Transition from Late rifting to middle dynamic foreland, southwestern U.S. and northwestern

Timothy F. Lawton1,†, Jeffrey M. Amato2, Sarah E.K. Machin1, John C. Gilbert3, and Spencer G. Lucas4 1Bureau of Economic Geology, Jackson School of Geosciences, The University of at Austin, Austin, Texas 78758, USA 2Department of Geological Sciences, State University, Las Cruces, New Mexico 88003, USA 3Office of Management and Enterprise Services, State of Oklahoma, 2401 North Lincoln Boulevard, Suite 118, Oklahoma City, Oklahoma 73105, USA 4New Mexico Museum of Natural History, 1801 Mountain Road, N.W., Albuquerque, New Mexico 87104, USA

ABSTRACT ­renewed subsidence, the middle Cretaceous, an extensional tectonic set- arc continued to supply volcanic-lithic ting of northern Sonora, southern , and Subsidence history and prov- to the Altar-Cucurpe basin, which by then southwestern New Mexico contrasted with re- enance of the Bisbee basin of southwestern was the foredeep of the foreland basin. Sand- gional tectonics to the north, where crustal short- New Mexico, southern Arizona, and north- stone of the Bootheel basin is more quartzose ening of the Sevier orogeny was well under way ern Sonora, Mexico, demonstrate basin evo- than the Altar-Cucurpe basin, but uncommon by time in southern Nevada, , and lution from an array of –Early sandstone beds contain neovolcanic lithic Wyoming (Armstrong, 1968; Heller et al., 1986; Cretaceous rift basins to a partitioned middle fragments and young zircon grains that were Lawton, 1994; DeCelles, 2004; Yonkee and Cretaceous retroarc foreland basin. The fore- transported to the basin as airborne ash. Lat- Weil, 2015). Flexural subsidence adjacent to the land basin contained persistent depocenters est Albian–early Cenomanian U-Pb tuff ages, Sevier orogenic belt created an adjacent retroarc that were inherited from the rift basin array detrital zircon maximum depositional ages foreland basin (sensu Ingersoll, 2012), termed and determined patterns of Albian–early ranging from ca. 102 Ma to 98 Ma, and am- the Cordilleran foreland basin, whose foredeep Cenomanian routing. Upper Ju- monite all demonstrate equivalence of lay along the western flank of the epicontinental rassic and –Aptian strata were middle Cretaceous proximal foreland strata (Jordan, 1981; Lawton, deposited in three narrow extensional basins, of the U.S.-Mexico border region with distal 1982, 1994; Robinson Roberts and Kirschbaum, termed the Altar-Cucurpe, Huachuca, and back-bulge strata of the Cordilleran foreland 1995; Currie, 1997, 1998). Dynamic topography Bootheel basins. Initially rapid Late Jurassic basin. Marine strata buried a former rift created by the subducted Farallon slab beneath subsidence in the basins slowed in the Early shoulder in southwestern New Mexico dur- western likely contributed to Cretaceous, then increased again from mid- ing late Albian to earliest Cenomanian time long-wavelength subsidence in the central and Albian through middle Cenomanian time, (ca. 105–100 Ma), prior to widespread trans- eastern parts of the basin (e.g., Cross, 1986; marking an episode of foreland subsidence. gression in central New Mexico (ca. 98 Ma). Pang and Nummedal, 1995; Nummedal, 2004). Sandstone composition and detrital zir- Lateral stratigraphic continuity across the In comparison, development of a retroarc fore- con provenance indicate different sediment former rift shoulder likely resulted from re- land basin in the U.S.–Mexico border region, as sources in the three basins and demonstrate gional dynamic subsidence following late Al- inferred from subsidence analysis of the Lower their continued persistence as depocenters bian collision of the Guerrero composite vol- Cretaceous section in southwestern New Mexico during Albian foreland basin development. canic terrane with Mexico and emplacement (Mack, 1987a), did not begin until late Albian Late Jurassic basins received sediment of the Farallon slab beneath the U.S.–Mexico time, or at least 20 m.y. later than farther north. from a nearby magmatic arc that migrated border region. Inferred dynamic subsidence Studies of foreland basin depositional history westward with time. Following a 10–15 m.y. in the foreland of southern Arizona and south- and tectonics have tended to focus on latitudes depositional hiatus, an con- western New Mexico was likely augmented in north of Las Vegas, Nevada, as demonstrated tinental margin arc supplied sediment to the Sonora by flexural subsidence adjacent to an by extensive reviews of basin history adjacent Altar-Cucurpe basin in Sonora as early as incipient thrust load driven by collision of the to the Sevier orogenic belt (Lawton, 1994; ca. 136 Ma, but local sedimentary and base- ­Guerrero superterrane. ­DeCelles, 2004; Yonkee and Weil, 2015). In ment sources dominated the Huachuca basin contrast, mechanisms of Late Jurassic to mid- of southern Arizona until catchment exten- INTRODUCTION dle Cretaceous sedimentary basin development sion tapped the arc source at ca. 123 Ma. The in the U.S.–Mexico border region of Arizona, Bootheel basin of southwestern New Mexico A fundamental plate tectonic reorganization ­Sonora, and southwestern New Mexico, and received sediment only from local basement took place during middle Cretaceous (Albian–­ their relationship with the Cordilleran foreland and recycled sedimentary sources with no Cenomanian) time in the southwest U.S.– basin, remain poorly understood because of less contemporary arc source evident. During Mexico border region immediately following extensive study and uncertainty regarding age Jurassic–Early Cretaceous continental rifting and depositional setting of the Upper Jurassic– (Bilodeau, 1982; Lawton and McMillan, 1999; Cretaceous section. Basin geometry and strati- †[email protected]. Dickinson and Lawton, 2001b). Prior to the graphic architecture are extensively concealed

GSA Bulletin; November/December 2020; v. 132; no. 11/12; p. 2489–2516; https://doi.org/10.1130/B35433.1; 13 figures; 5 tables; Data Repository item 2020158. published online 8 April 2020

© 2020 The Authors. Gold Open Access: This paper is published under the terms of the CC-BY license. 2489

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

by younger sedimentary and volcanic rocks son et al., 1986; Lawton and McMillan, 1999; ­potentially rendering inception of foreland his- and overprinted by subsequent deformation Lawton, 2000, 2004; Dickinson and Lawton, tory difficult to recognize. and magmatism (e.g., Soreghan, 1998; McKee 2001b). Basin geometry, forebulge develop- The Bisbee basin, an important archive of the et al., 2005; Amato et al., 2009; Mauel et al., ment, and the migration history of foreland sedimentary history of the U.S.–Mexico bor- 2011; González-León et al., 2011; Clinkscales basins formed on recently rifted lithosphere der region, initially formed as part of the Late and Lawton, 2018). In addition, the middle Cre- differ significantly from equivalent character- Jurassic Mexican Border rift system (Fig. 1; taceous foreland basin formed on continental istics of basins developed upon old lithosphere MBR, originally termed Mexican Borderland crust only newly extended during Jurassic– that behaves like an unbroken elastic plate (Fil- rift; Lawton and McMillan, 1999). The MBR Cretaceous rifting (Bilodeau, 1982; Dickin- dani and Hessler, 2005; Fosdick et al., 2014), extended from the McCoy basin of southern

Explanation USA North 40° N 85-100 Ma plutons 100-148 Ma plutons Map area Mexican Undifferentiated Sierra Nevada border rift 30° N batholith NE of San Andreas fault Gulf of Baja magnetic anomaly Mexico Figure 1. Location map of 20° N Mexico Intra-rift uplift Basement age localities discussed in text. province boundary 110° W 100° W 90° W 80° W Communities: A—Albuquer- que; Ba—Batopilas; En— Nevada Ensenada; EP—El Paso; Central Nevada 115° W 110° W 105° W LA—Los Angeles; LV—Las thrust belt n Vegas; P—Phoenix; SC—Sil- ver City; T—Tucson. Geologic Utah localities of study: Ar—Arizpe; WPt Sevier orogenic belt Arizona New Mexico CC—Clyde Canyon (Burro SJb Mountains); CM—Chirica- LV hua Mountains; CR—; Cu—Cucurpe; Ha— SY CordilleranColorado foreland basi Plateau Ha Hagan basin; HM—Huachuca Mountains; LHM—Little California Mogollon 35° N SAf A Hatchet Mountains; PM—Pel- Mojave LA Yavapai oncillo Mountains; SA—south- C.P. C.P. Mazatzal C.P. ern ; McC Highlands Fig. 11 SR—Saddlerock Canyon (Burro Mountains); SY—San SR Fence diagram P Ysidro; Tu—Tuape. Geologic CC SC SA abbreviations: ABf—Agua rift flank PM CR Blanca fault; Cah—Cana- EP Bisbee T CM nea high; Cht—Chihuahua En BootheLHM Bisbee trough; C.P.— ABf HM el Huachuca crustal province; McC—Mc- Gulf of California Altar Coy basin; Saf—San Andreas -CucurpeCah fault; SJb—; Caborca block Basin Grenville C.P. Cu WPt—Wheeler Pass thrust. Ar Pacific Tu 30° N Altar-Cucurpe, Bootheel, and rift flank(?) Huachuca labels indicate sub- Cht basins of Bisbee basin. Plutonic Sonora belts are igneous roots of mag- Baja California Inner flank of matic arcs discussed in text (adapted from Jacobson et al., Peninsula Guerrero composite Chihuahua 2011; Hildebrand and Whalen, terrane 2014). for Ba Mexico geographic reference only.

0 300 km

2490 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

California and southwestern Arizona (Fig. 1) 2016). As a result of an evident low volume of southeastward reintroduction of the subducted across southern Arizona, northern Sonora, and pre-Aptian Early Cretaceous magmatism, and Farallon slab beneath continental lithosphere of southwestern New Mexico, then southeastward possibly because of its distribution within the the southwestern­ corner of North America. to the Gulf of Mexico via the Chihuahua trough batholith, ca. 143–124 Ma zircon grains are un- of northern Mexico (Dickinson and Lawton, common in Cretaceous strata of the Cordilleran GEOLOGIC SETTING 2001b; Haenggi, 2002; Lawton, 2004; Mickus foreland basin in the (Laskowski et al., 2009; Stern and Dickinson, 2010; Spencer et al., 2013), in contrast with grain age distribu- Most strata in this study pertain to the Bis- et al., 2011). The MBR began to form in Late Ju- tions in the foreland basin of northern Mexico bee , defined from the Bisbee mining dis- rassic time as a northwest- and west-northwest– (Juárez-Arriaga et al., 2019). trict in southeastern Arizona (Fig. 1; Ransome, trending assemblage of rift basins created by the In this paper, we integrate new and published 1904). The type includes the combined effects of rollback of an oceanic slab stratigraphic, petrographic, and geochrono- Glance Conglomerate and Morita, Mural, and beneath southwestern North America (Lawton logic data for –middle Cenomanian Cintura Formations, which span latest Juras- and McMillan, 1999; Dickinson and Lawton, strata in southern Arizona, northern Sonora, and sic–late Early Cretaceous time (Ransome, 1904; 2001a; Fitz-Díaz et al., 2018) and separation of southwestern and northwestern New Mexico Stoyanow, 1949; Bilodeau and Lindberg, 1983; North and South America during the breakup of to interpret Late Jurassic through middle Cre- Dickinson et al., 1986). Following recognition Pangea (e.g., Pindell and Kennan, 2009). Sedi- taceous basin development of the southwestern that the Glance Conglomerate records synde- mentary basins of the MBR are alternatively in- U.S.–Mexico border region. Stratigraphic corre- positional normal faulting (Bilodeau, 1979, terpreted as an array of pull-apart basins formed lation, subcrop relations, petrography, and U-Pb 1982), the Bisbee basin and its eponymous fill along a throughgoing Middle–Late Jurassic detrital-zircon and ash-fall tuff ages permit an came to be regarded as products of continental (Anderson, 2015) or Late Jurassic (Anderson improved understanding of the sedimentary rifting (Bilodeau, 1979, 1982; Dickinson et al., and Silver, 2005) sinistral transcurrent fault, the history of southwestern North America and the 1986; Lawton and McMillan, 1999). Thinning Mojave-Sonora megashear, which is inferred to nascent Western Interior seaway and provide in- and pinch-out of the Glance Conglomerate in have accommodated the opening of the Gulf of sight into the tectonics of foreland history that the Mule Mountains north of Bisbee, Arizona, Mexico (Anderson and Nourse, 2005), Impor- followed crustal extension. We use the new dates indicate the presence of a topographic high of tantly, Late Jurassic–middle Cretaceous deposi- and previously published biostratigraphic ages incompletely defined distribution within the tion in the Bisbee basin continued beyond the to construct geohistory diagrams, or subsidence greater Bisbee basin (Fig. 1; Bilodeau, 1982; extensional history of the MBR, a theme we curves, for Upper Jurassic–Cenomanian strata Bilodeau et al., 1987). A part of this paleohigh, develop in this paper. in southwestern New Mexico and northern So- where it is expressed in the Mule Mountains, was West and southwest of the MBR in Mexico, nora. Published subsidence histories for Bisbee termed the Bisbee block (McKee et al., 2005). Late Jurassic–middle Cretaceous magmatism basin strata in southwestern New Mexico (Mack, Discovery of Upper Jurassic marine, con- took place without apparent interruption along 1987a) and Sonora (González-León, 1994) in- tinental, and mafic volcanogenic strata in nar- the North American continental margin. Mag- cluded only Early Cretaceous strata because row basins in northern Sonora, southeastern- matic history is inferred on the basis of zircon Upper Jurassic strata from New Mexico were most Arizona, and southwestern New Mexico grain ages in the adjacent retroarc foreland basin not yet described; in addition, correlative Juras- ­corroborated the previously posited extensional of northern Mexico, which are common in the sic strata in Sonora (e.g., Rangin, 1977) had not origin of the Bisbee basin (Lawton and Olm- range ca. 145–123 Ma, less common in the range yet been factored into the stratigraphic history stead, 1995; Lawton and Harrigan, 1998; Lu- ca. 122–110 Ma, and abundant in the range ca. of the basin. Improved biostratigraphic and geo- cas et al., 2001; Villaseñor et al., 2005; Mauel 105–88 Ma (Juárez-Arriaga et al., 2019). In the chronologic data published over the intervening et al., 2011). Marine and volcanic strata in the southwestern-most U.S. and along western Mex- 25 and new data presented here yield simi- Mountains of southeastern Arizona ico, Early Cretaceous arc magmatism occurred lar subsidence histories from the two locations. that overlie a thin (∼40 m) basal conglomer- in the Guerrero composite terrane, which col- The regionally consistent subsidence curves sug- ate interpreted as Glance Conglomerate were lided diachronously with North America during gest that strata of the Bisbee basin, as defined assigned to the Bisbee Group (Drewes et al., a debated time interval, probably sometime in above, record regional Late Jurassic–middle 1995), a stratigraphic convention that was re- the Aptian–Cenomanian (Dickinson and Law- Albian crustal extension and passive thermo- tained when the section was recognized as Ju- ton, 2001a; Centeno-García et al., 2008; Martini tectonic subsidence followed by rapid late Al- rassic (Fig. 2) (Lawton and Olmstead, 1995). et al., 2014). Magmatism in the Sierra Nevada bian–Cenomanian subsidence, likely of com- Bisbee Group nomenclature was later extended of eastern California, which is generally inferred bined flexural and dynamic origin. We infer a into southwestern New Mexico (Lawton and to have been caused by east-dipping subduction tectonic history in southwestern North America Harrigan, 1998; Lucas and Estep, 1998b), of the Farallon slab beneath the North American of rift basin subsidence that accompanied Late where the group presently includes the Upper plate, experienced a mid-Cretaceous magmatic Jurassic separation of the Guerrero composite Jurassic Broken Jug Formation and the Lower lull, after which it became active beginning at ca. volcanic terrane (abbreviated as Guerrero; e.g., Cretaceous Hell-to-Finish, U-Bar, and Mojado 123 Ma and peaked in volume between 100 Ma Dickinson and Lawton, 2001a) from the conti- Formations (Zeller, 1965, 1970; Lawton, 2004; and 90 Ma (Ducea, 2001; DeCelles et al., 2009). nental margin during formation of a marginal Lucas and Lawton, 2000; Lucas et al., 2001). A A compilation of 430 U-Pb ages from the Sierra backarc basin southward from the latitude of well-exposed Jurassic–Lower Cretaceous sec- Nevada batholith south of 37.25°N demonstrates Ensenada, Baja California (in present coordi- tion overlies carbonate rocks in the a nearly continuous range of ages from 248 Ma nates; Fig. 1). Rifting was succeeded by broad , northward of which to 76 Ma but contains only 11 ages (2.6% of the foreland subsidence during progressive closure and strata thin toward the total) in the range 145–124 Ma (Nadin et al., of the intervening marginal basin and middle Burro Mountains, where the middle Cretaceous 2016). Ages in the range 140–120 Ma lie along to late Albian collision of Guerrero with the Beartooth Member of the the western flank of the batholith (Nadin et al., continental margin, combined with progressive directly overlies granitoids

Geological Society of America Bulletin, v. 132, no. 11/12 2491

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

Geologic Time Scale Sonora SE Arizona SW New Mexico El Paso area NW New Mexico TS Assemblage Mancos Mancos Redbeds; Ma Cenomanian Mancos 5 Fluvial strata Late ? Buda/Del Rio Dakota 100 ? La Juana Conglomeratic Mojado strata Cintura Cintura Mojado 4 Albian DSMM Marine sandstone; Finlay Delta front 110 Lágrima Mural Mural U-Bar Benigno 3 Cuchillo Shelfal

Aptian 120 Eolian sandstone Morita Morita Hell-to-Finish Las Vigas Early 2 Turbidite Burro Canyon/ Cretaceous (part)

130 Jackpile Ramp carbonate Valanginian ? ? Basinal carbonate; La Colgada ? Calciturbidites 140 Duration of hiatus unknown Shoreface sand- in AZ and NM stone and laminated Navarrete carbonate 150 Onion Saddle Basalt Mbr Dolostone Cucurpe La Casita/ Morrison Kimmeridg. Malone Subaerial basalt Late 1 ?

Glance Cave Crystal 160 Jurassic (part) Oxfordian San Rafael Gp. Pillow basalt “Glance” Broken Jug

Middle Jurassic Permian Strata Permian Strata strata strata Calculated DZ MDA U-Pb zircon age on tuff

Figure 2. Correlation chart showing formations of tectonostratigraphic assemblages in Sonora, Arizona, New Mexico, and west Texas. Upper Albian–Cenomanian units for the El Paso region utilize nomenclature of Lucas et al. (2010); lower Albian and older units are gen- eralized from the Chihuahua trough to the south in Mexico (after Haenggi, 2002). after Cohen et al. (2013, updated). DSMM: Del Norte, Smeltertown, Muleros, and Mesilla Valley Formations, in ascending order. AZ—Arizona; MDA—maximum deposi- tional age; NM—New Mexico; Gp—Group.

and metamorphic rocks (Paige, 1916; Darton, Morita, Mural and Cintura, together with local vary among sub-basins within the broader basin: 1917; Lasky, 1936; Chafetz, 1982). Thinning of variants, are employed in Sonora (González- (TSA 1) Upper Jurassic volcanic-sedimentary the section takes place by erosional León, 1994; Jacques-Ayala, 1989, 1995). The strata that include deep marine, deltaic, fluvial, truncation of upper Paleozoic strata beneath the Bisbee basin in Sonora lies south of a horst block and alluvial deposits; (TSA 2) Valanginian–low- Lower Cretaceous section and by northward termed the Cananea high, which lacks a Juras- er Aptian continental redbeds and local shallow stepwise thinning of Lower Cretaceous strata sic–Lower Cretaceous section (Fig. 1; McKee marine deposits; (TSA 3) upper Aptian–middle at syndepositional normal faults (Bayona and and Anderson, 1998; McKee et al., 2005; Page Albian carbonate and siliciclastic strata; (TSA Lawton, 2003; Lawton, 2004; Machin, 2013). et al., 2010). South of the Cananea high, an Up- 4) upper Albian–lower Cenomanian siliciclastic The Early Cretaceous paleogeographic high in per Jurassic deepwater section with interbedded continental and shallow-marine strata; and (TSA the Burro Mountains is termed the Burro uplift mafic volcanic rocks termed the Cucurpe For- 5) middle–upper Cenomanian shale with inter- (Elston, 1958), a southeastward extension of the mation underlies the Morita Formation locally bedded ash-fall tuff, bentonite, and subordinate Mogollon highlands (Fig. 1; Bilodeau, 1986). and is equivalent in part to the Glance Forma- sandstone. The lithic types within these assem- A resulting southeast-trending paleogeographic tion (Figs. 1 and 2; Villaseñor et al., 2005; Mauel blages vary in detail, but each assemblage can be element, the Burro-Mogollon uplift, formed the et al., 2011; Peryam et al., 2012). Upper Jurassic recognized throughout the region. Late Jurassic–Early Cretaceous rift shoulder of strata of Sonora are not included in the Bisbee TSA 1 consists of marine shale, locally tuffa- the Bisbee basin (Bilodeau, 1982; Dickinson Group (Mauel et al., 2011). ceous turbidites, subaqueous and subaerial mafic et al., 1986; Mack, 1987a, 1987b). Bisbee Group volcanic rocks and hyaloclastite, and feldspathic strata are present on the international border be- Tectonostratigraphic Assemblages to volcanic-lithic sandstone (Fig. 2; Lawton and tween Chihuahua and New Mexico near El Paso, Olmstead, 1995; Mauel et al., 2011). Fossils of Texas (Fig. 1), where they consist of We divide Upper Jurassic–Lower Cretaceous Tethyan affinity indicate an Oxfordian–Titho- and subordinate sandstone intercalated with car- strata of the broadly defined Bisbee basin into nian age range and a connection with the Gulf bonate strata, which dominate the middle Cre- five regionally correlative tectonostratigraphic of Mexico (Olmstead et al., 1996; Olmstead taceous section of the Chihuahua trough (Mon- assemblages (TSAs; Fig. 2). Tectonostrati- and Young, 2000; Lucas et al., 2001; Villase- real and Longoria, 2000; Haenggi, 2002; Lucas graphic assemblages “record phases of basin ñor et al., 2005). Interbedded mafic volcanic et al., 2010). history during which the fundamental controls of rocks have ocean-island basalt chemistry, indi- Lower Cretaceous strata of northern Sonora tectonic setting, subsidence style, and basin ge- cating asthenospheric derivation (Lawton and are likewise included in the Bisbee Group. The ometry are relatively similar” (May et al., 2013, McMillan, 1999). In southern Arizona, the as- stratigraphic succession there resembles that p. 1403). The TSAs of this study consist of one semblage is instead represented by the locally of Arizona, and the formation names Glance, or more formations whose names in some cases thick alluvial Glance Conglomerate, which

