PII: S0895-9811(01) - Centro de Geociencias ::.. UNAM

Journal of South American Earth Sciences 14 (2001) 475±504
www.elsevier.com/locate/jsames
Tectonic setting and sandstone petrofacies of the Bisbee basin
(USA±Mexico)
William R. Dickinson a,*, Timothy F. Lawton b
a
b
Department of Geosciences, Box 210077, University of Arizona, Tucson, AZ 85721, USA
Department of Geological Sciences, Box 3AB, New Mexico State University, Las Cruces, NM 88003, USA
Received 1 September 2000; revised 1 February 2001; accepted 1 May 2001
Abstract
The Late Jurassic to mid-Cretaceous Bisbee basin spanning the USA±Mexico border was part of the Border rift system, which extended
into the continental block from the Gulf of Mexico in response to Cordilleran slab rollback. Rift initiation was marked by eruption of a
bimodal mid-Jurassic volcanic assemblage succeeding Early to Middle Jurassic arc volcanism. The core of the Bisbee basin is delineated by
Upper Jurassic marine strata of limited extent, and by more widespread syntectonic conglomerate of Late Jurassic to Early Cretaceous age
derived from intrabasinal fault blocks of synrift paleotopography. Subsequent thermotectonic subsidence of the rift belt induced Early
Cretaceous marine and nonmarine sedimentation spreading from the basin core to its ¯anks, with peak Aptian±Albian transgression marked
by deposition of platformal limestone. Late Cretaceous to Paleogene Laramide deformation disrupted the Bisbee deposystems. Bisbee
sandstones include arkosic petrofacies derived from intrabasinal fault blocks, lithic (volcaniclastic) petrofacies derived principally from the
coeval Alisitos arc to the southwest, subquartzose petrofacies derived mainly from the Mogollon paleohighland forming the northern rift
shoulder, and quartzose petrofacies also derived from the rift shoulder. The joint association of disparate petrofacies re¯ects the unusual
geotectonic setting of the Border rift belt. q 2001 Elsevier Science Ltd. All rights reserved.
Keywords: USA±Mexico border; Petrofacies; Bisbee deposystem
1. Introduction
The late Mesozoic Bisbee basin of southeastern Arizona,
southwestern New Mexico, and adjacent parts of Sonora and
Chihuahua is a unique tectonic element of the North American Cordillera. Local studies by geologists working on both
sides of the international border over the past two decades
allow an improved appraisal of basin setting and internal
anatomy.
The Bisbee basin developed as a result of mid-Jurassic
intracontinental rifting (Bilodeau, 1979, 1982), which
produced the Border rift system extending toward the northwest into continental crust from coeval oceanic crust in the
Gulf of Mexico (Lawton and Dickinson, 1999). We infer
that extensional deformation, which penetrated into the
continental block for a distance of ,1750 km from the
Rio Grande embayment at the Gulf margin, was initiated
by rollback of a subducted slab in the mantle beneath the
Cordilleran continental arc (Lawton and McMillan, 1999;
Dickinson and Lawton, 1999, 2001). This geodynamic
* Corresponding author. Tel./fax: 11-520-299-5220.
E-mail address: wrdickin@geo.arizona.edu (W.R. Dickinson).
explanation for the rifting event is speculative, but the existence of a rift belt including the Bisbee basin is observational. The Jurassic rift trend followed, however, a backarc
path roughly parallel to Jurassic arc volcanism in central
Mexico as far to the northwest as the eastern Bisbee
basin, and angled obliquely into the Cordilleran arc belt as
an intra-arc or post-arc feature in the western Bisbee basin
and farther to the northwest. The Bisbee basin, with overall
dimensions of 300 km £ 400 km, occupies the central
segment of the resulting paleorift connection between
Cordilleran and Caribbean realms (Dickinson et al., 1986).
Stratigraphic relations along the trend of the Border rift
system preserve the sedimentary record of progressive
marine invasion into the continent from the Gulf of Mexico.
Sandstones of the Bisbee basin display unfamiliar petrofacies associations, re¯ecting the unusual geotectonic
setting of the basin (Klute, 1987, 1991) and shedding light
on provenance relations within and surrounding the Border
rift system. After outlining the regional and subregional
tectonism and sedimentation associated with or related to
the Bisbee basin, we discuss Bisbee petrofacies as a contribution to sedimentary tectonics. We also brie¯y treat the
contrasting petrofacies of synorogenic post-Bisbee
0895-9811/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0895-981 1(01)00046-3
476
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 1. Regional tectonic setting of Bisbee and associated basins of the USA±Mexico border region. Bisbee core basin and Bisbee ¯ank basin from Fig. 3;
asterisks ( p ) denote exposures of Upper Jurassic marine strata along rifted keel of Bisbee basin. Extent of McCoy basin from Fig. 5. Con®guration of basins
and platforms in northeastern Mexico adapted after DeFord (1969), Smith (1970, 1981), Pingitore et al. (1983), Wilson et al. (1984), Cantu Chapa et al. (1985),
Padilla y SaÂnchez (1986), Wilson (1990), Zwanziger (1992) and Wilson and Ward (1993). CTB is Chihuahua tectonic belt (Laramide) ¯anking deformed
Chihuahua trough, and IC is Jurassic±Cretaceous Isla del Cuervo at the south end of Chihuahua trough. Jurassic islands fringing Gulf of Mexico margin after
MoraÂn-Zenteno (1994) for positions, and GoÈtte and Michalzik (1992) for names; Late Jurassic±Early Cretaceous La Mula island (LM) of Sabinas basin after
Jones et al. (1984). Aptian Cupido±Sligo and Albian Stuart City reef trends after Lehmann et al. (1999, 2000) and Wilson (1999) in Mexico, and Winker and
Buf¯er (1988) in Texas. North edge of Bisbee basin (limit of rift trough) modi®ed after Mack (1987a,b) and Dickinson et al. (1989). Cities: B Ð Brownsville;
Cb Ð Caborca; Ch Ð Ciudad Chihuahua; EP Ð El Paso; H Ð Hermosillo; L Ð Laredo; LC Ð Las Cruces; M Ð Monterrey; P Ð Phoenix; S Ð Saltillo;
To Ð Torreon; Tu Ð Tucson.
sandstones that were deposited within Late Cretaceous to
Paleogene Laramide sedimentary basins superposed upon
Upper Jurassic to mid-Cretaceous strata of the Bisbee basin.
