymond/Dimple) Ordovician-Cambrian (Ellenburger) 3000 8000 Cretaceous Leonardian Mississippian (Tesnus) Cambrian (Dagger Flat, Simpson Springs, and mid-Cambrian ) 2000 4000 1000 Middle Guadalupian (Altuda) Wolfcampian Mississippian-Ordovician (Simpson) Precambrian 8000 16,000 ft 0 Middle Permian (Vidrio) Pennsylvanian (Gaptank) Devonian-Ordovician 0 2000 4000 m

Figure 2. (A) Cross section of the Glass Mountains study area (pink line on B) displaying N-NW progradation of Wolfcampian–Guadalupian stratigraphic units during basin formation and fill (Hair- abian and Janson, 2016); (B) geologic map of the Marathon Region with sample locations; and (C) structural cross section of the broader Marathon Region derived from multiple well and seismic ties (Strickler and Reed, 1990). DZ—detrital zircon.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 569 by guest on 30 September 2021 Research Paper

Figure 3. Late Precambrian to middle Cambrian O rift-transform breakup of Rodinia, subsequent OA R thrust belt formation during Mississippian– OU Permian collision of Laurentia and Gondwana. BRR Alabama-Oklahoma Transform AF Abbreviations: A—Alabama promontory; AF—late Guadalupe Mtns. Paleozoic Appalachian thrust front; BRR—Blue Ridge Rift; D—Delaware aulacogen; L—Llano (or T Texas) promontory; M—Marathon embayment; MR—Marathon Rift; O—Oklahoma Basin (early D A OF OR Paleozoic); OA—Oklahoma aulacogen; OF—late Paleozoic Ouachita thrust front; OR—Ouachita M Rift; OU—Ouachita embayment; R—Redfoot rift (with early Paleozoic Mississippi Valley Basin); Texas TransformL MR T—Tobosa Basin (early Paleozoic). From Arbenz Rift (1989b) and Thomas et al. (2002). Thrust Fault Transform Protobasin 0 500 km Aulocogen

BALTICA BALTICA

NORTH NORTH AMERICA Cadomia AMERICA Avalonia Cadomia Maya Avalonia Carolina Iberia RHEIC OCEAN Florida/Suwanee Maya Sierra Madre Oaxaquia Florida Oaxaquia Sierra Madre Mixteca Chortis SOUTH Iberia Chortis SOUTH AMERICA AMERICA AFRICA AFRICA

350 Ma 300 Ma Figure 4. Paleogeographic reconstructions showing Laurentian, Gondwanan, and African continental pieces during Carboniferous reconstruction on the active Pacific margin with subduction beneath the Mixteca (Acatlán) and Oaxaquia terranes and Permian age construction of Pangea during continental collision of Laurentia and Gondwana (adapted from Keppie et al., 2008).

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 570 by guest on 30 September 2021 Research Paper

uplift known as the Central Basin Platform, and basin-bounding highs as seen N in the Northwest Shelf, Eastern Shelf, and Devils River Uplift (Fig. 5) (Ross, 1986; Yang and Dorobek, 1995; Ewing, 2016). By defining the morphology of the W E

Permian Basin, these basin-bounding fault zones and basement uplifts altered S its structural evolution and subsidence history while simultaneously changing the patterns of foreland deposition (Muehlberger and Tauvers, 1989; Yang and NW Shelf Dorobek, 1995). Yang and Dorobek (1995) characterize the Permian Basin as Eastern Shelf a composite foreland basin because of the complex system of tectonic and load-induced flexures that form the separate sub-basins and uplifts, which are Midland Central not prevalent in all foreland basin systems (Hagen et al., 1985; Flemings et al., Basin Delaware Basin 1986; Beck et al., 1988; Yang and Dorobek, 1995). Basin Platform

Stratigraphic Record of the Glass Mountains and Marathon Fold and Diablo Thrust Belt Platform Ozona Arch

Situated in the southernmost Delaware Basin, the Glass Mountains and Puckett-Gray Ranch Fault V Marathon fold and thrust belt contain outcrops of the entire syn- to early al Verde Basin post-collisional stratigraphic section (Figs. 2 and 6) (King, 1930, 1937, 1978; Marathon Orogenic Belt Marfa McBride, 1989). The deposits within this region record two stages of sedimen- Basin tation, which include (1) pre-collisional passive margin strata of the southern Laurentian continent, generally characterized by thin-bedded , shale, Devils River Uplift chert, and shaly sandstone (975–1100 m >170 m.y.) and (2) syn-orogenic strata, 0 50 mi

which accumulated rapidly to a thickness of 4200–5600 m during ~60 m.y. 0 50 km (Figs. 2 and 6) (King, 1937, 1978; Flawn et al., 1961; Thomson and McBride, 1964;

McBride, 1989; Hickman et al., 2009). The Upper Mississippian–​Lower Penn- Thrust Fault sylvanian Tesnus Formation is a shaly sandstone representing flysch deposits Central Basin Platform/Ozona Arch forming submarine fan and channel complexes sourced from the approach- ing Gondwanan continental terranes (McBride, 1989; Hickman et al., 2009). Figure 5. Mapped structural features of the Permian Basin region. Adapted from Yang and Do- Overlying the Tesnus are the Pennsylvanian Dimple Limestone and Haymond robek (1995). Formation. The Lower Pennsylvanian Dimple Limestone indicates a short period of waning tectonism in which its northern shelf equivalent fed carbonate turbid- ites southward (Hairabian and Janson, 2016). Deposited during ongoing tectonic activity, the sediment source history of the Haymond Formation remains contro- Previous Provenance Studies versial (King, 1958; Thomson and McBride, 1964; Denison et al., 1969; McBride, 1989; Gleason et al., 2007). The uppermost Pennsylvanian unit in the system is Long-standing interpretations (Fig. 7) have concluded that the majority of the Gaptank Formation. Regionally, coarse clastic fluvial and deltaic siliciclastic input to the Delaware Basin during middle Permian (Guadalupian) mark the lower Gaptank. In contrast, the Gaptank in the Glass Mountains is time was sourced from the Ancestral (ARM) of central Lau- composed of carbonate shelf deposits and phyloid algal mounds, indicative of rentia. Hull (1957) suggested that observed arkosic mineralogy of the sandy tectonic quiescence in the region (King, 1937; Hairabian and Janson, 2016). A few and silty units of the Artesia and Delaware Mountain groups of the Northwest miles north of these exposures are the Permian deposits of the Glass Moun- Shelf were evidence of north-to-south transport of sediment from a pluton- tains. The Glass Mountains expose a series of north-northwest–​prograding ic-metamorphic source such as the ARM and midcontinent region of Laurentia. mixed carbonate-siliciclastic deposits (Figs. 1 and 2). The 1500–2000 m of Fischer and Sarnthein (1988) emphasized similarities between the Delaware Wolfcampian–Guadalupian​ (early to middle Permian) stratigraphy of the Glass Basin (arid, low latitude, south of the northern trade-wind belt) to the Holocene–​ Mountains range in depositional setting from Wolfcampian proximal syn-oro- of northwest Africa, as well as generally north-south–trending genic conglomerates to Guadalupian carbonate -rimmed platform and slope paleocurrent data from dunes in the Grand Canyon region during late Paleozoic (King, 1930, 1937; Ross, 1963; Hairabian and Janson, 2016). time to be evidence of an ARM source for the Delaware Mountain group of the

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 571 by guest on 30 September 2021 Research Paper

Middle Permian (265-255 Ma) North Regional Stratigraphy International American Standard Standard Guadalupe Delaware Glass Mountains Mountains Basin and Marathon Basin Laurentia 259.1 Reef Trail Lamar ARM

Capitanian McCombs Capitan

Altuda Delaware Basin Capitan Rader Appalachian-Alleghanian Orogen Bell Canyon 265.1 Pinery Altuda

Traditional Capitanian Traditional ALSS1 Hegler

Wordian Manzanita Vidrio VSS1-4 Yucatan Marathon-Ouachita Orogen South Wells

Guadalupian Goat Appel Ranch Seep Oaxaquia Gondwana 268.8 Getaway Willis Ranch Chortis Cherry Canyon 1000 km Word Roadian WSS5 Blakey Deep Time Maps Brushy Canyon China Tank Initial sediment routing interpretation from outcrop paleocurrent data and petrographic analysis (Fischer and Sarntheirn, 1988; Kocurek and Kirkland, 1998) Cutoff Road Canyon 272.9 Avalon Sediment routing interpretation from DZ geochronologic studies Shale Cathedral Mountain (Soreghan and Soreghan, 2013; Xie et al., 2018)

Bone Skinner Victorio Peak Hess LSS1 Leonardian Spring Ranch Figure 7. Paleogeographic map of the Laurentian and Gondwanan continents during the mid- (Early Perm.) dle Permian. Arrows indicate sediment routing interpretations by previous studies within the Lenox Delaware Basin. Modified from Blakey Deep Time Maps (https://deeptimemaps.com/). ARM— Hills WCSS1-2 Ancestral Rocky Mountains. Wolfcamp Neal

Wolfcampian Ranch 298.9 Cisco Late from the of West Texas (Fig. 3) refuted an ARM source Gaptank Pennsylvannian Canyon for the Delaware Mountain Group. They documented dominant Paleozoic, 307.0 Neoproterozoic, and Mesoproterozoic DZ age modes and minor late Paleop- Strawn Haymond HF1-2 roterozoic ages, only the latter of which would be indicative of an ARM source E. and Mid. Pennsylvanian (Yavapai-Mazatzal ≈1825–1600 Ma). Instead, they suggest that measured U-Pb Atoka Dimple Ls age modes combined with Laurentian paleocurrent data point to a combination 323.2 Morrow ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ? of recycled Appalachian- and/or Ouachita-derived sources and Gondwanan Late Tesnus Mississippian TSS1-2 and Pan-African terranes south of the Ouachita-Marathon suture. Using DZ samples from the central and southern Delaware Basin, Anthony (2015) and Xie et al. (2018) suggested sediment was sourced from the Appalachian orogenic Figure 6. Stratigraphic column of the Guadalupe Mountains, Delaware Basin, and Glass Mountains and Marathon fold and thrust belt (this study). Detrital zircon sample locations are denoted by belt, Ouachita orogenic belt, and peri-Gondwanan terranes while also noting a white symbols within stratal intervals. Adapted from Barnaby et al. (2004), Olszewski and Erwin shift in provenance during the middle Permian, suggesting a continental-scale (2009), and Nestell et al. (2019). fluvial system spanning from the Appalachian orogenic belt westward across Pangea (Fig. 7). While the Guadalupe Mountains and Delaware Basin have been at the epicenter of stratigraphic and provenance studies, detailed work Guadalupe Mountains. Alternatively, Kocurek and Kirkland (1998) suggested in the Glass Mountains has historically lacked similar treatment. Gleason et al. that paleogeography and paleowinds would support eolian sediment transport (2007) analyzed detrital zircons from the Pennsylvanian Haymond Formation from the Whitehorse Group of the Anadarko Basin, Oklahoma. More recently, of the Marathon fold and thrust belt, suggesting that sources range from the through a combination of DZ U-Pb geochronology, compiled paleocurrent Laurentian craton, uplifted strata of the remnant-ocean basin, Gondwanan data, and paleoclimate data, Soreghan and Soreghan’s (2013) study of samples crustal sources, and Ouachita sources to the east. Gleason et al. (2007) went

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 572 by guest on 30 September 2021 Research Paper

further to interpret transport of this sediment as braid deltas and foredelta 100° 90° 80° Hearne submarine-ramp environments emanating from the closing Pangean suture Penokean Orogeny zone. The stratigraphic framework and biostratigraphy of the Glass Mountains (>2.5 Ga) Superior Province 45° are documented by a few studies (King, 1930, 1937; Rathjen, 1993; Haneef

Trans-Hudson Wyoming et al., 2000; Harris et al., 2000; Lambert et al., 2000; Wardlaw, 2000; Yang and Yancey, 2000; Lambert et al., 2002; Hairabian and Janson, 2016), but detailed Taconic/ provenance history of the clastic record has not been previously attempted. Yavapai Peri-Gondwana Acadian/ (1.8-1.7 Ga) Alleghanian Orogenies Cordilleran Accreted TerranesMiogeocline Mazatzal 35° (1.7-1.62 Ga)

Laurentian Grenville Potential Sediment Sources ARM

Delaware Archean and Paleoproterozoic (>1825 Ma) Basin Sabine Suwanee (650-510 Ma) Detrital zircons older than 1825 Ma could potentially be sourced from the Yucatan/Maya 25° northern extents of the Laurentian craton or from the Amazonian paleo-hin- (550, 420-410 Ma) Coahuila terland of Gondwana. The continent of Amazonia has a multi-stage history of agglutination with both Laurentia, during the assembly of Rodinia (along the Grenvillian margin of Laurentia), and Gondwana during the opening of Mixteca the Iapetus at ca. 600 Ma (Sánchez-Bettucci and Rapalini, 2002; Tohver et al.,

Peri-Gondwana Maroni-Itacaiunas (Acatlán) Oaxaquia

E. Mex Arc (2.25-1.95 Ga) 2002; Cordani et al., 2009). Amazonian basement provinces contain a wide Venturari-Tapajós range of Archean and Paleoproterozoic ages. The Amazonian central craton

Pan-African/Brasiliano (1.98-1.81 Ga) (Fig. 8) contains the Carajás region southwest of the Amazon Basin. Basement

Central (900-600 Ma) Amazonia Craton rocks within Carajás are dated between 3200 and 2600 Ma (Cordani et al., Rio Negro-Juruena 2009). The Maroni-Itacaiunas province occurs to the north and northeast of (1.78-1.55 Ga) (>2.6 Ga) the Central-Amazonian province with rocks dated between 2250 and 1950 Ma

Rondonia-San Ignacio (Fig. 8) (Cordani et al., 2009). Calc-alkaline, granite-gneiss complexes of the 0 1000 km (1.55-1.3 Ga) Ventuari-Tapajós and Rio Negro–​Juruena provinces formed between 1980 Sunsas (1.28-0.95 Ga) and 1810 Ma (Fig. 8) (Santos, 2003; Cordani and Teixeira, 2007; Cordani et al., 2009). These Amazonian provinces were potential sources of Archean and Archean and Paleoproterozoic (>1825 Ma) Paleoproterozoic age zircons both during Rodinian time as well as during the formation of Pangea (connection with the Gondwanan hinterland). Potential Late Paleoproterozoic (1825-1578 Ma) sources within the Laurentian continent include the Wyoming and Superior Early Mesoproterozoic (1578-1300 Ma) Provinces (3000 and 2700–2500 Ma) of the northern Laurentian continent (Xie et al., 2018). Grains of 1900–1800 Ma and 1850–1780 Ma from the Penokean Mesoproterozoic (1300-920 Ma) orogen and associated Wisconsin magmatic terrane and Trans-Hudson prov- Neoproterozoic and Early Cambrian (920-521 Ma) ince of the northern Laurentian craton (Fig. 8) were also capable of shedding sediment south toward the Permian Basin (Sims et al., 1989; Beck and Murthy, Paleozoic (521-263 Ma) 1991; Anderson and Morrison, 1992; Holm et al., 2007). Rifted Laurentian Margin

Rodinian Rift Volcanics (780-540 Ma) Late Paleoproterozoic (ca. 1825–1578 Ma) 1.49-1.34 Ga Granite-Rhyolite Province Late Paleoproterozoic age zircons match closest with the Ancestral Rocky Permo-Triassic Arc Volcanics (385-245 Ma) Mountains and Yavapai-Mazatzal (1800–1600 Ma) terranes of central Lauren- tia (Fig. 8). These first-cycle basement sources were exposed throughout the Figure 8. Map depicting main basement age provinces of North America and accreted terranes capable of shedding sediment to the Permian Basin. Adapted from Lawton et al. (2015b). ARM— Pennsylvanian and earliest Permian (Anderson and Morrison, 1992; Anderson Ancestal Rocky Mountains. et al., 1993; Holm et al., 2007; Giles et al., 2013; Lawton et al. 2015a). These