2492 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

thins across syndepositional normal faults near Sonora and the Huachuca basin indicates onlap to the Yavapai and Mazatzal basement provinces the town of Bisbee (Bilodeau, 1982; Bilodeau of Lower Cretaceous strata onto beveled, tilted (Fig. 1; ca. 1.79–1.64 Ga; Karlstrom et al., 2004; and Lindberg, 1983; Bilodeau et al., 1987). Ma- Paleozoic and Jurassic sections (Soreghan, 1998; Anderson and Silver, 2005; Amato et al., 2008). rine strata in Sonora grade to apparently alluvial Mauel et al., 2011). Elsewhere, Jurassic strata Proterozoic granitoids with ages of ca. 1.46 Ga conglomerate, which is also termed Glance, on were tightly folded prior to deposition of TSA 2 in the Burro Mountains of southwestern New the south flank of the Cananea high in northern (Peryam et al., 2012). Mexico intrude older (ca. 1.65 Ga) metamor- Sonora (Fig. 1; Peryam et al., 2012), indicating TSA 3 consists of interbedded carbonate and phic rocks (Amato et al., 2009; Amato et al., local high relief and syndepositional fault move- siliciclastic strata deposited in a carbonate ramp 2011). In the Little Hatchet Mountains of south- ment (e.g., McKee and Anderson, 1998; McKee setting that included large rudistid reefs, skeletal western New Mexico, basement rocks include et al., 2005). TSA 1 defines initial development carbonate platforms, delta front environments, granitoids with ages of ca. 1.1 Ga (Clinkscales, of three separate, southeast-trending rift basins, and mixed siliciclastic-carbonate shorelines 2011; Amato and Mack, 2012; Clinkscales and each with depositional connections via a system (Zeller, 1965, 1970; Archibald, 1987; Ferguson, Lawton, 2018). of extensional basins to the nascent Gulf of Mex- 1987; Warzeski; 1987; Jacques-Ayala, 1989; ico (e.g., Bilodeau, 1982; Lawton and McMillan, Lawton et al., 2004; González-León et al., 2008). Middle Mesozoic Magmatic Arcs 1999). We use the terms Altar-Cucurpe basin for Abundant fossils provide the stratigraphically the sub-basin south of the Cananea high (Mauel densest age control of the section (Lucas and The Peninsular Ranges and Sierra Nevada et al., 2011), Huachuca basin for the sub-basin in Estep, 1998b; Lawton et al., 2004; González- batholiths constitute the roots of Jurassic–Mid- southern Arizona, which contains the type Bis- León et al., 2008). TSA 3 strata have been inter- dle Cretaceous magmatic arcs of southwestern bee Group, and Bootheel basin for the sub-basin preted as a record of post-rift thermal subsidence North America that were active during deposi- in southeastern Arizona and southwestern New (Mack, 1987a, 1987b; González-León, 1994). tion in the Bisbee basin; thus, they represent Mexico. The Huachuca basin as defined here TSA 4 is a thick succession of fluvial and possible sources of sediment in the basin. The encompasses smaller basins previously termed shallow-marine siliciclastic strata represented Peninsular Ranges batholith (PRB) is zoned the Hereford basin of Late Jurassic age and the by the Mojado Formation of southwestern New from east to west in composition and age (Gas- upper Bisbee basin, in which strata of TSA 2–4, Mexico and the Cintura Formation of Arizona til, 1975; Silver and Chappell, 1988; Todd et al., described below, accumulated (McKee et al., and Sonora. These units overlie TSA 3 with a 1988; Kimbrough et al., 2001). The western zone 2005). The general extent of the Altar-Cucurpe sharp contact. The Mojado Formation has been includes the roots of an older arc system that is basin has also been termed the Sonora basin, interpreted as recording inception of a flexural composed of the Santiago Peak and Alisitos which is inferred to have accumulated sediment foreland basin in New Mexico (Mack, 1987a, arcs and exposed in the northern Baja California from Late Jurassic through time (Rodrí- 1987b). In Sonora, by contrast, the Cintura Peninsula and southern California. The Santiago guez-Castañeda, 2002), an interval that includes Formation is interpreted to record continued Peak arc, represented by an Upper Jurassic– post-Bisbee basin tectonics; therefore, we pre- post-extensional, thermo-tectonic subsidence Lower Cretaceous volcano-sedimentary succes- fer the more temporally restricted term of Mauel (González-León, 1994). TSA 4 also includes sion that has yielded Late Jurassic marine fossils et al. (2011). Because basin margins are not well marine shale, , and conglomerate of and U-Pb zircon ages ranging from 138 Ma to defined everywhere, and normal faults created the informal La Juana formation, which over- 120 Ma, extended from the Transverse Ranges numerous local half-graben structures within the lies the Cintura Formation in northern Sonora of southern California to the Agua Blanca fault basins (e.g., Soreghan, 1998), sub-basin extents (Mauel et al., 2011). directly south of Ensenada (∼32°N; Fig. 1; Wet- are depicted schematically in Figure 1. TSA 5 includes marine strata that overlie up- more et al., 2002, 2003). Inherited TSA 2 unconformably overlies Permian to per Albian (or, locally, earliest Cenomanian) cores in some zircon grains indicate a deposi- Jurassic strata in all depozones and consists of strata on a sharp contact in southwestern New tional connection with basement of southwestern dominantly fluvial continental strata encompass- Mexico (Cobban et al., 1989; Lucas et al., 1988, North America (Wetmore et al., 2002). ing channel sandstone complexes that are lo- 2000; Lucas and Lawton, 2005; Machin, 2013), Plutons of the Alisitos arc crop out from the cally conglomeratic and interbedded with thick but they are largely absent from southern Arizo- Agua Blanca fault to the central part of the Baja red siltstone intervals containing calcareous na and likely eroded before and during Laramide California Peninsula (∼28°N); although covered (Dickinson et al., 1986; Mack et al., deformation. This assemblage includes the Man- by younger volcanic rocks farther south, the plu- 1986; Mauel et al., 2011; Peryam et al., 2012). cos Formation, locally exposed in southwest- tons create a strong magnetic anomaly that ex- Interfingering alluvial fan and thick lacustrine ern New Mexico, and the tends to the tip of the Baja California Peninsula deposits are present in the complexly faulted in northwestern New Mexico, with a base that but does not clearly continue onto the Mexican northwestern part of the Huachuca basin, which correlates with the uppermost part of TSA 4 of mainland (Fig. 1; Langenheim and Jachens, contained several half-graben structures during southwestern New Mexico. TSA 5 permits cor- 2003). Associated strata contain Aptian–Albian Aptian time (Soreghan, 1998). Deformation of relation of late- Bisbee basin history with fossils and a narrow range of zircon ages (116– subjacent strata of TSA 1 varies significantly early deposition in the Western Interior seaway 115 Ma; Wetmore et al., 2002). Zircon grains with respect to that of TSA 2. In southwestern and Cordilleran foreland basin. lack inheritance (Wetmore et al., 2002), and New Mexico and southeastern Arizona, local detrital zircons in basinal sandstone older than discordance with underlying Paleozoic or Juras- Basement Rocks 110 Ma lack evidence of continental derivation, sic strata is slight, but regional relations indicate suggesting isolation from North America prior to abrupt changes of subjacent strata across high- Important to detrital zircon provenance in- that time (Alsleben et al., 2012). angle faults of Jurassic and Early Cretaceous terpretations, local basement rocks of northern The Alisitos arc was likely developed on oce- age (Lawton, 2000, 2004; Bayona and Lawton, Sonora, southeastern Arizona, and southwestern anic lithosphere and is alternatively interpreted 2003). Omission of older TSA 2 strata above New Mexico consist of conti- as an exotic oceanic arc (Dickinson and Law- the unconformity at neighboring localities in nental crust of southwestern Laurentia assigned ton, 2001a; Wetmore et al., 2003) or fringing

Geological Society of America Bulletin, v. 132, no. 11/12 2493

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

arc separated from North America by an ocean tism recorded mainly in the western part of the the general shape of the subsidence curves. We basin (Busby et al., 2006; Marsaglia et al., 2016; Sierra Nevada batholith (Nadin et al., 2016) that employ the current geologic time scale of the In- Boschman et al., 2018). Farther south in western occurred during continuing magmatic activity in ternational Commission on Stratigraphy (Cohen Mexico, the Guerrero terrane contains Upper Ju- the Alisitos arc, which is described above. The et al., 2013; updated). rassic–Lower Cretaceous volcanic and intrusive part of the second episode is rocks (Talavera-Mendoza et al., 2007; Mortensen thus, in part, correlative with the La Posta event VALANGINIAN–CENOMANIAN et al., 2008; Centeno-García et al., 2008, 2011; (Ducea, 2001; Kimbrough et al., 2001). Ash-fall STRATIGRAPHY OF NORTHERN Martini et al., 2009, 2011). Quartzarenite col- tuff beds generated during La Posta eruptions are SONORA, SOUTHEASTERN ARIZONA, lected from deformed deep-marine rocks near widely distributed in Cenomanian and AND SOUTHWESTERN NEW MEXICO the city of Zacatecas in north-central Mexico marine deposits of the Western Interior seaway has a maximum depositional age of 109 ± 3 Ma (Christiansen et al., 1994). We focus here on Lower and middle Creta- (early Albian) and contains a dominant detrital ceous siliciclastic strata of TSAs 2, 4, and 5, zircon age group ranging over ca. 145–108 Ma, METHODS although fossiliferous strata of TSA 3 are im- with age peaks at ca. 158 Ma, 137 Ma, 130 Ma, portant in regional correlation. TSA 1 has been and 117 Ma, that is interpreted as derived from Zircons were separated from ∼5–10 kg of described previously (Lawton and Olmstead, Guerrero (Ortega-Flores et al., 2016). The de- sample using standard crushing and mineral 1995; Lawton and Harrigan, 1998; Busby trital ages indicate that igneous rocks with ages separation techniques, including magnetic et al., 2005; Bassett and Busby, 2005; Mauel equivalent to the Early Cretaceous Santiago Peak separation and dense liquid settling in sodium et al., 2011). and Alisitos arcs are likely present in mainland polytungstate and methylene iodide. The high- Mexico as part of the Guerrero composite ter- density fraction was washed in acetic acid if car- Tectonostratigraphic Assemblage Two rane, consistent with interpretations of previous bonate minerals were present and in nitric acid workers that include the Santiago and Alisitos if sulfide minerals were present. Zircons were Rancho La Colgada and Morita Formations arcs in Guerrero (e.g., Dickinson and Lawton, hand-picked (for igneous samples) or randomly In northern Sonora, siliciclastic marine strata 2001a). The detrital record also demonstrates poured (for detrital samples) and mounted in ep- of the Rancho La Colgada Formation uncon- that Guerrero contains a suite of Lower Creta- oxy for analysis. formably overlie Upper Jurassic deepwater strata ceous magmatic rocks older than 117 Ma, in Laser ablation–multi-collector–inductively of the Cucurpe Formation and grade upsection the range of ca. 145–108 Ma. metasedi- coupled plasma mass spectrometry (LA–MC– to fluvial strata of the Morita Formation, which mentary rocks that make up the oldest known ICPMS) and single-collector LA–ICPMS dating ranges from 700 m to 1200 m thick (González- rocks of Guerrero contain detrital zircons that were conducted on detrital and igneous zircon León, 1994; Peryam et al., 2012). The Rancho indicate a connection with continental Mexico grains at the Arizona Laserchron Laboratory at La Colgada Formation is absent at some locali- prior to development of an oceanic basin, the the University of Arizona. Beam diameter was ties where fluvial Morita beds directly overlie Arperos basin, between Mexico and Guerrero, generally 20–30 µm. Errors on spot ages of in- the Jurassic section (Fig. 3; Mauel et al., 2011; likely in latest Jurassic time (Centeno-García dividual zircon grains are reported in the text Peryam et al., 2012). The contact between et al., 2008). and tables at 1σ, and we report weighted mean ­Jurassic and Cretaceous strata in northern Sono- The eastern zone of the PRB is defined by ages in the text and figures at the σ2 level. We ra has alternatively been considered transitional tonalite and low-K granodiorite intrusions report and plot 206Pb/238U ages for grains <1 Ga ­(Rodríguez-Castañeda, 1990, 1991). termed La Posta suite, named after an intrusion and 206Pb/207Pb ages for grains >1 Ga. Data are In southern Arizona, fluvial channel com- that spans the U.S.–Mexico border (Silver and presented on concordia diagrams using Isoplot plexes and red to purple siltstone, commonly Chappell, 1988; Walawender et al., 1990). The (Ludwig, 2012). with calcareous nodules, similarly dominate the voluminous La Posta suite constitutes ∼47% of New and published stratigraphic ages were Morita Formation (Fig. 3). At Bear Canyon in the the surface exposures in the Peninsular Ranges used to construct geohistory diagrams for south- southeastern Huachuca Mountains, red siltstone batholith and has U-Pb ages ranging from 98 Ma western New Mexico and north-central Sonora. and sandstone—interpreted here as the basal part to 92 Ma, which indicates voluminous magma The diagrams incorporate the decompaction of the Morita Formation—overlie thick-bedded, emplacement during that time interval (Kim- algorithim of Angevine et al. (1990) and show clast-supported Glance Conglomerate on a sharp brough et al., 2001). La Posta intrusions postdate subsidence history for complete stratigraphic contact. Siltstone and sandstone beds contain ir- an episode of shortening that deformed western sections in the two regions. Age inputs for the regular micrite nodules and interbeds of micrite- zone igneous rocks during the time interval 118– subsidence curves include biostratigraphic data, granule conglomerate 50–125 cm thick. This 105 Ma (Todd et al., 1988; Thomson and Girty, U-Pb ages on tuff beds, and maximum deposi- stratigraphic interval, ∼30 m thick, represents a 1994; Johnson et al., 1999). tional ages calculated from the youngest zircon succession of paleosols and conglomerate con- The California arc, represented by the Si- grain ages that overlap at 2σ error, using the taining micrite nodules reworked from the pa- erra Nevada batholith, had a somewhat differ- weighted mean algorithm of Isoplot (Ludwig, leosol horizons (McKee et al., 2005). Although ent magmatic history than the arcs of the Baja 2012). Errors for these different data types are on measured bed attitudes do not unambiguously California Peninsula (Ducea, 2001). Magmatism the order of ±1 m.y. (∼1% 2σ); errors for strati- demonstrate discordance between Glance and took place in two short-lived episodes, a Late graphic thicknesses are approximately ± 5%. Morita strata in Bear Canyon, the and Jurassic episode from 160 Ma to 150 Ma and The analytical method is sensitive to the thick- granule conglomerate interval thickens to great- a more voluminous Late Cretaceous episode ness of overburden inferred to have overlain the er than 50 m in outcrops 2100 m to the north- beginning at ca. 121 Ma and that became par- analyzed section, which is estimated from re- west, which suggests angular discordance at the ticularly marked from 100 Ma to 85 Ma (Ducea, gional stratigraphic relations, but this factor con- Glance-Morita contact. Previous workers have 2001; DeCelles et al., 2009; Nadin et al., 2016), sistently affects the degree of corrected compac- recognized the condensed pedogenic nature of with an intervening episode of reduced magma- tion in a particular section and does not change the Glance-Morita transition interval; some have

2494 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

A Altar-Cucurpe Basin B Huachuca Basin C Bootheel Basin near Cucurpe, Sonora SW Huachuca Mountains, AZ Little Hatchet Mountains, NM Cananea uplift Unnamed uplift meters Mural Limestone meters U-Bar Formation 500 115 ± 1 Ma (04SJQ5) 1000 Morita Formation Hell-to-Finish Formation Red nodular siltstone Incomplete Fluvial sandstone, siltstone, conglomerate meters 300 123 ± 1 Ma (11BC1) Fluvial volcanic-lithic ss 1014 ± 30 Ma (HF1) Fluvial quartzose ss 0 Basalt Member Shoreface deposits with burrows of Broken Jug Formation Fluvial sandstone & 198 ± 5 Ma (10LM11) 0 vf fmc conglomerate Angular unconformity 138 ± 2 Ma (9/4LA1) Glance Conglomerate vf fmc Shale Siltstone Sandstone Shale 134 ± 2 Ma (05SJQ1) Conglomerate Siltstone

250 ± 3 Ma (9/4LA5) Sandstone

Conglomerate Siltstone and fine-grained Conglomerate Rancho La Colgada Formation sandstone with pedogenic nodules Shoreface deposits Sedimentary Carbonaceous shale breccia 0 Angular unconformity Cucurpe Formation Sandstone Subaerial basalt flow

vf fmc Well sorted marine sandstone Carbonate nodules Shale Carbonate Siltstone Sandy limestone mudstone Sandstone Conglomerate 198 ± 5 Ma (10LM11) MDA in Ma of (DZ sample)

Figure 3. Generalized stratigraphic sections of tectonostratigraphic assemblage 2, indicating stratigraphic levels of detrital zircon (DZ) samples. (A) Composite section from vicinity of Rancho La Colgada, northern Sonora (generalized from Peryam et al., 2012). (B) Partial section, Bear Creek, southwestern Huachuca Mountains, Arizona (after Gilbert, 2012). (C) Measured section of Hell-to-Finish Forma- tion, central part of Little Hatchet Mountains, New Mexico (after Lucas and Lawton, 2000). Abbreviations: DZ—detrital zircon; MDA— maximum depositional age; sandstone grain size: vf—very fine; f—fine; m—medium; c—coarse. Age data are shown inTables 2 and 3.

assigned it to the Glance (McKee et al., 2005) (10LM11) and a channel sandstone bed of the Lawton, 2000). In the northern Little Hatchet and others to the Morita (Vedder, 1984; Klute, volcanic petrofacies (11BC1). Mountains, a section with faulted top and base is 1991). Strata overlying the condensed interval 580 m thick and consists of limestone- and chert- contain laterally continuous beds of fine- to me- Hell-to-Finish Formation pebble conglomerate, feldspathic sandstone, silt- dium-grained, well-sorted quartzose sandstone The Hell-to-Finish Formation of southwest- stone, and shale that is interpreted as alluvial-fan 5–10 m thick with abundant trough cross-beds, ern New Mexico overlies Jurassic basaltic and fluvial deposits (Mack et al., 1986). Inter- planar lamination, and abundantly burrowed bed rocks of the Broken Jug Formation on a sharp bedded conglomerate and fossiliferous sandy tops (Fig. 3). We interpret this part of the section but concordant contact in the central part of the limestone in correlative strata of the Peloncillo as shoreface deposits approximately correlative Little Hatchet Mountains (Lawton and Harrigan, Mountains on the Arizona–New Mexico border with the Rancho La Colgada Formation. Conti- 1998), whereas elsewhere it overlies Paleozoic are interpreted as fan-delta deposits (Bayona nental fluvial strata continue to a stratigraphic limestone (Gillerman, 1958; Zeller, 1965). The and Lawton, 2003). The Hell-to-Finish Forma- level 300 m above the base of the Morita, where formation consists of red siltstone and thin, up- tion has been interpreted as continental deposits the sandstone composition abruptly changes to ward-fining beds of conglomerate and sandstone with transverse alluvial sediment dispersal from a volcanic compositional petrofacies (Klute, with scoured bases. A stratigraphically complete local bounding uplifts and axial, southeast-ori- 1987). We collected samples for detrital zir- exposure of the unit in the central part of the Lit- ented fluvial systems (Lawton, 2004). A single con (DZ) analysis from the shoreface deposits tle Hatchet Mountains is 525 m thick (Lucas and sample of Hell-to-Finish sandstone (HF1) was

Geological Society of America Bulletin, v. 132, no. 11/12 2495

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

collected from the lower part of the unit in the Altar-Cucurpe basin near the end of the Albian, bedded sandstone of fluvial origin Fig. 4( ). The central Little Hatchet Mountains. a transgression recorded in the Bootheel basin by remainder of the section consists of interbed- siliciclastic strata. ded heterolithic feldspathoquartzose sandstone, Tectonostratigraphic Assemblage Four siltstone, and shale and upward coarsening and Mojado Formation thickening successions of burrowed sandstone, Cintura Formation The Mojado Formation of southwestern which we interpret as estuarine and tidally in- In northern Sonora and southeastern Arizona, New Mexico is a succession of middle Creta- fluenced shoreface deposits, respectively. Paleo- the uppermost formation of the Bisbee Group ceous sandstone, siltstone, and shale in the Big current measurements in the upper part of the is termed the Cintura Formation. In Sonora, it Hatchet and Little Hatchet Mountains (Zeller, unit indicate sediment transport to the southwest, gradationally overlies the Mural Formation and 1965, 1970) and adjacent ranges, including probably by ebb-tidal currents. consists of as much as 2000 m of interbedded the Peloncillo Mountains and Cookes Range, content and bracketing stratigraphic quartzose sandstone and red-weathering silt- where correlative strata were formerly termed relations indicate that the Mojado Formation stone, commonly with calcareous pedogenic the Johnny Bull and Sarten Formations (Giller- in the Cookes Range is late Albian to earliest nodules (Mauel et al., 2011). The entire forma- man, 1958; Mack et al., 1988; Lucas and Estep, Cenomanian. Upper Albian ammonites are pres- tion represents deposits of fluvial systems with 1998b; Lawton, 2004). The Mojado Formation ent in lower part of the marine section (Fig. 4), paleocurrent directions to the northeast and generally thins northward from its thickest oc- and lower Cenomanian molluscs characteristic southeast (González-León, 1994). In Sonora, currences in southwestern New Mexico and of the Del Rio and Buda Formations of Texas the informal La Juana formation of TSA 4, de- overlies progressively older rocks to the north. are present in the upper part of the section (Cob- scribed below, gradationally overlies the Cintura In the northern part of the Little Hatchet Moun- ban, 1987; Lucas et al., 1988; Lucas and Estep, Formation and correlates with marine intervals tains, the formation is 1245 m thick (Galemore, 1998b). Fossils suggest that the entire Mojado in the upper part of the Cintura and Mojado for- 1986); in the Cookes Range, a composite section Formation in the is late mations of the Bootheel basin. is ∼107 m thick (Fig. 4; Lucas et al., 1988). In Albian (Lucas and Estep, 1998b). The dominant- In Arizona, the Cintura Formation contains southwestern New Mexico localities, it over- ly marine section of the Cookes Range locality strata composed of upward-fining sandstone lies marine carbonate strata of the upper Albian has been interpreted to represent the downdip that grades to red and gray siltstone (Dickinson U-Bar Formation on a sharp, concordant con- marine shoreline equivalent of the fluvial sys- et al., 1986). This formation is dominated by flu- tact variably interpreted as gradational (Mack tem of the Little Hatchet Mountains (Galemore, vial strata in the western part of the Bisbee basin, et al., 1986, 1988) or disconformable (Lucas 1986; Mack et al., 1986); alternatively, the flu- in the vicinity of Tucson, and in the Huachuca and Lawton, 2000). In the Cookes Range, the vial succession of the Little Hatchet Mountains basin (Klute, 1991). It is considered to grade to Mojado Formation overlies lower Permian red may predate deposition of the Cookes Range mixed marine and fluvial strata in southeastern beds of the Abo Formation (Clemons, 1982); far- section, with the possible exception of the lower Arizona (Dickinson and Lawton, 2001b), but ther east, in the southern San Andres Mountains fluvial beds there. the presence of marine beds in the Huachuca north of El Paso, Texas (Fig. 1), a thin section We collected three detrital zircon samples basin, although expected, remains unconfirmed. (∼6 m) of Mojado overlies lower Permian strata from the upper part of the Mojado Formation, The Cintura Formation in the Chiricahua Moun- and is inferred to be overlain by ∼50 m of Upper two from the upper member in the Little Hatchet tains of the Bootheel basin consists of laterally Cretaceous Dakota Formation with large trough Mountains (11BQ17, 11BQ19), and one in the extensive sandstone bodies deposited in shore- cross beds, although the relative proportion of Cookes Range nine meters below the top of the face settings, with rounded quartzite clasts to the two units there remains debated (Lucas and formation (12BQ37); we also review DZ data for 15 cm in diameter locally in the upper part of Estep, 1998a, and references therein). two previously published samples, one from the the section. Although the formation there is not In the Little Hatchet Mountains, the Mojado lower part of the formation in the Little Hatchet well exposed, fluvial strata appear to be absent. Formation consists of three members (Gale- Mountains (5.12.15.10; Clinkscales and Law- The age of the Cintura Formation in Arizona is more, 1986; Mack et al., 1986, 1988), formal- ton, 2015) and one from a shoreface deposit controlled by its stratigraphic position above the ly termed the Fryingpan Spring, Sarten, and in the upper part of the Cintura Formation of middle Albian Mural Limestone and below Up- Rattlesnake Ridge members (Lucas and Estep, the Chiricahua Mountains (KBCR; Dickinson per Cretaceous Laramide strata (Dickinson and 1998b; Lucas and Lawton, 2000), the upper and et al., 2009) that is probably equivalent to the Lawton, 2001b). lower of which represent shoreface deposits. The Rattlesnake Ridge Member of the Mojado in Sarten Member consists of 650 m of lenticular, New Mexico. La Juana Formation trough cross-bedded sandstone beds and red or Near Tuape, Sonora, red siltstone of the upper green shale deposited by meandering rivers that Beartooth Member of Mojado Formation part of the Cintura Formation grades to calcar- flowed eastward to a marine shoreline near El (Beartooth Quartzite) eous shale, fossiliferous limestone, sandstone, Paso, Texas (Galemore, 1986; Mack et al., 1986, The Beartooth Quartzite, named by Paige and pebble conglomerate of La Juana forma- 1988; Lucas et al., 2010). As the members pass (1916), is now termed the Beartooth Member of tion. Pebbles consist of tan, gray, and laminated into adjacent ranges, they either pinch out or the Mojado Formation (Lucas and Estep, 1998b). dolostone; fossiliferous, intraclastic, and oolitic change , with the result that correlations Prior to geochronologic data presented here, limestone; 10%–30% chert, which is more abun- of fluvial strata with fossil-bearing marine facies Beartooth-Mojado equivalence was inferred dant in the smaller pebble fraction; and white, are uncertain; therefore, we restrict our usage to based on stratigraphic position (“Beartooth fine-grained arenite. Fossil molluscs re- the Mojado Formation, except where our new problem” of Lawton, 2004). In the Burro Moun- covered from La Juana formation near Arizpe data permit better correlation between different tains, at Saddlerock Canyon, ∼20 km west of Sil- (Fig. 1) indicate a late Albian age (R.W. Scott, facies tracts (Fig. 4). In the Cookes Range, the ver City, New Mexico (Figs. 1 and 4), it is 30 m written communication, 2001). As noted above, base of the Mojado contains 10–15 m of locally thick and directly overlies Proterozoic grano- La Juana represents marine incursion into the derived, limestone-clast conglomerate and cross- diorite (Fig. 4; Hedlund, 1980; Amato et al.,

2496 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

meters Burro Mountains ~80 km Cookes Range ~350 km 100 San Ysidro, NM

Ash B Offshore deposits

Acanthoceras amphibolum Dakota Formation (middle Cenomanian) Paguate Tongue Shoreface deposits

Clay Mesa Tongue 50 Offshore deposits meters Mancos Formation 94.3 ± 1.1 Ma Dakota Formation (12BQ35) Cubero Tongue 100 Calycoceras canitaurinum Conlinoceras tarrantense zone (early middle Mancos Formation 94.3 ± 1.0 Ma (lowermost upper Cenomanian; Cobban et al., 1989) Cenomanian; 95.7 Ma; Cobban et al., 2006) (19BQ44) Oak Canyon Mbr meters sp. 94.9 ± 0.9 Ma (19BQ45, tuff) 50 (7) Estuarine deposits 96.3 ± 1.3 Ma (12BQ47, tuff) (1) 97.7 ± 1.0 Ma (12BQ37, DZ) Ash A; 97.6 ± 1.3 Ma (12BQ41, tuff) (2) amphibolum (middle Ceno- Encinal Canyon Mbr Shoreface deposits Datum is Ash A or manian; 94.96 ± 0.50 Ma; Cobban et al., 2006) equivalent 99.4 ± 1.1 Ma (12BQ38, DZ) Tarrantoceras sellardsi 0 Unconformity Unconformity Jackpile Sandstone 101 ± 2 Ma (11BQ01, DZ) Lower shoreface and Estuarine deposits offshore deposits vf f m c 50

Beartooth Member Drakeoceras lasswitzi zone tone Shoreface deposits (mid-upper to upper Albian; Cobban, 1987) Shale

Pebble-boulder conglomerate Silts

Mojado Formation Sandstone 0 Nonconformity Proterozoic granodiorite sp. (mid-upper Albian; Conglomerat e vf f m c Section offset Cobban, 1987) Pebbly sandstone with Sandstone Tidal estuary deposits channel bodies Section offset Shale Carbonaceous or Conglomerate silty shale Siltstone Braided-fluvial deposits

Sandstone Heterolithic ss and siltstone Ash-rich shale Conglomerat e with lenticular beds 0 Unconformity Ash bed, with U-Pb Permian strata Heterolithic sandstone and shale zircon age and sample number vf f m c Sandstone with Ammonite occurrence cross-beds

Shale (7) Mean paleocurrent (number of measurements) Siltstone 97.7 ± 1.0 Ma sample and age in Ma for Sandstone (19BQ37, DZ) tuff (tuff), MDA in Ma for sandstone (DZ) Conglomerate

Figure 4. Fence diagram of Mojado, Dakota, and Mancos Formations, southwestern and central New Mexico. Burro Mountains section is a composite of Beartooth Member measured at Saddlerock Canyon and Mancos Formation measured at Clyde Canyon (locations are shown in Figure 1; Machin, 2013). Ammonites identified during this study are illustrated inFigure 5 . Ammonites from published sources are indi- cated in figure. DZ—detrital zircon; MDA—maximum depositional age. Age data are shown inTables 2 and 3.