For geochronology and chronostratigraphy, we follow the
timescales of Gradstein et al. (1994) for the Mesozoic and
Berggren et al. (1995) for the Cenozoic.
2. Regional tectonic setting
Tectonic trends delineating the keel of the Bisbee basin
extend westward to de®ne the ¯anks of the narrow McCoy
basin athwart the Colorado River, and merge eastward with
structures controlling basins that extend through northeastern Mexico to the ¯ank of the Gulf of Mexico oceanic basin.
2.1. Mexican platforms and basins
Pre-mid-Mesozoic basement of northeastern Mexico was
fragmented into multiple platforms and intervening basins
(Fig. 1) by extensional tectonism associated with Jurassic
opening of the oceanic Gulf of Mexico by sea¯oor spreading
between Texas and Yucatan beginning in Middle Jurassic
(Callovian) time (Marton and Buf¯er, 1994; Dickinson and
Lawton, 1999, 2001). The structural boundaries between
basins and platforms were reactivated or overprinted by
latest Cretaceous (Campanian±Maastrichtian) to early
Paleogene (Paleocene±Eocene) Laramide contractional
deformation. The late Mesozoic platforms and basins developed above two conjoined crustal blocks juxtaposed across
the Ouachita±Marathon suture (Fig. 1) in Early Permian
time: (1) Laurentian crust of the North American craton
and (2) Gondwanan crustal elements of eastern Mexico.
To the southwest of its exposed segment in west Texas,
the Ouachita±Marathon suture is masked beneath sedimentary cover of the Laramide Tornillo basin and the southern
end of the pre-Laramide Chihuahua trough, but passes
southward between the Aldama and Coahuila platforms
(Fig. 1).
The late Mesozoic array of platforms and basins in northeastern Mexico was inundated progressively from southeast
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
477
Fig. 2. Representative stratigraphic columns (central datum at Aptian±Albian boundary) showing regional WNW±ESE lithofacies gradient within Jurassic±
Cretaceous strata of the Border rift system (see Fig. 1 for geographic relations): A Ð McCoy basin after Harding and Coney (1985), Stone et al. (1987) and
Tosdal and Stone (1994); B±C±D Ð Bisbee basin (see Fig. 3 for distribution of facies belts), including western (nonmarine) facies tract (B) after Dickinson et
al. (1986, 1987) and central (mixed marine±nonmarine) facies tract (C±D) in southeast Arizona (C) after Dickinson et al. (1986, 1987) and lute (1991), and in
southwest New Mexico (D) after Mack et al. (1986), Lawton and Harrigan (1997, 1998) and Lucas and Lawton (2000); E Ð Chihuahua trough after CoÂrdoba
(1969), CoÂrdoba et al. (1970), DeFord and Haenggi (1970), Eaton et al. (1983), Araujo Mendieta and Casar GonzaÂlez (1987), Cantu Chapa (1993), MonrealSaavedra (1993) and Monreal and Longoria (1999); F Ð Sabinas basin after Cantu Chapa et al. (1985), Cuevas Leree (1985) and McFarlan and Menes (1991).
Basaltic volcanic and volcaniclastic strata (denoted `v') are interstrati®ed with Upper Jurassic marine strata within Martyr window of Chiricahua Mountains
(Fig. 4) in Arizona near New Mexico line (Lawton and Olmstead, 1995). Double-headed arrow denotes local occurrences of Albian±Cenomanian marginalmarine facies at top of Bisbee Group, and asterisks denote Aptian±Albian marginal-marine interval within otherwise nonmarine Shellenberger Canyon
Formation. La Gloria and Las Vigas Formations also include marginal-marine intervals. Key biostratigraphic boundaries (inferred where dashed or queried):
Cret/Jur Ð Cretaceous±Jurassic; Apt/Neoc Ð Aptian±Neocomian; Alb/Apt Ð Albian±Aptian; Ceno/Alb Ð Cenomanian±Albian (Upper±Lower Cretaceous boundary); LAR (unconformity or hachured line) Ð base of Laramide (Campanian±Maastrichtian) strata.
to northwest as the subsiding structural ¯ank of the Gulf of
Mexico was ¯ooded by marine waters (Padilla y SaÂnchez,
1986; Winker and Buf¯er, 1988; Cantu Chapa, 1998). Late
Jurassic transgression advanced up the Sabinas basin northeast of the Coahuila platform during Oxfordian time and up
the Chihuahua trough to the northwest during Kimmeridgian time (Salvador, 1987, 1991; Zwanziger, 1992; CantuÂ
Chapa, 1993). Transgression of the Mar Mexicano (Eguiluz
de AntunÄano and Campa Uranga, 1982; Eguiluz de AntunÄano, 1985) of central Mexico advanced simultaneously
along the western ¯ank of the Coahuila platform from
Oxfordian into Kimmeridgian time (Contreras Montero et
al., 1988). Successive positions of the Aptian Cupido±Sligo
and Albian Stuart City reef trends (Fig. 1), marking the shelf
edges of platformal carbonate systems, record the retrogradational landward shift of Early Cretaceous depositional
systems prior to widespread Late Cretaceous overlap of
platforms and intervening basins.
2.1.1. Chihuahua trough Ð Sabinas basin relations
The ¯oors of the Sabinas basin and the Chihuahua trough
were partly isolated from the expanding Gulf of Mexico and
Mar Mexicano, respectively, by subdued paleotectonic
elements that restricted access of marine waters, and both
basin keels are delineated by the subsurface distribution of
evaporites formed during intervals when water bodies were
ponded during basin evolution (DeFord and Haenggi, 1970;
GoÈtte and Michalzik, 1992). Late Jurassic islands screening
the Sabinas basin from the opening Gulf of Mexico (Fig. 1)
were inundated by the beginning of Cretaceous time
(McFarlan and Menes, 1991). The south end of the Chihuahua trough was partially separated from the Mar Mexicano
to the south, however, by a positive paleogeographic
element termed Isla del Cuervo located between the
Coahuila and Aldama platforms (Fig. 1) throughout Late
Jurassic and earliest Cretaceous (Neocomian) time (Zwanziger, 1992).