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 573 by guest on 30 September 2021 Research Paper

ages have also been documented in detrital zircons from a multitude of basin (Becker et al., 2006). Zircon analyses within the Pennsylvanian and Permian studies of Paleozoic strata along the Ouachita-Marathon suture zone as well strata of the Appalachian foreland basin, however, show that age modes of as north and northeast of the Permian Basin, giving the possibility of sedi- contemporary magmatism are typically underrepresented or absent from ment recycling (Dickinson and Gehrels, 2003; Becker et al., 2005; Becker et al., synorogenic strata (Thomas et al., 2004; Becker et al., 2005, 2006). 2006; Gehrels et al., 2011b; Giles et al., 2013; Link et al., 2014; Xie et al., 2016). The exotic terranes associated with the collision of south and southwest- ern Laurentia (present-day Alabama-Texas) with Gondwana are also capable of supplying zircons of this age range. Peri-Gondwanan terranes along the Early Mesoproterozoic (ca. 1578–1300 Ma) Ouachita-Marathon suture as well as Pennsylvanian–​Permian strata of the Yucatan/Maya, Coahuila, Oaxaquia, and Mixteca (Acatlán) terranes (Figs. 4 Zircons within the early Mesoproterozoic age range are typically represen- and 8) had been accreted and uplifted by Permian time (Sacks and Secor, 1990; tative of anorogenic igneous activity of the granite-rhyolite province of the Dickinson and Lawton, 2001, 2003; Poole et al., 2005; Soreghan and Soreghan, central North American plate, ranging in age between 1.49 and 1.34 Ga, and 2013; Xie et al., 2018). Though all of these now Mexican and Central American spanning from Labrador to California (Fig. 8) (Bickford et al., 1986; Hoffman terranes are understood to have agglutinated to the Gondwanan continent et al., 1989; Anderson and Morrison, 1992). The Rondonian–San​ Ignacio prov- during the formation of Pangea, accretion timing and positions of the ter- ince of Amazonia also contains accreted domains of intra-oceanic character, ranes remain subject to debate (Dickinson and Lawton, 2001; Murphy et al., dated between 1550 and 1300 Ma (Geraldes et al., 2001; Santos, 2003; Cordani 2004; Keppie et al., 2008; Martens et al., 2010; Soreghan and Soreghan, 2013). et al., 2009), granitoids dated between 1520 and 1470 Ma, and the Santa Helena As stated previously, the Tesnus Formation is a shaly sandstone represent- plutonic arc (1450–1420 Ma) (Cordani et al., 2009). Additionally, zircons of this ing pre-collisional flysch deposits in the form of submarine fan and channel age range are found in other Pennsylvanian and Permian age basinal strata deposits sourced from the approaching Gondwanan continent (McBride, 1989; (Dickinson and Gehrels, 2003; Becker et al., 2005, 2006; Gehrels et al., 2011b). Hickman et al., 2009). The whole of the Mississippian Tesnus was sourced from the Gondwanan continent, yet it does not contain any Neoproterozoic to early Cambrian zircons as have been observed in the Yucatan/Maya, Coahuila, and Mesoproterozoic (ca. 1300–920 Ma) Mixteca (Acatlán) terranes (Fig. 8). Rodinian syn-rift volcanic rocks represent the most likely source of sediment Mesoproterozoic age zircon grains are generally indicative of the Gren- that has been overlooked by previous studies. They are situated along the ville orogen, present in southern and eastern North America, Mexico, and southwestern margin of the Laurentian rifted margin. Rifting of the Rodinian Gondwana (Fig. 8) (Rainbird et al., 1992; Rainbird et al., 1997; Gillis et al., 2005; supercontinent occurred between ca. 825 Ma and 740 Ma with notable volca- Talavera-Mendoza et al., 2005; Soreghan and Soreghan, 2013; Xie et al., 2018). nic pulses from 780 to 540 Ma (Fig. 8) (Li et al., 2008; Dickinson and Gehrels, These grains typically make up a significant portion of the DZ population in 2009; Hanson et al., 2016). These ancient syn-rift volcanic rocks are restricted North American sandstones from the Neoproterozoic to Mesozoic (Dickinson to the distal southern Laurentian margin and were subsequently covered by and Gehrels, 2003; Moecher and Samson, 2006). strata deposited along the Laurentian passive margin (King, 1937; King, 1978; Flawn et al., 1961; McBride, 1989; Hickman et al., 2009). The basement of the Devils River Uplift to the east of the Delaware Basin preserves a portion of Neoproterozoic to early Cambrian (ca. 920–521 Ma) the distal Laurentian rifted margin (Figs. 1 and 5). This region contains a suc- cession of Grenvillian-age basement (1246–1121 Ma Rb/Sr age) (Nicholas and Because there are a number of potential sources, the Neoproterozoic to Waddell, 1989; Rodriguez et al., 2017) and a metasedimentary-metavolcanic early Cambrian age suite remains contentious regarding provenance within the unit of Cryogenian–Early​ Cambrian age (Nicholas and Rozendal, 1975; Denison Permian strata of the Delaware Basin (Soreghan and Soreghan, 2013; Xie et al., et al., 1977; Nicholas and Waddell, 1989; Thomas, 2011b; Rodriguez et al., 2017). 2018). The Appalachian orogenic belt traces the accretion of the (northern) Ava- lon terrane, (southern) Carolina terrane, and Suwannee (Florida) terrane, which were involved in the collision of Africa and North America during Alleghanian Paleozoic (ca. 521–263 Ma) tectonism (Fig. 4) (Opdyke et al., 1987; Mueller et al., 1994; Heatherington et al., 1996; Wortman et al., 2000; Hibbard et al., 2002; Dickinson and Gehrels, 2009; The Appalachian orogenic ranges and related Ordovician–Devonian​ igneous Park et al., 2010). Specifically, magmatism of 650–550 Ma only occurred in the rocks of the Taconic (490–440 Ma) and Acadian (390–350 Ma) tectonic regions accreted Avalon, Carolina, and Suwannee terranes (Heatherington et al., 1996). function as potential east and northeast sources for Paleozoic-age zircons The Pennsylvanian to Permian strata of the Appalachian Basin generally are within the Permian Basin (Hatcher, Jr. et al., 1989; Park et al., 2010; Gehrels thought to capture the erosional history of the Alleghanian orogenic wedge et al., 2011b). Portions of the Gondwanan Pan-African regions of Carolina

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 574 by guest on 30 September 2021 Research Paper

(590–535 Ma) and Suwannee (535–511 Ma), closer to the Appalachian oro- from multiple LA-ICP-MS sessions are tabulated in Supplemental File S11. For genic region, are also potential sources of Paleozoic age zircons (Soreghan all standard and unknown analyses, a 30 mm laser spot was used to create and Soreghan, 2013). Other potential Laurentian Paleozoic sources include the ~16-mm-deep ablation pits on un-polished, tape-mounted zircons, producing Franklinian orogen of the Arctic Canadian region and its actively eroding clas- depth profiles for each analyzed grain. The depth profiling technique enables tic wedge (450–370 Ma) (Patchett et al., 2004; Gehrels et al., 2011b; Soreghan resolution of multiple zircon growth zones (Moecher and Samson, 2006) evi- and Soreghan, 2013). dent from core and rim ages (Marsh and Stockli, 2015). Analyses were then Geochronological data from peri-Gondwanan and Pan-African terranes reduced using Iolite data reduction software and VizualAge (Paton et al., 2011; of Mixteca (Acatlán) (480–440 Ma; metamorphic ages of 416–388 Ma) and Petrus and Kamber, 2012), and no common Pb correction was applied. For Yucatan-Maya (418 Ma) closely match the observed DZ U-Pb age modes of the greater precision, the ages represented in analyses are 206Pb/238U ages for zir- Paleozoic-age components throughout the syn- and post-collisional strata of cons younger than 1200 Ma and 207Pb/206Pb ages for zircons older than 1200 Ma the Marathon fold and thrust belt and southern Delaware Basin (Steiner and (Gehrels et al., 2008; Marsh et al., 2019). All ages are quoted at 2σ absolute 206 238 207 235 TSS1 HF1 Walker, 1996; Keppie, 2004; Gillis et al., 2005; Steiner and Anderson, 2005; Tala- uncertainties. The discordance is calculated using the Pb/ U and Pb/ U vera-Mendoza et al., 2005; Keppie et al., 2006, 2008; Weber et al., 2006, 2008; ages if younger than 1200 Ma and the 206Pb/238U and 207Pb/206Pb ages if older Martens et al., 2010). The youngest grain ages within this study are <360 Ma, than 1200 Ma. 207Pb/206Pb and 206Pb/238U zircon ages with >30% and >10%–30% most likely linking them to adjacent volcanic arc activity. The Mississippian discordance, respectively, and >10% analytical uncertainty were discarded. TSS2 HF2 Las Delicias Arc of the Coahuila terrane (McKee et al., 1999; Lopez et al., 2001) In total, after filtering, 1721/1758 (~98%) precise and concordant zircon U-Pb and middle Permian East Mexican Arc likely represent the youngest possible ages were accepted and used for data interpretation. Mineral separation and volcanic activity within Marathon proximity. LA-ICP-MS analyses were completed at the UTChron Geo- and Thermochronol- ogy facilities at the Jackson School of Geosciences at the University of Texas WCSS1 LSS1 at Austin following the analytical procedures of Hart et al. (2016). Data visu- ■■ DETRITAL ZIRCON SAMPLING AND METHODOLOGY alization and plotting were performed using detritalPy (Sharman et al., 2018) python scripts (Fig. 9) (see Supplemental File S1 [footnote 1]). For graphical

WCSS2 WSS5 Sampling and visual simplicity, all DZ visualization figures of samples taken from the same stratigraphic interval are created using the combination of DZ ages across We collected and analyzed 13 samples across the Marathon fold and thrust the entire sample suite and represented by a single kernel density estimation belt and Glass Mountains stratigraphic sections (Figs. 1 and 6). Standard min- (KDE) plot (Fig. 9). For example, VSS1–4 is a single representation of the four eral separation techniques for all samples proceeded as follows: crushing and Vidrio Formation samples: VSS1, VSS2, VSS3, and VSS4. The only exceptions grinding of whole-rock samples, water table separation, bromoform heavy to this are Figures 10 and 11 in which each individual sample is broken out and 1 Supplemental Material. Zipped file containing Sup- liquid separation, Frantz magnetic separation, and methyl iodide heavy liquid given unique values for the purpose of MDA and multidimensional scaling plemental File S1: A PDF of detailed information re- garding the process of data representation and pro- separation. Heavy mineral separates were poured onto double-sided tape (MDS) visualization. MDA estimates (see Supplemental File S2) were derived cessing involved in the study and three secondary on epoxy resin mounts, zircon grains were then chosen at random for laser using the following rules: youngest single grain (YSG), weighted-mean age of standard concordia age plots, and a tabulated Excel ablation–​inductively coupled plasma mass spectrometry (LA-ICP MS) to obtain youngest cluster of two or more grain ages overlapping age at 1σ (YC1σ (2+)), table with secondary standard U-Pb measurements; Supplemental File S2: A PDF file of two sets of max- U-Pb ages (Hart et al., 2016). In order to ensure representation of grain com- and weighted-mean age of the youngest cluster of three or more grain ages imum depositional age (MDA) measurements of 13 ponents comprising >5% of the total sample population is achieved at 95% overlapping in age at 2σ (YC2σ (3+)) (Dickinson and Gehrels, 2009). However, individual samples (MDA plots gathered from detrital confidence and ensure accuracy of maximum depositional age (MDA), 120–140 for this study, we focus on the youngest single grain to emphasize the earliest zircon [DZ] U-Pb data and processed and plotted in grains were picked in each sample as per Vermeesch (2004). possible lag time between zircon crystallization and deposition. More infor- detritalPy [Sharman et al., 2018]); and Supplemental File S3: A tabulated Excel file with 10 tables: Table S1: mation regarding the plotting software and data representation can be found TSS1 and TSS2 sample U-Pb analyses; Table S2: HF1 in their respective sections of Supplemental File S1. and HF2 sample U-Pb analyses; Table S3: WCSS2 Analytical Methodology sample U-Pb analyses; Table S4: WCSS1 and LSS1 sample U-Pb analyses; Table S5: WSS5 and VSS4 sample U-Pb analyses; Table S6: VSS3 and ALSS1 Analyses were run on a PhotonMachine Analyte G.2 excimer laser with a Core-Rim Analysis sample U-Pb analyses; Table S7: VSS1 and VSS2 large-volume Helex sample cell and a Thermo Element2 ICP-MS. GJ1 was used sample U-Pb analyses; Table S8: Sample locations; Table S9: Sample sheet set up for detritalPy (Shar- as the primary reference standard (Jackson et al., 2004) and Pak1 (in-house Depth profiling of each individual zircon grain yielded 85 concordant core- man et al., 2018); and Table S10: Best age and 1σ thermal ionization mass spectrometry [TIMS] 206Pb/238U age of 43.03 ± 0.01), rim age relationships with three analyzed zircons having multiple concordant error of sample analyses set up for detritalPy (Shar- Plesovice zircon (337.13 ± 0.37; Sláma et al., 2008), and 91500 (1065 ± 0.4 Ma, rims. That is to say, of the 1721 detrital zircon U-Pb ages, 85 U-Pb age pairs man et al., 2018). Please visit https://doi.org​/10.1130​ /GES02108.S1 or access the full-text article on www​ Wiedenbeck et al., 1995) were used as secondary standards to ensure data (Fig. 12) were extracted from both cores of zircons as well as igneous or .gsapubs.org to view the Supplemental Material. quality and procedural integrity. Secondary standard data and weighted means metamorphic rims (Moecher and Samson, 2006; Marsh and Stockli, 2015).

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 575 by guest on 30 September 2021 Research Paper

100

80 ALSS1, N=(1, 144/144)

Guadalupian VSS1-4, N=(4, 569/589) 60 WSS5, N=(1, 138/146)

40 Leonardian LSS1, N=(1, 116/139) Wolfcampian WCSS1-2, N=(2, 285/298) Cumulative % 20 Pennsylvanian HF1-2, N=(2, 209/212)

Mississippian TSS1-2, N=(2, 227/230) 0 Guadalupian Altuda Formation ALSS1 10 Post-Tectonic

0 Figure 9. Cumulative probability density plot and Guadalupian Vidrio Formation VSS1-4 25 non-uniform kernel density estimation plots of each Post-Tectonic sample suite analyzed in this study. Highlighted col- or-coded groups designate broad-scale potential 0 detrital zircon source terranes and dominant age WSS5 10 range from each source (Fig. 8) (Lawton et al., 2015b). Guadalupian Word Formation Dark blue—Mississippian Tesnus Formation (TSS1 Post-Tectonic and TSS2); light blue—Pennsylvanian Haymond 0 Formation (HF1 and HF2); yellow—Wolfcampian Lenox Hills Formation (WCSS1 and WCSS2); or- LSS1 10 Leonardian Skinner Ranch Formation ange—Leonardian Skinner Ranch Formation (LSS1); Late Syn-Tectonic light red—Guadalupian Word Formation (WSS5); red—Guadalupian Vidrio Formation (VSS1, VSS2, 0 VSS3, and VSS4); dark red—Guadalupian Altuda WCSS1-2 Formation (ALSS1). All colors coordinate with the Wolfcampian Lenox Hills Formation 10 Syn-Tectonic given stratigraphic column (Fig. 6). 0 Detrital Zircon Count Pennsylvannian Haymond Formation HF1-2 10 Syn-Tectonic 0 TSS1-2 Mississippian Tesnus Formation 10 Pre-Tectonic 0 0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400 Age (Ma)

Archean and Paleoproterozoic (>1825 Ma) Neoproterozoic and Early Cambrian (920-521 Ma) Late Paleoproterozoic (1825-1578 Ma) Paleozoic (521-263 Ma) Early Mesoproterozoic (1578-1300 Ma) Rodinian Rift Volcanics (780-540 Ma) Mesoproterozoic (1300-920 Ma)

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 576 by guest on 30 September 2021 Research Paper

240 1:1 line

0 My 260 ALSS1: YC1σ ALSS1: YSG VSS4: YSG VSS3: YSG VSS4: YC1σ VSS3: YC1σ WSS5: YC1σ VSS2: YSG VSS1: YSG VSS2: YC1σ VSS1: YC1σ 37 My LSS1: YSG WSS5: YSG WCSS1: YC1σ 35 My WCSS1: YSG 280

Age (Ma) 39 My WCSS2: YC1σ WCSS2: YSG

HF1: YSG 300 HF1: YC1σ HF2: YSG HF2: YC1σ Figure 10. Maximum depositional age from detrital zircon U-Pb versus stratigraphic dep- 43 My ositional age from publication (Olszewski and Erwin, 2009; Glasspool et al., 2013; Richards, 2013; Nestell et al., 2019). TSS1: YSG TSS1: YC1σ TSS2: YSG 320 TSS2: YC1σ 55 My Stratigraphic Depositional 340

360 240 260 280 300 320 340 360 380 400 420 U-Pb Maximum Depositional Age (Ma)

A more comprehensive history of zircon provenance and recycling can be Marathon fold and thrust belt samples and nine Glass Mountain samples were gleaned by gathering multiple ages from single zircon grains. Specifically, condensed into seven sample suites (grouped based on multiple samples in the these data help fingerprint a narrower range of potential sources by incorpo- same stratigraphic unit) and are shown as KDE plots, histograms, and cumu- rating core-rim age relationships with previously published DZ data within lative probability density plots of Figure 9. Highlighted color-coded groups potential sediment source terranes. All core-rim data are from the Permian designate possible detrital zircon source terranes and dominant age range from strata of the Glass Mountains, while no such relationships were found from each source (Fig. 8) (Lawton et al., 2015b). On the basis of these possible source the zircon of the Marathon fold and thrust belt. When plotted, these analy- ages, detrital zircon U-Pb ages can be separated into six groups: Archean and ses cluster into five relative core and rim clusters, each with 1–3 subgroups Paleoproterozoic (>1825 Ma), late Paleoproterozoic (ca. 1825–1578 Ma), early (Fig. 12) representing potential zircon growth and recycling history related to Mesoproterozoic (ca. 1578–1300 Ma), Mesoproterozoic (ca. 1300–920 Ma), Neo­ specific sediment source terranes. protero­zoic to Early Cambrian (ca. 920–521 Ma), and Paleozoic (ca. 521–263 Ma).