2011). The basal part of the Beartooth consists of ber contains no fossils, correlation with some the Little Hatchet Mountains (Lucas and Law- pebble-to-boulder conglomerate 1.5 m thick that part of the Mojado Formation was inferred (e.g., ton, 2005), 60 m in the Cookes Range (Lucas fines upward to fine- to medium-grained sand- Mack, 1987a, 1987b; Lucas and Estep, 1998b; et al., 1988), and from 115 m to 160 m thick in stone. Conglomerate clasts include angular to Lawton, 2004) and now is confirmed by our new complete sections of the Burro Mountains and at rounded quartzarenite and chert pebbles, as well geochronologic data. exposures on the Gila River in westernmost New as distinctive granodiorite boulders as much as Mexico (Finnell, 1987; Lucas and Estep, 1998b). 30 cm in diameter. Most of the overlying sand- Tectonostratigraphic Assemblage 5 The texture of the shale, which contains abun- stone is laminated to apparently structureless dant thin beds of volcanic ash, varies among with uncommon burrows and shale intervals. A Mancos Formation localities. In the Little Hatchet Mountains, the pebble lag at the top of the unit is directly over- The Mancos Formation, or , shale is dark gray and subfissile. A 15 cm tuff lain by shale with thin interbeds of siltstone and overlies the Mojado Formation in the Little collected in the lower part of the shale (11BQ18) very fine-grained sandstone of the Mancos For- Hatchet Mountains and Cookes Range (Fig. 4; is a coarse-grained, vitric crystal ash-fall tuff mation, as indicated by our composite section Lucas et al., 1988; Lucas and Lawton, 2005; with fresh crystals 0.15–2.0 mm in diameter that from the Burro Mountains (Fig. 4). Cobban et al., 1989) and the Beartooth Member consist of 30%–35% biotite, 20% plagioclase, The Beartooth Member was deposited in a in the Burro Mountains, where the shale was pre- 15% inclusion-free quartz, and 5%–15% sani- nearshore marine environment characterized viously termed Colorado Formation (Hedlund, dine with Carlsbad twins. Vitric volcanic grains by mixed tidal and wave energy. Heterolithic 1980; Finnell, 1987). The basal contact at all consisting of microcrystalline, low-birefringent sandstone-siltstone indicates intermittent tidal localities is a sharp, transgressive unconformity crystal domains make up 20%–25% of the tuff currents (e.g., Nio and Yang, 1991). Hummocky overlain by a thin bed of poorly sorted sandstone and are interpreted as tuffaceous lapilli. The low- cross-stratification low in the section indicates a with pebbles and shell fragments. An incom- er six meters of the Mancos in the Cookes Range shoreface setting. Although the Beartooth Mem- plete section of the Mancos is ∼100 m thick in consist of laterally continuous, tan-weathering

Geological Society of America Bulletin, v. 132, no. 11/12 2497

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

platy beds, 5–15 cm thick, of dark brown shale with abundant white tuffaceous grains. Bedding surfaces contain abundant pits up to 0.25 mm in diameter where lapilli have weathered out. Am- monites and inoceramid bivalves are common on bedding surfaces (Cobban et al., 1989). The shale contains at least 11 conspicuous waxy or- ange bentonite beds 0.5–16 cm thick in the lower 6 m of the formation. The uppermost sampled bentonite (12BQ35) consists of altered tuffa- ceous or vitric material lacking obvious grain boundaries and containing ∼3% quartz and feld- spar crystals to ∼0.03 mm in diameter and ∼10% A B opaque oxide grains to ∼0.04 mm in diameter. Above this bentonite, the shale changes in char- F acter to laminated, blocky weathering brown to light gray shale and silty shale. Uncommon thin sandstone beds with current-ripple cross lamina- tion are present in the blocky shale. At Clyde Canyon in the Burro Mountains, the Mancos consists of gray to black fissile shale with thin interbeds of very fine-grained sandstone and silt- stone (Fig. 4). An ash-fall tuff beneath a thin bed C E (∼30 cm) of gray, sandy limestone was sampled (12BQ47) 14 m above the base of the forma- tion. We also collected a sample of Mancos H sandstone (12BQ21) for detrital zircon analysis at Clyde Canyon. Fossils indicate that the Mojado–Mancos con- G tact is diachronous in southwestern New Mexico. Middle Cenomanian ammonites assigned to the D Acanthoceras amphibolum zone were recovered during this study 4–5 m above the base of the Figure 5. Ammonoids from the Acanthoceras amphibolum zone of the Mancos Formation in Mancos at Clyde Canyon (Figs. 4 and 5) and are the Burro Mountains. (A–B) Acanthoceras amphibolum Morrow [New Mexico Museum of ∼70–90 m above the base in the Little Hatchet Natural History (NMMNH) locality 8840]; (C–D) Desmoceras sp. (NMMNH locality 8839). Mountains (Lucas and Lawton, 2005). Upper (E–H) Tarrantoceras sellardsi (Adkins) (NMMNH locality 8839). Cenomanian ammonites occur in the lower part of the Mancos Formation in the Cookes Range (Fig. 4; Cobban et al., 1989). Therefore, the 4 and 5 (Fig. 2), the formation correlates, in part, of the lower middle Cenomanian Conlinoceras lower part of the Mancos Formation is possibly with the upper part of the Mojado Formation and tarrantense ammonite zone, which is interpret- as old as early Cenomanian in the Little Hatchet overlies progressively younger Paleozoic and ed as 95.73 ± 0.49 Ma on the basis of calibrated Mountains, middle Cenomanian in the Burro Mesozoic strata northward across western New biostratigraphy and 40Ar/39 tuff ages elsewhere Mountains, and late Cenomanian in the Cookes Mexico and eastern Arizona (Dickinson et al., (Cobban, 1977; Cobban et al., 2006), and a re- Range. The Mancos Formation was deposited in 1989). At San Ysidro, near the southeastern gionally persistent ash-fall tuff, termed “ash A” an offshore setting below fairweather wave base. flank of the San Juan basin, the Encinal Can- (Owen and Head, 2001), that is present in the yon Member of the Dakota Formation, which Oak Canyon Member. We collected a sample of Dakota Formation consists of discontinuous fluvial sandstone medium-grained sandstone (12BQ38) from the The relationship of the Dakota Formation to bodies (Owen et al., 2007), discordantly over- Encinal Canyon Member 1 m above the uncon- the Mojado Formation is important in under- lies strata assigned to the Jackpile Sandstone, formity with the Jackpile Sandstone Member in standing the tectonic link between the Bisbee a fluvial sandstone widespread in the eastern the Hagan basin (Fig. 1) and a sample of ash A basin and Cordilleran foreland basin during San Juan basin and interpreted alternatively as (12BQ41) in the middle part of the Oak Canyon the middle Cretaceous; therefore, an important part of the Upper Jurassic Member at San Ysidro, New Mexico. Dakota locality is included in this study. In and or the Lower Cretaceous Burro Canyon Forma- adjacent to the San Juan basin of northwestern tion (Aubrey, 1998; Owen and Owen, 2003; SANDSTONE PETROLOGY New Mexico, the Dakota Formation interfingers Owen et al., 2007). The basal contact of the complexly with the Mancos Formation; together, Encinal Canyon Member represents a regional Sandstone composition of Lower Cretaceous the units represent broadly transgressive deposits unconformity on Lower Cretaceous and Juras- siliciclastic strata varies greatly within each of fluvial, estuarine, marginal marine, and off- sic strata (Aubrey, 1986; Anderson and Lucas, assemblage across the Bisbee basin (Fig. 6). shore environments (Fig. 4; Owen and Owen, 1996; Owen et al., 2007). The overlying Oak ­Sandstone of the Rancho La Colgada and 2003; Owen et al., 2005, 2007). Spanning TSAs Canyon and Cubero members contain fossils Morita Formations includes quartzo-lithic,

2498 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

Quartzose Morita Arizona Upper part of Mojado Figure 6. Sandstone compo- Volcanic-lithic Morita Arizona Qm Qm 11BQ01 sitional diagrams for Bisbee Morita Sonora x Cintura Sonora CI 9/4LA5 12BQ38 Rancho La Colgada Sonora Group sandstone samples. Solid 11BQ17 DZ sample Hell-to-Finish New Mexico 10LM11 12BQ37 black symbols indicate data 9/4LA1 DZ Sample of this study (Table 1). (A) Va- Quartzose petrofacies 9/4LA1 langinian–Aptian field of Klute (1987) 11BQ19 Volcanic petrofacies TC RO (TSA 2). Symbols for published field of Klute (1987) x data: White circles—González- Hell-to-Finish Fm. HF1 RO (Mack, 1987a) CB CB León (1994); gray squares Lower 100 m of DA 05CLC12 and diamonds—Peryam et al. x x Hell-to-Finish Fm. x (Mack, 1987a) (2012). Fields for Hell-to-Finish TA x xxx x x Formation from Mack (1987a). x x x xx 11BQ17 BU x MA 11BC1 UA MA (B) Albian sandstones (TSA F Lt F Lt 4). All samples from this study A Valanginian-Aptian Sandstones B Upper Albian Sandstones are upper Mojado equivalent strata (Table 1). Symbols for published data: x—González-León (1994); gray circles and compositional field—Mack (1987a). Bold lines delineate general provenance categories of Garzanti (2016): CB—continental block; MA—magmatic arc; RO—recycled orogen. Fine lines delineate provenance fields of Dickinson (1985): BU—basement uplift; TC—transitional continental; and CI—continental interior, indicating progressively more mature compositions of the continental block category; DA—dissected arc; TA—transitional arc; and UA—undissected arc, indicating progres- sively more lithic compositions of the magmatic arc category.

­feldspatho-quartzose, and litho-quartzose aren- framework; discontinuous twins in some grains Most sandstone of the Mojado, Beartooth, and ite (classification of Garzanti, 2016), with lithic suggest alteration of another grain type, prob- Dakota Formations is more quartzose than older sandstone dominated by volcanic lithic grain ably potassium , to albite (e.g., Walker, sandstone, but some sandstone intervals con- types (­ Table 1; Fig. 6A; González-León, 1994; 1984). Relative abundance of lithic fragments tain volcanic lithic grains. Quartzose samples Peryam et al., 2012). Conglomerate channel lags tends to increase upsection. One of our detrital are dominated by monocrystalline quartz and associated with quartzolithic sandstone in Sono- zircon samples (11BC1) plots near the volcanic contain subordinate polycrystalline quartz and ra locally contain as much as 75% sedimentary petrofacies field of Klute (1987). It contains a di- chert (Fig. 6B; Mack, 1987a). Potassium feld- clasts, including quartzite, chert, and limestone verse suite of volcanic lithic fragments, including spar is more abundant than plagioclase, which is pebbles, some containing crinoid fossils (Peryam microlitic (Lvm), felsitic (Lvf), and uncommon typically absent (Table 1), indicating that TSA et al., 2012). Quartzo-lithic sandstones fall within lathwork (Lvl) volcanic grains. The Arizona vol- 4 was not extensively albitized. An uncommon the compositional field of the quartzose petrofa- canic lithic sample and many Sonora Morita sam- compositional variant consists of quartzo-lithic cies of Klute (1987) that is defined in southeast- ples plot in the magmatic arc field of Garzanti sandstone with abundant potassium feldspar ern Arizona (Fig. 6A). The quartzose sandstones, (2016). In contrast, the Hell-to-Finish Formation and volcanic lithic fragments (Fig. 6B; Mack, with rounded to subrounded grains, contain is quartzo-lithic, with Qm and chert dominating 1987a). Some Qm grains are clear and inclu- abundant, inclusion-rich monocrystalline quartz the lithic fraction, and plots in continental block sion-free, and some are euhedral with pyrami- (Qm). Plagioclase makes up 10%–15% of the and recycled orogen provenance fields. dal terminations, although a few well-rounded

TABLE 1. RECALCULATED MODAL POINT-COUNT DATA FOR BISBEE BASIN SANDSTONES, U.S. AND MEXICO Sample Formation QtFL% QmFLt% LmLvLs% QmPK% Qt F L Qm F Lt Lm Lv Ls Qm P K Upper Albian Sandstones, New Mexico (this paper) 11BQ17 Mojado Formation Sarten Member 23 24 53 19 24 57 0 100 0 44 1 55 11BQ19 Mojado Formation Sarten Member 62 16 22 57 16 27 0 95 5 78 3 19 11BQ01 Beartooth Formation 99 1 0 99 1 1 NA NA NA 99 0 1 12BQ37 Mojado Formation Sarten Member 83 9 8 78 9 13 NA NA NA 90 0 10 12BQ38 Dakota Fm. Encinal Canyon Member 98 2 0 97 2 1 NA NA NA 98 0 2 Aptian Sandstones, Huachuca Mountains, Arizona (Gilbert, 2012) 10LM11 (DZ) Morita Formation Quartzose PF 85 11 3 82 11 5 NA NA NA 88 12 0 10LM12 Morita Formation Quartzose PF 86 11 3 83 11 4 NA NA NA 88 12 0 11BC1 (DZ) Morita Formation Volcanic PF 13 32 55 11 32 57 0 100 0 26 74 0 Aptian Sandstone, Little Hatchet Mountains, New Mexico (Gilbert, 2012) HF1 Hell-to-Finish Formation 81 1 18 46 1 53 0 0 18 98 2 0 Aptian Sandstones, Sonora (Peryam et al., 2012) 05SJQ3 La Colgada Formation 66 14 20 61 14 25 0 100 0 82 18 0 9/4LA5 (DZ) La Colgada Formation 89 2 9 87 2 11 NA NA NA 98 1 1 04SJQ1 (DZ) Morita Formation 61 11 28 56 11 33 0 99 1 84 16 0 9/4LA1 (DZ) Morita Formation 84 5 11 77 5 18 NA NA NA 94 4 2 05CLC12 (DZ) Morita Formation 34 16 50 30 16 54 0 100 0 66 34 0 Mean 69 11 20 63 11 26 0 85 3 81 13 6 SD 28 9 20 28 9 22 0 37 7 22 20 15 Notes: Qt, total quartz = Qm, monocrystalline quartz + Qp, polycrystalline quartz; rare chert is included Qp grain category. F, total feldspar = K, alkali feldspar + P, plagioclase. Lt = L, microcrystalline lithic fragments + Qp. DZ—detrital zircon; Lm—metamorphic lithic fragments; Ls—sedimentary lithic fragments; Lv—volcanic lithic fragments. NA—Not calculated for samples with QtFL%L<20%.

Geological Society of America Bulletin, v. 132, no. 11/12 2499

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

spherical grains are present. Potassium feldspar, types observed in conglomerate beds. La Juana zircon age groups in the detrital sample set. Ac- mostly sanidine, makes up 24% of the frame- sandstone also contains 10%–15% feldspar, both cordingly, observed age groups are assigned to work of the volcanic-lithic sandstone. Volcanic plagioclase and alkali feldspar, and uncommon eight time intervals defined by previous authors lithic fragments, including Lvf, Lvm, and vitric schistose white mica-quartz grains. The com- (e.g., Dickinson and Gehrels, 2009a; Gehrels volcanic grains (Lvv), compose more than half bination of pebbles and grain types suggests a et al., 2011) and augmented by additional cita- of the framework. The Lvv grains are low-bire- primary source in Paleozoic sedimentary rocks tions in the text: (1) grains (ca. 3000– fringence microcrystalline aggregates, some of and a subordinate basement source. Beds of vol- 2500 Ma) of the Wyoming craton; (2) Paleo- which contain devitrified glass shards. We inter- canic lithic sandstone are interbedded apparently proterozoic grains (ca. 2300–1800 Ma) of the pret these volcanic lithic fragments as neovolca- randomly with quartzose sandstone beds in the Wopmay and Trans-Hudson orogens; (3) late nic grains generated by contemporary volcanism Cintura of southeastern Arizona (Klute, 1987). Paleoproterozoic and Mesoproterozoic grains (e.g., Critelli and Ingersoll, 1995). (ca. 1780–1630 Ma) of the Yavapai and Mazat- The uppermost part of the Mojado Formation U-Pb GEOCHRONOLOGY zal basement provinces; (4) Mesoproterozoic in the Cookes Range and the Beartooth Member grains (ca. 1480–1350 Ma) of the Laurentian are quartzose sandstones, but variable proportions Detrital U-Pb ages of zircon grains indicate that -rhyolite province (Anderson, 1989); (5) of grain-sized domains consist of finely crystal- age populations changed from early to late Early late Mesoproterozoic and earliest Neoprotero- line mixtures of quartz and white mica or hematite Cretaceous time in our samples. Young zircon zoic grains (ca. 1250–980 Ma) of the Grenville and . These suggest the former presence of grains from some Lower Cretaceous sandstone orogen of eastern and southern Laurentia; (6) unstable lithic grains and feldspar, indicating that samples proved useful in calculating maximum (ca. 730–510 Ma) Pan-Afri- the quartzose composition is diagenetic. depositional ages (MDAs) that in many cases in- can and Suwannee terrane grains; (7) Paleo- Cintura sandstone of Sonora is consistently dicate near-depositional ages. U-Pb zircon ages of zoic grains (ca. 490–270 Ma) attributed to the more lithic than Mojado sandstone, with vary- tuff beds and fine-grained bentonites in the Man- Taconic, Acadian, and Alleghanian orogens of ing feldspar content (Fig. 6B; González-León, cos Formation help correlate the unit from south- the Appalachian region (Thomas, 2011); and 1994). Volcanic lithic grains make up 91% of the western to northwestern New Mexico. A summa- (8) Permian–Mesozoic grains (ca. 275–92 Ma) total grain population, followed by sedimentary ry of MDAs and tuff ages for our new analyses derived from Cordilleran arc sources (Barth and lithic grains (7%) and metamorphic lithic grains and published data is shown in Tables 2 and 3. Wooden, 2006; Riggs et al., 2013, 2016). (2%). Sandstone composition of the overlying Sample locations from this study are in Table 4. La Juana Formation differs markedly from that Valanginian-Aptian Strata of the Cintura Formation. Examination of two La Detrital Zircon U-Pb Ages TSA 2 contains zircon grains with ages Juana thin sections revealed feldspatho-quartzo- that can be attributed to local Proterozoic lithic composition with abundant chert and de- Most Laurentian basement age provinces basement, recycling of Paleozoic-Mesozoic trital carbonate grains consistent with pebble and Cordilleran arc sources are represented by sedimentary rocks, and arc sources (Fig. 7;

TABLE 2. SUMMARY OF U-Pb DETRITAL ZIRCON AGES FROM SANDSTONE Sample Area Unit Biostratigrapic age n MDA ± 2s MSWD YSG ± 1s Reference no. (Ma) (n for MDA) for MDA (Ma) 9/4LA5 Arroyo La Angostura Rancho La Colgada ND 88 250 ± 3 (8) 1. 3 136 ± 3 Peryam et al. (2012) 9/4LA1 Arroyo La Angostura Lower Morita ND 93 138 ± 2 (6) 1. 0 135 ± 3 Peryam et al. (2012) 05SJQ1 Arroyo San Joaquin Rancho La Colgada ND 83 134 ± 2 (6) 0.6 133 ± 1 Peryam et al. (2012) 04SJQ5 Arroyo San Joaquin Upper Morita late Aptian 93 115 ± 1 (57) 1. 8 105 ± 5 Peryam et al. (2012) 11BQ01 Burro Mountains Beartooth Member Mojado pre-middle 86 101 ± 2 (8) 0.3 99.1 ± 2.2 This study Cenomanian 12BQ21 Burro Mountains Mancos ND 86 94.2 ± 2.0 (3) 0.5 92.1 ± 2.5 This study KBCR Chiricahua Mtns. Cintura ND 97 219 ± 3 (7) 1. 2 108 ± 2 Dickinson et al. (2009) 12BQ37 Cookes Range Upper Mojado late Albian–late 75 97.7 ± 1.0 (13) 0.5 96.3 ± 1. 0 This study Cenomanian 10LM11 Huachuca Mtns. Lower Morita ND 82 198 ± 5 (8) 1. 5 133 ± 2 This study 11BC1 Huachuca Mtns. Morita ND 96 123 ± 1 (47) 0.5 120 ± 4 This study 9-4-3 La Cumarosa Basal Morita ND 92 139 ± 3 (8) 1. 8 132 ± 7 Peryam et al. (2012) 05CLC12 La Cumarosa Upper Morita ND 96 113 ± 3 (3) 0.9 111 ± 1 Peryam et al. (2012) HF1 Little Hatchet Mtns. Hell-to-Finish ND 65 1014 ± 30 (5) 1. 9 182 ± 10 This study 11BQ17 Little Hatchet Mtns. Upper Mojado ND 45 102 ± 1 (10) 0.4 99.4 ± 8.2 This study 11BQ19 Little Hatchet Mtns. Upper Mojado ND 72 99.1 ± 1.0 (29) 0.6 95.1 ± 2.0 This study 5.12.15.10 Little Hatchet Mtns. Lower Mojado ND 84 163 ± 12 (3) 1. 4 158 ± 4 Clinkscales and Lawton (2015) 12BQ38 Hagen Basin Encinal Canyon Mbr Dakota ND 86 99.4 ± 1.1 (13) 0.8 97.0 ± 2.6 This study Note: All ages are by laser ablation–multi-collector–inductively coupled plasma mass spectrometry. MSWD—mean square of weighted deviates; MDA—maximum depositional age; Mtns.—Mountains; n—number of individual analyses; ND—no data; YSG—youngest single zircon grain age (Dickinson and Gehrels, 2009b).

TABLE 3. SUMMARY OF U-Pb IGNEOUS ZIRCON AGES FROM TUFF LAYERS Sample no. Area Unit Biostratigraphic age n Age ± 2s MSWD Reference (Ma) 11BQ18 Little Hatchet Mountains Mancos early–middle Cenomanian 25 97.2 ± 1. 6 0.7 This study 12BQ35 Cookes Range Mancos late Cenomanian 18 94.3 ± 1. 1 0.3 This study 12BQ41 San Ysidro Oak Canyon Member Dakota early-middle Cenomanian 16 97.6 ± 1. 3 0.3 This study 12BQ47 Burro Mountains Mancos middle Cenomanian 16 96.3 ± 1. 3 1. 0 This study 19BQ44 Cookes Range Mancos late Cenomanian 39 94.3 ± 1. 0 3.0 This study 19BQ45 Cookes Range Mancos late Cenomanian 35 94.9 ± 0.9 1. 7 This study Note: All ages are by laser ablation–multi-collector–inductively coupled plasma mass spectrometry (LA–MC–ICPMS), except samples 19-BQ44 and 45, which are by LA–ICPMS; MSWD—mean square of weighted deviates; n—number of individual analyses.