The Coahuila platform and nearby La Mula island (Fig.
1), as well as the Isla del Cuervo, were overtopped by rising
marine waters in late Early Cretaceous (Aptian) time (Jones
et al., 1984; Zwanziger, 1992; Lehmann et al., 1999). The
northeastern ¯ank of the Coahuila platform is delineated by
the San Marcos fault (McKee et al., 1990), which was recurrently active during Late Jurassic and Early Cretaceous time
478
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
(McKee et al., 1984). Before the Coahuila platform was
submerged, clastic wedges of Upper Jurassic to Lower
Cretaceous strata were shed across the fault trend into the
Sabinas basin on the northeast (Smith, 1981; McKee et al.,
1984; Wilson et al., 1984), and also off the Coahuila platform into eastern Mexico to the southeast (Michalzik and
Schumann, 1994). Succeeding Albian strata, dominantly
carbonate platforms and associated reef-¯ank basinal
sequences, are approximately as thick (600±750 m) over
the Coahuila platform as within the adjacent Chihuahua
trough and Sabinas basin (Lehmann et al., 1999). The top
of the Aldama platform west of the Chihuahua trough was
still emergent during all or much of Aptian time, but was
¯ooded during Albian time (Monreal-Saavedra, 1993) when
the Diablo platform east of the Chihuahua trough was also
overtopped by sediment (DeFord, 1969).
At times during late Mesozoic basin evolution when the
interbasin sill between the Sabinas basin and the Chihuahua
trough (Fig. 1) was paleogeographically emergent, receiving nonmarine sediment, the Coahuila platform has been
described as the tip of the Coahuila peninsula, but the
same tectonic feature has been termed the Coahuila island
during times when the interbasin sill was submerged to
receive marine sediment (Smith, 1981). Similarly, the Sabinas basin has been designated paleogeographically as the
Sabinas gulf at times when the interbasin sill was emergent
(GoÈtte and Michalzik, 1992). The two basins jointly de®ne a
paleotectonic linkage along the trend of the Border rift belt
between the Gulf of Mexico and the Bisbee basin, and the
keel of the Bisbee basin was invaded by a Late Jurassic
marine incursion (Fig. 1) from the northern end of the
Chihuahua trough (Lawton and Olmstead, 1995; Lucas
and Lawton, 2000; Lucas et al., 2001). The Border rift
belt was separated structurally from the Mar Mexicano to
the southwest by ¯anking platforms and intervening sills
(Fig. 1).
In both the Sabinas basin and Chihuahua trough, complex
Late Jurassic and Early Cretaceous depositional systems
included nonmarine or marginal-marine to shallow-marine
clastic strata, and shelfal to basinal carbonate strata including Aptian±Albian reef complexes correlative with a prominent limestone interval within the Bisbee basin (Fig. 2).
Overlying Upper Cretaceous marine shales and limestones
of more regional facies tracts cap the older basinal successions in both cases. Subsidence analysis of the Sabinas basin
implies a stretching factor b of 1.6±1.8 (Cuevas Leree,
1985), and the broadly comparable thicknesses of Late
Jurassic to mid-Cretaceous strata within the Chihuahua
trough and along the deepest keel of the Bisbee basin
(Fig. 2) suggest crustal extension of similar magnitude.
2.1.2. Laramide tectonism
The eastern ¯ank of the northern Chihuahua trough was
severely deformed, as the Chihuahua tectonic belt (Fig. 1),
by Laramide deformation that everted the basin ®ll and
thrust it eastward over the ¯ank of the Diablo platform
(Hennings, 1994). The Laramide Tornillo basin (Fig. 1)
developed in the foreland of the thrust system (Lehman,
1986, 1991). Lower Cretaceous strata of the Chihuahua
trough change facies eastward and thin rapidly by onlap
of a pre-mid-Mesozoic substratum along the ¯ank of the
Diablo platform (Gries and Haenggi, 1970; Underwood,
1980; Drewes and Dyer, 1993). Although structural telescoping along the basin margin has accentuated the lateral
stratigraphic gradient, Lower Cretaceous strata exposed
where the structural front crosses the Rio Grande thin
from 3150 m on the southwest to only 250 m on the northeast, within a lateral span of just 60 km (Amsbury and
Reaser, 1988). By contrast, the Sabinas basin, lying farther
east with respect to the evolving Laramide structural front,
and roughly along Late Cretaceous tectonic strike from the
Tornillo basin (Fig. 1), is capped conformably by a thick
succession of latest Cretaceous (Campanian±Maastrichtian)
clastic strata (Sohl et al., 1991). This dominantly shaly Laramide succession is analogous to the prodeltaic Parras Shale,
which underlies the deltaic Difunta Group deposited south
of the Coahuila platform, and was likewise derived from
growing Laramide highlands farther west (McBride et al.,
1974). The area of the Bisbee basin lay within the belt of
Laramide deformation.
2.2. Chihuahua trough Ð Bisbee basin relations
Near the USA±Mexico border, Lower Cretaceous strata
at the northern end of the Chihuahua trough pass gradationally to the north and west into sequences of Bisbee Group
Fig. 3. Lateral extent and internal compartments of Bisbee basin (USA±Mexico). See Fig. 1 for regional setting. Black areas denote Bisbee Group exposures
(nearby outcrops grouped for clarity). Stippled areas denote mid-Cretaceous marine to marginal-marine strata (PF Ð Pinkard Formation; BF Ð Beartooth
Formation; SF Ð Sarten Formation) onlapping Bisbee rift shoulder marked by Mogollon paleohighland (eastern extent termed Burro uplift in New Mexico).