■■ RESULTS Pre-Permian (Mississippian–Pennsylvanian)​

In total, 1721 detrital zircon U-Pb ages were measured in this study. All The pre-Permian, syn-orogenic samples of the Marathon fold and thrust sample analytical data, sample locations, and detrital Py tables can be found belt (blue-colored samples in Figs. 1 and 9) exhibit two distinct detrital zircon in Supplemental File S3 [footnote 1]. Detrital zircon U-Pb data from the four age spectra for the Mississippian Tesnus and the Pennsylvanian Haymond

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 577 by guest on 30 September 2021 Research Paper

formations. Two samples from massive sandstone beds of the Mississippian Tesnus Formation (TSS1 and TSS2, n = 227) were taken along U.S. Highway 90, 0.2 east of Marathon, Texas, within the Marathon fold and thrust belt (Fig. 1). The 0.15 ALSS1 DZ U-Pb age spectrum (Fig. 9) exhibits a dominant Mesoproterozoic DZ U-Pb age component (51.1%), with a prominent age peak at ca. 1037 Ma. There exist 0.1 minor fractions of detrital zircons of Archean and Paleoproterozoic (10.6%), VSS3 Late Paleoproterozoic (14.5%), early Mesoproterozoic (12.3%), Paleozoic (11%) 0.05 VSS2 TSS2 HF2 ages, and a single Neoproterozoic age zircon of 588 ± 5.7 Ma. Two turbiditic WSS5 VSS4 0 sandstone samples from the Pennsylvanian Haymond Formation (HF1 and TSS1 VSS1 HF2, n = 209) from along U.S. Highway 90, east of Marathon, Texas, exhibit a -0.05 HF1 markedly different detrital zircon age spectrum with Neoproterozoic and Early WCSS2 Cambrian ages totaling 20.6%. The Mesoproterozoic age component continues WCSS1 -0.1 to dominate and accounts for 36.4% of measured detrital zircons, while the LSS1 Archean and Paleoproterozoic (14.8%), late Paleoproterozoic (7.2%), early Meso- -0.15 proterozoic (7.7%), and Paleozoic (13.4%) age components are minor fractions. -0.2

-0.2 -0.1 0 0.1 0.2 0.3 Early Permian (Wolfcampian–​Leonardian)

The younger, early Permian strata of the Glass Mountains (Fig. 1), north of 0.2 Marathon, Texas, overlie the deformed strata of the Marathon fold and thrust belt. Two Wolfcampian samples (WCSS1 and WCSS2, n = 297) were taken 0.15 ALSS1 Post-Collisional from the conglomeratic Lenox Hills Formation of Leonard Mountain (Fig. 1), and the Leonardian sample (LSS1, n = 138) was taken from the massive - 0.1 Pre- and Tectonic Stabilization stone turbidites of the Skinner Ranch Formation. These samples exhibit less VSS3 Syn-Collisional Syn-Collisional dramatic changes in detrital zircon age distributions compared to changes 0.05 VSS2 TSS2 WSS5 VSS4 HF2 observed between the Tesnus and Haymond samples. The DZ U-Pb age spectra 0 TSS1 of the Lenox Hills Formation exhibit minor age components of Archean and VSS1 HF1 Paleoproterozoic (9.4%), late Paleoproterozoic (5.4%), and early Mesoprotero- -0.05 WCSS2 zoic (5.4%). Paleozoic zircons make up 16.2% of total DZ U-Pb ages, while the Tesnus Deposition WCSS1 Distance/Disparity -0.1 Meso­protero­zoic (32.3%) and Neoproterozoic to Early Cambrian (31.3%) age LSS1 Orogenic Laurentian components represent the majority of detrital zircon ages with prominent age Unroofing -0.15 Margin Uplift peaks at ca. 1090 Ma and ca. 524 Ma, respectively. The DZ age spectrum of the Skinner Ranch Formation (LSS1) is similar to that of the Lenox Hills (WCSS1 -0.2 and WCSS2), with age components of Archean and Paleoproterozoic (9.4%), late Paleoproterozoic (2.2%), early Mesoproterozoic (5.1%), Mesoproterozoic -0.2 -0.1 0 0.1 0.2 0.3 (30.4%), Neoproterozoic and Early Cambrian (40.6%), and Paleozoic (12.3%). Dissimilarity

Figure 11. Multidimensional scaling (MDS) map of study samples. Greater distance rep- resents greater degrees of dissimilarity within the sample set as measured using likeness. Middle Permian (Guadalupian) (A) Uninterpreted MDS map showing the relative dissimilarity of samples. (B) Interpreted MDS map of samples, dark black lines show transition in pre-, syn-, and post-collisional The middle Permian (Guadalupian) samples were taken from mixed car- stages while blue arrows represent specific stages of sediment supply and regional pro- cesses. MDS formatted using Vermeesch (2013) Matlab code. Dark blue—Mississippian bonate-clastic turbidites of the Word (WSS5), Vidrio (VSS1, VSS2, VSS3, and Tesnus Formation (TSS1 and TSS2); light blue—Pennsylvanian Haymond Formation (HF1 VSS4), and Altuda formations (ALSS1) north of Marathon, Texas, and farthest and HF2); yellow—Wolfcampian Lenox Hills Formation (WCSS1 and WCSS2); orange—Leon- north of the sampled Glass Mountains section (Fig. 1). The single sample of ardian Skinner Ranch Formation (LSS1); light red—Guadalupian Word Formation (WSS5); red—Guadalupian Vidrio Formation (VSS1, VSS2, VSS3, and VSS4); dark red—Guadalupian the Word Formation (WSS5, n = 138) was taken from the turbidite outcrops Altuda Formation (ALSS1). All colors coordinate with the given stratigraphic column (Fig. 6). north of Leonard Mountain, and is lowest in the Guadalupian stratigraphic

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 578 by guest on 30 September 2021 Research Paper

2500 Core Rim A Peri-Gondwana PG or E. Mex Arc B Rodinia I/II or Pan-Africa PG, P-A, or E. Mex Arc 1:1 line C1 Grenville E. Mex Arc C2 Grenville PG C3 Grenville P-A or Rodinian Rifting C4 Grenville Grenville 2000 D1 Granite-Rhyolite or Amazonia PG, P-A, or Rodinia Rifting D2 Granite-Rhyolite or Amazonia Grenville E1 Yavapai-Mazatzal or Amazonia PG, or P-A E2 Yavapai-Mazatzal or Amazonia Grenville

ALSS1 LSS1 Figure 12. Plot of measured core and rim ages produced from U-Pb age measurements of detrital zircon depth VSS1-4 WCSS1-2 profiles. Alphabetic and numerically labeled boxes group 1500 together interpreted source terranes based on core-rim WSS5 relationships. A grain plotting close to the 1:1 line is indicative of similar crystallization age of core and rim components. For the filled dots: yellow—Wolfcampian Lenox Hills Formation (WCSS1 and WCSS2); orange— Age (Ma) D2 Leonardian Skinner Ranch Formation (LSS1); light C4 red—Guadalupian Word Formation (WSS5); red—Guada- Rim lupian Vidrio Formation (VSS1, VSS2, VSS3, and VSS4); 1000 dark red—Guadalupian Altuda Formation (ALSS1). All colors coordinate with the given stratigraphic column E2 (Fig. 6). PG—peri-Gondwana; P-A—Pan Africa.

C3 D1 B

E1 500 C2 A

C1

500 1000 1500 2000 2500 Core Age (Ma)

section. The DZ U-Pb spectra of the Word Formation exhibit an increase in were taken from turbidites within Road Canyon, farther north of VSS1 and Paleozoic age component (27.7%) and a decrease in Neoproterozoic to Early VSS2 (Fig. 1). The compiled age spectra (Fig. 9) exhibit similar components Cambrian age component (28.5%) relative to the Wolfcampian and Leonardian to that of the Word Formation, with high fractions of Paleozoic (27.1%), Neo- samples. The Mesoproterozoic age component accounts for 24.1% of DZ U-Pb proterozoic to Early Cambrian (21.6%), and late Mesoproterozoic (26.9%) age ages, and the early Mesoproterozoic (9.5%), late Paleoproterozoic (5.8%), and components, and low fractions of early Mesoproterozoic (9.5%), late Paleop- Paleoproterozoic and Archean (4.5%) age components maintain minor frac- roterozoic (7.7%), and Paleoproterozoic and Archean (7.2%) age components. tions of the age spectrum. The four Vidrio Formation samples (VSS1, VSS2, The Altuda Formation (ALSS1, n = 144) marks the youngest stratigraphic unit VSS3, and VSS4, n = 569) were taken in two separate locations. The first two sampled in this study. The sample was taken from a massive sandstone tur- samples (VSS1 and VSS2; n = 276) were taken above the Word sample, north bidite outcrop just northwest of Gilliand Canyon (King, 1930). The DZ U-Pb age of Leonard Mountain, while the second two samples (VSS3 and VSS4; n = 293) spectrum remains similar to the previous Word and Vidrio samples, but with

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 579 by guest on 30 September 2021 Research Paper

a noticeable enrichment in Paleozoic zircons (34.7%) and relative decrease in Grenvillian Cores amount of late Mesoproterozoic zircons (14.6%). Grenvillian zircons are ubiquitous in the Neoproterozoic–​Mesozoic sand- stones of North America (Dickinson and Gehrels, 2003; Moecher and Samson, ■■ ZIRCON CORE-RIM AND ANALYSIS OF MAXIMUM 2006), impeding provenance interpretations for these detrital age components. DEPOSITIONAL AGE However, depth profiling of zircons from this study (C core groups, Fig. 12) presents some revealing data. The oldest rims of the Grenvillian-core suite Here, we present a group-by-group breakout and interpretation of the are of contemporaneous Grenvillian age, shown in the C3 subgroup (Fig. 12). five separate core-rim age clusters. Each grouping is first separated by color Provenance of this group is difficult to decipher due to the potential for core on the basis of core age (x axis). Within these core age groupings, unique and rim generation within a Grenvillian province. However, subgroup C2 dis- subgroups have been further broken out according to rim ages (y axis). The plays Grenvillian cores with rims of Neoproterozoic–middle​ Paleozoic age. combination of a unique core and rim age sheds light on potential source and Magmatic overprinting of 650–550 Ma volcanism within rocks of the Avalon, recycling history (Fig. 12). Carolina, and Suwannee terranes (Heatherington et al., 1996; Samson et al., 2005; Moecher and Samson, 2006) could be potential sources for these zircons, but this is less likely for the Suwannee terrane because its DZ age spectra are Paleoproterozoic and Archean Cores not well represented by Grenvillian-age zircons (Mueller et al., 1994). A more likely source for these multi-age zircons would be the terranes of Coahuila Zircon core crystallization ages of Paleoproterozoic and Archean age could and Oaxaquia. Lopez et al. (2001) dated Grenvillian and Pan-African rocks have been sourced from a broad range of terranes within and attached to the within basement terranes of Coahuila and Oaxaquia. Zircon U-Pb analyses Laurentian and Gondwanan continents, but specifically the Amazonian terranes of the Grenville rocks presented grains of 1201 ± 60 Ma to 1232 ± 7 Ma with of the Gondwanan hinterland (Cordani et al., 2009) and Yavapai-Mazatzal, Wyo- concordia intercepts at ca. 580 Ma, and Pan-African rocks presented grains ming, Superior, and Penokean provinces of northern Laurentia (Anderson and of 580 ± 4 Ma. These were interpreted as being interrelated through melting Morrison, 1992). The E1 group, with core ages ranging ca. 1770–1560 Ma and of the Grenvillian basement during Pan-African–age​ volcanism, supporting rim ages ranging ca. 600–300 Ma, captures potential core ages from Amazonia our observation of similar core-rim relationships within single zircons (sub- or Yavapai-Mazatzal but has much younger rim ages. These rim ages are most group C2, Fig. 12). Additionally, these ages line up well with volcanism from likely representative of peri-Gondwanan or Pan-African origin, suggesting a the rifted Laurentian margin (780–560 Ma), supporting the interpretation that combined core-rim provenance history of southern or eastern origin. The single during orogenesis, the continental margin would have been included into zircon grain within E2 (Fig. 12) contains a 1641 ± 33 Ma core with a 921 ± 71 Ma the accreted wedge. Subgroup C2, which plots within subgroup C3, has rim rim. A core of this age would typically be grouped with the Yavapai-Mazatzal ages of 422 ± 17 Ma and 432 ± 24 Ma. These rim ages are characteristic of the province of Laurentia; however, the Grenvillian-age rim correlates with the peri-Gondwanan terranes of Mixteca (Acatlán) (480–440 Ma; metamorphic younger Grenville-age basement of Oaxaquia as presented by Gillis et al. (2005). ages of 416–388 Ma) and Yucatan-Maya (418 Ma).

Early–Middle Mesoproterozoic Cores Neoproterozoic–Early Paleozoic Cores

Zircon core-group D falls between the early–​middle Mesoproterozoic age Group B contains zircon cores of Neoproterozoic–​early Paleozoic ages ranges, a range that could be represented by either the granite-rhyolite prov- and rims of Neoproterozoic–late Paleozoic ages. The core ages are likely to inces of the North American interior craton (Bickford et al., 1986; Hoffman et al., be of Rodinian volcanic origin (Li et al., 2008) or potentially from Pan-African 1989; Anderson and Morrison, 1992) or Amazonian basement and recycled terranes (Heatherington et al., 1996; Lopez et al., 2001; Samson et al., 2005; terranes (Geraldes et al., 2001; Santos, 2003; Cordani et al., 2009). Subgroups Moecher and Samson, 2006). Notably, the top right-hand cluster of core-rim D2 and D1 of the plotted core-rim ages (Fig. 12) show concordant rims of ages within group B trend close to the 1:1 line, indicating secondary anatec- Ectasian–​Tonian and Tonian–​Carboniferous age ranges, respectively. Group tic growth and crystallization, relatively soon after initial crystallization. This D2 contains Grenvillian age rims, which could potentially be related to either kind of growth behavior could be indicative of the first intermittent volcanic Laurentian or Gondwanan magmatic or metamorphic activity. Zircons within pulses following the breakup of Rodinia. On the other hand, the cluster at the the subgroup D1 have a broad range of rim ages with the youngest at 343 ± bottom left-hand corner of group B contains younger cores, which appear to 17 Ma, an age that potentially adheres to volcanic activity or tectonism within have Pan-African sourcing and a secondary growth related to the younger, the peri-Gondwanan terranes. peri-Gondwanan terranes.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 580 by guest on 30 September 2021 Research Paper

Paleozoic Cores Pennsylvanian–Permian strata, while minimum lag times deduced from the youngest zircon components are nearly constant at ~30–40 m.y. Due to correla- Core-rim relationships with purely Paleozoic cores are shown in group A tion between plutonic arc exhumation and calculated lag time, we interpret the (Fig. 12). These are the youngest core-rim relationships and are almost certainly zircons with lag times of 30–40 m.y. to have been derived from arc magmatic indicative of the peri-Gondwanan terrane and/or volcanic arc activity within rocks emplaced along the Gondwanan convergent margin during closure of the peri-Gondwanan, Coahuila, or Oaxaquian terranes (Fig. 8). We suggest rim the Rheic Ocean in Carboniferous time (Torres et al., 1999; Keppie et al., 2008). growth to be indicative of zircon recycling and inclusion in the volcanic arc In summary, based on the differences between true depositional age and zir- activity within the Gondwanan hinterland and subsequent transport during con-based MDA, there are three dominant sources of zircons in the southern the collision of Gondwana and Laurentia. Permian Basin: (1) multi-cycle zircons eroded and recycled from Proterozoic to early Paleozoic or younger Gondwanan or Laurentian strata; (2) plutonic sources erosionally unroofed within the hinterland arc related to Rheic closure; Maximum Depositional Ages and (3) first-cycle volcanic zircons derived from active volcanic arc magmatism in the middle Permian. In particular, the presence of Carboniferous arc mag- The youngest zircon grain age or youngest age mode is a powerful method matic zircons with relatively constant lag times is noteworthy because they for determining maximum depositional ages (MDAs), particularly in the unequivocally point to derivation from the Gondwanan convergent margin absence of biostratigraphic age constraints (Dickinson and Gehrels, 2009). and to protracted unroofing of this magmatic arc. Furthermore, the presence This approach is predominantly limited by the generation of new, first-cycle of these ages suggests essentially continuous Devonian–Carboniferous arc volcanic zircons and their airborne or near-instantaneous fluvial input into magmatism related to Rheic closure. This is noteworthy given the sparse the basinal strata. These MDA estimates not only provide chronostratigraphic record of this magmatism along the Gondwanan margin from Florida to NE control, but also important insights into the presence or absence of local and/or Mexico, including the Las Delicias Arc (Steiner and Walker 1996; McKee et al., regional volcanic or plutonic sources as well as erosional lag-time constraints 1999; Lopez et al., 2001; Keppie, 2004; Keppie et al., 2008; Soreghan and (i.e., the difference between MDA and depositional age) for independently Soreghan, 2013). dated strata (Rahl et al., 2007). MDA interpretations by means of the youngest grain ages in a stratigraphic unit have been used to determine the lag time between zircon crystallization and delivery into a sedimentary basin. This ■■ PROVENANCE EVOLUTION observation assists with the differentiation of volcanic, plutonic, or recycled and multi-cycle zircons and their sources (e.g., Dickinson and Gehrels, 2009; The detrital zircons from the Marathon fold and thrust belt and Glass Xu et al., 2017). While volcanic and plutonic zircons are commonly character- Mountains represent a wide range of crystallization ages and a complex ized by lag times of <1 and <20–40 m.y., respectively, multi-cycle zircons often history of sediment recycling (Fig. 9). By comparing the DZ U-Pb signa- exhibit lag times >>50 m.y. tures with respect to deposition during pre-, syn-, or post-collisional time Integrating MDA values with coeval biostratigraphic and sequence strati- windows, we are able to interpret and reconstruct proximal basin sedi- graphic information from previous regional studies in the Glass Mountains mentary evolution during progressive collision. Further, incorporation of and Marathon fold and thrust belt provides insight into regional volcanic other provenance studies regionally helps to establish a broader under- influence. Specifically, this comparison exhibits the presence of active volcanic standing of how sediment dispersal from the southern orogenic highlands arc magmatism, the unroofing of magmatic belts in the hinterland drainage and distal hinterland influenced sediment dispersal during foreland basin system, as well as the long-term recycling of pre-contractional sedimentary formation and fill. sequences in the source terrane. Figure 10 displays the MDAs of all 13 sam- A few key provenance trends can be inferred from the DZ U-Pb analyses ples plotted against their independent best stratigraphic ages, using both in this study: (1) appearance of Neoproterozoic–Cambrian zircon grains in the the YSG and YC1σ (2+) approaches (Dickinson and Gehrels, 2009; Xu et al., Pennsylvanian Haymond Formation points toward basin inversion and the 2017). Here, the independent best stratigraphic ages are based on regional uplift and exhumation of volcanic units related to Rodinian rifting; (2) upsection biostratigraphic and sequence stratigraphic data (Olszewski and Erwin, 2009; decrease in Grenvillian (ca. 1300–920 Ma) and increase in Paleozoic zircons Glasspool et al., 2013; Richards, 2013; Nestell et al., 2019). The results of this denote a steady provenance shift from a dominantly orogenic highland source analysis show that volcanic zircons with negligible lag times did not arrive to one of sediment sources deeper in the Gondwanan hinterland; and (3) lack in the basinal system until the middle Permian (Guadalupian), marking the of drastic changes in provenance during the early to middle Permian supports input and onset of Cordilleran or East Mexico Arc volcanism at ca. 260–270 Ma the argument for waning tectonism within the collisional domain during this (Dickinson et al., 2000; Dickinson and Lawton, 2001). Prior to middle Permian time period as does overlap of strata of those ages onto deformed rocks of time, however, no zero-lag-time volcanic zircons were recorded in any of the the orogenic front in the Glass Mountains (Fig. 2).