2500 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

TABLE 4. LOCATIONS FOR SAMPLES FROM THIS STUDY Table DR11). Samples from the lower parts Sample no. Area Latitude Longitude of the sections in Sonora and Arizona and the (°N) (°W) Hell-to-Finish sample of New Mexico contain 10LM11 Huachuca Mountains, Arizona 31.378254 110.363993 11BC1 Huachuca Mountains, Arizona 31.375117 110.363641 prominent age peaks in the Yavapai-Mazatzal, HF1 Little Hatchet Mountains, New Mexico 31.809687 108.440031 granite-rhyolite, and Grenville time intervals 11BQ01 Burro Mountains, New Mexico 32.742698 108.506862 11BQ17 Little Hatchet Mountains, New Mexico 31.871333 108.471974 (Fig. 7). The Hell-to-Finish includes only a few 11BQ18 Little Hatchet Mountains, New Mexico 31.863364 108.465345 percent late Paleozoic and Mesozoic grains, 11BQ19 Little Hatchet Mountains, New Mexico 31.861318 108.462664 12BQ21 Burro Mountains, New Mexico 32.759379 108.691189 but other units contain ages that span the late 12BQ35 Cookes Range, New Mexico 32.486724 107.713957 Paleozoic to earliest Cretaceous, ranging from 12BQ37 Cookes Range, New Mexico 32.486537 107.712164 12BQ38 Hagan Basin, New Mexico 35.297162 106.320873 ca. 278–133 Ma. Volcanic lithic samples from 12BQ41 San Ysidro, New Mexico 35.501755 106.834950 12BQ47 Burro Mountains, New Mexico 32.757302 108.686179 the upper part of the Morita in Sonora and 19BQ44 Cookes Range, New Mexico 32.486499 107.713950 Arizona are distinguished by uncommon pre- 19BQ45 Cookes Range, New Mexico 32.486127 107.713618 Jurassic grains, with the bulk of the grain ages forming a bimodal distribution of Jurassic and 1647 upper Hell-to-Finish Early Cretaceous ages in the age range of ca. (HF-1; Aptian) 191–113 Ma. The volcanic lithic sample from n = 65 Arizona contains 46 grains (of 97 grains ana- 1235 1458 lyzed) with overlapping grain ages in the range

SW NM of ca. 127–120 Ma. Six of seven La Colgada and Morita samples 122 Morita volcanic petrofacies with abundant volcanic lithic fragments yielded (11BC1; late Aptian) calculated Early Cretaceous MDAs ranging from n = 96 148 139 Ma to 113 Ma (Valanginian–latest Aptian; Table 2). Samples at the top of the Morita section in Sonora, collected directly beneath fossilifer- 198 231 1428 Morita quartzose petrofacies 1639 ous upper Aptian beds of the Mural Limestone 1738 (10LM11; Valanginian?) (Fig. 2; Lawton et al., 2004), yielded late Aptian

SE Arizona n = 82 MDAs of 115 ± 1 Ma and 113 ± 3 Ma (Table 2). Previously published age data for the lower part of the section in Sonora (Peryam et al., 2012) 115 upper Morita (late Aptian) yielded MDAs ranging from ca. 250 Ma to (N= 2; n = 189) ca.134 Ma. Although no biostratigraphic data 165 exist to confirm that the MDAs there represent depositional ages, their consistent Valanginian age (139 ± 3 to 134 ± 2 Ma; Table 2), appar- 154 lower Morita (pre-Aptian) ent consistency between sections, and inferred (N = 2; n = 184 plus one grain at 3366 Ma) proximity to a magmatic arc that began prior to 278 1184 those ages and continued beyond them imply 1050 1441 1688 that they represent approximate depositional Sonora ages. We infer that the Rancho La Colgada– 257 134 La Colgada (Valanginian) Morita succession was deposited between ca. (N = 2; n = 170 plus one grain at 3196 Ma) 136 Ma and 115 Ma (Valanginian–late Aptian) and that a hiatus of 10–15 m.y. exists between 1050 1431 1668 youngest preserved strata of the Cucurpe Forma- tion (ca. 149 Ma; Mauel et al., 2011) and basal Lower Cretaceous strata. Physical stratigraphic evidence discussed above likewise indicates the presence of an unconformity. Calculated MDAs for the Morita Formation of the Hua- chuca basin are 198 ± 5 Ma for the quartzose sample (10LM11) at the base of the section and opmay / Wyoming Suwanee – W Composite 123 ± 1 Ma for the volcanic lithic sandstone Pan-Africa n rans-Hudson vapai-Mazatzal Cordilleran ar c T (11BC1) that is 300 m higher in the section Grenville orogen Granite-Rhyolite Ya Appalachian orogen

0 500 1000 1500 2000 2500 3000 1GSA Data Repository item 2020158, LA–MC– ICPMS and LA–ICPMS U-Pb zircon analyses of Age, Ma detrital and igneous samples (three tables in Excel format), is available at http://www.geosociety.org/ Figure 7. Detrital zircon probability density plots of TSA 2 samples. N—number of samples; datarepository/2020 or by request to editing@ n—number of analyses. geosociety.org.

Geological Society of America Bulletin, v. 132, no. 11/12 2501

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

99 231 Mancos, Burro Mtns (11BQ21) cal correlation with the La Colgada and Morita (Cenomanian) sections and its compositional similarity to the 165 n = 86 volcanic-lithic Morita in Sonora. The MDA of 1416 1688 the Hell-to-Finish sample, 1014 ± 30 Ma, indi-

SW NM cates that young zircons were not supplied to the 97 Encinal Canyon Mbr Dakota (12BQ38) Bootheel basin during Aptian time. (Cenomanian) 212 n = 86 Albian-Cenomanian Strata Detrital zircon age populations of TSAs 4 and 1156 1023 NW NM 5, represented by the Mojado, Cintura, Dakota, and Mancos Formations, fall into two categories 216 upper Cintura, Chiricahua Mtns (KBCR) (Fig. 8; Table DR3a): (1) Samples with diverse (Albian) and abundant Proterozoic ages, subordinate n = 96 plus 1 grain at 3056 Ma 442 1054 Paleozoic ages, and intermediate numbers of 1449 Mesozoic ages; and (2) samples with fewer old Arizona grains and a high proportion of Mesozoic ages. Sarten Mbr Mojado, Cookes Range (12BQ37) 216 Category 1 samples average 69% Archean and (late Albian) Proterozoic grains, with appreciable proportions 1195 1412 1685 n = 75 1016 of Yavapai-Mazatzal (10%), granite-rhyolite (10%), and Grenville (30%) ages, 11% Paleo- zoic grains, and 20% Mesozoic grains, whereas 115 category 2 samples average 39% Proterozoic 415 grains, 7% Paleozoic grains, and 53% Mesozoic 1174 1474 221 1068 grains. Archean grains are rare to uncommon in samples of both categories. Category 1 samples include the lower Mojado in the Little Hatchet 103 Rattlesnake Ridge Mbr Mojado, LHM (11BQ19) Mountains (5.12.15.10), the Beartooth Member (late Albian) in the Burro Mountains (11BQ01), the upper part n = 71 plus 1 grain at 3320 Ma of the Mojado in the Cookes Range (12BQ37), 1678 1750 and the upper Cintura Formation in the Chirica- hua Mountains (KBCR). Category 2 samples 106 168 Ratttlesnake Ridge Mbr Mojado, LHM (11BQ17) include the upper Mojado samples in the Little

SW New Mexico (late Albian) Hatchet Mountains (11BQ17, 11BQ19), the 1738 n = 45 Dakota Formation in northwestern New Mexico (12BQ38), and the Mancos Formation sandstone in the Burro Mountains (12BQ21). 207 Fryingpan Spring Mbr Mojado, LHM (5.12.15.10) Samples of the upper part of the Mojado For- 1054 (Albian) 1158 mation yielded young grains that provide MDAs 160 1647 n = 84 of late Albian–early Cenomanian age (Table 2). The calculated MDA in the Cookes Range (12BQ37) is 98 ± 1 Ma. MDAs of the samples in the Little Hatchet Mountains, both of which contain abundant neovolcanic grains (Table 1), are 102 ± 1 Ma and 99 ± 1 Ma. The Beartooth Member (sample 11BQ01) yielded an MDA of 101 ± 2 Ma, and the Mancos sandstone from

opmay / high in the formation exposed at Clyde Canyon Wyoming Suwanee– W Composite Pan-Africa n yielded an MDA of 94 ± 2 Ma. The Encinal rans-Hudson vapai-Mazatzal Cordilleran ar c T Grenville orogen

Granite-Rhyolite Canyon Member of the Dakota Formation in Ya Appalachian orogen northwestern New Mexico yielded an MDA of 99 ± 1 Ma. The MDAs are consistent with pub- 0 500 1000 1500 2000 2500 3000 lished biostratigraphic ages from the same sec- tions and indicate correlation of the upper part of U-Pb Age, Ma the Mojado with the Dakota Formation.

Figure 8. Detrital zircon probability density plots of TSA 4 and 5 samples. U-Pb Ages of Volcanic Ash Beds

(Table 2; Fig. 3). The MDA from the base of MDAs from multiple samples, the MDA of the Six tuffs in the lower part of the Mancos the section is pre-depositional based on strati- upper sample likely approximates its inferred Formation and one from the Dakota Formation graphic relationships. Although unsupported by depositional age (early Aptian) based on physi- were analyzed with LA–ICPMS. U-Pb zircon

2502 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

Mean: 97.2 ± 1.6 Ma 0.08 MSWD = 0.7 11BQ-18

110 0.07 Pb 206

100 / 0.06 Pb Pb Age (Ma) 207 206 90 0.05

U/ 120 110 100 90 238 11BQ-18: Little Hatchet Mtns. Ash 0.04 80 50 54 58 62 66 70 74 78 238U/206Pb

106 0.09 Mean: 97.6 ± 1.3 Ma 12BQ-41 MSWD = 0.3 102 0.08 Figure 9. U-Pb igneous ages Pb 98 0.07 of tuffs in the Little Hatchet 206

/ Mountains, Burro Mountains,

Pb Age (Ma) and at San Ysidro in northwest- 94 Pb 0.06 ern New Mexico. MSWD— 206 207 mean square of weighted U/ 0.05 90 deviates. 108 104 100 96 92 88 238 12BQ-41: SanYsidro Ash 0.04 86 58 62 66 70 74 238U/206Pb 108 Mean: 96.3 ± 1.3 Ma 12BQ-47 104 MSWD = 1.0 0.07

100 Pb 0.06 206 96 / Pb Age (Ma) Pb

206 92 0.05 207

U/ 112 108 104 100 96 92 88 88 238 12BQ-47: Burro Mtns. Ash 0.04 84 56 60 64 68 72 76 238U/206Pb

ages range from 97.6 Ma to 94.3 Ma (Fig. 4; 10-cm-thick transgressive lag at the base of the GEOHISTORY ANALYSIS Tables 3, DR3b). The oldest tuff, ash A, in the Mancos, yielded an age of 94.9 ± 0.9 Ma (mean Oak Canyon Member of the Dakota Forma- square of weighted deviates [MSWD] = 1.7, Geohistory diagrams constructed using strati- tion (12BQ41) yielded an age of 97.6 ± 1.3 Ma n = 35). Only five grains were outside of this graphic data from southwestern New Mexico (Fig. 9). The coarse-grained tuff near the base population, two around 100 Ma, one at 312 Ma, and north-central Sonora reveal broadly similar of the Mancos in the Little Hatchet Mountains one at 425 Ma, and one Proterozoic grain at subsidence history through Late Jurassic–Late (11BQ18) yielded an age of 97.2 ± 1.6 Ma. The 1138 Ma. The next highest sample (19BQ44) Cretaceous time (Fig. 11). Each geohistory plot bentonite bed in the Burro Mountains yielded an yielded an age of 94.3 ± 1.0 Ma (MSWD = 3.0, (Fig. 11) has two curves, one for total subsid- age of 96.3 ± 1.3 Ma. n = 39). The high MSWD indicates that this ence of the decompacted section with water The lower part of the Mancos Formation, group of ages likely includes multiple popula- depth included that is termed the total subsid- which directly overlies the Mojado Formation in tions, but we were unable to determine whether ence curve, and another for subsidence with the the Cookes Range, has at least 11 orange benton- the older or younger zircons are more represen- isostatic effects of sediment and water load re- ite layers, which are interpreted as ash-fall tuffs, tative of the eruption age, and thus we report moved. The latter curve represents the tectonic over a stratigraphic interval of 6 m and ranging the average for all of the grains younger than subsidence caused by extrinsic mechanisms, in thickness from 1 cm to 15 cm. We dated three one clearly older grain at 290 Ma. The high- such as lithospheric thinning or flexural loading of these layers (Fig. 10). A sample of the lower- est sampled tuff (12BQ35) returned an age of (Heller et al., 1986; Angevine et al., 1990; Xie most layer (19BQ45), which directly overlies a 94.3 ± 1.1 Ma (MSWD = 0.30, n = 18). and Heller, 2009).

Geological Society of America Bulletin, v. 132, no. 11/12 2503

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

116 Mean: 94.3 ± 1.1 Ma 0.14 112 MSWD = 0.3 12BQ35 108 meters 0.12 104 Pb 100 0.10 206 /

Pb Age (Ma) 96

30 Pb 0.08 206 92 207 U/ 88 0.06

238 84 12BQ35 (Cookes Range tuff) 120 110 100 90 80 0.04 50 54 58 62 66 70 74 78 238U/206Pb 0.060 101 Mean: 94.3 ± 1.0 Ma 19BQ44 MSWD = 3.0 99 0.056

97 Pb 0.052 206

95 20 / Figure 10. U-Pb igneous ages Pb Age (Ma) 93 Pb 0.048 102 98 94 90 and stratigraphy of bentonites, 206 91 207 Cookes Range. MSWD—mean U/ 0.044 square of weighted deviates.

23 8 89 19BQ44 (Cookes Range tuff) 0.040 87 62 64 66 68 70 72 74 238U/206Pb 110 0.064 Mean: 94.9 ± 0.9 Ma 19BQ45 106 MSWD = 1.7 0.060

102 10 Pb 0.056 20 6 98 / 0.052 Pb Age (Ma) Pb 0.048 20 6 94 112 108 104 100 20 7 96 92 88 U/ 90 0.044 23 8 19BQ45 (Cookes Range tuff) 0.040 86 56 60 64 68 72 76 238U/206Pb 12BQ37: Cookes Range Sandstone DZ MDA 97.7 ± 1.4 Ma 0

We analyzed sections in the two areas that Mountains of southeastern Arizona (Lawton and lier, an unconformity with a hiatus of as much as maximize stratigraphic completeness and offer Olmstead, 1995; Olmstead and Young, 2000), 15 m.y. separates Jurassic and Lower Cretaceous the most extensive biostratigraphic and geo- and local corals, which resemble ­Oxfordian strata in Sonora, and a similar lapse in sedimen- chronologic data for calibrating stratigraphic forms from the subsurface of Arkansas (Lu- tation can be inferred from reduced subsidence age. The New Mexico analysis derives from cas et al., 2001); therefore, the inception of the in the New Mexico curve. Given uncertainties complete sections in the central and northern steep curve in New Mexico, where Jurassic that might result from lithostratigraphic cor- Little Hatchet Mountains (Table 5), spanning strata directly overlie Permian strata, is poorly relation and errors in unit thicknesses, the two from Oxfordian through middle Cenomanian controlled. Deposition may have begun later in subsidence histories are remarkably similar. The time, a time interval of ∼66 m.y., as compared the Late Jurassic, as suggested by a late Oxford- concave-upward, declining Jurassic–middle Al- to a 26 m.y. interval analyzed by Mack (1987a) ian– age (ca. 159–156 Ma) for bian curves generally resemble subsidence his- (Aptian-Albian; Cohen et al., 2013; updated). the base of the Jurassic section in the Chihuahua tories that result from synrift stretching to post- The Sonora analysis represents a composite of trough (e.g., Haenggi, 2002). This would shorten rift thermotectonic subsidence of extensional sections near Tuape, Sonora, which lies 15 km the time span represented by the New Mexico basins or passive margins, where the amount of south of Cucurpe (Fig. 1). Ages of stratigraphic curve stated above by ∼4 m.y. rapid initial tectonic subsidence and sediment intervals of both curves employ a combination of The Late Jurassic–earliest Cretaceous subsid- accumulation is determined by the amount of biostratigraphic data, MDAs from detrital zircon ence history of both localities shows initially crustal stretching (Steckler and Watts, 1978; data, and U-Pb ages of ash beds from published rapid but decreasing rates of tectonic subsid- Xie and Heller, 2009). In contrast, the rapid sources and this study. In the Little Hatchet ence during Late Jurassic through middle Aptian late Albian–Cenomanian subsidence resembles Mountains, the ages of Jurassic strata are based (Fig. 11A) or middle Albian (Fig. 11B) time, fol- flexural subsidence curves from foreland basins principally on correlation with a Kimmeridg- lowed by increased subsidence rates in late Al- (Xie and Heller, 2009). The inflection point be- ian–Tithonian Jurassic section in the Chiricahua bian time, recorded by TSA 4. As discussed ear- tween subsidence rates of TSA 3 and TSA 4 in

2504 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

Age, Ma Age, Ma 180160 140 120 100 80 60 180 160 140 120100 80 60 ...... Middle . Middle .

Aptian Albian ron. Aptian Albian ron. Paleogene thon thon xford. xford. Jurassic alan. Jurassic alan. Con. Con. Berr Berr Barr Cen. Barr Cen. Haut. Haut. Maas Maas V V Kimm Kimm Tu Tu O O Santon Santon Ti Ti Camp Camp Broken Hell-To- Mojado- La U-Bar Laramide Cucurpe Mural Mo- Laramide Jug Finish Mancos Colgada- LJ 0 0 Morita

1 1

2 2

3 3

4 4 Subsidence, km 123 4,5

5 5 4,5 Tectonic Subsidence 123 6 Total Subsidence 6

A New Mexico Age Category 7 7 Correlation Tuff/Volcanic U-Pb Tuff/Volcanic 40Ar/39Ar 8 MDA Biostratigraphy 9 B Sonora

Figure 11. Geohistory diagrams indicating Late Jurassic through Late Cretaceous subsidence history of southwestern New Mexico and northern Sonora. (A) Southwestern New Mexico, using stratigraphic section of Little Hatchet Mountains (Lucas and Lawton, 2000; Clink- scales and Lawton, 2015) with new ages for Mojado and Mancos Formations presented in this work. Tracked horizon is top Paleozoic. (B) North-central Sonora, employing stratigraphic section in vicinity of Cucurpe (Mauel et al., 2011; González-León et al., 2011, 2017). Tracked horizon is top Middle Jurassic. Data points unconnected by lines indicate prolonged hiatuses. MDA—maximum depositional age. For de- tails of ages and stratigraphic units, see Table 5. Numbers in plot refer to tectonostratigraphic assemblages.

Sonora is particularly well defined Fig. 11B( ), enance relationships in turn yield insight into corroborate published biostratigraphic data and whereas TSA 3 marks an intermediate increase a temporal transition from separate rift basins provide new ages for unfossiliferous continental in subsidence rate between TSA 2 and TSA 4 to a partitioned foreland integrated with the and high-energy marine deposits (Figs. 2–4 and in New Mexico (Fig. 11A). Possible subsidence Cordilleran foreland-basin system. We discuss Fig. 11). Our geochronological ages are consis- mechanisms are elaborated upon further in the correlations among the three discrete basins of tent with their stratigraphic positions in indi- Discussion. the greater Bisbee basin, interpret sandstone vidual sections (Figs. 3 and 4), and despite the provenance and sediment dispersal systems inherent problems of comparing biostratigraphic DISCUSSION during basin evolution, and propose a link be- ages calibrated with 40Ar/39Ar ages (e.g., Scott, tween basin history and tectonics of the conti- 2014), our age errors overlap with interpreted U-Pb ages derived from detrital sandstone nental margin. fossil zone boundary ages, as discussed below. samples and tuffs across the U.S.–Mexico Ages of Upper Jurassic marine strata, gener- ­border region provide improved correlation of Regional Stratigraphic Correlation ally of deepwater origin, in the Altar-Cucurpe Upper Jurassic–middle Cretaceous stratal suc- and Bootheel basins, are well known from U-Pb cessions. Similar subsidence histories derived New and published U-Pb geochronological and biostratigraphic data (Olmstead and Young, from correlative strata with contrasting prov- data augmented by new fossil identifications 2000; Lucas et al., 2001; Villaseñor et al., 2005;

Geological Society of America Bulletin, v. 132, no. 11/12 2505

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

TABLE 5. DATA FOR SUBSIDENCE CURVES, SOUTHWEST NEW MEXICO AND NORTHERN SONORA Unit Age Data type Thickness Location References (Ma) (m) New Mexico, Little Hatchet Mountains Overburden 33 Ignimbrite Ar-Ar 1500 Playas, New Mexico, USA McIntosh and Bryan (2000) Upper Skunk Ranch 70 ± 1 Tuff U-Pb 75 SW of Playas Peak Jennings et al. (2013) Middle Skunk Ranch 70 ± 1 Tuff U-Pb 183 SW of Playas Peak Jennings et al. (2013) Lower Skunk Ranch 71 ± 1 Tuff U-Pb 480 SW of Playas Peak Jennings et al. (2013) Upper Ringbone 73 ± 1 MDA 393 S of Playas Peak Clinkscales and Lawton (2015) Middle Kr 73 ± 1 Tuff U-Pb 1095 SE of Playas Peak Clinkscales and Lawton (2015) Lower Kr 74 Correlation 120 SE of Playas Peak Clinkscales and Lawton (2015) Mancos Shale 97.2 ± 1. 6 Tuff U-Pb 100 Howells Well syncline Lucas and Lawton (2005); This study Mojado Formation (top) 99.1 ± 1. 7 MDA 1245 Howells Well syncline Galemore (1986); This study U-Bar Formation, Still Ridge Member (top) 105 Biostratigraphic 441 S of Hachita Peak Lucas and Lawton (2000) U-Bar Formation, Old Hachita Member 108 Biostratigraphic 83 S of Hachita Peak Lucas and Lawton (2000) U-Bar Formation, Carbonate Hill Member 112 Biostratigraphic 474 S of Hachita Peak Lucas and Lawton (2000) Hell-to-Finish Formation, Winkler Ranch Mbr. 119 Correlation 325 S of Hachita Peak Lucas and Lawton (2000) Hell-to-Finish Formation, Stone Cabin Member 140 Corrlelation 200 S of Hachita Peak Lucas and Lawton (2000) Broken Jug Formation, Basalt Member 146 Correlation 230 E flank Hachita Peak Lucas and Lawton (2000) Broken Jug Formation, Upper Cgl Member 149 Biostratigraphic 200 E flank Hachita Peak Lucas et al. (2001) Broken Jug Formation, Fine-Grained Member 153 Correlation 380 E flank Hachita Peak Lucas and Lawton (2000) Broken Jug Formation, Lower Cgl Member 155 Correlation 200 E flank Hachita Peak Lucas and Lawton (2000) Broken Jug Formation, Dolostone Member 161 Correlation 200 E flank Hachita Peak Lucas and Lawton (2000) Sonora Baucarit () 21 Biostratigraphic 2100 NE Sonora González-León et al. (2011) Tarahumara Formation 57 Volcanic U-Pb 1250 NE Sonora González-León et al. (2017) Cabullona Group 73 MDA 1000 Cabullona Basin González-León et al. (2011) Cocóspera Formation 77 Correlation 2000 NE Sonora González-León et al. (2017) La Juana Formation 100 Biostratigraphic 300 Cañada La Juana Lawton et al. (2004); Mauel et al. (2011) Cintura Formation 101 Correlation ∼2000 Vicinity of Tuape Lawton et al. (2004); Mauel et al. (2011) Mural Limestone, Mesa Quemada Member 104 (top) Correlation 300 Cerro La Ceja Lawton et al. (2004) Mural Limestone, Cerro La Espina Member 105 (top) Biostratigraphic 30 Cerro La Ceja Lawton et al. (2004) Mural Limestone, Cerro La Puerta Member 106 (top) Biostratigraphic 125 Cerro La Puerta Lawton et al. (2004) Mural Limestone, Los Coyotes Member 111 (top) Biostratigraphic 125 Cañada Los Coyotes Lawton et al. (2004) Mural Limestone, Tuape Shale Member 112 (top) Biostratigraphic 100 Cañada Los Coyotes Lawton et al. (2004) Mural Limestone, Cerro La Ceja Member 114 (top) Biostratigraphic 175 Cañada Los Coyotes Lawton et al. (2004) Morita Formation (top = base Mural) 115 Tuff U-Pb 950 Arroyo La Angostura Peryam et al. (2012 Morita Formation (base) ca. 133 MDA Bear Gulch, Arizona, USA Gilbert (2012); Peryam et al. (2012) Rancho La Colgada Formation 134 MDA 150 Rancho La Colgada Peryam et al. (2012) Cucurpe Formation (top) 150 Tuff U-Pb 750 Sierra Cucurpe Mauel et al. (2011) Cucurpe Formation (middle) 152 Biostratigraphic Sierra Cucurpe Villaseñor et al. (2005) Cucurpe Formation (base) 154 MDA 750 Sierra Cucurpe Mauel et al. (2011) Note: Cgl—conglomerate; MDA—maximum depositional age.