Lines with arrows show trends of panel diagrams (Fig. 4), with numbers at arrowheads indicating general locations of control columns (1±6, Fig. 4 top; 7±12,
Fig. 4 bottom). Locality key: A Ð Arivaca area; BH Ð Big Hatchet Mtns.; CCC Ð Cerros Cabeza Colgada; CdO Ð Cerro de Oro area; CeM Ð Cerro
Mayo; CH Ð Canelo Hills; CM(W) Ð Chiricahua Mtns. (Martyr window); CM(R) Ð Chiricahua Mtns. (Rucker Canyon horst); DC Ð Dos Cabezas Mtns.;
DM Ð Dragoon Mtns.; EM Ð Empire Mtns.; EP Ð East Potrillo Mtns.; GM Ð Galiuro Mtns.; HM Ð Huachuca Mtns.; LH Ð Little Hatchet Mtns.; LM Ð
Lone Mtn. (Sierra San Jose); MM Ð Mule Mtns.; PdP Ð Planchas de Plata area; PaM Ð Patagonia Mtns.; PeM Ð Pedregosa Mtns.; PiM Ð Pajarito Mtns.;
PoM Ð Peloncillo Mtns.; SAn Ð Sierra Anibacachi; SAz Ð Sierra Azul; SC/R Ð Santa Catalina/Rincon Mtns.; SdC Ð Sierra del Caloso; SEC Ð Sierra El
Chanate; SP Ð Sierra Perilla; SR Ð Sierra Rica; SRM Ð Santa Rita Mtns.; T Ð Tuape area; TM Ð Tucson Mtns.; VM Ð Victorio Mtns.; WM Ð
Whetstone Mtns. Towns and hamlets (italics): An Ð Animas; AP Ð Agua Prieta; Ar Ð Arizpe; At Ð Altar; Ba Ð Bavispe; Be Ð Benson; Ca Ð Cananea;
Co Ð Columbus; Cu Ð Cucurpe; De Ð Deming; Do Ð Douglas; H Ð Hachita; L Ð Lordsburg; Na Ð Nacozari; No Ð Nogales; SA Ð Santa Ana; Sa Ð
Sasabe; Se Ð Sells; SV Ð Sierra Vista; To Ð Tombstone; Tu Ð Tucson; W Ð Willcox.
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
479
480
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
typical of the eastern end of the Bisbee basin (Mack et al.,
1986; Brown and Dyer, 1987; Seager and Mack, 1998;
Mack et al., 1998). In southeasternmost Arizona and southwesternmost New Mexico, the Bisbee Group in the core of
the Bisbee basin displays an internal stratigraphy that has
two key hallmarks (Fig. 2): (1) a markedly diachronous
(Grijalva-Noriega, 1996) basal conglomerate, the Glance
Conglomerate of Arizona and Sonora, and (2) a marine
limestone interval, of Aptian±Albian age, intercalated
midway through overlying basin ®ll that is otherwise
composed of nonmarine to marginal-marine clastic strata.
The limestone-rich medial deposystem, termed the Mural
Limestone in Arizona and adjacent Sonora and the U-Bar
Formation in New Mexico, marks the maximal marine
transgression into the Bisbee basin (Hayes, 1970). A lateral
change in stratigraphic nomenclature is traditionally placed
at the state boundary between Arizona and New Mexico
(Ferguson, 1987), but some workers have carried the
Arizona terminology across the state line (Drewes and Thorman, 1980a,b; Drewes, 1986), and more local formational
names are still used in some areas near the state line
(Bayona and Lawton, 2000). A less-prominent transgressive
phase of sedimentation near the Albian±Cenomanian
boundary is represented by marine to littoral deposits locally
preserved at the top of the Bisbee Group (Fig. 2).
Along the keel of the Bisbee basin, extending east-west
across southeastern Arizona and the `boot-heel' of southwesternmost New Mexico (Bilodeau and Lindberg, 1983),
the basal Glance Conglomerate is a syntectonic deposit
(Bilodeau, 1978) varying locally in thickness by orders of
magnitude (20±2000 m). The thickest conglomerate is
preserved within relict grabens and half-grabens, with thinner successions overtopping adjacent horsts and tilt blocks.
The oldest Glance sections underlie fossiliferous Upper
Jurassic beds (Lawton and Olmstead, 1995; Olmstead and
Young, 2000), but younger Glance intervals are separated
from Aptian±Albian limestones of the Bisbee Group by
abbreviated sections of shale-rich Lower Cretaceous strata
that are locally only 10±20 m thick. The Glance conglomeratic interval, which locally intertongues with overlying
Bisbee strata, is accordingly inferred to range in age from
Late Jurassic to Early Cretaceous.
We interpret depocenters containing either thick Glance
Conglomerate or overlying marine Jurassic beds, or both, as
local fault-controlled extensional depressions formed by
mid-Jurassic rifting of an underlying Jurassic arc assemblage or nearby basement during slab rollback beneath a
dormant magmatic arc. Volcanic rocks of a bimodal igneous
suite associated with arc rifting locally underlie, and in
places are intercalated with, the Glance Conglomerate or
overlying Upper Jurassic marine strata (Krebs and Ruiz,
1987; Lawton and McMillan, 1999). Younger strata of the
Bisbee Group accumulated during thermotectonic subsidence of the rift belt and its margins over a time span
extending to the middle of the Cretaceous. The Mogollon
paleohighland of central Arizona, and its southeastern
extension along the Burro uplift of New Mexico, formed a
high-standing rift shoulder ¯anking the Bisbee basin on the
north (Bilodeau, 1986; Dickinson et al., 1989). The rift
shoulder lay along Early Cretaceous tectonic strike with
the Diablo and Burro platforms northeast of the Chihuahua
trough and Sabinas basin, respectively (Fig. 1).
2.3. Bisbee core and ¯ank basins
Southward across Sonora from the keel of the Bisbee
basin where syntectonic stratal assemblages are prominent,
characteristic post-Glance formations of the Bisbee Group
are widespread, including the distinctive Mural Limestone
(GonzaÂlez-LeoÂn and Jacques-Ayala, 1990; GrijalvaNoriega, 1991; Monreal, 1995), but overlie uniformly thin
intervals (,100 m) of Glance Conglomerate (Monreal et al.,
1994). We term this part of the basin the `Bisbee ¯ank
basin,' as opposed to the `Bisbee core basin' farther north
(Figs. 1 and 3). Along the western fringe of the Bisbee ¯ank
basin, the Mural Limestone grades into marginal-marine
clastic strata of the Arroyo SaÂsabe Formation (JacquesAyala, 1989) derived from volcanic sources that probably
lay within the Aptian±Albian Alisitos arc of present-day
Baja California (Jacques-Ayala, 1995). Southward into
Sonora, post-Glance formations of the Bisbee Group overstep the area of Glance deposition to rest upon a contrasting
basal Bisbee lithologic unit of interbedded marine shale and
limestone (Cerro de Oro Formation) of Early Cretaceous
(Aptian) age (GonzaÂlez-LeoÂn and Jacques-Ayala, 1988;
Monreal, 1994; GonzaÂlez-LeoÂn and Lucas, 1995). In the
type section at the southern limit of its outcrops (Figs. 3
and 4), the Cerro de Oro Formation unconformably overlies
Proterozoic-Paleozoic sedimentary rocks, with no intervening Mesozoic strata present (GonzaÂlez-LeoÂn and JacquesAyala, 1988; GonzaÂlez-LeoÂn and Lucas, 1995).