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 581 by guest on 30 September 2021 Research Paper

Mississippian–Pennsylvanian Provenance Evolution pre-collisional nature of the Tesnus Formation. Specifically, the peri-Gond- wanan terranes of Mixteca (Acatlán) (480–440 Ma; metamorphic ages of Previous DZ provenance studies of the Permian Basin focused their efforts 416–388 Ma) and Yucatan-Maya (418 Ma) function as the closest match to on resolving the age spectra of the Delaware Mountain Group, which includes the DZ U-Pb age modes of the Tesnus age spectra (Steiner and Walker, 1996; the Brushy, Cherry, and Bell Canyon formations of the Guadalupe Mountains Keppie, 2004; Gillis et al., 2005; Steiner and Anderson, 2005; Talavera-Men- and Delaware Basin (Soreghan and Soreghan, 2013; Anthony, 2015; Xie et al., doza et al., 2005; Keppie et al., 2006, 2008; Weber et al., 2006, 2008; Martens 2018). While these studies made significant contributions to the understanding et al., 2010). By applying these age domains and collisional timing based on of provenance during basin fill of the Delaware Basin, they only capture prov- paleogeographic and tectonic reconstructions of the peri-Gondwanan terranes enance during post-tectonic sedimentation. Results from this study evaluate (Keppie, 2004; Centeno-García et al., 2005; Weber et al., 2008) previous prove- the evolutionary component of southern provenance during basin formation nance studies argue for initial denudation of the peri-Gondwanan terranes to and fill. By defining the evolution of DZ age spectra over an extensive period be during the middle Permian (Soreghan and Soreghan, 2013; Anthony, 2015; of tectonism and sediment transport, results from this study help to resolve Xie et al., 2018). Alternatively, based on the measured Paleozoic DZ ages from long-term trends in provenance observed both locally and regionally. the Tesnus Formation, denudation and incorporation of sediment from these terranes were likely to have initiated in the Carboniferous, during migration of the Gondwanan continent. Tesnus Formation

A wide range of age modes is present in the Tesnus DZ spectra (Fig. 9) with Haymond Formation a notably high Grenvillian (ca. 1300–920 Ma) age mode and absence of a middle Neoproterozoic–Cambrian (ca. 920–521 Ma) age mode. Because of its presence Deposited during ongoing tectonism, the sediment source history of the in the rest of the samples throughout this study, the lack of Neoproterozoic– Haymond Formation has been controversial (Denison et al., 1969; King, 1958; Cambrian zircon grains is conspicuous. We interpret the Tesnus Formation to McBride, 1989; Gleason et al., 2007). As stated previously, we can effectively be completely sourced from the southern Gondwanan continental assemblage. use the Tesnus Formation DZ U-Pb age spectra as a baseline for the younger Using the Tesnus as a baseline for provenance evolution helps in understand- strata sampled, and, based on shifts in provenance, we are able to make ing provenance shifts within the younger strata of the Marathon fold and thrust inferences regarding evolution of the collisional and foreland system. In this belt and Glass Mountains. case, the Haymond DZ U-Pb age spectra exhibit a marked change in age Grenville Signature. The Grenville orogen, present in southern and east- population with the appearance of the previously absent Neoproterozoic– ern North America, Mexico, and the Gondwanan continents (Rainbird et al., Cambrian age mode. It is important to note that this is the only significant 1992; Rainbird et al., 1997; Soreghan and Soreghan, 2013), typically makes up provenance difference between the Tesnus and Haymond DZ spectra. This a significant portion of the DZ population in North American sandstones from change in provenance denotes the inversion and exhumation of the distal Neoproterozoic to Mesozoic (Dickinson and Gehrels, 2003; Moecher and Sam- Laurentian margin during Pennsylvanian collision with Gondwana. Further, son, 2006), adding difficulty to resolution of true source for these age modes. this defines the Haymond age spectra as a transitional unit between pre- and Grenvillian zircons within the Tesnus are no different, constituting >50% of syn-orogenic units. the DZ population. However, as stated in a previous section, depth profil- Rifting of the Rodinian supercontinent occurred between ca. 825 Ma and ing of Grenvillian zircons can make constraining their original source region 740 Ma with notable volcanic pulses from 780 to 540 Ma (Li et al., 2008; Dick- possible. Gillis et al. (2005) presented a compilation of DZ U-Pb ages from inson and Gehrels, 2009; Hanson et al., 2016). Strata deposited along the the Oaxaquian crystalline basement terrane and Paleozoic strata of southern southern passive margin of Laurentia covered these ancient syn-rift volcanic Mexico, which showed strong age peaks of 981 Ma and 993 Ma, respectively. rocks (King, 1937, 1978; Flawn et al., 1961; McBride, 1989; Hickman et al., 2009). When compared to Grenville ages of southwestern and northeastern North Therefore, the most viable method to explain inclusion of these volcanic units America, the Oaxaquian DZ U-Pb ages are consistently younger (<1100 Ma) into younger strata of the Marathon fold and thrust belt (and eventually the (Gross et al., 2000; Stewart et al., 2001; Eriksson et al., 2003; Gillis et al., 2005). southern Delaware Basin) would be through uplift of the deep-seated con- For this reason, we attribute the most prominent Grenvillian DZ peak within tinental margin. We propose the formation of a duplex zone between the the Tesnus age spectra (1037 Ma) to the Oaxaquian source terrane during Gondwanan continent and Laurentian margin (e.g., Devils River Uplift) (Fig. 13), pre-collisional sediment transport. to exhume syn-rift volcanic rocks. Therefore, the Haymond Formation is most Paleozoic Signature. Zircons of Paleozoic age are the youngest among likely composed of a combination of southern remnant-ocean basin sediment Tesnus age spectra with a prominent peak at ca. 422 Ma. Peri-Gondwa- similar to that of the Tesnus Formation and uplifted strata of the ancient Lau- nan terranes (Fig. 8) are the most likely sediment sources because of the rentian passive margin during the onset of continental collision.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 582 by guest on 30 September 2021 Research Paper

Late Guadalupian ca. 263 Ma 0 100 200 mi 0 100 200 300 km Inferred eolian transport Erosion of Orogenic Mass, N-NW Clinoform Progradation from Soreghan and Soreghan (2013) N Early—Middle Permian S Laurentia Gondwana Midland Basin N-NW progradation Erosion transport of Marathon Hinterland stabilization of clinoforms Fold and Thrust Belt strata and drainage organization Delaware Basin Leonardian—Guadalupian Final thrusting prior to Exhumation and denudation of Neoproterozoic clinoforms tectonic quescence to Early Cambrian rift volcanics

Ouachita-Marathon trace

Hinterland drainage stabilization Active Las Delicias Arc

Early Wolfcampian ca. 295 Ma 0 100 200 mi Laurentian Margin Duplexing, Orogen Formation 0 100 200 300 km Middle Pennsylvanian—Early Permian N S Laurentia Gondwana Delaware Midland Basin Basin Incorporation of rifted Backarc volcanism, Foreland deformation Laurentian margin through compressive thrusting, and of L. Paleozoic units crustal duplex zones Retroarc basin formation

Hinterland reverse thrust faults E. Mexico Arc/Las Delicias Arc Proximal orogenic Extinct arc wedge erosion

Duplexing of ancient Active Las Laurentian margin Delicias Arc e.g. Devils River Uplift

Pennyslvanian Des Moines ca. 308 Ma 0 100 200 mi 0 100 200 300 km

Continental Collision N Early—Middle Pennsylvanian S Laurentia Gondwana

Overriding of L. Paleozoic Laurentian strata Continued collision and recycling of remnant ocean deposits (Rheic Ocean)

Decollement surface Early foreland basin system

35 km Haymond deposition 35 km 50 km Forearc basin

Las Delicias Arc

Mississippian Chesterian ca. 325 Ma 0 100 200 mi 0 100 200 300 km Tesnus Deposition N Late Mississippian—Early Pennsylvanian S Laurentia Gondwana Laurentia Laurentian passive N-S deposition margin deposition of Tesnus Fm. Initiation of backarc volcanism Tobosa Basin Accretionary Wedge Laurentian rifted margin Gondwanan Forearc Tesnus Fm. Basin

35 km 50 km 35 km

Laurentian passive margin 110 km Subducting Laurentian slab

Tesnus deepwater flysch

Figure 13. Tectonostratigraphic reconstruction of the evolving Laurentian-Gondwanan margin throughout the Mississippian–middle Permian based on detrital zircon U-Pb data accompanied with coeval plan view paleogeographic maps. Dark-brown shading is indicative of high relief, eroding landscapes, while light-brown shading is indicative of low-relief, depositional landscapes. Light blue to dark blue represents varying water depth with light blue being shallow and dark blue be- ing deep. Strata colors in the cross-sectional view are age equivalent with stratigraphic units of Figure 6. Paleogeographic figures adapted from Blakey Deep Time Maps. (A) Pre-collisional deposition of Tesnus Formation on the leading edge of approaching Gondwanan continent, deposition of thin-bedded limestone, shale, and chert. (B) Continental collision during the Pennsylvanian. (C) Collision of Laurentia and Gondwana, incorporation of rifted Laurentian crust through crustal duplexing. (D) Post-collisional figure representing tectonic quiescence, sediment supply from deeper within the Gondwanan hinterland, supporting progradation of stable clinoforms along the southern margin of the Delaware Basin. See text for further explanation.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 583 by guest on 30 September 2021 Research Paper

Glass Mountains, Central Delaware Basin Early–Middle Permian Deposits: Unroofing and Tectonic Quiescence and Guadalupe Mountains MDS The latest syn-orogenic deposits of Wolfcampian age and potentially first GlM: ALSS1 post-orogenic deposits of Leonardian age display DZ U-Pb age spectra similar 0.6 SEDB: JCT4946 to that of the Pennsylvanian Haymond Formation. We interpret this similarity to be indicative of ongoing unroofing of the orogenic highlands. The deposition 0.4 SEDB: JCT6068 GlM: WSS5 of 4200–5600 m of siliciclastic strata during Wolfcampian time correlates well with the rapid erosion of the massive accreted wedge adjacent to the basin (King, 1937; Flawn et al., 1961; Thomson and McBride, 1964; King, 1978). The 0.2 GlM: VSS1-4 age spectra observed within the Wolfcampian samples are consistent with the GIM: WCSS1-2 interpretation of Thomas et al. (2019) on potential sourcing of zircons from SEDB: JCT5872 GlM: LSS1 the Coahuila block of Mexico. Additionally, the Leonardian Skinner Ranch 0.0 SEDB: JCT6446 Formation is a sandy turbidite facies, much more distal than the Wolfcampian GdM: Yates conglomerates of the Lenox Hills Formation; yet the sediment source appears

GdM: Rader to remain the same. -0.2 Transfer from the unroofing sequence seen during the Wolfcampian– GlM: HF1-2 Leonardian depositional period to that of post-tectonic sediment transport is captured within the middle Permian deposits, but it is far more subtle than the

Distance/Disparity -0.4 GdM: BcSand previously observed provenance shifts. Tracking the Grenvillian and Paleozoic DZ U-Pb age components upsection shows this subtle, progressive shift in provenance with a decrease in the Grenvillian age component and increase -0.6 GdM: BcSilt in the Paleozoic age component (Fig. 9). This trend matches the expected DZ

GlM: TSS1-2 age population observed within the Mexican terranes of the peri-Gondwanan -0.8 hinterland (McKee et al., 1999; Lopez et al., 2001). Therefore, we interpret this provenance shift to be indicative of the third and final stage of proximal prove- nance evolution, consisting of tectonic quiescence within the collisional domain -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 and catchment stabilization resulting in sediment sourcing from deeper in the Dissimilarity hinterland (Fig. 13D). By this point in time, the large orogenic wedge had been denuded to the point that sediment sourced deeper within the hinterland was Archean and Paleoproterozoic (>1825 Ma) possible and could feed the N-NW progradation of the Guadalupian (middle Permian) deposits. Late Paleoproterozoic (1825-1578 Ma)

Early Mesoproterozoic (1578-1300 Ma)

Mesoproterozoic (1300-920 Ma) ■■ MULTIDIMENSIONAL SCALING OF THE MISSISSIPPIAN– PERMIAN STRATIGRAPHY Neoproterozoic and Early Cambrian (920-521 Ma)

Paleozoic (521-263 Ma) Multidimensional scaling of the detrital zircon data is employed to compare the similarity of the provenance signature of different stratigraphic units and Figure 14. Multidimensional scaling (MDS) map from all Delaware Basin samples. to investigate the provenance evolution through time from before, during, and Data compiled from this study, Soreghan and Soreghan (2013), and Anthony (2015). Abbreviations: GlM—Glass Mountains sample; GdM—Guadalupe after continental collision. Intersample MDS plots (Vermeesch, 2013) (Figs. 11 Mountains sample; SEDB—southeastern Delaware Basin. Color key: dark blue— and 14) display dissimilarity between the DZ U-Pb age spectra. Sample clus- Mississippian Tesnus Formation (TSS1 and TSS2); light blue—Pennsylvanian tering indicates age spectra similarity, and samples plotting at distance are Haymond Formation (HF1 and HF2); yellow—Wolfcampian Lenox Hills Formation (WCSS1 and WCSS2); orange—Leonardian Skinner Ranch Formation (LSS1); light indicative of a higher degree of dissimilarity between age spectra. MDS analy- red—Guadalupian Word Formation (WSS5) and samples of coeval age from other sis of the Marathon fold and thrust belt and Delaware Basin sample set exhibits studies; red—Guadalupian Vidrio Formation (VSS1, VSS2, VSS3, and VSS4) and abrupt transitions in both the dominant DZ age modes as well as the entire DZ samples of coeval age from other studies; dark red—Guadalupian Altuda Forma- tion (ALSS1) and samples of coeval age from other studies. All colors coordinate U-Pb age spectrum (Fig. 11). The MDS plot shows three distinct stages of the with the given stratigraphic column (Fig. 6). evolution of the Marathon fold and thrust belt and southern Delaware Basin