Mauel et al., 2011). The age range of the alluvial graphic position above the Glance Conglomerate by late Albian time. Lower parts of thick Mo- Glance Conglomerate of the Huachuca basin is and correlation with the Sonoran stratigraphic jado sections in the Bootheel basin did not yield less well controlled, as it is bracketed between sections. Early Cretaceous sediment accumula- near-depositional zircon ages, but upper Moja- 172 ± 2 Ma, the age of an ignimbrite at the base tion similarly followed a prolonged depositional do samples from the Little Hatchet Mountains, of the conglomerate in the Huachuca Mountains hiatus or slow sedimentation indicated by an in- Cookes Range, and and Burro Mountains have (Gilbert, 2012), and ca. 136 ± 3 Ma, the inferred terval of paleosols and granule conglomerate. In MDAs ranging from 102 ± 1 Ma to 98 ± 1 Ma, age of the basal Morita Formation in Bear Can- New Mexico, where a concordant contact with which indicate correlation with the basal part yon. The bracketing data permit a Late Jurassic the subjacent Broken Jug Formation demon- (Encinal Canyon Member) of the Dakota Forma- age and are consistent with a post-170 Ma Ju- strates a post-Jurassic age for the Hell-to-Finish tion at San Ysidro (MDA = 99 ± 1 Ma; Table 2; rassic age inferred for the Glance Conglomer- ­Formation but does not reveal presence or ab- Figs. 4 and 11). ate from exposures in the Santa Rita Mountains, sence of a hiatus, the age of earliest Cretaceous U-Pb tuff ages and biostratigraphic data from 45 km northwest of the Huachuca Mountains, on deposition in the Bootheel basin remains poorly TSAs 4 and 5 indicate that the upper part of the basis of regional stratigraphic relations (Bas- known (Lucas and Lawton, 2000); nevertheless, the Mojado, the Mancos of southwestern New sett and Busby, 2005). superjacent Aptian carbonate strata indicate a Mexico, and lower part of the Dakota Forma- Although uncertainty remains regarding true general correlation with the TSA 2 units of the tion in NW New Mexico constitute a diachron- depositional ages of the lower part of the Creta- other basins (Fig. 2). ous but genetically related and widespread com- ceous section, discussed above, MDAs suggest Cintura and Mojado strata of TSA 4 were de- plex of upper Albian–Cenomanian estuarine and that deposition began ca. 136 Ma in the Altar- posited during late Albian–earliest Cenomanian shallow-marine deposits (Tables 2 and 3; Figs. 4 Cucurpe basin, where initial Early Cretaceous time. Fossiliferous middle Albian beds of subja- and 11). U-Pb ages of 97.6 ± 1.3 Ma from ash deposition of the Rancho La Colgada Formation cent Mural and U-Bar strata in the region support A at San Ysidro, 97.2 ± 1.6 Ma from a tuff in followed a hiatus of ∼10–15 m.y. and took place the older age of TSA 4 at localities where fossils the Mancos in the Little Hatchet Mountains, and in shallower marine conditions. Local absence are not yet known from the lower part of TSA 96.3 ± 1.3 Ma from a tuff in the Burro Moun- of the Rancho La Colgada at the base of the sec- 4 (Lucas and Estep, 1998b; Lucas and Lawton, tains indicate equivalence of marine shelf depos- tion suggests onlap of Cretaceous strata onto 2000; Lawton et al., 2004; González-León et al., its there with estuarine and shoreface deposits in Jurassic rift-basin sections tilted during the hia- 2008). Ammonites in the lower part of the Mo- the Cookes Range (Figs. 4 and 11). tus. Initial Cretaceous marine incursion into the jado Formation of the Cookes Range (Fig. 4) and The age of ash A at San Ysidro is consistent formerly continental Huachuca basin, recorded bivalves in the La Juana Formation of Sonora with its position 20 m below the mid-Cenoma- by shoreface deposits in the Morita Formation, demonstrate that middle Cretaceous marine in- nian Conlinoceras tarrantense ammonite zone took place at about the same time based on strati- cursion had taken place throughout the region (Cobban, 1977), which has a 96.2 Ma lower

2506 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

South North Central Little Northern Little Peloncillo Cookes Burro San Ysidro Hatchet Hatchet Mountains Range Mountains (Colorado Plateau) Mountains Mountains ~350 km Mancos Formation Dakota/Mancos Fm. Mancos Formation Datum is Ash A or equivalent (~97.6 Ma) 97.7 ± 1.0 Ma (DZ) 94.9 ± 0.9 Ma (tuff) 96.3 ± 1.3 Ma (tuff) 97.6 ± 1.3 Ma (Ash A) 97.2 ± 1.6 Ma (tuff) 101 ± 2 Ma (DZ) 99.4 ± 1.1 Ma fault 99.1 ± 1.0 Ma (DZ) (DZ) Rattlesnake Mojado Formation Permian Lower Ridge Mbr Marine strata ? strata Beartooth Mbr. Cretaceous Fluvial strata U-Bar on Proterozoic Formation Sarten granodiorite Mbr Hell-to-Finish Formation Fryingpan Mojado Formation Marine Spring Mbr Permian strata

U-Bar Formation km fault 025 Fluvial strata Sandstone and siltstone, mostly marine Conglomerate Limestone with rudistid bioherms Hell-to- 0 Finish Basalt Limestone Formation Mudstone with tuff beds Sandy dolostone meters Calcareous siltstone Paleozoic strata Marine shale Broken Jug Lower Cretaceous normal Formation 1000 Normal fault with Mesozoic offset, fault with reverse Laramide (Late Jurassic) inferred reactivation Conglomeratic debrites Unconformity

Permian strata

Figure 12. Southwest–northeast stratigraphic cross section of Bisbee Group, Dakota Formation, and Mancos Formation from Little Hatchet Mountains in southwestern New Mexico to San Ysidro in northwestern New Mexico. Line of section is shown in Figure 1. Datum at ca. 96.7 Ma is interpreted from a combination of biostratigraphic and geochronologic data. See text for discussion. Modified from Machin (2013).

age boundary (Scott, 2014). U-Pb and biostrati- Sandstone Provenance and Sediment Salt Wash fluvial fan of the Morrison Formation graphic ages of uppermost Albian–lowermost Routing (Owen et al., 2015), derived from the Mogol- Cenomanian marine deposits of the region thus lon highlands or the southern end of the Cordil- demonstrate not only previously posited north- Changing provenance relationships in Ju- leran fold and thrust belt (Dickinson and Geh- ward stratal onlap onto the Burro uplift (e.g., rassic–middle Cretaceous strata accompanied rels, 2008a), and the basal part of the Cucurpe Mack, 1987a), but also widespread marine de- basin evolution and provide evidence for local Formation in the Altar-Cucurpe basin, derived position from southwesternmost to northwest- ­sediment sources that served as topographic bar- from subjacent Middle Jurassic eolian sandstone ern New Mexico (Fig. 12). The short time inter- riers between the basins. The rapidly subsiding (Mauel et al., 2011). Similar grain-age distribu- val (94.9 ± 0.9 Ma to 94.3 ± 1.0 Ma) suggested Late Jurassic Altar-Cucurpe basin received sedi- tions in deformed Tithonian turbidites of the by U-Pb ages of tuff beds in the lower Mancos ment rich in pyroclastic material from active vol- forearc region, including the Peñasquitos and deposition in the Cookes Range agrees with canic centers in a westward-migrating magmatic Eugenia Formations (Kimbrough et al., 2014), late Cenomanian ammonites interbedded with arc (Fig. 13A; Mauel et al., 2011), but volcanic also suggest recycling of the southwestern edge the bentonites (Fig. 4; Cobban et al., 1989). detritus did not reach the Huachuca basin from of the former erg, likely uplifted on the southern Rapid deposition during the transgressive part ca. 172 Ma (Gilbert, 2012) until it appeared in end of the Central Nevada thrust belt as early of the Mancos Formation may have resulted the Morita Formation at ca. 123 Ma; the Juras- as 160 Ma (Giallorenzo et al., 2018) and routed from voluminous volcaniclastic input to the sic Bootheel basin never received arc-derived through a low-standing arc (e.g., Busby-Spe- region during the 98–92 Ma La Posta mag- detritus. Sediment derived from recycled Lower ra, 1988). matic event of southern California and the Baja and Middle Jurassic eolianite of the southwest- Valanginian–Aptian sandstone of TSA 2 California Peninsula (Kimbrough et al., 2001; ern United States, which contains grain ages that varies in lithic content and detrital zircon ages Ortega-Rivera, 2003), but we cannot presently include distinctive Grenville-aged Proterozoic among basins, demonstrating persistence of sep- discriminate between potential contemporary (1180–980 Ma), Neoproterozoic-early Paleozo- arate sediment sources and of the basins them- volcanic sources represented by the Peninsular ic (680–510 Ma), and Paleozoic (490–350 Ma) selves. Although pebbles of quartzite, fossilife- and the Sierra Nevada batholiths (e.g., Ducea, ages (Dickinson and Gehrels, 2009a), is pres- reous limestone, and chert in some conglomerate 2001; DeCelles et al., 2009). Our Mancos tuff ent in many Upper Jurassic units of the region lags indicate local Paleozoic sources, Sonoran ages all fall within the range of 40Ar/39Ar tuff and suggests export of eolianite-derived sedi- rocks contain an upsection increase in volcanic- ages (97.5–91 Ma) reported from the Buda and ment in several directions during Late Jurassic lithic content (Fig. 6A). Zircon ages in Rancho Eagle Ford Formations in west Texas and lie in time (Fig. 13A). Important stratigraphic units La Colgada and lower Morita samples from So- the age range of Oceanic 2 (El- containing the distinctive combination of grain nora and the basal Morita sample in Arizona sug- drett et al., 2014). ages include upper Kimmeridgian strata of the gest that most Phanerozoic zircon grains were

Geological Society of America Bulletin, v. 132, no. 11/12 2507

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

110° W 105° W B 115° W 105° W A SWf Low-flux segment WPt Cordilleran foreland

basin ce a e tr Sevier orogenic belt ulg LV LV b Ha re Morrison CM fo SY o paleodrainage Formation f magmatic ar Mogollon - Bu A 35° N Approx. A 35° N ri BC ft paleodrainage

rro c Inferred rift AZ catchment shoulder P should P trench extended to arc er by ~123 Ma T T EP EP Pacific LA LA Santiago P Ocean Huac h W Cah uca Ch Pb estw Ct ? eak arc Marine deposystem Al ard -migratingtar- 30° N 30° N Cucu rp Ap Continental deposystem e ABf Ct Pacific Basement high Ocean bac k-arc bas Alis 0 300 km m in(?) 0 300 km Opening of it agmatic ar os Arperos basin Ba ar TSA 1: Late Oxfordian– TSA 2: Valanginian– c Tithonian middle Aptian Eb ? ~152–145 Ma c ~136–120 Ma

lt 115° W 110° W 105° W 115° W 110° W 105° W C D

Cordilleran foreland e trac basin ce tra lge Sevier orogenic belt Sevier orogenic be lge LV u LV x. forebu reb o . fo SY SY Appr A 35° N California A 35° N Remnant Approx Dakota valley Cordilleran foreland Western Interior Mojave incision a basin (Mancos) seaway Desert topgraphy rc P P ? Pacific Ash delivered to Ocean EP incipient foreland EP by westerly LA LA ? Pacific Inferred shoreline Ocean 30° N 30° N La Juana migration ? Northern Mexico 0 300 km fold transgression Shelfal deposystem fold belt (Del Rio) seaway

Early emplacement phase of ? Main emplacement phase of Peninsular Ranges batholith ? belt Peninsular Ranges batholith ("La Posta" event) 0 300 km Guerrero accretion Guerrero TSA 4: Middle Albian– TSA 5: Middle–late suture early Cenomanian Cenomanian ~105–99 Ma ~98–94 Ma

Figure 13. Schematic paleogeography of Late Jurassic to Middle Cretaceous rift to foreland basin transition and inferred sediment dis- persal, indicated by blue arrows. Communities: A—Albuquerque; Ba—Batopilas; En—Ensenada; EP—El Paso; LA—Los Angeles; LV— Las Vegas; P—Phoenix; SC—Silver City; T—Tucson. Geologic localities of study: Ar—Arizpe; CC—Clyde Canyon (Burro Mountains); CM—Chiricahua Mountains; CR—Cookes Range; Cu—Cucurpe; Ha—Hagan basin; HM—Huachuca Mountains; LHM—Little Hatchet Mountains; PM—Peloncillo Mountains; SA—southern San Andres Mountains; SR—Saddlerock Canyon (Burro Mountains); SY—San Ysidro; Tu—Tuape. U.S. palinspastic reconstruction after Saleeby (2003); Baja California Peninsula restored after preferred reconstruction of Fletcher et al. (2007); Arizona and New Mexico south of Colorado Plateau and Sonora restored for ∼50% mid- extension (Gans, 1997). (A) Late Jurassic assemblage of backarc rift basins formed following westward migration of Middle–Late Jurassic arc across Mexico,­

2508 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

derived principally from Permian, Triassic, and ment derived from the Cretaceous arc likewise such as the Wheeler Pass thrust of the Central Jurassic rocks, including late Paleozoic–Tri- increased over time (Figs. 6A and 7). An Aptian Nevada thrust belt (Fig. 13C; Giallorenzo et al., assic plutons and widespread Early–Middle influx of arc detritus did not take place in the 2018). age peaks in the TS4 strata Jurassic plutons and ignimbrites in Sonora Bootheel basin, where the Hell-to-Finish Forma- are likewise explained by recycling of Triassic and the Mojave Desert region of southwestern tion contains no volcanic-lithic fragments and units such as the Upper Triassic Chinle Forma- Arizona and southern California (Figs. 7 and few zircon grains with Mesozoic ages. Rather, tion (Chinle Group of Lucas, 2004), which con- 13B; Riggs et al., 1993, 2013, 2016; Barth and feldspathic sediment containing Proterozoic zir- tains abundant Late Triassic zircon grains (Dick- Wooden, 2006; Arvizu et al., 2009; Mauel et al., con grains was derived from basement exposed inson and Gehrels, 2008; Riggs et al., 2013). The 2011; González-León et al., 2011). Local eolian as early as Aptian time on the rift shoulder north Triassic units were also uplifted in thrust sheets quartzarenite interbedded with Middle Jurassic of the basin (Fig. 7, sample HF-1; Mack et al., of southern Nevada. Alternatively, Late Triassic ignimbrites in Sonora (Leggett, 2009; Mauel 1986; Mack, 1987a). Contrasting sandstone plutons in the Mojave Desert region (Barth and et al., 2011) likely yielded Yavapai-Mazatzal, composition and lack of arc-derived sediment Wooden, 2006; Riggs et al., 2013) could have Granite-Rhyolite, Grenville, and Paleozoic ages confirm that the Bootheel basin in southwestern provided grains of the appropriate ages. observed in the older samples (Fig. 7). Zircon New Mexico was separate from other sub-basins Recycling of Jurassic eolianite strata that lay grains in the range of ca. 149–133 Ma and an- of the Bisbee basin. on the rift shoulder of the Bisbee basin has been desite clasts with ages ranging from ca. 143 Ma Volcanic sources continued to dominate sedi- suggested as a source for sediment of the Cintura to 125 Ma (Peryam et al., 2012) in the lower part ment of TSA 4 in Sonora and likely in southeast- and Mojado Formations (Dickinson et al., 2009), of the section in Sonora indicate that an Early ern Arizona, but composition and zircon age dis- but our analysis suggests alternative sources far- Cretaceous continental margin arc was a princi- tributions in the Bootheel basin indicate a shift ther to the west, in the southern end of the Se- pal source for the Altar-Cucurpe basin. to sources in older sedimentary strata (Fig. 8; vier orogenic belt (Fig. 13C). By Late Jurassic Volcanic-lithic composition and interpreted Dickinson et al., 2009; Clinkscales and Lawton, time, the eolian sandstones were already eroded near-depositional Aptian zircon MDAs higher 2015). Quartzose composition, well-rounded well back from the rift shoulder, as is demon- in the Morita section of southern Arizona dem- quartz grains, and paleoflow directions in the strated by progressive overlap of the Upper Ju- onstrate that volcanogenic sediment flooded Mojado Formation were interpreted to indicate rassic Morrison Formation onto Triassic strata the Huachuca basin by mid-Aptian time (ca. dominantly sedimentary source rocks, probably in central New Mexico (New Mexico Bureau 123 Ma; Fig. 7). As in Sonora, abundant grains part of a fold-thrust orogen that lay west of the of Geology and Mineral Resources, 2003). By in the range of ca. 127–120 Ma imply that sedi- basin (Mack, 1987a). Eastward sediment disper- middle Cretaceous time the Dakota Formation ment was derived from the Santiago Peak arc, sal, despite apparently contradictory zircon age had onlapped Permian strata along the Mogollon which occupied the continental margin begin- modes that suggest ultimate eastern Laurentian Rim in Arizona, the site of the former rift shoul- ning ca. 138 Ma (Wetmore et al., 2003). Abrupt origin, require that the enormous volume of com- der (Reynolds, 1988; Dickinson et al., 1989). influx of arc-derived sediment with a wide range positionally mature detritus in the Cintura and Presence of zircon derived from local basement of Early Cretaceous ages into the basin suggests Mojado Formations of the Bootheel basin was in the locally arkosic Hell-to-Finish Formation extension of the fluvial catchment into the arc routed from the west, likely from Jurassic eoli- further indicates extensive unroofing of the rift (Fig. 13B). In Sonora, the proportion of sedi- anite strata in thrust sheets in southern Nevada, shoulder prior to deposition of TSA 4 strata. The

adapted from Mauel et al. (2011) and Fitz-Díaz et al. (2018 and references therein). Distribution of marine facies and Aldama Platform (Ap) in Chihuahua are adapted from Haenggi (2002). Note inferred connection of continental Huachuca basin to Chihuahua trough (Ct). Extensive Lower to Middle Jurassic eolianite of southwestern U.S. and northwestern Mexico was recycled to backarc settings, such as Salt Wash fluvial fan (SWf), and forearc settings including Peñasquitos basin (Pb) and Eugenia basin (Eb) (e.g., Kimbrough et al., 2014). South- ern edge of Bisbee basin is drawn near north edge of Caborca block (Fig. 1), but rift may have merged southward with inferred backarc basin recorded by interbedded Upper Jurassic marine hyaloclastite and volcanic rocks in the vicinity of Batopilas, Chihuahua (Ba; Lyons, 2016). See text for discussion and additional references. (B) Valanginian–early Aptian backarc rift basins and continental-margin Santiago Peak arc; evidence for a proto-Alisitos arc to the south lies in abundant Lower Cretaceous grains (ca. 150–128 Ma) derived from it in Upper Cretaceous sandstone of northeastern and central Mexico (e.g., Lawton et al., 2009; Juárez-Arriaga et al., 2019). Headward extension of southern Arizona drainage into arc is evidenced by the appearance of volcanic-lithic sandstones in the Morita section. Mogollon–Burro rift shoulder formed headwaters of north-flowing Cedar Mountain (CM) and Burro Canyon (BC) drainages. Forebulge location is after Currie (2002) and Dickinson (2018). Aptian marked maximum extent of fluvial deposition in Chihuahua trough (Las Vigas Formation; Haenggi, 2002). (C) Closure of Arperos basin, accretion of Guerrero terrane, superimposition of eastern zone of Peninsular Ranges arc on suture, and continued sediment delivery from arc to Cintura Formation of northern Sonora and southern Arizona. Cintura and Mojado Forma- tions in Bootheel basin received sediment from eolianite from the southern end of the Cordilleran fold and thrust belt. Subduction polarity of Arperos marginal basin is intepreted as west dipping due to absence of Albian arc on mainland Mexico and emplacement of Guerrero onto Mexico (e.g., Martini et al., 2013; Fitz-Díaz et al., 2018). Onlap of Mojado delta and related marine deposits onto rift shoulder began in late Albian time. Dashed dispersal arrows in southern New Mexico represent schematic incised valleys of lower Dakota Formation (Bauer, 1989) eroded on north flank of remnant rift shoulder. Shoreline position is early Cenomanian maximum progradation of Dakota Forma- tion (adapted from Robinson Roberts and Kirschbaum, 1995). Ash dispersal delivered neovolcanic sediment to upper Mojado sandstone of Bootheel basin in southwestern New Mexico. Developing or remnant topographic highs are required to explain compositional differ- ences between former rift basins. (D) Incursion of Mancos shelfal marine deposits onto rift shoulder and into former Bisbee basin. Arrows indicate directions of marine transgression; dashed line indicates maximum landward extent of shoreline ca. 94 Ma, adapted in U.S. from Robinson Roberts and Kirschbaum (1995), highly schematic in Mexico where record is largely eroded.