We infer that the lateral lithologic transition (Monreal et
al., 1994; Monreal, 1995) from Glance Conglomerate to
Cerro de Oro Formation at the base of the Bisbee Group
heralded approach to the southern margin of the Bisbee
basin as the Cerro de Oro Formation onlapped pre-Bisbee
strata. The pre-Bisbee substratum beneath the southern rim
of the Bisbee ¯ank basin is part of the displaced Caborca
block (Stewart et al., 1990, 1997), which evidently formed a
sill separating the Bisbee basin from tectonic elements lying
farther to the south.
The extent and magnitude of rifting beneath the Bisbee
¯ank basin is uncertain, with contrasting interpretations
possible. On the one hand, a conglomeratic sequence underlying the Cerro de Oro Formation near Tuape (Fig. 3) in
north-central Sonora and a thick metaconglomerate succession exposed farther west near Altar (Fig. 3) may represent
early phases of Glance Conglomerate deposition (not shown
on Fig. 4) and mark the trend of a rift trough beneath the
Bisbee ¯ank basin subparallel to the rifted keel of the Bisbee
core basin. The prominence of felsic volcanic clasts in the
conglomerates at both Tuape and Altar is suggestive of at
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
481
Fig. 4. Summary panel diagrams of stratigraphic relationships within Bisbee Group: top, Bisbee core basin in Arizona and New Mexico (AZ/NM is state line);
bottom, Bisbee ¯ank basin in Sonora. Lateral spacing of control columns (Fig. 3) not to scale. Central datum at Aptian±Albian Mural/U-Bar marine limestone
interval (or equivalent). Unit designations: (a) Upper Jurassic marine strata (`v' denotes basaltic volcanic±volcaniclastic interbeds): Jmc Ð Crystal Cave Fm.
(Chiricahua Mtns.); Jmb Ð Broken Jug Fm. (Little Hatchet Mtns.). (b) Jurassic±Cretaceous Glance Conglomerate and marine Cretaceous equivalents (`v'
denotes ma®c and silicic volcanic interbeds): JKg Ð Glance Cg. (Kgc where inferred to be entirely Cretaceous); Kco Ð Cerro de Oro Fm. (c) Pre-mid-Aptian
(post-Glance but pre-Mural/U-Bar limestones) Lower Cretaceous strata: Kac Ð Apache Cyn. Fm. (lacustrine); Khf Ð Hell-to-Finish Fm. (includes Glanceequivalent conglomerate at base); Kma Ð Morita Fm.; Kwc Ð Willow Cyn. Fm. (d) Aptian±Albian marine limestone interval (and marginal-marine
equivalent): Kas Ð Arroyo SaÂsabe Fm.; Kmu Ð Mural Ls.; Kub Ð U-Bar Fm. (e) Post-mid-Albian (post-Mural/U-Bar limestones) Lower Cretaceous strata
(capped locally by mid-Cretaceous Albian±Cenomanian marine to marginal-marine deposits denoted by dashed line near top): Kci Ð Cintura Fm.; Kmo Ð
Mojado Fm.; Ksc Ð Shellenberger Canyon Fm. (basal part below Mural/U-Bar equivalent); Ktu Ð Turney Ranch Fm. Adapted after Mack et al. (1986),
Dickinson et al. (1986, 1989), Bilodeau et al. (1987), Mack (1987a,b), GonzaÂlez-LeoÂn and Jacques-Ayala (1988), Jacques-Ayala (1989, 1992a,b, 1995),
Monreal et al. (1994), Monreal (1995), GonzaÂlez-LeoÂn and Lucas (1995), Lawton and Olmstead (1995), Lawton and Harrigan (1997, 1998), Seager and Mack
(1998), McKee and Anderson (1998) and Lucas and Lawton (2000).
least local pre-Bisbee rifting, and accompanying quartzite
clasts imply high-standing exposures of pre-Mesozoic rocks
somewhere nearby. On the other hand, the lesser overall
thickness of the Bisbee Group in the Bisbee ¯ank basin in
comparison to its thickness in the Bisbee core basin, by a
factor of two (Fig. 4), suggests that initial rift extension and
subsequent thermotectonic subsidence of the Bisbee ¯ank
basin were less than for the Bisbee core basin. A signi®cant
fraction of the net subsidence within the Bisbee ¯ank basin
can probably be attributed to ¯exural loading of lithosphere
as thick sediment accumulated within the Bisbee core basin
to the north.
A summary of stratal relations near Tuape and Altar indicates the nature of interpretive ambiguities for the structural
482
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
condition of the substratum beneath the Bisbee ¯ank basin:
1. Near Tuape, the Cerro de Oro Formation (Monreal et al.,
1994; Monreal, 1995; Bacuchi Formation of RodrõÂguezCastanÄeda, 1991) is reported to rest gradationally (RodrõÂguez-CastanÄeda, 1988, 1990, 1991) on ,1000 m of
unfossiliferous clastic strata (Dos Naciones Formation
of RodrõÂguez-CastanÄeda, 1991), which include conglomeratic horizons that might represent inter®ngering lenses
of Glance Conglomerate. The undated strata rest gradationally in turn, however, upon marine Upper Jurassic
volcaniclastic strata of arc af®nity that have yielded
Oxfordian to middle Kimmeridgian ammonites (Rangin,
1977; Araujo Mendieta and Estavillo GonzaÂlez, 1987;
RodrõÂguez-CastanÄeda, 1991). Farther north near Cucurpe
(Fig. 3), correlative Upper Jurassic volcaniclastic strata
pass downward into an Early to Middle Jurassic arc
succession widespread to the west (Palafox et al.,
1992). Although the conglomeratic beds overlying the
arc assemblage near Tuape may represent a fragmentary
record of syn-Bisbee rifting, they can be regarded instead
as a nonmarine late Kimmeridgian and younger cap on an
unrifted arc edi®ce, which was in time onlapped progressively by the Lower Cretaceous Cerro de Oro Formation
as Bisbee sedimentation spread southward from the rifted
keel of the Bisbee core basin.