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 584 by guest on 30 September 2021 Research Paper

foreland deposition: (1) pre-collisional deposition of the Tesnus Formation; testing previous hypotheses for sediment sourcing during the deposition of (2) transition into syn-collisional deposition of the Haymond Formation with the Guadalupian-age Delaware Mountain Group (Fig. 14). Here, we compare Laurentian margin uplift; and (3) tectonic quiescence and hinterland catch- our collected data from the southern Delaware Basin with the data of those ment maturity. authors from the northern and southeastern Delaware Basin (Figs. 14 and 15). The Tesnus Formation samples plot as a distinct grouping that is most Soreghan and Soreghan (2013) interpreted DZ data from the Brushy and dissimilar from the youngest (Guadalupian) post-collisional samples. As Bell Canyon members of the Guadalupe Mountain region to indicate sediment previously stated, we interpret the Tesnus to be of pre-collisional, southern sources from the proximal Ouachita foreland as well as sources south of the origin, and the degree of dissimilarity seen in the MDS analysis bolsters the accreted Ouachita margin. Anthony (2015) used DZ data from core in the argument that the Tesnus is unique in being a baseline for how a purely south- southeastern Delaware Basin, specifically from the Cherry and Bell Canyon ern-sourced DZ profile would look. The samples of the Haymond Formation members, and his interpretation aligns with Soreghan and Soreghan (2013). and Wolfcampian–Leonardian samples mark a sharp provenance difference Our documentation of provenance from coeval deposits from the most south- related to the orogenic activity and inversion of the Laurentian passive margin erly portion of the Delaware Basin provides an opportunity to evaluate these and, hence, the mixing of Gondwanan and Laurentian margin sources. The shift interpretations. While potential distal transport of Ouachita regional sources from the Tesnus Formation to the Haymond Formation appears to be marked is likely for both of these cases, the importance lies more with volumes con- by the sudden input of Laurentian passive margin elements, including a higher tributed. During margin inversion, large volumes of sediment were shed into concentration of older Grenvillian zircons (>1100 Ma) and Neoproterozoic zir- the southern Permian Basin (4200–5600 m in some places) (King, 1937, 1978; cons. Though it should be noted that while the Rodinian rift volcanics display Flawn et al., 1961; Thomson and McBride, 1964; McBride, 1989; Hickman et al., a distinct DZ age signature (780–560 Ma) that can be tied to the ancient Lau- 2009). This begs the question of where south-north transport of sediment ends rentian margin (Li et al., 2008; Hanson et al., 2016), there exists no complete and sediment transport from other regions surrounding the basin begins. Com- data set documenting the DZ signature of the Cambrian–Devonian Laurentian parison of samples from Soreghan and Soreghan (2013), Anthony (2015), and passive margin strata. These data will help in understanding the DZ mixing this study provides insight into this issue (Table 1; Figs. 14 and 15) (DZ age data component between the two colliding plates. Subsequently, from the Haymond can be found in Supplemental File S3 [footnote 1]). Figure 14 shows that the Formation to the Early Permian Lenox Hills Formation (Wolfcampian), the MDS sediment in the Glass Mountains is more similar to the southeastern portion of shows a relatively smooth trajectory from initial inclusion of the Laurentian the Delaware Basin than that of the northern Delaware Basin. The disparities in passive margin into the orogenic wedge and initial sediment dispersal from distance and DZ age spectra are highlighted in Figure 15, in which the number the orogenic highlands (Figs. 13C and 13D). of Neoproterozoic-age zircons within each sample suite decreases from south In MDS space, the DZ signature of strata deposited during post-collisional to north. It is important to remember that samples within our study effectively tectonic quiescence and their provenance evolution appear to differ strongly define the southern provenance signal for the entire formation period of the from the syn-orogenic trajectory (Fig. 11). The Word (WSS5), Vidrio (VSS1, Permian Basin as well as post-collisional deposition of Guadalupian age strata VSS2, VSS3, and VSS4), and Altuda (ALSS1) formations appear to plot as a in the southern Delaware Basin. Thus, higher similarity between samples of steady provenance shift from oldest to youngest, with the exception of VSS1 this study and those of Anthony (2015) and dissimilarity with those of Soreghan plotting closer to the Wolfcampian–Leonardian (WCSS1 and WCSS2, and LSS1, and Soreghan (2013) indicate a more diverse range of sources in the northern respectively) samples than the stratigraphically older Word (WSS5) sample. portion of the basin and more pronounced southern signal in some of the cen- These trends and trajectories observed in MDS space are clear and intuitively tral and south-central portions of the Delaware Basin (Anthony, 2015). More reflect the overall tectonic evolution of the system from convergence, to con- work is needed, likely from core within the northern portion of the basin and tinental collision, to post-collisional deposition (Fig. 11B). denser upstream sampling of regional sediment feeders from the NW and NE. However, we argue that the Guadalupian section within the southern portion of the basin was being fed almost entirely by proximal Marathon sources as ■■ DISCUSSION opposed to Ouachita region transport.

Integration of Regional Delaware Basin Provenance Mississippian–Permian Tectonics The results from the collisional domain and proximal foreland may be informed by comparison with samples collected regionally across the Del- Pre-tectonic strata of the Laurentian passive margin consisted of variable aware Basin. Soreghan and Soreghan (2013), Anthony (2015), and Xie et al. mechanically weak, thin-bedded intervals and strong chert layers, ultimately (2018) each performed provenance analyses with the use of DZ U-Pb geo- leading to the formation of the complex structural features observed in the chronology in different parts of the Delaware Basin. These studies focus on Marathon fold and thrust belt (Hickman et al., 2009). Studies have documented

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 585 by guest on 30 September 2021 Research Paper

100 0 100 km

80

Midland Basin 60 NorthwestShelf Central Basin 40

Platform Cumulative % Delaware Diablo Platform Glass Mountains Samples, N=(6, 850/878) Basin 20 Central/Southern Delaware Basin Samples, N=(4, 446/447) Guadalupe Mountains Samples, N=(4, 357/357) Val Verde 0 TX MX Basin Marathon- 0 200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400 This study Age (Ma) Anthony (2015) Soreghan and Soreghan (2013)

Figure 15. Cumulative distribution function (CDF) and kernel density estimation (KDE) intersample comparison of detrital zircon age spectra from Guadalupian (middle Permian) units across the Delaware Basin. Glass Mountains data (green line) are from this study, and CDF and KDE plots are composed of detrital zircon (DZ) from samples: WSS5, VSS1, VSS2, VSS3, VSS4, and ALSS1. Southeastern Delaware Basin data (red line) are from Anthony (2015) and comprise DZ from samples: JCT-6446, JCT-6068, JCT-5872, and JCT-4946. Guadalupe Mountains data (blue line) are from Soreghan and Soreghan (2013) and are composed of DZ from samples: BCsand, BCsilt, Bell Canyon (Rader), and Bell Canyon (Yates).

the unique mechanical stratigraphy and structural style of the fold and thrust This is a clear indication of the pre-collisional nature of the Tesnus Formation belt (DeMis, 1983; Tauvers, 1988; Muehlberger and Tauvers, 1989; Hickman and demonstrates that, while subduction of the Laurentian oceanic crust was et al., 2009). Our results clearly indicate temporal evolution of the Marathon possible, incorporation of even distal portions of the Laurentian marginal fold and thrust belt through tracking of sediment provenance evolution of sediment was unlikely. the pre-, syn-, and post-collisional deposits of Mississippian–Permian age. Incorporation of Neoproterozoic zircons during the deposition of the Hay- Notably, these tectonostratigraphic units can be separated into three distinct mond Formation suggests incorporation of the Laurentian rifted margin into phases of tectonic evolution and subsequent sediment transport and deposi- the uplifting orogenic wedge during the Pennsylvanian (Fig. 13C). Hickman et al. tion: (1) pre-tectonic deposition of the Mississippian Tesnus Formation; (2) uplift (2009) noted that the multiple décollement horizons within the pre-tectonic and exhumation of the passive Laurentian margin and ensuing unroofing of section of Marathon strata led to formation of a large-scale duplex zone, a likely the orogenic wedge; and (3) post-tectonic margin stabilization coupled with mechanism that formed the Devils River uplift (e.g., Nicholas and Rozendal, catchment maturity and hinterland sediment transport (Fig. 13). 1975; Nicholas and Waddell, 1989). This idea is bolstered by the appearance During early to middle Paleozoic time, a relatively thin marine sequence of what we interpret to be zircons from the volcanic rocks extruded during was deposited along the Laurentian passive margin (Fig. 13A), with indication Rodinian rifting phases of the Laurentian margin (780–540 Ma). This potential of a north or northwest source of clastic sediment supply (King, 1978; McBride, source is unaccounted for by previous regional provenance studies (Soreghan 1989). Mississippian deposition of the Tesnus Formation in the form of fly- and Soreghan, 2013; Anthony, 2015; Xie et al., 2018). The similarity in DZ U-Pb sch deposits is the first record of the approaching Gondwanan continental age spectra of the Wolfcampian–Leonardian strata is indicative of constant margin as displayed by its unique DZ U-Pb age spectra (Fig. 9). These age unroofing of the accreted orogenic wedge. Knowledge of the sheer volume spectra lack the entire section of Neoproterozoic-age zircons (Cryogenian– of sediment deposited over this time interval (King, 1937; Flawn et al., 1961; earliest Cambrian) with the exception of a single grain dated at 588 ± 5.7 Ma. Thomson and McBride, 1964; King, 1978; McBride, 1989; Hickman et al., 2009)

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 586 by guest on 30 September 2021 Research Paper

TABLE 1. STRATIGRAPHIC STAGE, FORMATION, LITHOLOGY, LATITUDE AND LONGITUDE, LOCATION, AND DATA SOURCE OF EACH SAMPLE FROM THE PERMIAN BASIN, WEST TEXAS, USA Sample ID Stage Formation Lithology Latitude Longitude Location Data source (°) (°) ALSS1 Guadalupian Altuda Siltstone and/or very fine-grained sandstone 30.383844 −103.304593 Southern Delaware Basin This study JCT-4946 Guadalupian Bell Canyon Siltstone and/or very fine-grained sandstone 31.304495 −103.145150 South Eastern Delaware Basin Anthony (2015) Bell Canyon (Yates) Guadalupian Bell Canyon Siltstone 32.282550 −104.363613 Northern Delaware Basin Soreghan and Soreghan (2013) Bell Canyon (Rader) Guadalupian Bell Canyon Sandstone 31.931592 −104.731086 Northern Shelf Soreghan and Soreghan (2013) VSS4 Guadalupian Vidrio Siltstone and/or very fine-grained sandstone 30.396649 −103.243585 Southern Delaware Basin This study

VSS3 Guadalupian Vidrio Siltstone and/or very fine-grained sandstone 30.396649 −103.243585 Southern Delaware Basin This study VSS2 Guadalupian Vidrio Siltstone and/or very fine-grained sandstone 30.381872 −103.231476 Southern Delaware Basin This study VSS1 Guadalupian Vidrio Siltstone and/or very fine-grained sandstone 30.381794 −103.231211 Southern Delaware Basin This study JCT-5872 Guadalupian Cherry Canyon Siltstone and/or very fine-grained sandstone 31.304495 −103.145150 South-eastern Delaware Basin Anthony (2015)

JCT-6068 Guadalupian Cherry Canyon Siltstone and/or very fine-grained sandstone 31.304495 −103.145150 South-eastern Delaware Basin Anthony (2015) JCT-6446 Guadalupian Cherry Canyon Siltstone and/or very fine-grained sandstone 31.304495 −103.145150 South-eastern Delaware Basin Anthony (2015) WSS5 Guadalupian Word Fine-grained sandstone 30.382164 −103.231126 Southern Delaware Basin This study BCsilt Guadalupian Brush Canyon Siltstone 32.848303 −104.839861 Northern Delaware Basin Soreghan and Soreghan (2013)

BCsand Guadalupian Brushy Canyon Sandstone 31.848303 −104.839861 Northern Delaware Basin Soreghan and Soreghan (2013) LSS1 Leonardian Skinner Ranch Sandstone 30.326233 −103.235566 Southern Delaware Basin This study WCSS2 Wolfcampian Lenox Hills Sandstone and/or conglomerate 30.323535 −103.234836 Southern Delaware Basin This study WCSS1 Wolfcampian Lenox Hills Sandstone and/or conglomerate 30.323490 −103.234193 Southern Delaware Basin This study

HF2 Pennsylvanian Haymond Sandstone 30.212437 −102.988188 Marathon fold and thrust belt This study HF1 Pennsylvanian Haymond Sandstone 30.208736 −102.984907 Marathon fold and thrust belt This study TSS2 Mississippian Tesnus Sandstone 30.204687 −102.973623 Marathon fold and thrust belt This study TSS1 Mississippian Tesnus Sandstone 30.204225 −102.967016 Marathon fold and thrust belt This study

combined with our knowledge of a consistent orogenic sediment source leads sediment supplied to the foreland basin during this time. Of the increasingly us to believe that the large volumes of syn-orogenic sediment deposited into sparse Grenvillian-age zircons, a few core-rim relationships exist (Fig. 12). the proximal foreland basin are dominated by a source in the orogenic wedge, Grenvillian cores with Pan-African rims ranging from 583 to 422 Ma suggest potentially over long periods of time (~60 m.y.). that Gondwanan terranes, especially those of Oaxaquia and Coahuila were The thickness of post-tectonic (Guadalupian) stratigraphy of the proximal likely candidates for sediment supplied to the foreland basin during this time. foreland basin is far more modest than that of the syn-orogenic deposits Finally, the youngest DZ U-Pb age modes within the Guadalupian age spectra (Fig. 13D). The DZ U-Pb age spectra of these samples suggest regional tec- are >300 Ma, indicating sediment sourcing from volcanic arc terranes, specif- tonic quiescence, aligning with a change in depositional environment during ically the Las Delicias Arc located within the Coahuila terrane (Figs. 13B and transgression of the southern margin of the Delaware Basin (King, 1930; Hair- 13C) (Lopez et al., 2001; Thomas et al., 2019). abian and Janson, 2016). Upsection increase in Paleozoic ages and decrease Provenance studies within the collisional domain and foreland basin in Grenvillian (Mesoproterozoic) ages are indicative of catchment maturation, have been performed in a variety of basins globally, especially in the Andean and drainage stabilization within the peri-Gondwanan hinterland and poten- (retroarc) and Himalayan regions (DeCelles et al., 1998; Horton et al., 2010; tially deeper into the accreted Pan-African terranes through time. This is also Gehrels et al., 2011a; Bande et al., 2012; Colleps et al., 2018). This study resolves supported by the core-rim relationships extracted through depth profiling a provenance evolution of both spatial and temporal significance prior to of the Guadalupian zircons. A few core-rim relationships show Grenvillian collision, within collisional domain during continental collision, and during cores with Pan-African rims (Fig. 12), which suggest that Gondwanan ter- post-collisional sediment dispersal to the resultant foreland basin setting. It ranes, especially those of Oaxaquia and Coahuila, were likely candidates for is intuitive that the most high-relief, proximal region of the fold-thrust belt

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 587 by guest on 30 September 2021 Research Paper

would supply the majority of sediment to the adjacent foreland basin system, which, due to universal abundance in both Gondwanan and Laurentian source and we have demonstrated that is the case in the Glass Mountains of the terranes, have been particularly difficult to interpret in past regional DZ studies. southern Delaware Basin. Moreover, and curiously enough, there are similar- Elucidating an extensive history of zircon transport and recycling during ities in the dominant age modes of detrital zircons throughout the basin (e.g., multiple stages of tectonism coupled with the ability to tie these data to their Soreghan and Soreghan, 2013). However, decades of sedimentological and original sources improves our understanding of potential source to sink rela- stratigraphic analysis of the depositional fairways suggest a dominant north- tionships within the basin. By tracking provenance trends across long-term west to south paleoflow direction, at least for the Guadalupian strata (Rossen (107–108 yr) sedimentation and tectonic deformation, we observed a regional and Sarg, 1987; Gardner et al., 2003; Pyles et al., 2010). Thus, the ubiquitous change in provenance through the final stages of continental collision. Spe- provenance signal in the Permian Basin does not necessarily reflect direct sed- cifically, the three main provenance stages observed were: (1) pre-tectonic iment input from Laurentian source terranes. Indeed, Soreghan and Soreghan deposition; (2) uplift and exhumation of a passive margin and ensuing unroof- (2013) and Kocurek and Kirkland (1998), among others, interpret convoluted ing of the orogenic wedge; and (3) post-tectonic margin stabilization coupled sediment-routing pathways from initial source to ultimate sink. So, although with catchment maturity and hinterland sediment transport. Similar collisional Permian Basin deposits reflect dominantly southerly provenance, they likely events are ubiquitous across the geologic record. This work could help to took a more circuitous pathway to deposition, especially in the northwestern guide others studying sediment sourcing and evolution of the orogenic sys- Delaware Basin. Future work in the Delaware Basin and beyond should focus tem of foreland basins. on retracing the sediment-routing pathways from predominantly peri-Gond- wanan source terranes to the northern basin margins. Our analyses of the earliest pre-collisional strata in the southern Delaware Basin helped us to ACKNOWLEDGMENTS trace sediment dispersal and provenance evolution. Difficulties may stem from We thank the UTChron Geo- and Thermochronology Laboratory and Lisa Stockli. We also thank the sponsors of the Quantitative Clastics Laboratory (http://beg.utexas.edu/qcl)​ and Dr. Scott gaps in the geologic record, but, where possible, this method can elucidate a Tinker for support. We thank the ranchers and landowners of the outcrops we sampled, namely: record of sediment sourcing from areas that have undergone extensive and Gordon Buscher of the Boss Ranch, Bruce Blakemore of the Primavera Ranch, and Brad Kelly complex tectonic deformation. and Merrill Dana of the Iron Mountain Ranch. Finally, we would like to thank Geosphere Science Editor Dr. David Fastovsky and Associate Editor Dr. Cathy Busby as well as reviewers Dr. Glenn R. Sharman and Dr. Andrew Leier for their time, consideration, and valuable input. ■■ CONCLUSIONS REFERENCES CITED By dating the crystallization age of individual detrital zircon grains, we are Anderson, J., and Morrison, J., 1992, The role of anorogenic granites in the Proterozoic crustal able to correlate detrital zircon ages with regional paleogeography of potential development of North America, in Condie, K.C., ed., Developments in Precambrian Geology: sediment source terranes to interpret provenance. By sampling strata from the Volume 10: Elsevier, p. 263–299, https://​doi​.org​/10​.1016​/S0166​-2635​(08)70121​-X. Anderson, J., Bender, E., Anderson, R., Bauer, P., Robertson, J., Bowring, S., Condie, K., Denison, pre-, syn-, and post-orogenic region of the late Paleozoic Delaware Basin, we R., Gilbert, M., and Grambling, J., 1993, Transcontinental Proterozoic provinces, in Reed, J.C., were able to define a type section of provenance from a southern source across Jr., Bickford, M.E., , R.S., Link, P.K., Rankin, D.W., Sims, P.K., and Van Schmus, W.R., a dynamic tectonic system. Here, we present 1721 new detrital zircon (DZ) U-Pb eds., Precambrian: Conterminous U.S.: Geological Society of America, The Geology of North America, v. C-2, p. 171–334, https://​doi​.org​/10​.1130​/DNAG​-GNA​-C2​.171. ages and 85 separate age core-rim relationships from 13 samples within the Anthony, J.M., 2015, Provenance of the Middle Permian, Delaware Mountain Group: Delaware Marathon fold and thrust belt and Glass Mountains of West Texas. To analyze basin, southeast and West Texas: Texas Christian University. provenance, we supplement our new data with published DZ geochronologic Arbenz, J., 1989a, Ouachita thrust belt and Arkoma basin, in Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita Orogen in the United States: Geological data sets of regional provenance studies from older and contemporaneous Society of America, The Geology of North America, v. F-2, p. 621–634, https://doi​ ​.org​/10​.1130​ basin-filling episodes along the Appalachian-Ouachita suture zone as well as /DNAG​-GNA​-F2​.621. within the Laurentian cratonic interior. Key findings include: (1) appearance of Arbenz, J.K., 1989b, The Ouachita system, in Bally, A.W., and Palmer, A.R., eds., The Geology of Neoproterozoic–Cambrian zircon grains in the Pennsylvanian Haymond Forma- North America—An overview: Geological Society of America, The Geology of North America, v. A, p. 371–396, https://​doi​.org​/10​.1130​/DNAG​-GNA​-A​.371. tion indicates basin inversion and the uplift and exhumation of volcanic units Bande, A., Horton, B.K., Ramírez, J.C., Mora, A., Parra, M., and Stockli, D.F., 2012, Clastic deposition, related to Rodinian rifting; (2) an upsection decrease in Grenvillian (ca. 1300– provenance, and sequence of Andean thrusting in the frontal Eastern Cordillera and Llanos 920 Ma) and increase in Paleozoic zircons denotes a steady provenance shift foreland basin of Colombia: Geological Society of America Bulletin, v. 124, no. 1–2, p. 59–76, https://​doi​.org​/10​.1130​/B30412​.1. from a dominantly orogenic highland source to sediment sources deeper in the Barnaby, R.J., Oetting, G.C., and Gao, G., 2004, Strontium isotopic signatures of oil-field waters: Gondwanan hinterland; and (3) lack of drastic changes in provenance during the Applications for reservoir characterization: American Association of Petroleum Geologists early to middle Permian supports the argument for tectonic quiescence within Bulletin, v. 88, no. 12, p. 1677–1704, https://​doi​.org​/10​.1306​/07130404002. Beck, J.W., and Murthy, V.R., 1991, Evidence for continental crustal assimilation in the Hemlock the collisional domain during this time period. Additionally, depth profiling Formation flood basalts of the early Proterozoic Penokean Orogen, Lake Superior region: and core-rim relationships aided in tracing sources for Grenvillian-age zircons, U.S. Department of the Interior, U.S. Geological Survey Bulletin 1904-1, p. 11–19, p. 110–125.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 588 by guest on 30 September 2021 Research Paper