Geological Society of America Bulletin, v. 132, no. 11/12 2509

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

presence of Late Triassic grains in Cintura and Late Jurassic–Middle Cretaceous flexural subsidence following emplacement of a Mojado samples is not well explained by deriva- Subsidence History tectonic load (e.g., Mack, 1987a). Although we tion from the eolianites; instead, the early Meso- still cannot specify a particular load, pebbles of zoic grains were likely derived from basement Subsidence of the Late Jurassic–Middle Cre- limestone, dolostone, chert, and granite in La sources in the Mojave Desert or Triassic strata taceous sedimentary basins followed broadly Juana formation at the top of TSA 4 in Sonora in the Cordilleran thrust belt. similar historical trajectories in Sonora and suggest nearby late Albian uplift of the local Pa- Compositional and detrital zircon similarities southwestern New Mexico (Figs. 11A and leozoic section, probably on the Caborca block of our Dakota sample to the Mojado Formation 11B). As described above, both basins experi- (Figs. 1 and 13C). The shape of the decompacted (Fig. 8) can likewise be explained by deriva- enced rapid Late Jurassic subsidence rates that subsidence curves at both localities resembles tion from Triassic strata and Jurassic eolianite. waned during Berriasian–early Aptian time those that result from flexure of an elastic plate However, the location of our Dakota sample far and accelerated during the Albian; however, (Fig. 11; e.g., Xie and Heller, 2009), but the dis- north of the former rift shoulder makes it more neither of the subsidence curves follows trends tribution of subsidence across the resulting fore- difficult to determine actual sediment-dispersal diagnostic of standard basin histories, and this land is quite broad, extending >300 km from pathways, whether directly from the Cordilleran likely indicates superposition of simultaneous the Caborca block (Fig. 13C), and it does not thrust belt or more local sources. Recycling of tectonic factors (e.g., Xie and Heller, 2009). We conform well with predictions of a narrow basin Dakota sediment by erosion of previously depos- infer that rapid Late Jurassic subsidence in the developed on a recently broken plate (Fosdick ited foreland basin strata from a time-equivalent Altar-Cucurpe and Bootheel basins (Fig. 11) re- et al., 2014). We suggest that broadly distrib- forebulge that ran northeast, parallel with the sulted from crustal extension that accompanied uted subsidence across the older basin array and front of Sevier orogenic belt, as suggested by westward migration of the Late Jurassic arc and burial of the former rift shoulder resulted from Aubrey (1998), is consistent with paleocurrent subduction zone (Lawton and McMillan, 1999; dynamic subsidence after emplacement of the data and not precluded by the zircon data (Figs. 8 Dickinson and Lawton, 2001a; Fitz-Díaz et al., subducted Farallon slab beneath the U.S.–Mexi- and 13C). An alternative possibility—that zircon 2018). In this scenario, initial rapid subsidence co border region, as discussed further below. grains of the Dakota Formation were recycled, resulted from synrift crustal stretching, with in part, directly from subjacent units within the slightly greater tectonic subsidence in Sonora Sedimentary Basin Evolution and beveled Paleozoic–Triassic stratigraphic section than in New Mexico a likely result of greater Continental Margin Tectonics on the north flank of the former rift shoulder—is crustal thinning with increased proximity to the suggested by local variability of grain-age distri- arc. Initial crustal extension resulted in local Concomitant evolution of sedimentary basins butions and locally abundant Triassic grains in basins and uplifts typical of rifted settings (e.g., of the southwestern U.S.–Mexico border region Dakota samples from south-central New Mexico Gawthorpe and Leeder, 2000). and the continental margin arc in southernmost (Stopka, 2017). Several aspects of the Jurassic–Aptian tran- California and the Baja California Peninsula Ash plumes transported zircon grains di- sition suggest that subsidence was not a conse- suggests a common link with the history of the rectly to the Bootheel basin and northwestern quence of simple rifting: (1) Neither subsidence Mexican Guerrero composite volcanic terrane New Mexico to be deposited in the Mojado curve (Fig. 11) has a smooth transition from (Fig. 13). The extensional phase of basin de- and Dakota Formations in late Albian–early synrift stretching to post-rift thermal subsid- velopment was coeval with Late Jurassic–Early Cenomanian time (Fig. 13C). This conclusion ence; (2) evidence for a depositional hiatus ­exists Cretaceous retroarc shortening that formed the is derived from the following observations: (1) between Upper Jurassic and Lower Cretaceous Central Nevada and Sevier thrust belts north of Abundant neovolcanic grains are present in stra- stratigraphic sections in the Altar-Cucurpe and Las Vegas, Nevada (DeCelles, 2004; Yonkee and ta interbedded with otherwise quartzose sand- Huachuca basins and can be inferred in the Weil, 2015; Giallorenzo et al., 2018). As shorten- stone in the Mojado and Dakota Formations; (2) Bootheel basin; and (3) at some localities in ing to the north created the retroarc Cordilleran neovolcanic sandstones yield calculated MDAs Sonora, Upper Jurassic strata were folded prior foreland basin (Currie, 1998; DeCelles, 2004), consistent with published biostratigraphic age to deposition of superjacent Lower Cretaceous a migrating Jurassic magmatic arc sweeping assessments where biostratigraphy is available strata. These characteristics of the Late Juras- westward across northern and central Mexico (Table 2, Fig. 4); (3) sandstone MDAs and tuff sic–Early Cretaceous stratigraphic transition had just passed the site of the Late Jurassic Bis- ages young stratigraphically at all localities support inferences of Late Jurassic transcurrent bee rift basins (Fig. 13A; Mauel et al., 2011). (Fig. 12); and (4) younger U-Pb tuff ages in su- movement along basin-bounding faults (e.g., Recent models for the origin of Guerrero posit perjacent Mancos strata are likewise consistent Anderson and Nourse, 2005; Bassett and Busby, the opening of a marginal oceanic Arperos basin with ammonite biostratigraphy. Neither a devel- 2005; Busby et al., 2005; McKee et al., 2005, between mainland Mexico and the westward- oping forebulge nor a relict rift-flank source can and references in those papers); nevertheless, we migrating arc in latest Jurassic–Aptian time explain the presence of syndepositional zircon do not view such transcurrent deformation as re- (Fig. 13B) (Martini et al., 2014; Fitz-Díaz et al., ages in the Dakota sample. A mixture of euhe- sulting from a throughgoing transform fault such 2018). The opening phase of the Arperos basin dral quartz, sanidine, tuffaceous fragments, and as the Mojave–Sonora megashear, but rather as a thus corresponded in time with synrift extension young zircon grains suggests that the volcanic- widely distributed response to northwest–south- and decreasing Early Cretaceous subsidence lithic sandstones were reworked directly from east separation of North and South America dur- rates in at least two of the Bisbee sub-basins. intrabasinal ash beds. Following transgression ing opening of the and Gulf of Mexico Although the likelihood of regional transcurrent of low-energy shelfal conditions into the region, basins (e.g., Pindell and Kennan, 2009). Long- faulting that might have resulted from continen- the ash falls were preserved as tuffs and ben- distance separation of sediment sources and tal separation of North and South America (e.g., tonites of the Mancos Formation and younger sinks is not supported by the provenance data Pindell and Kennan, 2009) cannot be discounted, parts of the complexly intertongued Mancos and described above. the proximity of a westward migrating arc (e.g., Dakota Formations at San Ysidro, New Mexico The middle- to late-Albian increase in subsid- Fitz-Díaz et al., 2018) that provided sediment to (Figs. 4 and 13D). ence rate likely resulted in part from onset of Jurassic and Early Cretaceous backarc basins

2510 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

is here regarded as the primary basin-forming progressive suturing of Guerrero may explain three basins persisted as separate depocenters. mechanism. apparent inconsistencies in the timing of the Sandstone petrography and U-Pb detrital zir- The Arperos basin subsequently closed in post-rift increased subsidence (Fig. 11). We con ages indicate that sediment sources of the central and southern Mexico in Cenomanian infer that the well-defined, late Albian subsid- individual basins varied from south to north. A time (Martini et al., 2014), and limited paleo- ence event in Sonora resulted from collision continental margin arc provided abundant sedi- magnetic and geochronological data suggest no of Guerrero with North America (e.g., Martini ment to the Altar-Cucurpe basin of northern So- lateral translation of Guerrero relative to North et al., 2013; Palacios-García and Martini, 2014). nora from latest Jurassic through Aptian time, America between the opening and closing of Collision-induced shortening of basinal rocks whereas arc-derived sediment did not arrive in the marginal basin (Boschman et al., 2018). and North American platformal strata yielded the Huachuca basin of southern Arizona until ca. Ongoing magmatism from ca. 145–124 Ma in locally derived pebbles in La Juana formation 123 Ma and failed to arrive at all in the Bootheel the Mexican segment of the arc system, consist- of the Sonora section, which occupied the flex- basin of southwestern New Mexico and south- ing of the Alisitos arc of the Guerrero terrane ural foredeep (Figs. 11B and 13C). In contrast, a eastern Arizona. Beginning in the latter part of and the Santiago Peak arc in southern Califor- specific tectonic load has yet to be identified in a late Aptian–middle Albian period of carbonate nia and the northern Baja California Peninsula, southwestern New Mexico adjacent to the thick deposition, subsidence rates increased across the is recorded in detrital zircon analyses of Upper Mojado Formation; instead, that section is part region to form a foreland basin partitioned by Cretaceous sandstones of northern and central of a broad subsiding region that preserved the the depocenters that persisted on the sites of the Mexico (Lawton et al., 2009; Juárez-Arriaga former basins as separate depocenters with their older basins. Foreland basin culmination in late et al., 2019). Detritus in the Morita Formation own provenance and sediment routing character- Albian–earliest Cenomanian time was marked of Sonora records time-equivalent magmatism istics. Widespread sedimentation eventually bur- by rapid subsidence across the former rifted re- of the Santiago Peak arc in southern California ied the former rift margin in southwestern New gion and burial of the rift flank. (Fig. 7; e.g., Wetmore et al., 2002, 2003). The Mexico (Fig. 12) and accompanied resumed Sediment provenance demonstrates that the same time interval marked a magmatic decline sediment accommodation following a hiatus of broad foreland basin remained structurally par- in the California arc of the Sierra Nevada (Du- at least 20 m.y. in northwestern New Mexico titioned. Arc-derived detritus continued to reach cea, 2001; DeCelles et al., 2009) and a shift of (Fig. 2). The Albian–Cenomanian subsidence the proximal Sonoran part of the basin, which pluton emplacement to the western flank of the event was a long-wavelength phenomenon that likely constituted a foredeep adjacent to an arc- batholith (Nadin et al., 2016). The contrasting resembles modeled dynamic topography result- proximal thrust belt. At the same time, quartz- along-strike magmatic histories of Guerrero and ing from initial emplacement of a slab subducted ose detritus eroded from Jurassic eolianite strata the western United States mirrored the respec- into the asthenosphere (e.g., Gurnis, 1992). In- in the southern part of the Cordilleran fold and tive, contemporary along-strike transition from deed, inversion of the local basins to form po- thrust belt was delivered southeastward along Late Jurassic–Early Cretaceous backarc exten- tential tectonic loads by crustal shortening, as the strike of the partitioned foreland to south- sion to retroarc shortening. recorded by Bisbee Group clasts in synorogenic western New Mexico. Foreland basin subsidence Sedimentary basin development and prov- conglomerate, did not take place until Turonian in Sonora likely had a flexural component, but enance relationships of the U.S.–Mexico border time in northern Sonora (González-León et al., long-wavelength subsidence across the foreland region outlined above thus evolved together with 2011) and time in southwestern New region reaching as far into the continent as north- opening and closing phases of the marginal basin Mexico (Lawton, 2000; Clinkscales and Lawton, western New Mexico is interpreted as a dynamic model, and this provides a mechanism for ob- 2015). The later inversion-related deposition is topographic effect induced by introduction of the served patterns of basin subsidence and sediment indicated by the Laramide subsidence events in Farallon slab beneath northwestern Mexico and sources. Pre-Neogene restoration of the Lower Figure 11. We therefore interpret emplacement the southwestern United States. Basin evolution Cretaceous arc rocks of southern California of the Farallon slab beneath southwestern North from rift to foreland basin paralleled Early Cre- and the Baja California Peninsula, according America following closure of the Arperos basin taceous opening and closing of a back-arc mar- to estimates of extension in northern Sonora as the principal mechanism of regional late Al- ginal basin that extended the length of modern (Gans, 1997) and dextral offset across the Gulf bian to middle Cenomanian dynamic subsidence Mexico east of the Guerrero composite volca- of California (Fletcher et al., 2007), places the of the U.S.–Mexico border region. nic terrane. continental Santiago Peak arc ∼100 km from Our stratigraphic analysis demonstrates that the northwestern end of the Altar-Cucurpe ba- CONCLUSIONS strata historically included in the Bisbee Group sin. The Santiago Peak arc could have provided contain at least one significant region-wide un- sediment, including andesite pebbles, to the New U-Pb zircon ages on tuffs, combined conformity that separates Jurassic and Lower nearby extensional backarc basin even as the with subsidence analysis and consideration of Cretaceous strata. In addition, individual forma- Alisitos arc pivoted away from North America provenance history based on sandstone petrog- tions or combinations of formations that consti- at the Agua Blanca fault to become progressively raphy and U-Pb zircon detrital ages, elucidate tute five tectonostratigraphic assemblages de- more isolated from the continental rift basins as Late Jurassic–middle Cretaceous sedimentary fined here record deposition in the U.S.–Mexico the marginal basin opened. Continued Valangin- evolution of the Bisbee basin and its link to border region as it evolved from lithospheric ian–middle Aptian extension accompanied mar- magmatic arc systems on the continental mar- extension to crustal shortening. Changing ba- ginal basin development; this is indicated by the gin of the southwestern U.S.–Mexico border sin genesis from Jurassic through earliest Late subsidence curves from Sonora and New Mexico region. Three rapidly subsiding back-arc rift Cretaceous time suggests a need to reevaluate (Figs. 11 and 13B). basins, with subsidence possibly augmented the formal definition of the Bisbee Group and The marginal basin model predicts that clo- by a strike-slip component of faulting during its stratigraphic components. A possible nomen- sure of the Arperos basin began sometime be- separation of North and South America, formed clatural solution would be to raise the Bisbee tween mid-Aptian and mid-Albian time. Dia- in Late Jurassic time. In Valanginian–middle Group to supergroup rank, a “formal assemblage chronous closure of the marginal basin and Aptian time, subsidence rates decreased as the of related or superposed groups, or of groups

Geological Society of America Bulletin, v. 132, no. 11/12 2511

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

and formations. Such units have proved useful Anderson, T.H., 2015, Jurassic (170–150 Ma) basins: The Bilodeau, W.L., 1986, The Mesozoic Mogollon High- tracks of a continental-scale fault, the Mexico–Alas- lands, Arizona; an Early Cretaceous rift shoulder: in regional and provincial syntheses” (North ka megashear, from the Gulf of Mexico to Alaska, The Journal of Geology, v. 94, p. 724–735, https://doi​ American Commission on Stratigraphic No- in Anderson, T.H., Didenko, A.N., Johnson, C.L., .org/10.1086/629077. menclature, 2005, p. 1570). Jurassic and Lower Khanchuk, A.I., and MacDonald, J.H., Jr., eds., Late Bilodeau, W.L., and Lindberg, F.A., 1983, Early Cretaceous Jurassic Margin of —A Record of Faulting tectonics and sedimentation in southern Arizona, south- Cretaceous strata, separated by a readily identifi- Accommodating Plate Rotation: Geological Society western New Mexico, and northern Sonora, Mexico, able regional unconformity, could be assigned to of America Special Paper 513, p. 107–188, https://doi​ in Reynolds, M. W., and Dolly, E. D., eds., Mesozoic different groups within the Bisbee Supergroup. .org/10.1130/2015.2513(03). Paleogeography of West-central United States: , Anderson, T.H., and Nourse, J.A., 2005, Pull-apart basins at Colorado, Society of Economic Paleontologists and releasing bends of the sinistral Late Jurassic Mojave- Mineralogists, Rocky Mountain Section, p. 173–188. ACKNOWLEDGMENTS Sonora fault system, in Anderson, T.H., Nourse, J.A., Bilodeau, W.L., Kluth, C.F., and Vedder, L.K., 1987, Re- McKee, J.W., and Steiner, M.B., eds., The Mojave- gional stratigraphic, sedimentologic and tectonic rela- George Gehrels and the staff at the Arizona Laser- Sonora Megashear Hypothesis: Development, As- tionships of the Glance Conglomerate in southeastern Chron Center, supported by EAR-1649254, helped sessment, and Alternatives: Geological Society of Arizona, in Dickinson, W.R., and Klute, M.A., eds., acquire zircon data. Lawton thanks Chris Clinkscales America Special Paper 393, p. 97–122, https://doi​ Mesozoic Rocks of Southern Arizona and Adjacent for discussions of Bisbee Group stratigraphy and .org/10.1130/2005.2393(1103). Areas: Tucson, Arizona, Arizona Geological Society structure. This study originated as part of the Master’s Angevine, C.L., Heller, P.L., and Paola, C., 1990, Quantita- Digest 18, p. 229–256. thesis work of Sarah E.K. Machin at New Mexico tive sedimentary basin modeling: American Association Boschman, L.M., Molina Garza, R.S., Langereis, C.G., of Petroleum Geologists Continuing Education Course and van Hinsbergen, D.J.J., 2018, Paleomagnetic State University; she was supported in part by grants Notes 32, 133 p. constraints on the kinematic relationship between the from the Geological Society of America and the New Archibald, L.E., 1987, Stratigraphy and sedimentology of the Guerrero terrane (Mexico) and North America since Mexico Geological Society. The manuscript was im- Bisbee Group in the Whetstone Mountains, southeast- Early Cretaceous time: Geological Society of America proved by insightful, thorough reviews by Carl Jacob- ern Arizona, in Dickinson, W.R., and Klute, M.A., eds., Bulletin, v. 130, p. 1131–1142, https://doi​.org/10.1130/ son and an anonymous reviewer. Mesozoic Rocks of Southern Arizona and Adjacent B31916.1. Areas: Tucson, Arizona, Arizona Geological Society Busby, C.J., Bassett, K.N., Steiner, M.B., and Riggs, N.R., REFERENCES CITED Digest 18, p. 273–282. 2005, Climatic and tectonic controls on Jurassic intra- Armstrong, R.L., 1968, Sevier orogenic belt in Nevada arc basins related to northward drift of North America, Alsleben, H., Wetmore, P.H., Gehrels, G.E., and Paterson, and Utah: Geological Society of America Bulletin, in Anderson, T.H., Nourse, J.A., McKee, J.W., and S.R., 2012, Detrital zircon ages in Palaeozoic and Me- v. 79, p. 429–458, https://doi​.org/10.1130/0016- Steiner, M.B., eds., The Mojave-Sonora Megashear sozoic basement assemblages of the Peninsular Ranges 7606(1968)79[429:SOBINA]2.0.CO;2. Hypothesis: Development, Assessment, and Alterna- batholith, Baja California, Mexico: Constraints for Arvizu, H.E., Iriondo, A., Izaguirre, A., Chávez-Cabello, G., tives: Geological Society of America Special Paper depositional ages and provenance: International Geol- Kamenov, G.D., Solís-Pichardo, G., Foster, D.A., and 393, p. 359–376, https://doi​.org/10.1130/0-8137-2393- ogy Review, v. 54, p. 93–110, https://doi​.org/10.1080/​ Lozano-Santa Cruz, R., 2009, Rocas graníticas pérmi- 0.359. 00206814.2010.509158. cas en la Sierra Pinta, NW de Sonora, México: Mag- Busby, C., Adams, B.F., Mattinson, J.M., and Deoreo, S., Amato, J.M., and Mack, G.H., 2012, Detrital zircon geo- matismo de subducción asociado al inicio del margen 2006, View of an intact oceanic arc, from surficial to chronology from the Bliss continental activo del SW de Norteamérica: Revista mesozonal levels: Cretaceous Alisitos arc, Baja Cali- Sandstone, New Mexico: Evidence for contrasting Mexicana de Ciencias Geológicas, v. 26, p. 709–728. fornia: Journal of Volcanology and Geothermal Re- Grenville-age and Cambrian sources on opposite sides Aubrey, W.M., 1986, The nature of the Dakota-Morrison search, v. 149, no. 1-2, p. 1–46, https://doi.org/10.1016/​ ​ of the Transcontinental Arch: Geological Society of boundary, southeastern San Juan basin, in Turner- j.jvolgeores.2005.06.009. America Bulletin, v. 124, p. 1826–1840, https://doi​ Peterson, C.E., Santos, E.S., and Fishman, N.S., eds., Busby-Spera, C.J., 1988, Speculative tectonic model for .org/10.1130/B30657.1. A Basin Analysis Case Study: Morrison Formation, the early Mesozoic arc of the southwest Cordilleran Amato, J.M., Boullion, A.O., Serna, A.M., Sanders, A.E., Grants Uranium Region, New Mexico: American As- United States: Geology, v. 16, p. 1121–1125, https:// Farmer, G.L., Gehrels, G.E., and Wooden, J.L., 2008, sociation of Petroleum Geologists Studies in Geology doi​.org/10.1130/0091-7613(1988)016<1121:STMFTE Evolution of the Mazatzal Province and the timing of 22, p. 93–104. >2.3.CO;2. the Mazatzal orogeny; insights from U-Pb geochronol- Aubrey, W.M., 1998, A newly discovered, widespread fluvial Centeno-García, E., Guerrero-Suastegui, M., and Talavera- ogy and geochemistry of igneous and metasedimentary facies and unconformity marking the Upper Jurassic/ Mendoza, O., 2008, The Guerrero composite terrane rocks in southern New Mexico: Geological Society Lower Cretaceous boundary, Colorado Plateau: Modern of western Mexico: Collision and subsequent rifting in of America Bulletin, v. 120, p. 328–346, https://doi​ Geology, v. 22, p. 209–233. a supra-subduction zone, in Draut, A., Clift, P.D., and .org/10.1130/B26200.1. Barth, A.P., and Wooden, J.L., 2006, Timing of magma- Scholle, D.W., eds., Formation and Applications of the Amato, J.M., Lawton, T.F., Mauel, D.J., Leggett, W.J., tism following initial convergence at a passive mar- Sedimentary Record in Arc Collision Zones: Geologi- González-León, C.M., Farmer, G.L., and Wooden, J.L., gin, southwestern U.S. Cordillera, and ages of lower cal Society of America Special Paper 436, p. 279–308, 2009, Testing the Mojave–Sonora megashear hypothe- crustal magma sources: The Journal of Geology, v. 114, https://doi​.org/10.1130/2008.2436(13). sis: Evidence from Paleoproterozoic igneous rocks and p. 231–245, https://doi​.org/10.1086/499573. Centeno-García, E., Busby, C.J., Busby, M., and Gehrels, deformed Mesozoic strata in Sonora, Mexico: Geology, Bassett, K.N., and Busby, C.J., 2005, Tectonic setting of G.E., 2011, Evolution of the Guerrero composite ter- v. 37, p. 75–78, https://doi​.org/10.1130/G25240A.1. the Glance Conglomerate along the Sawmill Canyon rane along the Mexican margin, from extensional Amato, J.M., Heizler, M.T., Boullion, A.O., Sanders, A.E., fault zone, southern Arizona: A sequence analysis fringing arc to contractional continental arc: Geologi- Toro, J., McLemore, V.T., and Andronicos, C.L., 2011, of an intra-arc strike-slip basin, in Anderson, T.H., cal Society of America Bulletin, v. 123, p. 1776–1797, Syntectonic 1.4 Ga magmatism and rapid cooling Nourse, J.A., McKee, J.W., and Steiner, M.B., eds., https://doi​.org/10.1130/B30057.1. of a gneiss dome in the southern Mazatzal Province: The Mojave-Sonora Megashear Hypothesis: Develop- Chafetz, H.S., 1982, The Upper Cretaceous Beartooth Sand- Burro Mountains, New Mexico: Geological Society ment, Assessment, and Alternatives: Geological Society stone of southwestern New Mexico: A transgressive of America Bulletin, v. 123, p. 1720–1744, https://doi​ of America Special Paper 393, p. 377–400, https://doi​ deltaic complex on silicified paleokarst: Journal of .org/10.1130/B30337.1. .org/10.1130/0-8137-2393-0.377. Sedimentary Research, v. 52, p. 157–169. Anderson, J.L., 1989, Proterozoic anorogenic of the Bauer, R.D., 1989, Depositional environments, sediment Christiansen, E.H., Kowallis, B.J., and Barton, M.D., 1994, southwestern United States, in Jenney, J.P., and Reyn- dispersal, and provenance of the Dakota Sandstone, Temporal and spatial distribution of volcanic ash in Me- olds, S.J., eds., Geologic Evolution of Arizona: Tuc- , south-central New Mexico [M.S. sozoic sedimentary rocks of the Western Interior: An son, Arizona, Arizona Geological Society Digest 17, thesis]: Las Cruces, New Mexico, New Mexico State alternative record of Mesozoic magmatism, in Caputo, p. 211–238. University, 64 p. M.V., Peterson, J.A., and Franczyk, K.J., eds., Meso- Anderson, O.J., and Lucas, S.G., 1996, Stratigraphy and Bayona, G., and Lawton, T.F., 2003, Fault-proximal strati- zoic Systems of the Rocky Mountain Region: USA: depositional environments of Middle and Upper Juras- graphic record of episodic extension and oblique Denver, Colorado, Rocky Mountain Section Society of sic rocks, southeastern San Juan basin, New Mexico, inversion, Bisbee basin, southwestern New Mexico, Economic Paleontologists and Mineralogists, p. 73-94. in Goff, F., Kues, B.S., Rogers, M.A., McFadden, L.S., USA: Basin Research, v. 15, p. 251–270, https://doi​ Clemons, R.E., 1982, Geology of quadrangle, and Gardner, J.N., eds., Region: New .org/10.1046/​j.1365-2117.2003.00199.x. Luna County, New Mexico: New Mexico Bureau of Mexico Geological Society Guidebook 47, p. 205–210. Bilodeau, W.L., 1979, Early Cretaceous tectonics and deposi- Mines and Mineral Resources Geologic Map 51, scale Anderson, T.A., and Silver, L.T., 2005, The Mojave-Sonora tion of the Glance Conglomerate, southeastern Arizona 1:24,000. megashear–Field and analytical studies leading to the [Ph.D. thesis]: Stanford, California, Stanford Univer- Clinkscales, C.A., 2011, Structural geology, Laramide stra- conception and evolution of the hypothesis, in An- sity, 145 p. tigraphy, and kinematic history of the northern Little derson, T.H., Nourse, J.A., McKee, J.W., and Steiner, Bilodeau, W.L., 1982, Tectonic models for Ear- Hatchet Mountains, Grant and Hidalgo Counties, south- M.B., eds., The Mojave-Sonora Megashear Hypothesis: ly Cretaceous rifting in southeastern Ari- western New Mexico [M.S. thesis]: Las Cruces, New Development, Assessment, and Alternatives: Geo- zona: Geology, v. 10, p. 466–470, https://doi​ Mexico, New Mexico State University, 174 p. logical Society of America Special Paper 393, p. 1–50, .org/10.1130/0091-7613(1982)10<466:TMFECR>2.0 Clinkscales, C.A., and Lawton, T.F., 2015, Timing of Late https://doi​.org/10.1130/0-8137-2393-0.1. .CO;2. Cretaceous shortening and basin development, Little