2. Near Altar, Nourse (2001) has reported that a metaconglomerate sequence ,2000 m thick and enclosing interstrati®ed volcanic rocks lie gradationally beneath
previously mapped Glance Conglomerate at the base of
the Bisbee Group, and regards the undated older succession as lower Glance Conglomerate. His conclusions
contrast markedly with previous interpretations inferring
that the metaconglomerates are Late Cretaceous in age
and interstrati®ed with Laramide volcanic rocks
(Jacques-Ayala et al., 1990). As metamorphism of the
strata is constrained (Nourse, 2001) to the interval 71±
51 Ma (Maastrichtian to early Eocene), available
geochronology cannot resolve the difference of opinion,
which should be addressed by closer attention to ®eld
relations (Nourse, 2001). Information on the geochemical character of the interstrati®ed volcanic rocks might
also permit a distinction to be drawn between rift-related
Glance and arc-related Laramide volcanism.
2.4. Lampazos shelf and slope
Farther to the southeast in Sonora, a succession of limestone and shale equivalent in age but not in lithology to the
Bisbee Group is exposed in erosional windows beneath
younger strata near and southeast of Lampazos (Palafox
and MartõÂnez, 1985; Bartolini and Herrera U, 1986; GonzaÂlez-LeoÂn, 1988; Scott and GonzaÂlez-LeoÂn, 1991; Minjarez
Sosa, 1991; Grijalva-Noriega, 1991; Baron-Szabo and
GonzaÂlez-LeÂon, 1999; Monreal and Longoria, 2000a). The
Aptian±Albian marine sequence represents carbonate shelf
facies, or platform and slope facies, but no depositional base
is exposed. We denote its depositional setting as the Lampazos shelf, inferred to lie along the northern ¯ank of the Mar
Mexicano west of the Aldama platform but south of the
Bisbee basin (Fig. 1). The Bisbee ¯ank basin is thereby
viewed as a sill-like region lying west of the Aldama platform and separating the Bisbee core basin from the Lampazos shelf (Fig. 1). Because depositional systems of the
Lampazos shelf and Bisbee basin were laterally contiguous,
the two distinctive stratal associations have been grouped
together within a more inclusive `Sonora basin' in past
usage (GonzaÂlez-LeoÂn and Jacques-Ayala, 1990), and both
have also been regarded as part of an expanded Bisbee basin
(GonzaÂlez-LeoÂn, 1994). The Espinazo del Diablo Formation
at Lampazos is lithologically similar to the Mural Limestone and apparently re¯ects progradation of an Aptian±
Albian carbonate platform from the Bisbee ¯ank basin
across the Lampazos shelf during maximum regional transgression.
The limestone-rich succession of the Lampazos shelf
southeast of the Bisbee basin displays stratal af®nities
with coeval sections in the Chihuahua trough (GonzaÂlezLeoÂn, 1988; Monreal, 1996; Mora Villalobos, 1997;
Monreal and Longoria, 2000a,b) but was seemingly separated from the latter by the intervening Aldama platform
(Valencia GoÂmez, 1994). In easternmost exposures near
Arivechi and Sahuaripa (GonzaÂlez-LeoÂn and JacquesAyala, 1990; Grijalva-Noriega, 1991), limestone and shale
of the Lampazos shelf-slope sequence overlie conglomeratic beds (`Conglomerado de Zarapuchi'), which are not
contiguous with the Glance Conglomerate but resemble
nonmarine to marginal-marine facies (Campbell, 1980) of
the broadly coeval Las Vigas Formation in the Chihuahua
trough. We infer that coarse Zarapuchi detritus was shed
westward from the high-standing Aldama platform over
the same Early Cretaceous (Neocomian) time frame during
which comparable Las Vigas conglomeratic beds were
deposited east of the Aldama platform.
Overall paleogeographic relations of the Bisbee basin
imply that marine ¯ooding proceeded across the Lampazos
shelf into the Bisbee ¯ank basin from the south, as well as
westward into the Bisbee core basin from the northern end
of the Chihuahua trough. As the western limit of the Aldama
platform is masked by continuous Cenozoic ignimbrite
cover of the Sierra Madre Occidental, the Bisbee ¯ank
basin may represent a subdued western continuation of the
Aldama platform. Maximum Aptian±Albian transgression
within the Bisbee basin coincided approximately with ®nal
inundation of the Aldama platform, suggesting that access
of marine waters into the Bisbee basin was facilitated by
foundering of the Aldama platform to the southeast. Shallow-marine Aptian±Albian strata of the Lampazos shelf
sequence may have close counterparts to the east in buried
coeval strata capping the Aldama platform beneath Tertiary
volcanic cover and linking the Lampazos shelf with upper
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
stratigraphic horizons of the Chihuahua trough. The thickness (,1775 m) of marine Aptian±Albian strata at Lampazos (Scott and GonzaÂlez-LeoÂn, 1991; Monreal and
Longoria, 2000a) is less than the thickness (,2800 m) of
comparable Aptian±Albian facies in the Chihuahua trough
(Fig. 2), but greater than the thickness (,1150 m) of coeval
strata (GonzaÂlez-LeoÂn and Lucas, 1995) at the southeastern
limit of the Bisbee ¯ank basin (Fig. 4).