Beck, R.B., Vondra, C.F., Filkins, J.E., and Olander, J.D., 1988, Syntectonic sedimentation and Dickinson, W.R., and Lawton, T.F., 2003, Sequential intercontinental suturing as the ultimate control Laramide basement thrusting, Cordilleran foreland: Timing of deformation, in Schmidt, C.J., for Pennsylvanian Ancestral Rocky Mountains deformation: Geology, v. 31, no. 7, p. 609–612, and Perry, W.J., Jr., eds., Interaction of the Rocky Mountain Foreland and the Cordilleran https://​doi​.org​/10​.1130​/0091​-7613​(2003)031​<0609:​SISATU>2​.0​.CO;2. Thrust Belt: Geological Society of America Memoir 171, p. 465–488, https://doi​ .org​ /10​ ​.1130​ Dickinson, W.R., Soreghan, M., and Gehrels, G., 2000, Geodynamic interpretation of Paleozoic /MEM171​-p465. tectonic trends oriented oblique to the Mesozoic Klamath-Sierran continental margin in Cali- Becker, T.P., Thomas, W.A., Samson, S.D., and Gehrels, G.E., 2005, Detrital zircon evidence of fornia, in Soreghan, M.J., and Gehrels, G.E., eds., Paleozoic and Triassic Paleogeography and Laurentian crustal dominance in the lower Pennsylvanian deposits of the Alleghanian clastic Tectonics of Western Nevada and Northern California: Geological Society of America Special wedge in eastern North America: Sedimentary Geology, v. 182, no. 1–4, p. 59–86, https://doi​ ​ Paper 347, p. 209–246, https://​doi​.org​/10​.1130​/0​-8137​-2347​-7​.209. .org​/10​.1016​/j​.sedgeo​.2005​.07​.014. Dutton, S.P., Flanders, W.A., and Barton, M.D., 2003, Reservoir characterization of a Permian Becker, T.P., Thomas, W.A., and Gehrels, G.E., 2006, Linking late Paleozoic sedimentary prove- deep-water sandstone, East Ford field, Delaware basin, Texas: American Association of Petro- nance in the Appalachian basin to the history of Alleghanian deformation: American Journal leum Geologists Bulletin, v. 87, no. 4, p. 609–627. of Science, v. 306, no. 10, p. 777–798, https://​doi​.org​/10​.2475​/10​.2006​.01. Eriksson, K.A., Campbell, I.H., Palin, J.M., and Allen, C.M., 2003, Predominance of Grenvillian Bickford, M., Van, W., and Zietz, I., 1986, Proterozoic history of the midcontinent region of North magmatism recorded in detrital zircons from modern Appalachian rivers: The Journal of America: Geology, v. 14, no. 6, p. 492–496, https://​doi​.org​/10​.1130​/0091​-7613​(1986)14​<492:​ Geology, v. 111, no. 6, p. 707–717, https://​doi​.org​/10​.1086​/378338. PHOTMR>2​.0​.CO;2. Ewing, T.E., 2016, Texas through Time: Lone Star Geology, Landscapes, and Resources: The Uni- Cawood, P.A., and Buchan, C., 2007, Linking accretionary orogenesis with supercontinent assembly: versity of Texas at Austin, Bureau of Economic Geology Udden Series No. 6, 431 p. Earth-Science Reviews, v. 82, no. 3–4, p. 217–256, https://doi​ .org​ /10​ .1016​ /j​ .earscirev​ .2007​ .03​ .003.​ Fedo, C.M., Sircombe, K.N., and Rainbird, R.H., 2003, Detrital zircon analysis of the sedimentary Cawood, P.A., McCausland, P.J., and Dunning, G.R., 2001, Opening Iapetus: Constraints from the record: Reviews in Mineralogy and Geochemistry, v. 53, no. 1, p. 277–303, https://doi​ ​.org​ Laurentian margin in Newfoundland: Geological Society of America Bulletin, v. 113, no. 4, /10​.2113​/0530277. p. 443–453, https://​doi​.org​/10​.1130​/0016​-7606​(2001)113​<0443:​OICFTL>2​.0​.CO;2. Fischer, A.G., and Sarnthein, M., 1988, Airborne silts and dune-derived in the Permian of Centeno-García, E., Gehrels, G., and Talavera-Mendoza, O., 2005, Zircon provenance of Triassic the Delaware Basin: Journal of Sedimentary Research, v. 58, no. 4, p. 637–643. (Paleozoic?) turbidites from central and western Mexico: Implications for the early evolution Flawn, P.T., Goldstein, A., Jr., King, P.B., and Weaver, C.E., 1961, The Ouachita System: Bureau of of the Guerrero arc: Geological Society of America Abstracts with Programs, v. 37, no. 4, p. 64. Economic Geology: University of Texas Publication 6120, 401 p. Colleps, C.L., McKenzie, N.R., Stockli, D.F., Hughes, N.C., Singh, B.P., Webb, A.A.G., Myrow, P.M., Flemings, P.B., Jordan, T.E., and Reynolds, S.A., 1986, Flexural analysis of two broken foreland Planavsky, N.J., and Horton, B.K., 2018, Zircon (U‐Th)/He Thermochronometric Constraints on basins: The Late Cenozoic Bermejo Basin and the Early Cenozoic Green River Basin: American Himalayan Thrust Belt Exhumation, Bedrock Weathering, and Cenozoic Chemistry: Association of Petroleum Geologists Bulletin, v. 70, p. 591. Geochemistry, Geophysics, Geosystems, v. 19, no. 1, p. 257–271. Galley, J.E., 1958, Oil and geology in the Permian basin of Texas and New Mexico, in Weeks, Cordani, U.G., and Teixeira, W., 2007, Proterozoic accretionary belts in the Amazonian Craton, in L.G., ed., Habitat of Oil: American Association of Petroleum Geologists Special Publication, Hatcher, R.D., Jr., Carlson, M.P., McBride, J.H., and Martínez Catalán, J.R., eds., 4-D Frame- p. 395–446, https://​doi​.org​/10​.1306​/SV18350C16. work of Continental Crust: Geological Society of America Memoir 200, p. 297–320, https://​ Gardner, M.H., Borer, J.M., Melick, J.J., Mavilla, N., Dechesne, M., and Wagerle, R.N., 2003, Strati- doi​.org​/10​.1130​/2007​.1200​(14). graphic process-response model for submarine channels and related features from studies of Cordani, U.G., Teixeira, W., D’Agrella-Filho, M., and Trindade, R., 2009, The position of the Ama- Permian Brushy Canyon outcrops, West Texas: Marine and Petroleum Geology, v. 20, no. 6–8, zonian Craton in supercontinents: Gondwana Research, v. 15, no. 3–4, p. 396–407, https://​doi​ p. 757–787, https://​doi​.org​/10​.1016​/j​.marpetgeo​.2003​.07​.004. .org​/10​.1016​/j​.gr​.2008​.12​.005. Gehrels, G., 2012, Detrital zircon U‐Pb geochronology: Current methods and new opportunities, DeCelles, P.G., and Giles, K.A., 1996, Foreland basin systems: Basin Research, v. 8, no. 2, p. 105–123, in Busby, C., and Azor, A., eds., Tectonics of Sedimentary Basins: Recent Advances: Blackwell https://​doi​.org​/10​.1046​/j​.1365​-2117​.1996​.01491​.x. Publishing, p. 45–62, https://​doi​.org​/10​.1002​/9781444347166​.ch2. DeCelles, P.G., Gehrels, G.E., Quade, J., and Ojha, T., 1998, Eocene–early Miocene foreland basin Gehrels, G., 2014, Detrital zircon U-Pb geochronology applied to tectonics: Annual Review of Earth development and the history of Himalayan thrusting, western and central Nepal: Tectonics, and Planetary Sciences, v. 42, p. 127–149, https://doi​ .org​ /10​ .1146​ /annurev​ -earth​ -050212​ -124012.​ v. 17, no. 5, p. 741–765, https://​doi​.org​/10​.1029​/98TC02598. Gehrels, G., Kapp, P., DeCelles, P., Pullen, A., Blakey, R., Weislogel, A., Ding, L., Guynn, J., Martin, DeMis, W.D., 1983, Geology of the Hell’s Half Acre, Marathon Basin, Texas [M.A. thesis]: Univer- A., and McQuarrie, N., 2011a, Detrital zircon geochronology of pre‐ strata in the Tibetan‐ sity of Texas at Austin, 130 p. Himalayan orogen: Tectonics, v. 30, no. 5, https://​doi​.org​/10​.1029​/2011TC002868. Denison, R., Burke, W., Otto, J., Hetherington, E., and Stone, C., 1977, Age of igneous and meta- Gehrels, G.E., Valencia, V.A., and Ruiz, J., 2008, Enhanced precision, accuracy, efficiency, and spatial morphic activity affecting the Ouachita foldbelt: Little Rock, Arkansas, Arkansas Geological resolution of U‐Pb ages by laser ablation–multicollector–inductively coupled plasma–mass Commission, Proceedings, Symposium on the Geology of the , v. 1, spectrometry: Geochemistry, Geophysics, Geosystems, v. 9, no. 3, https://doi​ ​.org​/10.1029​ p. 25–40. /2007GC001805. Denison, R.E., Kenny, G.S., Burke, W.H., Jr., and Hetherington, E.A., Jr., 1969, Isotopic ages of Gehrels, G.E., Blakey, R., Karlstrom, K.E., Timmons, J.M., Dickinson, B., and Pecha, M., 2011b, igneous and metamorphic boulders from the Haymond Formation (Pennsylvanian), Mara- Detrital zircon U-Pb geochronology of Paleozoic strata in the Grand Canyon, Arizona: Litho- thon Basin, Texas, and their significance: Geological Society of America Bulletin, v. 80, no. 2, sphere, v. 3, no. 3, p. 183–200, https://​doi​.org​/10​.1130​/L121​.1. p. 245–256, https://​doi​.org​/10​.1130​/0016​-7606​(1969)80​[245:​IAOIAM]2​.0​.CO;2. Geraldes, M.C., Van Schmus, W.R., Condie, K.C., Bell, S., Teixeira, W., and Babinski, M., 2001, Dickinson, W.R., 1988, Provenance and sediment dispersal in relation to paleotectonics and paleo- Proterozoic geologic evolution of the SW part of the Amazonian Craton in Mato Grosso geography of sedimentary basins, in Kleinspehn, K.L., and Paola, C., eds., New Perspectives state, Brazil: Precambrian Research, v. 111, no. 1–4, p. 91–128, https://​doi​.org​/10​.1016/S0301​ ​ in Basin Analysis: New York, New York, Springer, p. 3–25. -9268​(01)00158​-9. Dickinson, W.R., and Gehrels, G.E., 2003, U-Pb ages of detrital zircons from Permian and Jurassic Giles, J.M., Soreghan, M.J., Benison, K.C., Soreghan, G.S., and Hasiotis, S.T., 2013, Lakes, loess, eolian sandstones of the , USA: Paleogeographic implications: Sedimentary and paleosols in the Permian Wellington Formation of Oklahoma, USA: Implications for Geology, v. 163, no. 1–2, p. 29–66, https://​doi​.org​/10​.1016​/S0037​-0738​(03)00158​-1. paleoclimate and paleogeography of the Midcontinent: Journal of Sedimentary Research, Dickinson, W.R., and Gehrels, G.E., 2009, Use of U–Pb ages of detrital zircons to infer maximum v. 83, no. 10, p. 825–846, https://​doi​.org​/10​.2110​/jsr​.2013​.59. depositional ages of strata: A test against a Colorado Plateau Mesozoic database: Earth and Gillis, R.J., Gehrels, G.E., Ruiz, J., and de Dios Gonzaléz, L.A.F., 2005, Detrital zircon provenance Planetary Science Letters, v. 288, no. 1–2, p. 115–125, https://doi​ .org​ /10​ .1016​ /j​ .epsl​ .2009​ .09​ .013.​ of Cambrian–Ordovician and Carboniferous strata of the Oaxaca terrane, southern Mexico: Dickinson, W.R., and Lawton, T.F., 2001, Carboniferous to Cretaceous assembly and fragmentation Sedimentary Geology, v. 182, no. 1–4, p. 87–100, https://​doi​.org​/10​.1016​/j​.sedgeo​.2005​.07​.013. of Mexico: Geological Society of America Bulletin, v. 113, no. 9, p. 1142–1160, https://​doi​.org​ Glasspool, I., Wittry, J., Quick, K., Kerp, H., and Hilton, J., 2013, A preliminary report on a Wolfcam- /10​.1130​/0016​-7606​(2001)113​<1142:​CTCAAF>2​.0​.CO;2. pian age floral assemblage from the type section for the Neal ranch Formation in the Glass

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 589 by guest on 30 September 2021 Research Paper