2512 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

Hatchet Mountains, southwestern New Mexico, USA— Dickinson, W.R., and Gehrels, G.E., 2009a, U-Pb ages of Fosdick, J.C., Graham, S.A., and Hilley, G.E., 2014, Influ- Implications for regional Laramide tectonics: Basin detrital zircons in Jurassic eolian and associated sand- ence of attenuated lithosphere and sediment loading on Research, v. 27, p. 453–472, https://doi​.org/10.1111/ stones of the Colorado Plateau: Evidence for transconti- flexure of the deep-water Magallanes retroarc foreland bre.12083. nental dispersal and intraregional recycling of sediment: basin, Southern Andes: Tectonics, v. 33, p. 2505–2525, Clinkscales, C.A., and Lawton, T.F., 2018, Mesozoic–Paleo- Geological Society of America Bulletin, v. 121, no. 3-4, https://doi​.org/10.1002/2014TC003684. gene structural evolution of the southern U.S. Cordille- p. 408–433, https://doi​.org/10.1130/B26406.1. Galemore, J.A., 1986, Stratigraphy, depositional environ- ra as revealed in the Little and Big Hatchet Mountains, Dickinson, W.R., and Gehrels, G.E., 2009b, Use of U–Pb ments, and provenance of the Mojado Formation southwest New Mexico: Geosphere, v. 14, p. 162–186, ages of detrital zircons to infer maximum deposi- (middle Albian–early Cenomanian), southwestern New https://doi​.org/10.1130/GES01539.1. tional ages of strata: A test against a Colorado Pla- Mexico [M.S. thesis]: Las Cruces, New Mexico, New Cobban, W.A., 1977, Fossil mollusks of the Dakota Sand- teau Mesozoic database: Earth and Planetary Science Mexico State University, 103 p. stone and intertongued Mancos Shale of west-central Letters, v. 288, p. 115–125, https://doi​.org/10.1016/​ Gans, P.B., 1997, Large-magnitude Oligo- extension New Mexico: New Mexico Geological Society Guide- j.epsl.2009.09.013. in southern Sonora: Implications for the tectonic evolu- book, v. 28, p. 213–220. Dickinson, W.R., and Lawton, T.F., 2001a, Carbonifer- tion of northwest Mexico: Tectonics, v. 16, p. 388–408, Cobban, W.A., 1987, Ammonite faunas of the Sarten Sand- ous to Cretaceous assembly and fragmentation of https://doi​.org/10.1029/97TC00496. stone (Cretaceous), Luna County, New Mexico: U.S. Mexico: Geological Society of America Bulletin, Garzanti, E., 2016, From static to dynamic provenance Geological Survey Bulletin 1641-B. v. 113, p. 1142–1160, https://doi​.org/10.1130/0016- analysis—Sedimentary petrology upgraded: Sedimen- Cobban, W.A., Hook, S.C., and Kennedy, W.J., 1989, Upper 7606(2001)113<1142:CTCAAF>2.0.CO;2. tary Geology, v. 336, p. 3–13, https://doi​.org/10.1016/​ Cretaceous rocks and ammonite faunas of southwestern Dickinson, W.R., and Lawton, T.F., 2001b, Tectonic set- j.sedgeo.2015.07.010. New Mexico: New Mexico Bureau of Mines and Min- ting and sandstone petrofacies of the Bisbee basin Gastil, R.G., 1975, Plutonic zones in the Peninsular Ranges eral Resources Memoir 45, 137 p. (USA-Mexico): Journal of South American Earth of southern California and northern Baja California: Ge- Cobban, W.A., Walaszczyk, I., Obradovich, J.D., and McK- Sciences, v. 14, p. 475–504, https://doi​.org/10.1016/ ology, v. 3, p. 361–363, https://doi​.org/10.1130/0091- inney, K.C., 2006, A USGS zonal table for the Upper S0895-9811(01)00046-3. 7613(1975)3<361:PZITPR>2.0.CO;2. Cretaceous middle Cenomanian- of the Dickinson, W.R., Klute, M.A., and Swift, P.N., 1986, The Gawthorpe, R.L., and Leeder, M.R., 2000, Tectono-sedi- Western Interior of the United States based on ammo- Bisbee basin and its bearing on late Mesozoic paleo- mentary evolution of active extensional basins: Basin nites, inoceramids, and radiometric ages: U.S. Geologi- geographic and paleotectonic relations between the Research, v. 12, p. 195–218. cal Survey Open-File Report OF 2006-1250, 46 p. Cordilleran and Caribbean regions, in Abbott, P.A., Gehrels, G.E., Blakey, R., Karlstrom, K.E., Timmons, J.M., Cohen, K.M., Finney, S.C., Gibbard, P.L., and Fan, J.-X., ed., Cretaceous Stratigraphy, Western North America: Dickinson, B., and Pecha, M., 2011, Detrital zircon 2013, Updated, The ICS International Chronostrati- Bakersfield, California: Society of Economic Paleon- U-Pb geochronology of Paleozoic strata in the Grand graphic Chart: Episodes, v. 36, p. 199–204. tologists and Mineralogists Book 46, p. 51–62. Canyon, Arizona: Lithosphere, v. 3, p. 183–200, https:// Critelli, S., and Ingersoll, R.V., 1995, Interpretation of Dickinson, W.R., Fiorillo, A.R., Hall, D.L., Monreal, R., Po- doi​.org/10.1130/L121.1. neovolcanic versus palaeovolcanic sand grains: An tochnik, A.R., and Swift, P.N., 1989, Cretaceous strata Giallorenzo, M.A., Wells, M.L., Yonkee, W.A., Stockli, example from Miocene deep-marine sandstone of the of southern Arizona, in Jenney, J.P., and Reynolds, S.J., D.F., and Wernicke, B.P., 2018, Timing of exhumation, Topanga Group (Southern California): Sedimentol- eds., Geologic evolution of Arizona: Tucson, Arizona, Wheeler Pass thrust sheet, southern Nevada and Cali- ogy, v. 42, p. 783–804, https://doi​.org/10.1111/​j.1365- Arizona Geological Society Digest 17, p. 447–461. fornia: Late Jurassic to middle Cretaceous evolution 3091.1995.tb00409.x. Dickinson, W.R., Lawton, T.F., and Gehrels, G.E., 2009, of the southern Sevier fold-and-thrust belt: Geological Cross, T.A., 1986, Tectonic controls of foreland basin subsid- Recycling detrital zircons: A case study from the Cre- Society of America Bulletin, v. 130, p. 558–579, https:// ence and Laramide style deformation, western United taceous Bisbee Group of southern Arizona: Geology, doi​.org/10.1130/B31777.1. States, in Allen, P.A., and Homewood, P., eds., Fore- v. 37, p. 503–506, https://doi.org/10.1130/G25646A.1​ . Gilbert, J.C., 2012, Age and provenance of the Glance Con- land Basins: International Association of Sedimen- Drewes, H., du Bray, E.A., and Pallister, J.S., 1995, Geo- glomerate, Morita Formation, and equivalent strata tologists Special Publication 8, p. 15–39, https://doi​ logic map of the Portal quadrangle and vicinity, Cochise from U-Pb geochronology, southeastern Arizona and .org/10.1002/9781444303810.ch1. County, southeast Arizona: U.S. Geological Survey southwestern New Mexico [M.S. thesis]: Las Cruces, Currie, B.S., 1997, of nonma- Map I-2450, scale 1:24,000. New Mexico, New Mexico State University, 181 p. rine Jurassic-Cretaceous rocks, central Cordilleran Ducea, M.N., 2001, The California arc: Thick granitic batho- Gillerman, E., 1958, Geology of the central Peloncillo foreland-basin system: Geological Society of Amer- liths, eclogitic residues, lithospheric-scale thrusting, and Mountains, Hidalgo County, New Mexico, and Cochise ica Bulletin, v. 109, no. 9, p. 1206–1222, https://doi​ magmatic flare-ups: GSA Today, v. 11, no. 11, p. 4–10, County, Arizona: New Mexico State Bureau of Mines .org/10.1130/0016-7606(1997)109<1206:SSONJC>2 https://doi.org/10.1130/1052-5173(2001)011​ <0004:TC and Mineral Resources Bulletin 57, 152 p. .3.CO;2. ATGB>2.0.CO;2. González-León, C.M., 1994, Early Cretaceous tectono-sed- Currie, B.S., 1998, Upper Jurassic-Lower Cretaceous Mor- Eldrett, J.S., Minisini, D., and Bergman, S.C., 2014, Decou- imentary evolution of the southwestern margin of the rison and Cedar Mountain Formations, NE Utah-NW pling of the carbon cycle during Ocean Anoxic Event Bisbee basin: Revista Mexicana de Ciencias Geológi- Colorado: Relationships between nonmarine deposition 2: Geology, v. 42, p. 567–570, https://doi​.org/10.1130/ cas, v. 11, p. 139–146. and early Cordilleran foreland basin development: Jour- G35520.1. González-León, C.M., Scott, R.W., Löser, H., Lawton, T.F., nal of Sedimentary Research, v. 68, p. 632–652, https:// Elston, W.E., 1958, Burro uplift, northeastern limit of Robert, E., and Valencia, V.A., 2008, Upper Aptian– doi​.org/10.2110/jsr.68.632. sedimentary basin of southwestern New Mexico and Lower Albian Mural Formation: Stratigraphy, biostra- Currie, B.S., 2002, Structural configuration of the Early southeastern Arizona: The American Association of tigraphy and depositional cycles on the Sonoran Shelf, Cretaceous Cordilleran foreland-basin system and Petroleum Geologists Bulletin, v. 42, p. 2513–2517. northern México: Cretaceous Research, v. 29, p. 249– Sevier thrust belt, Utah and Colorado: The Journal Ferguson, R.C., 1987, Bisbee Group (Lower Cretaceous) car- 266, https://doi​.org/10.1016/​j.cretres.2007.06.001. of Geology, v. 110, no. 6, p. 697–718, https://doi​ bonates, Guadalupe Canyon area, southeastern Arizona, González-León, C.M., Solari, L., Solé, J., Ducea, M.N., .org/10.1086/342626. in Dickinson, W.R., and Klute, M.A., eds., Mesozoic Lawton, T.F., Bernal, J.P., González Becuar, E., Gray, Darton, N.H., 1917, Description of the Deming quadrangle: Rocks of Southern Arizona and Adjacent Areas: Tuc- F., López Martínez, M., and Lozano Santacruz, R., U.S. Geological Survey, Geological Atlas, Deming fo- son, Arizona, Arizona Geological Society Digest 18, 2011, Stratigraphy, geochronology, and geochemistry lio 207, 15 p. p. 365–370. of the Laramide magmatic arc in north-central Sonora, DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Fildani, A., and Hessler, A.M., 2005, Stratigraphic record Mexico: Geosphere, v. 7, p. 1392–1418, https://doi​ Cordilleran thrust belt and foreland basin system, west- across a retroarc basin inversion: Rocas Verdes–Ma- .org/10.1130/GES00679.1. ern U.S.A: American Journal of Science, v. 304, no. 2, gallanes Basin, Patagonian Andes, Chile: Geological González-León, C.M., Solari, L.A., and Madhavaraju, J., p. 105–168, https://doi​.org/10.2475/ajs.304.2.105. Society of America Bulletin, v. 117, p. 1596–1614, 2017, Stratigraphy, geochronology and regional tec- DeCelles, P.G., Ducea, M.N., Kapp, P., and Zandt, G., 2009, https://doi​.org/10.1130/B25708.1. tonic setting of the Late Cretaceous (ca. 82–70 Ma) Cyclicity in Cordilleran orogenic systems: Nature Geosci- Finnell, T.L., 1987, Geologic map of the Cliff quadrangle, Cabullona basin, Sonora, Mexico: Journal of South ence, v. 2, p. 251–257, https://doi​.org/10.1038/ngeo469. Grant County, New Mexico: U.S. Geological Survey American Earth Sciences, v. 80, p. 494–511, https://doi​ Dickinson, W.R., 1985, Interpreting provenance from detrital Miscellaneous Investigations Series, Map I-1768, scale .org/10.1016/​j.jsames.2017.09.037. modes of sandstones, in Zuffa, G.G., ed., Provenance of 1:24,000. Gurnis, M., 1992, Rapid continental subsidence following Arenites: Dordrecht, Netherlands, D. Reidel, p. 333– Fitz-Díaz, E., Lawton, T.F., Juárez-Arriaga, E., and Chávez- the initiation and evolution of subduction: Science, 361, https://doi​.org/10.1007/978-94-017-2809-6_15. Cabello, G., 2018, The Cretaceous–Paleogene Mexican v. 255, p. 1556–1558, https://doi​.org/10.1126/sci- Dickinson, W.R., 2018, Notes on the Cretaceous and orogen: Structure, basin development, magmatism and ence.255.5051.1556. Laramide of the Colorado Plateau: Tucson, Arizona, tectonics: Earth-Science Reviews, v. 183, p. 56–84, Haenggi, W.T., 2002, Tectonic history of the Chihuahua Arizona Geological Survey, Contributed Report CR- https://doi​.org/10.1016/​j.earscirev.2017.03.002. trough, Mexico and adjacent USA, Part II: Meso- 18-G, 32 p. Fletcher, J.M., Grove, M., Kimbrough, D., Lovera, O., and zoic and Cenozoic: Boletín de la Asociacion Ge- Dickinson, W.R., and Gehrels, G.E., 2008, Sediment deliv- Gehrels, G.E., 2007, Ridge-trench interactions and the ológica Mexicana, v. 55, no. 1, p. 38–94, https://doi​ ery to the Cordilleran foreland basin: Insights from Neogene tectonic evolution of the Magdalena shelf and .org/10.18268/BSGM2002v55n1a4. U-Pb ages of detrital zircons in Upper Jurassic and southern Gulf of California: Insights from detrital zir- Hedlund, D.C., 1980, Geologic map of the Redrock SE quad- Cretaceous strata of the Colorado Plateau: American con U-Pb ages from the Magdalena fan and adjacent rangle, Grant County, New Mexico: Reston, Virginia, Journal of Science, v. 308, p. 1041–1082, https://doi​ areas: Geological Society of America Bulletin, v. 119, U.S. Geological Survey Miscellaneous Field Studies, .org/10.2475/10.2008.01. p. 1313–1336, https://doi​.org/10.1130/B26067.1. Map MF-1265, scale 1:25,000.

Geological Society of America Bulletin, v. 132, no. 11/12 2513

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

Heller, P.L., Bowdler, S.S., Chambers, H.P., Coogan, J.C., Rocks of Southern Arizona and Adjacent Areas: Tuc- Mexico Geological Society Special Publication 11, Hagen, E.S., Shuster, M.W., Winslow, N.S., and son, Arizona, Arizona Geological Society Digest 18, p. 137–152. Lawton, T.F., 1986, Time of initial thrusting in the p. 263–272. Lucas, S.G., and Estep, J.W., 1998a, Cretaceous stratigraphy Sevier orogenic belt, Idaho-Wyoming and Utah: Geol- Klute, M.A., 1991, Sedimentology, sandstone petrofacies, and biostratigraphy in the southern San Andres Moun- ogy, v. 14, p. 388–392, https://doi​.org/10.1130/0091- and tectonic setting of the late Mesozoic Bisbee basin, tains, Doña Ana County, New Mexico, in Mack, G.H., 7613(1986)14<388:TOITIT>2.0.CO;2. southeastern Arizona [Ph.D. thesis]: Tucson, Arizona, Austin, G.S., and Barker, J. M., eds., Las Cruces Coun- Hildebrand, R.S., and Whalen, J.B., 2014, Arc and slab-fail- University of Arizona, 268 p. try II: New Mexico Geological Society Guidebook 49, ure magmatism in Cordilleran batholiths II—The Creta- Langenheim, V.E., and Jachens, R.C., 2003, Crustal struc- p. 187–196. ceous Peninsular Ranges batholith of southern and Baja ture of the Peninsular Ranges batholith from magnetic Lucas, S.G., and Estep, J.W., 1998b, and California: Geoscience , v. 41, no. 4, p. 399– data: Implications for Gulf of California rifting: Geo- biostratigraphy of the Lower–middle Cretaceous Bis- 458, https://doi​.org/10.12789/geocanj.2014.41.059. physical Research Letters, v. 30, p. 1597, https://doi​ bee Group, southwestern New Mexico, USA, in Lucas, Ingersoll, R.V., 2012, Tectonics of sedimentary basins, with .org/10.1029/2003GL017159. S.G., Kirkland, J.I., and Estep, J.W., eds., Lower and revised nomenclature, in Busby, C., and Azor Pérez, Laskowski, A.K., DeCelles, P.G., and Gehrels, G.E., 2013, Middle Cretaceous Terrestrial Ecosystems: Albuquer- A., eds., Tectonics of Sedimentary Basins: Recent Ad- Detrital zircon geochronology of Cordilleran retroarc que, New Mexico, New Mexico Museum of Natural vances: Oxford, Blackwell Publishing Ltd., p. 3–43. foreland basin strata, western North America: Tec- History and Science Bulletin 14, p. 39–55. Jacobson, C.E., Grove, M., Pedrick, J.N., Barth, A.P., Marsa- tonics, v. 32, p. 1027–1048, https://doi​.org/10.1002/ Lucas, S.G., and Lawton, T.F., 2000, Stratigraphy of the glia, K.M., Gehrels, G.E., and Nourse, J.A., 2011, Late tect.20065. Bisbee Group (Jurassic–Cretaceous), Little Hatchet Cretaceous-early Cenozoic tectonic evolution of the Lasky, S.G., 1936, Geology and Ore Deposits of the Bayard Mountains, New Mexico, in Lawton, T.F., McMillan, southern California margin inferred from provenance Area, Central Mining District: Socorro, New Mexico: N.J., and McLemore, V.T., eds., Southwest Passage-A of trench and forearc : Geological Society U.S. Geological Survey Bulletin 870, 144 p. Trip Through the Phanerozoic: New Mexico Geological of America Bulletin, v. 123, p. 485–506, https://doi​ Lawton, T.F., 1982, Lithofacies correlations within the Upper Society Guidebook 51, p. 175–194. .org/10.1130/B30238.1. Cretaceous Indianola Group, central Utah, in Nielson, Lucas, S.G., and Lawton, T.F., 2005, Upper Cretaceous marine Jacques-Ayala, C., 1989, Arroyo Sásabe Formation (Aptian- D.L., ed., Overthust Belt of Utah: Utah Geological As- strata in the Little Hatchet Mountains, southwestern New Albian), northwestern Sonora, Mexico—Marginal sociation Publication 10, p. 199–213. Mexico: New Mexico Geology, v. 27, p. 63–69. marine sedimentation in the Sonora back-arc basin: Lawton, T.F., 1994, Tectonic setting of Mesozoic sedimen- Lucas, S.G., Kues, B.S., Hayden, S.N., Allen, B.D., Ki- Universidad Nacional Autónoma de México, Instituto tary basins, Rocky Mountain region, United States, in etzke, K.K., Williamson, T.E., Sealey, P., and Pence, de Geología: Revista, v. 8, p. 171–178. Caputo, M.V., Peterson, J.A., and Franczyk, K.A., eds., R., 1988, Cretaceous stratigraphy and biostratigraphy, Jacques-Ayala, C., 1995, Paleogeography and provenance of Mesozoic Systems of the Rocky Mountain Region, Cooke’s Range, Luna County, New Mexico, in Mack, the Lower Cretaceous Bisbee Group in the Caborca- USA: Tulsa, Oklahoma, Society for Sedimentary Ge- G.H., Lawton, T.F., and Lucas, S.G., eds., Cretaceous Santa Ana area, northwestern Sonora, in Jacques-Ayala, ology (SEPM), p. 1–25. and Laramide Tectonic Evolution of Southwestern New C., González-León, C.M., and Roldán-Quintana, J., eds., Lawton, T.F., 2000, Inversion of Late Jurassic–Early Creta- Mexico: New Mexico Geological Society Guidebook Studies on the Mesozoic of Sonora and Adjacent Ar- ceous extensional faults of the Bisbee basin, southeast- 39, p. 143–167. eas: Geological Society of America Special Paper 301, ern Arizona and southwestern New Mexico, in Lawton, Lucas, S.G., Estep, J.W., Boucher, L.D., and Anderson, p. 79–98, https://doi.org/10.1130/0-8137-2301-9.79​ . T.F., McMillan, N.J., and McLemore, V.T., eds., South- B.G., 2000, Cretaceous Stratigraphy, Biostratigraphy Jennings, G.R., Lawton, T.F., and Clinkscales, C.A., 2013, west Passage-A Trip through the Phanerozoic: New and Depositional Environments near Virden, Hidalgo Late Cretaceous U-Pb tuff ages from the Skunk Ranch Mexico Geological Society Guidebook 51, p. 95–102. County, New Mexico: New Mexico Museum of Natural Formation and their implications for age of Laramide Lawton, T.F., 2004, Upper Jurassic and Lower Cretaceous History and Science Bulletin 16, p. 107–117. deformation, Little Hatchet Mountains, southwest- strata of southwestern New Mexico and northern Chi- Lucas, S.G., Zeigler, K.E., Lawton, T.F., and Filkorn, H.F., ern New Mexico, U.S.A: Cretaceous Research, v. 43, huahua, Mexico, in Mack, G.H., and Giles, K.A., eds., 2001, Late Jurassic invertebrate fossils from the Little p. 18–25, https://doi.org/10.1016/​ j.cretres.2013.02.001​ . The —A Geologic History: Hatchet Mountains, southwestern New Mexico: New Johnson, S.E., Tate, M.C., and Fanning, C.M., 1999, Socorro, New Mexico, New Mexico Geological Soci- Mexico Geology, v. 23, p. 16–20. New geologic mapping and SHRIMP U-Pb zir- ety, p. 153–168. Lucas, S.G., Krainer, K., Spielmann, J.A.S., and Durney, K., con data in the Peninsular Ranges batholith, Baja Lawton, T.F., and Harrigan, P.J., 1998, Broken Jug Forma- 2010, Cretaceous stratigraphy, paleontology, petrogra- California, Mexico: Evidence for a suture?: Geol- tion: Redefinition of lower part of Bisbee Group, Little phy, depositional environments, and cycle stratigraphy ogy, v. 27, p. 743–746, https://doi​.org/10.1130/0091- Hatchet Mountains, Hidalgo County, New Mexico: at Cerro de Cristo Rey, Doña Ana County, New Mexico: 7613(1999)027<0743:NGMASU>2.3.CO;2. New Mexico Geology, v. 20, p. 69–77. New Mexico Geology, v. 32, p. 103–130. Jordan, T.E., 1981, Thrust loads and foreland basin evolu- Lawton, T.F., and McMillan, N.J., 1999, Arc abandonment as Ludwig, K.R., 2012, User’s Manual for Isoplot 3.75: A Geo- tion, Cretaceous, western United States: The American a cause for passive continental rifting: Comparison of chronological Toolkit for Microsoft Excel: Berkeley Association of Petroleum Geologists Bulletin, v. 65, the Jurassic Mexican Borderland rift and the Cenozoic Geochronology Center Special Publication 5, 75 p. p. 2506–2520. : Geology, v. 27, p. 779–782, https:// Lyons, J.I. Jr., 2016, Arc-related Mesozoic basins of northern Juárez-Arriaga, E., Lawton, T.F., Ocampo-Díaz, Y.Z.E., doi​.org/10.1130/0091-7613(1999)027<0779:AAAAC Mexico: Their origin, tectonic inversion and influence Stockli, D.F., and Solari, L., 2019, Sediment prove- F>2.3.CO;2. on ore localization [Ph.D. thesis]: Austin, Texas, The nance, sediment-dispersal systems, and major arc-mag- Lawton, T.F., and Olmstead, G.A., 1995, Stratigraphy University of Texas at Austin, 455 p. matic events recorded in the Mexican foreland basin, and structure of the lower part of the Bisbee Group, Machin, S.E.K., 2013, Solving the Beartooth problem: Us- North-Central and Northeastern Mexico: International northeastern Chiricahua Mountains, Arizona, USA, in ing U-Pb zircon geochronology to investigate the sig- Geology Review, v. 61, no. 17, p. 2118–2142, https:// Jacques-Ayala, C., González-León, C.M., and Roldán- nificance of mid-Cretaceous sandstones in New Mexico doi​.org/10.1080/00206814.2019.1581848. Quintana, J., eds., Studies on the Mesozoic of Sonora [M.S. thesis]: Las Cruces, New Mexico, New Mexico Karlstrom, K.E., Amato, J.M., Williams, M.L., Heizler, M., and Adjacent Areas: Geological Society of America State University, 141 p. Shaw, C., Read, A., and Bauer, P., 2004, Proterozoic Special Paper 301, p. 21–39, https://doi​.org/10.1130/0- Mack, G.H., 1987a, Mid-Cretaceous (late Albian) change evolution of the New Mexico region: A synthesis, in 8137-2301-9.21. from rift to retroarc foreland basin in southwestern Mack, G.H., and Giles, K.A., eds., The Geology of New Lawton, T.F., González-León, C.M., Lucas, S.G., and Scott, New Mexico: Geological Society of America Bulle- Mexico: A Geologic History: New Mexico Geological R.W., 2004, Stratigraphy and sedimentology of the tin, v. 98, p. 507–514, https://doi​.org/10.1130/0016- Society Special Publication 11, p. 1–34. upper Aptian-lower Albian Mural Limestone (Bis- 7606(1987)98<507:MLACFR>2.0.CO;2. Kimbrough, D.L., Smith, D.P., Mahoney, J.B., Moore, bee Group) in northern Sonora, Mexico: Cretaceous Mack, G.H., 1987b, Tectonic, eustatic, and climatic control T.E., Gastil, R.G., Ortega-Rivera, A., and Fanning, Research, v. 25, p. 43–60, https://doi​.org/10.1016/​ on Cretaceous (Aptian to Turonian) sedimentation in C.M., 2001, Forearc basin sedimentary response j.cretres.2003.09.003. southwestern New Mexico, in Dickinson, W.R., and to rapid Late Cretaceous batholith emplacement Lawton, T.F., Bradford, I.A., Vega, F.J., Gehrels, G.E., and Klute, M.A., eds., Mesozoic rocks of southern Arizona in the Peninsular Ranges of southern and Baja Amato, J.M., 2009, Provenance of Upper Cretaceous– and adjacent areas: Tucson, Arizona, Arizona Geologi- California: Geology, v. 29, p. 491–494, https://doi​ Paleogene sandstones in the foreland basin system of cal Society Digest 17, p. 371–379. .org/10.1130/0091-7613(2001)029<0491:FBSRTR>2 the Sierra Madre Oriental, northeastern Mexico, and Mack, G.H., Kolins, W.B., and Galemore, J.A., 1986, Lower .0.CO;2. its bearing on fluvial dispersal systems of the Mexican Cretaceous stratigraphy, depositional environments, Kimbrough, D.L., Abbott, P.L., Balch, D.C., Bartling, S.H., Laramide Province: Geological Society of America and sediment dispersal in southwestern New Mexico: Grove, M., Mahoney, J.B., and Donohue, R.F., 2014, Bulletin, v. 121, p. 820–836, https://doi​.org/10.1130/ American Journal of Science, v. 286, p. 309–331, Upper Jurassic Peñasquitos Formation—Forearc basin B26450.1. https://doi​.org/10.2475/ajs.286.4.309. western wall rock of the Peninsular Ranges batholith, Leggett, W.J., 2009, Stratigraphy, sedimentology and geo- Mack, G.H., Galemore, J.A., and Kaczmarek, E.L., 1988, in Morton, D.M., and Miller, F.K., eds., Peninsular chronology of Lower and Middle Jurassic rocks near The Cretaceous foreland basin in southwestern New Ranges Batholith, Baja California and Southern Cali- Rancho San Martin del Rincon, North-central Sonora, Mexico, in Mack, G.H., Lawton, T.F., and Lucas, S.G., fornia: Geological Society of America Memoir 211, Mexico [M.S. thesis]: Las Cruces, New Mexico, New eds., Cretaceous and Laramide Tectonic Evolution of p. 625–643, https://doi​.org/10.1130/2014.1211(1119). Mexico State University, 203 p. Southwestern New Mexico: New Mexico Geological Klute, M.A., 1987, Tectonic significance of sandstone petro- Lucas, S.G., 2004, The Triassic and Jurassic systems in New Society Guidebook 39, p. 135–141. facies within the Bisbee basin of southeastern Arizona, Mexico, in Mack, G.H., and Giles, K.A., eds., The Marsaglia, K.M., Barone, M., Critelli, S., Busby, C., and in Dickinson, W.R., and Klute, M.A., eds., Mesozoic Geology of New Mexico: A Geologic History: New Fackler-Adams, B., 2016, Petrography of volcaniclas-