2.5. Bisbee basin±McCoy basin relations
Along tectonic trend with the Bisbee basin within the
Border rift belt, but isolated by a wide expanse
(,125 km) in central Arizona where no Mesozoic strata
are exposed, the narrow McCoy basin of exclusively
nonmarine upper Mesozoic strata is exposed on both sides
of the Colorado River in southwestern Arizona and southeastern California (Fig. 1). A mid-Mesozoic uplift lying
north of the McCoy basin (Reynolds et al., 1989a) was
probably the western extension of the Mogollon paleohighland (Robison, 1980). The lower part of the basin ®ll,
termed in its entirety the McCoy Mountains Formation
(Fig. 2), occupies the same stratigraphic position as the
Bisbee Group farther east, and likewise rests depositionally
upon an assemblage of silicic volcanic rocks (`quartz
porphyry'). The latter, capping a magmatic arc assemblage,
may represent the distal limit of rift-related volcanism along
the Border rift belt at its farthest extent from the Gulf of
Mexico. The upper McCoy Mountains Formation, above an
intraformational unconformity, is a younger succession of
syntectonic strata associated with Late Cretaceous thrusting.
3. Bisbee basin stratotectonics
Exposures of the Bisbee Group occur as erosional
remnants and inliers distributed irregularly within `island
mountains' separated by interconnected alluviated basins
across the broken landscape of the block-faulted Basin
and Range province. The northwesterly `grain' displayed
by clusters of Bisbee exposures (Fig. 3) re¯ects the dominant trend of Cenozoic basin-range fault systems. Although
the outcrop pattern precludes tracing Bisbee strata continuously for more than a few kilometers along strike, because
most of the Mesozoic basin ®ll is masked beneath younger
cover or has been eroded, no signi®cant segment of the
Bisbee basin is entirely hidden in the subsurface. The
basin substratum, offset by pre-Bisbee or syn-Glance
normal faults, includes varied Paleozoic sedimentary,
Mesozoic volcanic, and Precambrian basement rocks (Bilodeau et al., 1987; Dickinson et al., 1987; GonzaÂlez-LeoÂn and
Jacques-Ayala, 1990). Strata directly overlying the Bisbee
Group range in age from Late Cretaceous through Cenozoic.
All pre-mid-Cenozoic units were deformed by Laramide
(Cretaceous±Paleogene) thrusts and folds. Younger extensional deformation included both mid-Cenozoic (Oligo-
483
cene±Miocene) detachment faulting and post-midMiocene block faulting (Dickinson, 1991).
The Bisbee basin was widely disrupted by post-midCretaceous deformation during Laramide contractional
tectonism, viewed as either retroarc or intra-arc as Laramide
igneous activity swept inland into the continental block in
response to subduction of oceanic lithosphere at progressively shallower angles beneath the Cordilleran region
(Dickinson, 1991). Laramide strata of Late Cretaceous to
Paleogene age were deposited either disconformably or with
angular unconformity above the Bisbee Group to thicknesses of typically 1500±2500 m in multiple local syntectonic basins commonly bounded or fragmented by
syndepositional thrusts (Seager and Mack, 1986; Hayes,
1987; Mack and Clemons, 1988; GonzaÂlez-LeoÂn and
Jacques-Ayala, 1988; Dickinson et al., 1989; Lawton and
Clemons, 1992; Lawton et al., 1993; Monreal et al., 1994;
GonzaÂlez-LeoÂn and Lawton, 1995; GonzaÂlez-LeoÂn et al.,
2000). Remnants of Bisbee basin ®ll are present beneath
Laramide cover within some mountain ranges, but they
also occur in other ranges where Laramide strata were not
deposited or have subsequently been removed by erosion.
Within the extent of the Bisbee core basin, syntectonic
Laramide sedimentation did not begin until Campanian or
Maastrichtian time, contemporaneous with the subregional
onset of Laramide magmatism (Dickinson, 1991). Within
the area of the Lampazos shelf (Minjarez Sosa et al., 1985;
Minjarez Sosa, 1991), and perhaps within the extent of the
Bisbee ¯ank basin as well (GonzaÂlez-LeoÂn et al., 1992),
post-Bisbee contractional deformation may have begun
earlier, in Turonian or even Cenomanian time. As both the
Bisbee ¯ank basin and the Lampazos shelf lay closer to the
continental margin of Mexico prior to Neogene opening of
the Gulf of California, retroarc thrusting and basin formation in advance of migratory arc magmatism may have in¯uenced their tectonic evolution before impacting the Bisbee
core basin.
3.1. Sub-Bisbee volcanic assemblages
Jurassic volcanic assemblages beneath and within the
southwestern part of the Bisbee basin record a transition
from arc to rift magmatism where the Border rift belt interacted with the Cordilleran magmatic arc.
3.1.1. Arc magmatism
Lower to Middle Jurassic volcanic rocks of southern
Arizona and northern Sonora represent the southeastern
prolongation of a mid-Mesozoic magmatic arc along the
Cordilleran margin (Busby-Spera, 1988; Tosdal et al.,
1989; Busby-Spera et al., 1990). Although discordance of
U±Pb data makes age interpretations a challenge, the time
span of arc magmatism in southern Arizona (Asmerom et
al., 1990; Riggs and Haxel, 1990; Riggs et al., 1993) ranged
from no later than the beginning of the Jurassic until at least
175 Ma (Middle Jurassic near the beginning of Bajocian
484
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
time). Sedimentological con®rmation of the age of the
Lower to Middle Jurassic volcanic assemblage is provided
by intercalated eolian quartzarenites that are correlative
with Jurassic erg deposits of the Colorado Plateau to the
north (Bilodeau and Keith, 1986; Riggs and Haxel, 1990;
Riggs et al., 1993). The delivery of eolian sand from sources
to the north into the arc terrane implies that the Mogollon
paleohighland was nonexistent or subdued prior to midJurassic time (Bilodeau and Keith, 1986), when it was
uplifted to form the rift shoulder of the Bisbee basin (Bilodeau, 1986; Lucas et al., 2001). The inland edge of the
Jurassic arc assemblage angles obliquely, northwest to
southeast, across the Bisbee basin (Fig. 3), leaving continental basement beneath the northeastern Bisbee basin in a
backarc position unaffected by arc magmatism prior to postBisbee Laramide events.