Mountains, Texas: The Carboniferous–Permian Transition: New Mexico Museum of Natural Horak, R.L., 1985, Trans-Pecos tectonism and its effects on the Permian Basin: Structure and History and Science Bulletin, v. 60, p. 98–102. Tectonics of Trans-Pecos Texas: Midland, Texas, West Texas Geological Society, p. 81–87. Gleason, J.D., Gehrels, G.E., Dickinson, W.R., Patchett, P.J., and Kring, D.A., 2007, Laurentian sources Horton, B.K., Saylor, J.E., Nie, J., Mora, A., Parra, M., Reyes-Harker, A., and Stockli, D.F., 2010, for detrital zircon grains in turbidite and deltaic sandstones of the Pennsylvanian Haymond Linking sedimentation in the northern Andes to basement configuration, Mesozoic exten- Formation, Marathon assemblage, west Texas, USA: Journal of Sedimentary Research, v. 77, sion, and Cenozoic shortening: Evidence from detrital zircon U-Pb ages, Eastern Cordillera, no. 11, p. 888–900, https://​doi​.org​/10​.2110​/jsr​.2007​.084. Colombia: Geological Society of America Bulletin, v. 122, no. 9–10, p. 1423–1442, https://doi​ ​ Graham, S.A., Ingersoll, R.V., and Dickinson, W.R., 1976, Common provenance for lithic grains .org​/10​.1130​/B30118​.1. in Carboniferous sandstones from Ouachita Mountains and Black Warrior Basin: Journal of Hull, J.P., Jr., 1957, Petrogenesis of Permian Delaware Mountain Sandstone, Texas and New Mexico: Sedimentary Research, v. 46, no. 3, p. 620–632. American Association of Petroleum Geologists Bulletin, v. 41, no. 2, p. 278–307. Gross, E.L., Stewart, J.H., and Gehrels, G.E., 2000, Detrital zircon geochronology of Neoproterozoic Ingersoll, R.V., 2012, Tectonics of sedimentary basins, with revised nomenclature, in Busby, C., to Middle Cambrian miogeoclinal and platformal strata: Northwest Sonora, Mexico: Geofísica and Azor, A., eds., Tectonics of sedimentary basins: Recent advances: Blackwell Publishing, Internacional, v. 39, no. 4, p. 295–308. p. 1–43, https://​doi​.org​/10​.1002​/9781444347166​.ch1. Hagen, E.S., Shuster, M.W., and Furlong, K.P., 1985, Tectonic loading and subsidence of intermon- Jackson, S.E., Pearson, N.J., Griffin, W.L., and Belousova, E.A., 2004, The application of laser tane basins: Wyoming foreland province: Geology, v. 13, no. 8, p. 585–588, https://doi​ ​.org​/10​ ablation–inductively coupled plasma–mass spectrometry to in situ U-Pb zircon geochronology: .1130​/0091​-7613​(1985)13​<585:​TLASOI>2​.0​.CO;2. Chemical Geology, v. 211, no. 1–2, p. 47–69, https://​doi​.org​/10​.1016​/j​.chemgeo​.2004​.06​.017. Haneef, M., Rohr, D., and Wardlaw, B., 2000, A deep water turbidity origin for the Altuda Formation Janson, X., and Hairabian, J., 2016, Guadalupian Stratigraphy and Slope Systems in the Southern (Capitanian, Permian), Northwest Glass Mountains, Texas: Smithsonian Contributions to the Delaware Basin, Glass Mountains: Reservoir Characterization Research Laboratory Guide- Earth Sciences, v. 32, p. 319–332. book, 138 p. Hanson, R.E., Roberts, J.M., Dickerson, P.W., and Fanning, C.M., 2016, Cryogenian intraplate Jordan, T.E., Flemings, P.B., and Beer, J.A., 1988, Dating thrust-fault activity by use of foreland-basin magmatism along the buried southern Laurentian margin: Evidence from volcanic clasts in strata, in Kleinspehn, K.L., and Paola, C., eds., New Perspectives in Basin Analysis: Frontiers Ordovician strata, Marathon uplift, west Texas: Geology, v. 44, no. 7, p. 539–542, https://doi​ ​ in Sedimentary Geology: New York, New York, Springer, p. 307–330. .org​/10​.1130​/G37889​.1. Keller, G., Lidiak, E., Hinze, W., and Braile, L., 1983, The role of rifting in the tectonic development Harms, J., Williamson, C., and Mazzullo, S., 1988, Deep-water density current deposits of the of the midcontinent, USA, in Morgan, P., and Baker, B.H., eds., Developments in Geotecton- Delaware Mountain Group, Delaware Basin, in Mazzullo, S.J., ed., The Geologic Evolution ics, v. 19, Processes of Continental Rifting: Elsevier, p. 391–412, https://doi​ ​.org​/10​.1016/B978​ ​ of the Permian Basin: Midland, Texas, SEPM (Society for Sedimentary Geology), Permian -0​-444​-42198​-2​.50028​-6. Basin Section, p. 34–36. Keppie, J.D., 2004, Terranes of Mexico revisited: A 1.3 billion year odyssey: International Geology Harris, M., Lehrmann, D., and Lambert, L., 2000, Comparison of the depositional environments and Review, v. 46, no. 9, p. 765–794, https://​doi​.org​/10​.2747​/0020​-6814​.46​.9​.765. physical stratigraphy of the Cutoff Formation (Guadalupe Mountains) and the Road Canyon Keppie, J.D., Nance, R., Fernández-Suárez, J., Storey, C.D., Jeffries, T.E., and Murphy, J.B., 2006, Detrital Formation (Glass Mountains): Lowermost Guadalupian (Permian) of West Texas, in Proceedings zircon data from the Eastern Mixteca Terrane, Southern Mexico: Evidence for an Ordovician– The Guadalupian symposium: Smithsonian Contributions to the Earth Sciences, v. 32, p. 127–152. Mississippian continental rise and a Permo-Triassic clastic wedge adjacent to Oaxaquia: Hart, N.R., Stockli, D.F., and Hayman, N.W., 2016, Provenance evolution during progressive rifting International Geology Review, v. 48, no. 2, p. 97–111, https://doi​ .org​ /10​ .2747​ /0020​ -6814​ .48​ .2​ .97.​ and hyperextension using bedrock and detrital zircon U-Pb geochronology, Mauléon Basin, Keppie, J.D., Dostal, J., Murphy, J.B., and Nance, R.D., 2008, Synthesis and tectonic interpreta- western Pyrenees: Geosphere, v. 12, no. 4, p. 1166–1186, https://​doi​.org​/10​.1130​/GES01273​.1. tion of the westernmost Paleozoic Variscan orogen in southern Mexico: From rifted Rheic Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., 1989, The Appalachian-Ouachita orogen in margin to active Pacific margin: Tectonophysics, v. 461, no. 1–4, p. 277–290, https://doi​ ​.org​ the United States: Boulder, Colorado, Geological Society of America, The Geology of North /10​.1016​/j​.tecto​.2008​.01​.012. America, v. F-2, https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2. King, P.B., 1930, The Geology of the Glass Mountains, Texas: The University of Texas Bulletin, Haughton, P., Todd, S., and Morton, A., 1991, Sedimentary provenance studies, in Morton, A.C., no. 3038, 420 p. Todd, S.P., and Haughton, P.D.W., eds., Developments in Sedimentary Provenance Studies: King, P.B., 1937, Geology of the Marathon region, Texas: U.S. Geological Survey Professional Paper Geological Society of London Special Publication 57, no. 1, p. 1–11. 187, 148 p., https://​doi​.org​/10​.3133​/pp187. Heatherington, A., Mueller, P., and Nutman, A., 1996, Neoproterozoic magmatism in the Suwan- King, P.B., 1958, Problems of Boulder Beds of Haymond Formation, Marathon Basin, Texas: DIS- nee terrane: Implications for terrane correlation, in Nance, R.D., and Thompson, M.D., eds., CUSSION: American Association of Petroleum Geologists Bulletin, v. 42, no. 7, p. 1731–1734. Avalonian and Related Peri-Gondwanan Terranes of the Circum–North Atlantic: Geological King, P.B., 1978, Tectonics and sedimentation of the Paleozoic rocks in the Marathon region, west Society of America Special Paper 304, p. 257–268, https://​doi​.org​/10​.1130​/0​-8137​-2304​-3​.257. Texas: Tectonics and Paleozoic facies of the Marathon geosyncline, West Texas: Society of Eco- Hibbard, J.P., Stoddard, E.F., Secor, D.T., and Dennis, A.J., 2002, The Carolina Zone: Overview of nomic Paleontologists and Mineralogists, Permian Basic Section, Publication, no. 78–17, p. 5–38. Neoproterozoic to Early Paleozoic peri-Gondwanan terranes along the eastern flank of the Kocurek, G., and Kirkland, B.L., 1998, Getting to the source: Aeolian influx to the Permian Dela- southern Appalachians: Earth-Science Reviews, v. 57, no. 3-4, p. 299–339, https://​doi​.org​/10​ ware basin region: Sedimentary Geology, v. 117, no. 3–4, p. 143–149, https://​doi​.org​/10​.1016​ .1016​/S0012​-8252​(01)00079​-4. /S0037​-0738​(98)00024​-4. Hickman, R.G., Varga, R.J., and Altany, R.M., 2009, Structural style of the Marathon thrust belt, Lambert, L.L., Lehrmann, D.J., Harris, M.T., Wardlaw, B., Grant, R., and Rohr, D., 2000, Correlation west Texas: Journal of Structural Geology, v. 31, no. 9, p. 900–909, https://​doi​.org​/10​.1016​/j​ of the Road Canyon and Cutoff formations, West Texas, and its relevance to establishing an .jsg​.2008​.02​.016. international Middle Permian (Guadalupian) series, in Proceedings, The Guadalupian Sym- Hills, J.M., 1984, Sedimentation, tectonism, and hydrocarbon generation in Delaware basin, west posium: Smithsonian Contributions to the Earth Sciences, v. 32, p. 153–183. Texas and southeastern New Mexico: American Association of Petroleum Geologists Bulletin, Lambert, L.L., Wardlaw, B.R., Nestell, M.K., and Nestell, G.P., 2002, Latest Guadalupian (middle v. 68, no. 3, p. 250–267. Permian) conodonts and foraminifers from West Texas: Micropaleontology, v. 48, p. 343–364, Hoffman, P.F., Bally, A., and Palmer, A., 1989, Precambrian geology and tectonic history of North https://​doi​.org​/10​.2113​/48​.4​.343. America, in Bally, A.W., and Palmer, A.R., eds., The Geology of North America—An overview: Lawton, T., Hunt, G., and Gehrels, G., 2010, Detrital zircon record of thrust belt unroofing in Lower Geological Society of America, The Geology of North America, v. A, p. 447–512, https://​doi​ Cretaceous synorogenic conglomerates, central Utah: Geology, v. 38, no. 5, p. 463–466, https://​ .org​/10​.1130​/DNAG​-GNA​-A​.447. doi​.org​/10​.1130​/G30684​.1. Holm, D., Anderson, R., Boerboom, T., Cannon, W., Chandler, V., Jirsa, M., Miller, J., Schneider, D.A., Lawton, T., Buller, C., and Parr, T., 2015a, Provenance of a Permian erg on the western margin of Schulz, K., and Van Schmus, W., 2007, Reinterpretation of Paleoproterozoic accretionary bound- Pangea: Depositional system of the (late Leonardian) Castle Valley and White Rim aries of the north-central United States based on a new aeromagnetic-geologic compilation: sandstones and subjacent Cutler Group, Paradox Basin, Utah, USA: Geosphere, v. 11, no. 5, Precambrian Research, v. 157, no. 1–4, p. 71–79, https://doi​ .org​ /10​ .1016​ /j​ .precamres​ .2007​ .02​ .023.​ p. 1475–1506, https://​doi​.org​/10​.1130​/GES01174​.1.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 590 by guest on 30 September 2021 Research Paper

Lawton, T., Pindell, J., Beltran-Triviño, A., Juárez-Arriaga, E., Molina-Garza, R., and Stockli, D., 2015b, Olszewski, T.D., and Erwin, D.H., 2009, Change and stability in Permian brachiopod commu- Late Cretaceous–Paleogene Foreland Sediment-Dispersal Systems in Northern and Eastern nities from western Texas: Palaios, v. 24, no. 1, p. 27–40, https://doi​ .org​ /10​ .2110​ /palo.2008​ ​ Mexico: Interpretations From Preliminary Detrital-Zircon Analysis: American Association of .p08-061r.​ Petroleum Geologists Search and Discovery, article 30423, 37 p. Opdyke, N.D., Jones, D.S., MacFadden, B.J., Smith, D.L., Mueller, P.A., and Shuster, R.D., 1987, Li, Z.-X., Bogdanova, S., Collins, A., Davidson, A., De Waele, B., Ernst, R., Fitzsimons, I., Fuck, Florida as an exotic terrane: Paleomagnetic and geochronologic investigation of lower Paleo- R., Gladkochub, D., and Jacobs, J., 2008, Assembly, configuration, and break-up history of zoic rocks from the subsurface of Florida: Geology, v. 15, no. 10, p. 900–903, https://doi​ ​.org​ Rodinia: A synthesis: Precambrian Research, v. 160, no. 1–2, p. 179–210, https://doi​ ​.org/10​ ​ /10​.1130​/0091​-7613​(1987)15​<900:​FAAETP>2​.0​.CO;2. .1016​/j​.precamres​.2007​.04​.021. Park, H., Barbeau, D.L., Jr., Rickenbaker, A., Bachmann-Krug, D., and Gehrels, G., 2010, Applica- Link, P.K., Mahon, R.C., Beranek, L.P., Campbell-Stone, E.A., and Lynds, R., 2014, Detrital zircon tion of foreland basin detrital-zircon geochronology to the reconstruction of the southern provenance of Pennsylvanian to Permian sandstones from the Wyoming craton and Wood and central Appalachian orogen: The Journal of Geology, v. 118, no. 1, p. 23–44, https://​doi​ River Basin, Idaho, USA: Rocky Mountain Geology, v. 49, no. 2, p. 115–136, https://​doi​.org​ .org​/10​.1086​/648400. /10​.2113​/gsrocky​.49​.2​.115. Patchett, P., Embry, A., Ross, G., Beauchamp, B., Harrison, J., Mayr, U., Isachsen, C., Rosenberg, Lopez, R., Cameron, K., and Jones, N.W., 2001, Evidence for Paleoproterozoic, Grenvillian, and E., and Spence, G., 2004, Sedimentary cover of the Canadian Shield through Mesozoic time Pan-African age Gondwanan crust beneath northeastern Mexico: Precambrian Research, reflected by Nd isotopic and geochemical results for the Sverdrup Basin, Arctic Canada: The v. 107, no. 3–4, p. 195–214, https://​doi​.org​/10​.1016​/S0301​-9268​(00)00140​-6. Journal of Geology, v. 112, no. 1, p. 39–57, https://​doi​.org​/10​.1086​/379691. Marsh, J.H., and Stockli, D.F., 2015, Zircon U-Pb and trace element zoning characteristics in an ana- Paton, C., Hellstrom, J., Paul, B., Woodhead, J., and Hergt, J., 2011, Iolite: Freeware for the visuali- tectic granulite domain: Insights from LASS-ICP-MS depth profiling: Lithos, v. 239, p. 170–185, sation and processing of mass spectrometric data: Journal of Analytical Atomic Spectrometry, https://​doi​.org​/10​.1016​/j​.lithos​.2015​.10​.017. v. 26, no. 12, p. 2508–2518, https://​doi​.org​/10​.1039​/c1ja10172b. Marsh, A.D., Parker, W.G., Stockli, D.F., and Martz, J.W., 2019, Regional correlation of the Sonsela Petrus, J.A., and Kamber, B.S., 2012, VizualAge: A novel approach to laser ablation ICP‐MS U‑Pb Member (Upper Triassic Chinle Formation) and detrital U-Pb zircon data from the Sonsela geochronology data reduction: Geostandards and Geoanalytical Research, v. 36, no. 3, p. 247– Sandstone bed near the Sonsela Buttes, northeastern Arizona, USA, support the presence of a 270, https://​doi​.org​/10​.1111​/j​.1751​-908X​.2012​.00158​.x. distributive fluvial system: Geosphere, v. 15, p. 1128–1139, https://doi​ .org​ /10​ .1130​ /GES02004​ .1.​ Poole, F.G., Perry, W.J., Madrid, R.J., and Amaya-Martínez, R., 2005, Tectonic synthesis of the Martens, U., Weber, B., and Valencia, V.A., 2010, U/Pb geochronology of Devonian and older Ouachita-Marathon-Sonora orogenic margin of southern Laurentia: Stratigraphic and struc- Paleozoic beds in the southeastern Maya block, Central America: Its affinity with peri-Gond- tural implications for timing of deformational events and plate-tectonic model, in Anderson, wanan terranes: Geological Society of America Bulletin, v. 122, no. 5–6, p. 815–829, https://​ T.H., et al., eds., The Mojave Megashear Hypothesis: Development, Assessment, and Alter- doi​.org​/10​.1130​/B26405​.1. natives: Geological Society of America Special Paper 393, p. 543–596, https://doi.org​/10.1130​ McBride, E., 1989, Stratigraphy and sedimentary history of pre-Permian Paleozoic rocks of the Mara- /0-8137​-2393​-0.543. thon uplift, in Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita Pyles, D.R., Jennette, D.C., Tomasso, M., Beaubouef, R.T., and Rossen, C., 2010, Concepts learned orogen in the United States: Geological Society of America, The Geology of North America, from a 3D outcrop of a sinuous slope channel complex: Beacon Channel Complex, Brushy v. F-2, p. 603–620, https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2​.603. Canyon Formation, West Texas, USA: Journal of Sedimentary Research, v. 80, no. 1, p. 67–96, McKee, J.W., Jones, N.W., Anderson, T.H., Bartolini, C., Wilson, J., and Lawton, T., 1999, Late https://​doi​.org​/10​.2110​/jsr​.2010​.009. Paleozoic and early Mesozoic history of the Las Delicias terrane, Coahuila, Mexico, in Barto­ Rahl, J.M., Ehlers, T.A., and van der Pluijm, B.A., 2007, Quantifying transient erosion of orogens lini, C., Wilson, J.L., and Lawton, T.F., eds., Mesozoic Sedimentary and Tectonic History of with detrital thermochronology from syntectonic basin deposits: Earth and Planetary Science North-Central Mexico: Geological Society of America Special Paper 340, p. 161–189, https://​ Letters, v. 256, p. 147–161, https://​doi​.org​/10​.1016​/j​.epsl​.2007​.01​.020. doi​.org​/10​.1130​/0​-8137​-2340​-X​.161. Rainbird, R.H., Hearnan, L.M., and Young, G., 1992, Sampling Laurentia: Detrital zircon geochro- Moecher, D.P., and Samson, S.D., 2006, Differential zircon fertility of source terranes and natural nology offers evidence for an extensive Neoproterozoic river system originating from the bias in the detrital zircon record: Implications for sedimentary provenance analysis: Earth and Grenville orogen: Geology, v. 20, no. 4, p. 351–354, https://doi​ .org​ /10​ .1130​ /0091​ -7613​ (1992)020​ ​ Planetary Science Letters, v. 247, no. 3–4, p. 252–266, https://doi​ .org​ /10​ .1016​ /j​ .epsl​ .2006​ .04​ .035.​ <0351:​SLDZGO>2​.3​.CO;2. Muehlberger, W., and Tauvers, P., 1989, Marathon fold-thrust belt, west Texas, in Hatcher, R.D., Rainbird, R.H., McNicoll, V., Theriault, R., Heaman, L., Abbott, J., Long, D., and Thorkelson, D., Jr., et al., eds., The Appalachian-Ouachita Orogen in the United States: Boulder, Colorado: 1997, Pan-continental river system draining Grenville Orogen recorded by U-Pb and Sm-Nd Geological Society of America, The Geology of North America, v. F-2, p. 673–680, https://doi​ geochronology of Neoproterozoic arenites and mud rocks, northwestern Canada: The .org​/10.1130​/DNAG-F2.673. Journal of Geology, v. 105, no. 1, p. 1–17, https://​doi​.org​/10​.1086​/606144. Mueller, P.A., Heatherington, A.L., Wooden, J.L., Shuster, R.D., Nutman, A.P., and Williams, I.S., Rathjen, J.D., 1993, Microfacies and depositional environment of the Word Formation (Perm- 1994, Precambrian zircons from the Florida basement: A Gondwanan connection: Geology, ian) Glass Mountains, Texas: Geological Society of America, Abstracts with Programs, v. 25, v. 22, no. 2, p. 119–122, https://​doi​.org​/10​.1130​/0091​-7613​(1994)022​<0119:​PZFTFB>2​.3​.CO;2. no. 1, p. 40. Murphy, J.B., Pisarevsky, S.A., Nance, R.D., and Keppie, J.D., 2004, Neoproterozoic–Early Paleozoic Reed, T., and Strickler, D., 1990, Structural geology and petroleum exploration of the Marathon evolution of peri-Gondwanan terranes: Implications for Laurentia-Gondwana connections: thrust belt, West Texas, in Proceedings Marathon Thrust Belt: Structure, Stratigraphy, and International Journal of Earth Sciences, v. 93, no. 5, p. 659–682, https://doi​ .org​ /10​ .1007​ /s00531​ ​ Hydrocarbon Potential, in Higgins, L., ed., West Texas Geological Society and Permian Basin -004​-0412​-9. Section SEPM (Society for Sedimentary Geology) Field Seminar, p. 3–64. Nestell, M.K., Nestell, G.P., and Wardlaw, B.R., 2019, Integrated fusulinid, conodont, and radiolar- Richards, B.C., 2013, Current status of the International Carboniferous time scale, in Lucas, S.G., ian biostratigraphy of the Guadalupian (Middle Permian) in the Permian Basin region, USA, et al., eds., The Carboniferous–Permian Transition: New Mexico Museum of Natural History in Ruppel, S.C., ed., 2019, Anatomy of a Paleozoic Basin: The Permian Basin, USA (vol. 1): and Science Bulletin, v. 60, p. 348–353. The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 285: Rodriguez, E., Dickerson, P., and Stockli, D., 2017, New Zircon U-Pb Age Constrain of the Origin AAPG Memoir 118, 412 p., https://​doi​.org​/10​.23867​/RI0285​-1. of Devil’s River Uplift (SW Texas) and Insights into the Late Proterozoic and Paleozoic Evo- Nicholas, R., and Waddell, D., 1989, The Ouachita system in the subsurface of Texas, Arkansas, and lution of the Southern Margin of Laurentia: Proceedings, American Geophysical Union Fall Louisiana, in Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita Meeting, abstract #EP538-1732. Orogen in the United States: Geological Society of America, The Geology of North America, Ross, C.A., 1963, Standard Wolfcampian Series (Permian), Glass Mountains, Texas: Geological v. F-2, p. 661–672, https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2​.661. Society of America Memoir 88, 204 p., https://​doi​.org​/10​.1130​/MEM88. Nicholas, R.L., and Rozendal, R.A., 1975, Subsurface positive elements within Ouachita foldbelt Ross, C.A., 1986, Paleozoic evolution of southern margin of Permian basin: Geological Society in Texas and their relation to Paleozoic cratonic margin: American Association of Petroleum of America Bulletin, v. 97, no. 5, p. 536–554, https://​doi​.org​/10​.1130​/0016​-7606(1986)97​ ​<536:​ Geologists Bulletin, v. 59, no. 2, p. 193–216. PEOSMO>2​.0​.CO;2.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 591 by guest on 30 September 2021 Research Paper