2514 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Rift to foreland transition, Bisbee basin

tic rocks in intra-arc volcano-bounded to fault-bounded Nadin, E.S., Saleeby, J., and Wong, M., 2016, Thermal evo- Mesozoic–Cenozoic tectonic history of northern Sierra basins of the Rosario segment of the Lower Cretaceous lution of the Sierra Nevada batholith, California, and Los Ajos and adjacent areas, Sonora, Mexico: Journal Alisitos oceanic arc, Baja California, Mexico: Sedi- implications for strain localization: Geosphere, v. 12, of South American Earth Sciences, v. 29, p. 557–571, mentary Geology, v. 336, p. 138–146, http://dx.doi​ no. 2, p. 377–399, https://doi.org/10.1130/GES01224.1​ . https://doi​.org/10.1016/​j.jsames.2009.11.008. .org/110.1016/​j.sedgeo.2015.1011.1008, https://doi​ New Mexico Bureau of Geology and Mineral Resources, Paige, S., 1916, Description of the Silver City quadrangle: .org/10.1016/​j.sedgeo.2015.11.008. 2003, Geologic map of New Mexico, scale 1:500,000. U.S. Geological Survey Atlas, Silver City folio, 19 p. Martini, M., Ferrari, L., López-Martínez, M., Cerca-Mar- Nio, S.-D., and Yang, C.-S., 1991, Diagnostic attributes of Palacios-García, N.B., and Martini, M., 2014, From back-arc tínez, M., Valencia, V.A., and Serrano-Durán, L., 2009, clastic tidal deposits: A review, in Smith, D.G., Re- rifting to arc accretion: The Late Jurassic–Early Creta- Cretaceous–Eocene magmatism and Laramide defor- inson, G.E., Zaitlin, B.A., and Rahmani, R.A., eds., ceous evolution of the Guerrero terrane recorded by a mation in southwestern Mexico: No role for terrane Clastic Tidal Sedimentology: Canadian Society of Pe- major provenance change in sandstones from the Sierra accretion, in Kay, S.M., Ramos, V.A., and Dickinson, troleum Geologists Memoir 16, p. 3–28. de los Cuarzos area, central Mexico: International Ge- W.R., eds., Backbone of the Americas: Shallow Sub- North American Commission on Stratigraphic Nomen- ology Review, v. 56, p. 1377–1394, https://doi​.org/10.1 duction, Plateau Uplift and Ridge and Terrane Colli- clature, 2005, North American Stratigraphic Code: 080/00206814.2014.938367. sion: Geological Society of America Special Paper 204, The American Association of Petroleum Geologists Pang, M., and Nummedal, D., 1995, Flexural subsidence p. 151–182, https://doi​.org/10.1130/2009.1204(07). Bulletin, v. 89, no. 11, p. 1547–1591, https://doi​ and basement tectonics of the Cretaceous interior ba- Martini, M., Mori, L., Solari, L., and Centeno-García, E., .org/10.1306/07050504129. sin, United States: Geology, v. 23, p. 173–176, https:// 2011, Sandstone provenance of the Arperos Basin Nummedal, D., 2004, Tectonic and eustatic controls on Up- doi​.org/10.1130/0091-7613(1995)023<0173:FSABTO (Sierra de Guanajuato, central Mexico): Late Jurassic– per Cretaceous stratigraphy of northern New Mexico, in >2.3.CO;2. Early Cretaceous back-arc spreading as the foundation Mack, G.H., and Giles, K.A., eds., The Geology of New Peryam, T.C., Lawton, T.F., Amato, J.M., González León, of the Guerrero terrane: The Journal of Geology, v. 119, Mexico: A Geologic History: New Mexico Geological C.M., and Mauel, D.J., 2012, Lower Cretaceous strata p. 597–617, https://doi​.org/10.1086/661989. Society Special Publication 11, p. 169–182. of the Sonora Bisbee Basin: A record of the tectono- Martini, M., Solari, L., and Camprubi, A., 2013, Kinematics Olmstead, G.A., and Young, K., 2000, Late Jurassic am- magmatic evolution of northwestern Mexico: Geologi- of the Guerrero terrane accretion in the Sierra de Gua- monites from the northeastern Chiricahua Mountains, cal Society of America Bulletin, v. 124, p. 532–548, najuato, central Mexico: New insights for the structural southeast Arizona: New Mexico Geology, v. 22, p. 1–7. https://doi​.org/10.1130/B30456.1. evolution of arc–continent collisional zones: Interna- Olmstead, G.A., Cornell, W.C., and Waanders, G.L., 1996, Pindell, J.L., and Kennan, L., 2009, Tectonic evolution of tional Geology Review, v. 55, no. 5, p. 574–589, https:// Oxfordian–Kimmeridgian (Jurassic) dinoflagellate the Gulf of Mexico, Caribbean and northern South doi​.org/10.1080/00206814.2012.729361. cysts from the Chiricahua Mountains, Arizona: Paly- America in the mantle reference frame: An update, in Martini, M., Solari, L., and López-Martínez, M., 2014, Cor- nology, v. 20, p. 221–230, https://doi​.org/10.1080/019 James, K.H., Lorente, M.A., and Pindell, J.L., eds., The relating the Arperos Basin from Guanajuato, central 16122.1996.9989477. Origin and Evolution of the Caribbean Plate: Geologi- Mexico, to Santo Tomás, southern Mexico: Implica- Ortega-Flores, B., Solari, L.A., and Escalona-Alcázar, cal Society of London Special Publication 328, p. 1–55, tions for the paleogeography and origin of the Guer- F.D.J., 2016, The Mesozoic successions of west- https://doi​.org/10.1144/SP328.1. rero terrane: Geosphere, v. 10, p. 1385–1401, https:// ern Sierra de Zacatecas, Central Mexico: Prov- Rangin, C., 1977, Sobre la presencia del Jurásico Superior doi​.org/10.1130/GES01055.1. enance and tectonic implications: Geological con amonitas en Sonora septentrional: Universidad Mauel, D.J., Lawton, T.F., González-León, C.M., Iriondo, A., Magazine, v. 153, p. 696–717, https://doi.org/10.1017/​ Nacional Autonóma de México, Instituto de Geología: and Amato, J.M., 2011, Stratigraphy and age of Upper S0016756815000977. Revista, v. 1, p. 1–4. Jurassic strata in north-central Sonora, Mexico: South- Ortega-Rivera, A., 2003, Geochronological constraints on Ransome, F.L., 1904, The geology and ore deposits of the western Laurentian record of crustal extension and tec- the tectonic history of the Peninsular Ranges batho- Bisbee quadrangle, Arizona: U.S. Geological Survey tonic transition: Geosphere, v. 7, p. 390–414, https://doi​ lith of Alta and Baja California: Tectonic implications Professional Paper 21, 167 p, https://doi​.org/10.3133/ .org/10.1130/GES00600.1. for western México, in Johnson, S.E., Paterson, S.R., pp21. May, S.R., Gray, G.G., Summa, L.L., Stewart, N.R., Geh- Fletcher, J.M., Girty, G.H., Kimbrough, D.L., and Reynolds, S.J., 1988, Geologic map of Arizona: Tucson, rels, G.E., and Pecha, M.E., 2013, Detrital zircon geo- Martín-Barajas, A., eds., Tectonic Evolution of North- Arizona, Arizona Geological Survey Map 26, scale chronology from the Bighorn Basin, Wyoming, USA: western Mexico and the Southwestern USA: Geologi- 1:1,000,000. Implications for tectonostratigraphic evolution and cal Society of America Special Paper 374, p. 297–335, Riggs, N.R., Mattinson, J.M., and Busby, C.J., 1993, Corre- paleogeography: Geological Society of America Bul- https://doi​.org/10.1130/0-8137-2374-4.297. lation of Jurassic eolian strata between the magmatic letin, v. 125, p. 1403–1422, https://doi​.org/10.1130/ Owen, A., Nichols, G.J., Hartley, A.J., Weissmann, G.S., and arc and the Colorado Plateau: New U-Pb geochrono- B30824.1. Scuderi, L.A., 2015, Quantification of a distributive flu- logic data from southern Arizona: Geological Society McIntosh, W.C., and Bryan, C., 2000, Chronology and geo- vial system: The Salt Wash DFS of the Morrison For- of America Bulletin, v. 105, p. 1231–1246, https:// chemistry of the Boot Heel volcanic field, New Mexico, mation, SW U.S.A: Journal of Sedimentary Research, doi​.org/10.1130/0016-7606(1993)105<1231:COJES in Lawton, T.F., McMillan, N.J., and McLemore, V.T., v. 85, p. 544–561, https://doi​.org/10.2110/jsr.2015.35. B>2.3.CO;2. eds., Southwest Passage: a Trip Through the Phanero- Owen, D.E., and Head, C.F., 2001, Summary of the sequence Riggs, N.R., Reynolds, S.J., Lindner, P.J., Howell, E.R., zoic, New Mexico Geological Society, Guidebook 51, stratigraphy of the Dakota Sandstone and adjacent Barth, A.P., and Walker, J.D., 2013, The Early Meso- p. 157–174. units, San Juan basin, northwestern New Mexico and zoic Cordilleran arc and Late Triassic paleotopography: McKee, J.W., McKee, M.B., and Anderson, T.H., 2005, Me- southwestern Colorado, in Broadhead, R., Cather, M., The detrital record in Upper Triassic sedimentary suc- sozoic basin formation, mass-gravity sedimentation, and Brister, B.S., eds., Proceedings for low permeabil- cessions on and off the Colorado Plateau: Geosphere, and inversion in northeastern Sonora and southeastern ity and underdeveloped natural gas reservoirs of New v. 9, p. 602–613, https://doi​.org/10.1130/GES00860.1. Arizona, in Anderson, T.H., Nourse, J.A., McKee, J.W., Mexico: Socorro, New Mexico, New Mexico Bureau of Riggs, N.R., Oberling, Z.A., Howell, E.R., Parker, W.G., and Steiner, M.B., eds., The Mojave-Sonora Megashear Mines and Mineral Resources, p. 2–19. Barth, A.P., Cecil, M.R., and Martz, J.W., 2016, Sourc- Hypothesis: Development, Assessment, and Alterna- Owen, D.E., and Owen, D.E.J., 2003, Stratigraphy of the Da- es of volcanic detritus in the basal , tives: Geological Society of America Special Paper 393, kota Sandstone and intertongued Mancos Shale along southwestern Laurentia, and implications for the Early p. 481–507, https://doi​.org/10.1130/2005.2393(18). the southern flank of the San Juan basin, west-central Mesozoic magmatic arc: Geosphere, v. 12, p. 439–463, McKee, M.B., and Anderson, T.H., 1998, Mass-gravity de- New Mexico, in Lucas, S.G., Semken, S.C., Beglof, https://doi​.org/10.1130/GES01238.1. posits and structures in the Lower Cretaceous of So- W., and Ulmer-Scholle, D, eds., Geology of the Zuni Robinson Roberts, L.N., and Kirschbaum, M.A., 1995, Pa- nora, Mexico: Geological Society of America Bulletin, Plateau: New Mexico Geological Society Guidebook leogeography of the Late Cretaceous of the Western v. 110, p. 1516–1529, https://doi​.org/10.1130/0016- 54, p. 325–330. Interior of middle North America- distribution and 7606(1998)110<1516:MGDASI>2.3.CO;2. Owen, D.E., Forgas, A.M., Miller, S.A., Stelly, R.J., and sediment accumulation: U.S. Geological Survey Pro- Mickus, K., Stern, R.J., Keller, G.R., and Anthony, E.Y., Owen, D.E., Jr., 2005, Surface and subsurface stratig- fessional Paper 1561, 115 p, https://doi​.org/10.3133/ 2009, Potential field evidence for a volcanic rifted raphy of the , Dakota Sand- pp1561. margin along the Texas Gulf Coast: Geology, v. 37, stone, and intertongued Mancos Shale of the Chama Rodríguez-Castañeda, J.L., 1990, Relaciones estructurales p. 387–390, https://doi​.org/10.1130/G25465A.1. basin, New Mexico, in Lucas, S.G., Zeigler, K.E., Lu- in la parte centroseptentrional del estado de Sonora: Monreal, R., and Longoria, J.F., 2000, Stratigraphy and eth, V.W., and Owen, D.E., eds., Geology of the Chama Universidad Nacional Autónoma de México, Instituto structure of the Lower Cretaceous of Lampazos, So- Basin: New Mexico Geological Society Guidebook 56, de Geología: Revista, v. 9, p. 51–61. nora, (northwest Mexico) and its relationship to the p. 218–226. Rodríguez-Castañeda, J.L., 1991, Mesozoic stratigraphy in Gulf Coast succession: The American Association of Owen, D.E., Siemers, C.T., and Owen, D.E., Jr., 2007, Da- north-central Sonora, Mexico: Boletín del Departamen- Petroleum Geologists Bulletin, v. 84, p. 1811–1831. kota and adjacent Morrison and lower Mancos stra- to de Geología: Universidad de Sonora, Hermosillo, Mortensen, J.K., Hall, B.V., Bissig, T., Friedman, R.M., tigraphy (Cretaceous and Jurassic) in the Holy Ghost v. 8, p. 13–27. Danielson, T., Oliver, J., Rhys, D.A., Ross, K.V., Spring Quadrangle, Land of Pinchouts, Jemez and Zia Rodríguez-Castañeda, J.L., 2002, Tectónica Cretácica y Ter- and Gabites, J.E., 2008, Age and paleotectonic set- Indian reservations, New Mexico, in Kues, B.S., Kel- ciaria en la margen suroeste del alto de Cananea, So- ting of volcanogenic massive sulfide deposits in the ley, S.A., and Lueth, V.W., eds., Geology of the Jemez nora norte-central [Ph.D. thesis]: Mexico City, Mexico, Guerrero terrane of central Mexico: Constraints from Region II: New Mexico Geological Society Guidebook Universidad Nacional Autonoma de México, 217 p. U-Pb age and Pb isotope studies: Economic Geology, 58, p. 188–194. Saleeby, J.B., 2003, Segmentation of the Laramide slab– v. 103, p. 117–140, https://doi​.org/10.2113/gsecon- Page, W.R., Gray, F., Iriondo, A., Miggins, D., Blodgett, R.B., evidence from the southern Sierra Nevada region: Geo- geo.103.1.117. Maldonado, F., and Miller, R.J., 2010, Stratigraphy and logical Society of America Bulletin, v. 115, p. 655–668,

Geological Society of America Bulletin, v. 132, no. 11/12 2515

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021 Lawton et al.

https://doi.org/10.1130/0016-7606(2003)115​ <0655:SO insights for the tectonic evolution of the southwestern M.F., eds., Mesozoic Rocks of Southern Arizona and TLSF>2.0.CO;2. North America during the late Mesozoic: Geological Adjacent Areas: Tucson, Arizona, Arizona Geological Scott, R.W., 2014, A Cretaceous chronostratigraphic data- Society of America Bulletin, v. 119, p. 1052–1065, Society Digest 18, p. 335–363. base: Construction and applications [Notebooks on https://doi​.org/10.1130/B26016.1. Wetmore, P.H., Schmidt, K.L., Paterson, S.R., and Herzig, Geology]: Carnets de Géologie, v. 14, no. 2, p. 15–37, Thomas, W.A., 2011, Detrital-zircon geochronology and C., 2002, Tectonic implications for the along-strike https://doi​.org/10.4267/2042/53522. sedimentary provenance: Lithosphere, v. 3, p. 304–308, variation of the Peninsular Ranges batholith, south- Silver, L.T., and Chappell, B., 1988, The Peninsular Ranges https://doi​.org/10.1130/RF.L001.1. ern and Baja California: Geology, v. 30, p. 247–250, batholith: An insight into the Cordilleran batholiths of Thomson, C.N., and Girty, G.H., 1994, Early Cretaceous https://doi.org/10.1130/0091-7613(2002)030​ <0247:TI southwestern North America: Royal Society of Edin- intra-arc ductile strain in Triassic–Jurassic and Creta- FTAS>2.0.CO;2. burgh Transactions: Earth Science, v. 79, p. 105–121. ceous continental margin arc rocks, Peninsular Ranges, Wetmore, P.H., Herzig, C., Alsleben, H., Sutherland, M., Soreghan, M.J., 1998, Facies distributions within an ancient California: Tectonics, v. 13, p. 1108–1119, https://doi​ Schmidt, K.L., Schultz, P.W., and Paterson, S.R., 2003, asymmetric lake basin: The Apache Canyon Forma- .org/10.1029/94TC01649. Mesozoic tectonic evolution of the Peninsular Ranges tion, Bisbee basin, Arizona, in Pitman, J.K., and Car- Todd, V.R., Erskine, B.G., and Morton, D.M., 1988, Meta- of southern and Baja California, in Johnson, S.E., Pa- roll, A.R., eds., Modern and Ancient Lake Systems: Salt morphic and tectonic evolution of the northern Penin- terson, S.R., Fletcher, J.M., Girty, G.H., Kimbrough, Lake City, Utah, Utah Geological Association Guide- sular Ranges batholith, southern California, in Ernst, D.L., and Martin-Barajas, A., eds., Tectonic Evolution book 26, p. 163–190. W.G., ed., Metamorphism and Crustal Evolution of the of Northwestern Mexico and the Southwestern USA: Spencer, J.E., Richard, S.M., Gehrels, G.E., Gleason, J.D., Western United States; Rubey Volume VII: Englewood Geological Society of America Special Paper 374, and Dickinson, W.R., 2011, Age and tectonic setting of Cliffs, New Jersey, Prentice Hall, p. 894–937. p. 93–116, https://doi​.org/10.1130/0-8137-2374-4.93. the Mesozoic McCoy Mountains Formation in western Vedder, L.K., 1984, Stratigraphic relationships between the Xie, X., and Heller, P.L., 2009, Plate tectonics and basin sub- Arizona, USA: Geological Society of America Bul- Late Jurassic Canelo Hills volcanics and the Glance sidence history: Geological Society of America Bulle- letin, v. 123, p. 1258–1274, https://doi​.org/10.1130/ Conglomerate, southeastern Arizona [M.S. thesis]: tin, v. 121, p. 55–64, https://doi​.org/10.1130/B26398.1. B30206.1. Tucson, Arizona, University of Arizona, 129 p. Yonkee, W.A., and Weil, A.B., 2015, Tectonic evolution Steckler, M.S., and Watts, A.B., 1978, Subsidence of the Villaseñor, A.B., González-León, C.M., Lawton, T.F., and Ab- of the Sevier and Laramide belts within the North Atlantic-type continental margin off New York: Earth erhan, M., 2005, Upper Jurassic ammonites and bivalves American Cordillera orogenic system: Earth-Science and Planetary Science Letters, v. 41, p. 1–13, https:// from the Cucurpe Formation, Sonora (Mexico): Revista Reviews, v. 150, p. 531–593, https://doi​.org/10.1016/​ doi​.org/10.1016/0012-821X(78)90036-5. Mexicana de Ciencias Geológicas, v. 22, p. 65–87. j.earscirev.2015.08.001. Stern, R.J., and Dickinson, W.R., 2010, The Gulf of Mexico Walawender, M.J., Gastil, R.G., Clinkenbeard, J.P., McCor- Zeller, R.A., 1965, Stratigraphy of the Big Hatchet Moun- is a Jurassic backarc basin: Geosphere, v. 6, p. 739–754, mick, W.V., Eastman, B.G., Wernicke, R.S., Wardlaw, tains Area: New Mexico Bureau of Mines and Mineral https://doi​.org/10.1130/GES00585.1. M.S., Gunn, S.H., and Smith, B.M., 1990, Origin and Resources Memoir 16, 128 p. Stopka, C.J., 2017, U-Pb detrital zircon geochronology, sand- evolution of the zoned La Posta-type plutons, eastern Zeller, R.A. Jr., 1970, Geology of the Little Hatchet Moun- stone modal composition, and paleoflow trends from Up- Peninsular Ranges batholith, southern and Baja Cali- tains, Hidalgo and Grant Counties, New Mexico: New per Cretaceous (Cenomanian–Campanian) nonmarine fornia, in Anderson, J.L., ed., The Nature and Origin Mexico Bureau of Mines and Mineral Resources Bul- strata in southern New Mexico [M.S. thesis]: Las Cruces, of Cordilleran Magmatism: Geological Society of letin 96, 22 p. New Mexico, New Mexico State University, 269 p. America Memoir 174, p. 1–18, https://doi​.org/10.1130/ Stoyanow, A., 1949, Lower Cretaceous stratigraphy of south- MEM174-p1. Science Editor: Rob Strachan eastern Arizona: Geological Society of America Mem- Walker, T.R., 1984, Diagenetic albitization of potassium Associate Editor: Emily Finzel oir 38, 169 p., https://doi​.org/10.1130/MEM38. feldspar in arkosic sandstones: Journal of Sedimentary Manuscript Received 19 July 2019 Talavera-Mendoza, O., Ruiz, J., Gehrels, G.E., Valencia, Petrology, v. 54, p. 3–16. Revised Manuscript Received 22 December 2019 V.A., and Centeno-García, E., 2007, Detrital zircon Warzeski, E.R., 1987. Revised stratigraphy of the Mural Manuscript Accepted 19 February 2020 U/Pb geochronology of southern Guerrero and west- Limestone: A Lower Cretaceous carbonate shelf in ern Mixteca arc successions (southern Mexico): New Arizona and Sonora, in Dickinson, W.R., and Klute, Printed in the USA

Erratum

ERRATUM: Transition from Late Jurassic rifting to middle Cretaceous dynamic foreland, southwestern U.S. and northwestern Mexico Timothy F. Lawton, Jeffrey M. Amato, Sarah E.K. Machin, John C. Gilbert, and Spencer G. Lucas

ORIGINAL ARTICLE: 2020, https://doi.org/10.1130/B35433.1. First published 9 April 2020

ERRATUM PUBLICATION: 2020. First published 15 July 2020

In Figure 1, “Fig. 11 Fence diagram” should instead read “Fig. 12 Fence diagram.”

2516 Geological Society of America Bulletin, v. 132, no. 11/12

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/132/11-12/2489/5206817/2489.pdf by guest on 25 September 2021