Rossen, C., and Sarg, J., 1987, Sedimentology and regional correlation of a basinally restricted Thomas, W.A., Astini, R.A., and Bayona, G., 2002, Ordovician collision of the Argentine Precor- deepwater siliciclastic wedge: Brushy Canyon Formation–Cherry Canyon Tongue (Lower dillera with Gondwana, independent of Laurentian Taconic orogeny: Tectonophysics, v. 345, Guadalupian), Delaware basin: American Association of Petroleum Geologists Bulletin, v. 71, no. 1–4, p. 131–152, https://​doi​.org​/10​.1016​/S0040​-1951​(01)00210​-4. no. CONF-870606. Thomas, W.A., Becker, T.P., Samson, S.D., and Hamilton, M.A., 2004, Detrital zircon evidence of Sacks, P.E., and Secor, D.T., 1990, Kinematics of late Paleozoic continental collision between Lau- a recycled orogenic foreland provenance for Alleghanian clastic-wedge sandstones: The rentia and Gondwana: Science, v. 250, no. 4988, p. 1702–1705, https://doi​ ​.org​/10​.1126​/science​ Journal of Geology, v. 112, p. 23–37, https://​doi​.org​/10​.1086​/379690. .250​.4988​.1702. Thomas, W.A., Gehrels, G.E., Lawton, T.F., Satterfield, J.I., Romero, M.C., and Sundell, K.E., 2019, Samson, S.D., D’Lemos, R.S., Miller, B.V., and Hamilton, M.A., 2005, Neoproterozoic palaeoge- Detrital zircons and sediment dispersal from the Coahuila terrane of northern Mexico into ography of the Cadomia and Avalon terranes: Constraints from detrital zircon U-Pb ages: the Marathon foreland of the southern Midcontinent: Geosphere, v. 15, no. 4, p. 1102–1127, Journal of the Geological Society of London, v. 162, no. 1, p. 65–71, https://​doi​.org​/10​.1144​ https://​doi​.org​/10​.1130​/GES02033​.1. /0016​-764904​-003. Thomson, A., and McBride, E.F., Summary of the geologic history of the Marathon geosyncline: Sánchez-Bettucci, L., and Rapalini, A.E., 2002, Paleomagnetism of the Sierra de Las Animas Com- Proceedings, Permian Basin Section, SEPM (Society for Sedimentary Geology) Field Trip plex, southern Uruguay: Its implications in the assembly of western Gondwana: Precambrian Symposium and Guidebook Publication 1964, v. 64, p. 52–60. Research, v. 118, no. 3, p. 243–265, https://​doi​.org​/10​.1016​/S0301​-9268​(02)00114​-6. Thomson, K.D., Stockli, D.F., Clark, J.D., Puigdefàbregas, C., and Fildani, A., 2017, Detrital zircon Santos, J.D., 2003, Geotectônica dos escudos das Guianas e Brasil-Central: Geologia, Tectônica (U‐Th)/(He‐Pb) double‐dating constraints on provenance and foreland basin evolution of the e Recusrsos Minerais do Brasil (texto, mapas & SIG): Brasília, Serviço Geológico do Brasil– Ainsa Basin, south‐central Pyrenees, Spain: Tectonics, v. 36, no. 7, p. 1352–1375, https://doi​ ​ CPRM/MME, p. 169–226. .org​/10​.1002​/2017TC004504. Sharman, G.R., Graham, S.A., Grove, M., Kimbrough, D.L., and Wright, J.E., 2015, Detrital zircon Tohver, E., van der Pluijm, B., Van der Voo, R., Rizzotto, G., and Scandolara, J., 2002, Paleogeog- provenance of the Late Cretaceous–Eocene California forearc: Influence of Laramide low-an- raphy of the Amazon craton at 1.2 Ga: Early Grenvillian collision with the Llano segment of gle subduction on sediment dispersal and paleogeography: Geological Society of America Laurentia: Earth and Planetary Science Letters, v. 199, no. 1, p. 185–200, https://doi​ ​.org/10​ ​ Bulletin, v. 127, no. 1–2, p. 38–60, https://​doi​.org​/10​.1130​/B31065​.1. .1016​/S0012​-821X​(02)00561​-7. Sharman, G.R., Sharman, J.P., and Sylvester, Z., 2018, detritalPy: A Python‐based Toolset for Torres, R., Ruiz, J., Patchett, P.J., Grajales, J.M., Bartolini, C., Wilson, J.L., and Lawton, T.F., 1999, Visualizing and Analyzing Detrital Geo‐Thermochronologic Data: The Depositional Record, A Permo-Triassic continental arc in eastern Mexico: Tectonic implications for reconstructions v. 4, p. 202–215, https://​doi​.org​/10​.1002​/dep2​.45. of southern North America, in Bartolini, C., Wilson, J.L., and Lawton, T.F., eds., Mesozoic Sims, P., Schmus, W.V., Schulz, K., and Peterman, Z., 1989, Tectono-stratigraphic evolution of the Sedimentary and Tectonic History of North-Central Mexico: Geological Society of America Early Proterozoic Wisconsin magmatic terranes of the Penokean Orogen: Canadian Journal Special Paper 340, p. 191–196, https://​doi​.org​/10​.1130​/0​-8137​-2340​-X​.191. of Earth Sciences, v. 26, no. 10, p. 2145–2158, https://​doi​.org​/10​.1139​/e89​-180. Vermeesch, P., 2004, How many grains are needed for a provenance study?: Earth and Planetary Sláma, J., Košler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood, M.S.A., Science Letters, v. 224, no. 3–4, p. 441–451, https://​doi​.org​/10​.1016​/j​.epsl​.2004​.05​.037. Morrish, G.A., Nasdala, L., Norberg, N., Schaltegger, U., Schoene, B., Tubrett, M.N., and White- Vermeesch, P., 2013, Multi-sample comparison of detrital age distributions: Chemical Geology, house, M.J., 2008, Plesovice zircon—A new natural reference material for U-Pb and Hf isotopic v. 341, p. 140–146, https://​doi​.org​/10​.1016​/j​.chemgeo​.2013​.01​.010. microanalysis: Chemical Geology, v. 249, p. 1–35, https://doi​ .org​ /10​ .1016​ /j​ .chemgeo​ .2007​ .11​ .005.​ Viele, G., and Thomas, W., 1989, Tectonic synthesis of the Ouachita orogenic belt, in Hatcher, R.D., Soreghan, G.S., and Soreghan, M.J., 2013, Tracing Clastic Delivery To the Permian Delaware Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita orogen in the United Basin, USA: Implications For Paleogeography and Circulation In Westernmost Equatorial States: Geological Society of America, Geology of North America, v. F-2, p. 695–728, https://​ Pangea Tracing Clastic Dispersal to the Westernmost Pangean Suture: Journal of Sedimentary doi​.org​/10​.1130​/DNAG​-GNA​-F2​.695. Research, v. 83, no. 9, p. 786–802, https://​doi​.org​/10​.2110​/jsr​.2013​.63. Walper, J.L., 1982, Plate tectonic evolution of the Fort Worth Basin: Petroleum Geology of the Steiner, M., and Walker, J.D., 1996, Late Silurian plutons in Yucatan: Journal of Geophysical Fort Worth Basin and Bend Arch Area: American Association of Petroleum Geologists/ Research. Solid Earth, v. 101, no. B8, p. 17,727–17,735, https://​doi​.org​/10​.1029​/96JB00174. Geological Society database, p. 237–251. Steiner, M.B., and Anderson, T., 2005, Pangean reconstruction of the Yucatan Block: Its Permian, Wardlaw, B.R., 2000, Guadalupian conodont biostratigraphy of the Glass and Del Norte Triassic, and Jurassic geologic and tectonic history, in Anderson, T.H., Nourse, J.A., McKee, Mountains, in Proceedings The Guadalupian Symposium, Smithsonian Institution Press, J.W., and Steiner, M.B., eds., The Mojave-Sonora Megashear Hypothesis: Development, p. 37–88. assessment, and alternatives: Geological Society of America Special Paper 393, p. 457–480, Weber, B., Schaaf, P., Valencia, V.A., Iriondo, A., and Ortega-Gutiérrez, F., 2006, Provenance ages of https://​doi​.org​/10​.1130​/0​-8137​-2393​-0​.457. late Paleozoic sandstones (Santa Rosa Formation) from the Maya block, SE Mexico: Implica- Stewart, J.H., Gehrels, G.E., Barth, A.P., Link, P.K., Christie-Blick, N., and Wrucke, C.T., 2001, Detrital tions on the tectonic evolution of western Pangea: Revista Mexicana de Ciencias Geologicas, zircon provenance of Mesoproterozoic to Cambrian arenites in the western United States and v. 23, p. 262–276. northwestern Mexico: Geological Society of America Bulletin, v. 113, no. 10, p. 1343–1356, Weber, B., Valencia, V.A., Schaaf, P., Pompa-Mera, V., and Ruiz, J., 2008, Significance of provenance https://​doi​.org​/10​.1130​/0016​-7606​(2001)113​<1343:​DZPOMT>2​.0​.CO;2. ages from the Chiapas Massif Complex (southeastern Mexico): Redefining the Paleozoic Talavera-Mendoza, O., Ruiz, J., Gehrels, G.E., Meza-Figueroa, D.M., Vega-Granillo, R., and Cam- basement of the Maya Block and its evolution in a peri-Gondwanan realm: The Journal of pa-Uranga, M.F., 2005, U-Pb geochronology of the Acatlán Complex and implications for the Geology, v. 116, no. 6, p. 619–639, https://​doi​.org​/10​.1086​/591994. Paleozoic paleogeography and tectonic evolution of southern Mexico: Earth and Planetary Wiedenbeck, M., Allé, P., Corfu, F., Griffin, W.L., Meier, M., Oberli, F., Von Quadt, A., Roddick, J.C., Science Letters, v. 235, no. 3–4, p. 682–699, https://​doi​.org​/10​.1016​/j​.epsl​.2005​.04​.013. and Spiegel, W., 1995, Three natural zircon standards for U‐Th‐Pb, Lu‐Hf, trace element and Tauvers, P.R., 1988, Basement-influenced deformation in the Marathon fold-thrust belt, west Texas: REE analyses: Geostandards Newsletter, v. 19, no. 1, p. 1–23, https://​doi​.org​/10​.1111​/j​.1751​ The Journal of Geology, v. 96, no. 5, p. 577–590, https://​doi​.org​/10​.1086​/629253. -908X​.1995​.tb00147​.x. Thomas, W.A., 2006, Tectonic inheritance at a continental margin: GSA Today, v. 16, no. 2, p. 4–11, Wortman, G.L., Samson, S.D., and Hibbard, J.P., 2000, Precise U-Pb zircon constraints on the earliest https://​doi​.org​/10​.1130​/1052​-5173​(2006)016​[4:​TIAACM]2​.0​.CO;2. magmatic history of the Carolina terrane: The Journal of Geology, v. 108, no. 3, p. 321–338, Thomas, W.A., 2011a, Detrital-zircon geochronology and sedimentary provenance: Lithosphere, https://​doi​.org​/10​.1086​/314401. v. 3, no. 4, p. 304–308, https://​doi​.org​/10​.1130​/RF​.L001​.1. Xie, X., O’Connor, P.M., and Alsleben, H., 2016, Carboniferous sediment dispersal in the Thomas, W.A., 2011b, The Iapetan rifted margin of southern Laurentia: Geosphere, v. 7, no. 1, Appalachian–​Ouachita juncture: Provenance of selected late Mississippian sandstones in p. 97–120, https://​doi​.org​/10​.1130​/GES00574​.1. the Black Warrior Basin, Mississippi, United States: Sedimentary Geology, v. 342, p. 191–201, Thomas, W.A., 2014, A mechanism for tectonic inheritance at transform faults of the Iapetan https://​doi​.org​/10​.1016​/j​.sedgeo​.2016​.07​.007. margin of Laurentia: Geoscience Canada, v. 41, no. 3, p. 321–344, https://doi​ ​.org​/10​.12789​ Xie, X., Anthony, J.M., and Busbey, A.B., 2018, Provenance of Permian Delaware Mountain Group, /geocanj​.2014​.41​.048. central and southern Delaware Basin, and implications of sediment dispersal pathway near

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 592 by guest on 30 September 2021 Research Paper

the southwestern terminus of Pangea: International Geology Review, v. 61, p. 1–20, https://​ Yang, K.-M., and Dorobek, S.L., 1995, The Permian basin of west Texas and New Mexico: Tectonic history doi​.org​/10​.1080​/00206814​.2018​.1425925. of a ”composite” foreland basin and its effects on stratigraphic development: Stratigraphic evolution Xu, J., Stockli, D.F., and Snedden, J.W., 2017, Enhanced provenance interpretation using com- of foreland basins: SEPM (Society for Sedimentary Geology) Special Publication, v. 52, p. 149–174. bined U–Pb and (U–Th)/He double dating of detrital zircon grains from lower Miocene strata, Yang, Z., and Yancey, T., 2000, Fusulinid biostratigraphy and paleontology of the Middle Permian proximal Gulf of Mexico Basin, North America: Earth and Planetary Science Letters, v. 475, (Guadalupian) strata of the Glass Mountains and Del Norte Mountains, west Texas: Smith- p. 44–57, https://​doi​.org​/10​.1016​/j​.epsl​.2017​.07​.024. sonian Contributions to the Earth Sciences, v. 32, p. 185–259.

GEOSPHERE | Volume 16 | Number 2 Soto-Kerans et al. | Permian foreland basin provenance Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/16/2/567/4968494/567.pdf 593 by guest on 30 September 2021