3.1.2. Rift magmatism
Before and during initial Bisbee Group sedimentation,
Jurassic arc magmatism of dominantly andesitic character
was succeeded by a bimodal phase of volcanism marking
the transition to tectonics of the Border rift system. PreGlance silicic ignimbrites were emplaced as intracaldera
bodies and out¯ow sheets of the Canelo Hills Volcanics
(Lipman and Hagstrum, 1992), erupted from calderas in
the Canelo Hills and Huachuca Mountains (Fig. 3). Waning
phases of the silicic volcanism also emplaced thin out¯ow
sheets of ignimbrite within the Glance Conglomerate of the
Canelo Hills (Bilodeau et al., 1987; Dickinson et al., 1987).
Farther west in the southern Santa Rita Mountains (Fig. 3),
felsic to intermediate volcanic and volcaniclastic rocks are
similarly interstrati®ed within Glance Conglomerate (Bilodeau, 1979; Bilodeau et al., 1987), mapped locally as the
Temporal and Bathtub formations (Drewes, 1971). Ma®c
alkalic to tholeiitic lavas were erupted in grabens and
half-grabens along the central keel of the Bisbee core
basin (Lawton and McMillan, 1999), where they are interstrati®ed with Glance Conglomerate in the Huachuca
Mountains and with Upper Jurassic marine strata that overlie the Glance Conglomerate in the Martyr window of the
Chiricahua Mountains (Fig. 3).
Silicic ignimbrites forming the Canelo Hills Volcanics
and interstrati®ed tuffs within the overlying Glance
Conglomerate display geochemical properties that are characteristic of extensional tectonic environments (Krebs and
Ruiz, 1987): (a) low chondrite-normalized La/Lu ratios
re¯ecting comparative enrichment in light rare-earth
elements (LREE); (b) a correspondingly ¯at curve of chondrite-normalized REE (rare-earth element) abundances
(especially for the LREE segment of the curve); and (c) a
strongly negative europium (Eu) anomaly. In these respects,
their geochemical signature is analogous to that of igneous
assemblages from extensional domains such as the modern
Rio Grande rift, the active east African rifts, the McDermitt
caldera and Coso volcanic ®eld of the Neogene Basin and
Range province, and backarc assemblages of the Scotia Sea
and Papua New Guinea. By contrast, magmatic arc assemblages commonly display strongly sloping REE curves (no
LREE enrichment) and more subdued Eu anomalies. Ta±
Th±Hf plots for the ignimbrites show af®nities with both arc
and backarc suites, as expected for a rifted-arc setting. The
ma®c lavas that are interstrati®ed with strata of the lower
Bisbee Group display e Nd values of 13± 1 5 (Lawton and
McMillan, 1999), indicative of asthenospheric sources and
fully compatible with eruption within an evolving rift belt.
We conclude that the bimodal volcanic assemblage present
immediately below and within basal horizons of the Bisbee
Group displays a petrologic character in keeping with the
inferred magmatic environment of an arc structure incipiently rifted by slab rollback.
3.1.3. Arc-rift transition
Age control for the transition from arc volcanism to rift
sedimentation is still imprecise. Available K±Ar ages
(Marvin et al., 1978) for the Canelo Hills Volcanics of the
rifted-arc assemblage below the Glance Conglomerate
range from 177 to 169 Ma (Bajocian), but individual
reported dates have high uncertainties (^6±8 Ma). An
axial pluton intruded beneath the resurgent dome of an
ignimbrite caldera in the Huachuca Mountains (Lipman
and Hagstrum, 1992) has yielded a K±Ar age of
167 ^ 6 Ma (Marvin et al., 1978), which is appropriately
near the younger end of the apparent age range for the
sub-Bisbee ignimbrites. Basaltic lavas in the Chiricahua
Mountains (Fig. 3) of Arizona both underlie and overlie
marine shales that have yielded Kimmeridgian (154±
151 Ma) ammonites (Lawton and Olmstead, 1995;
Olmstead and Young, 2000), and lie only 150 m conformably above fossiliferous Upper Jurassic strata in the Little
Hatchet Mountains (Fig. 3) of New Mexico (Lucas and
Lawton, 2000; Lucas et al., 2001). Ignimbrites interbedded
with Glance Conglomerate in the Canelo Hills (Kluth et al.,
1982) and the Santa Rita Mountains (Asmerom et al., 1990)
have yielded essentially identical Rb±Sr isochrons of
151 ^ 2 and 151 ^ 5 Ma (,Kimmeridgian±Tithonian
boundary), but the isochrons may not re¯ect eruptive age
because metasomatic alteration of the dated rocks is severe
(Krebs and Ruiz, 1987). The Rb±Sr ages are compatible,
however, with less precise K±Ar ages of 147±149 Ma
(Marvin et al., 1978), with uncertainties of ^6±11 Ma.
Interpretations are further clouded by the likelihood that
evolution of the Jurassic magmatic arc was in¯uenced by
intra-arc extension throughout much of its history (BusbySpera, 1988). In southeastern Arizona, the transition during
the approximate interval 175±170 Ma (Bajocian) from
dominantly andesitic volcanism associated with stratocones
to dominantly silicic volcanism associated with caldera
collapse (Asmerom et al., 1990; Riggs and Busby-Spera,
1990) can be viewed either as continued arc evolution or
as the onset of pre-Bisbee rifting. For example, the immense
Cobre Ridge caldera (,170 Ma) in the Pajarito Mountains
(Fig. 3) may have formed simply by volcano-tectonic
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
collapse of an arc edi®ce (Riggs and Busby-Spera, 1991),
but the in¯uence of regional extension on local caldera
collapse cannot in our view be excluded.
The youngest well-dated Jurassic granitic suite in southern Arizona was emplaced in the latter part of Middle Jurassic time during the interval 165±160 Ma (Tosdal et al.,
1989). We infer from the presence of Glance Conglomerate
underlying fossiliferous marine Kimmeridgian strata in the
Chiricahua Mountains, which lie northeast of the Jurassic
magmatic arc assemblage (Fig. 3), that the Border rift belt
had propagated westward, from a backarc position, to reach
the continental ¯ank of the Cordilleran magmatic arc shortly
thereafter, by approximately 160±155 Ma (Oxfordian). The
observed Rb±Sr isochrons of 151 Ma (Kimmeridgian±
Tithonian) obtained from ignimbrites intercalated within
the Glance Conglomerate lying stratigraphically above the
arc assemblage in southern Arizona are compatible with that
inference. In the Huachuca Mountains (Fig. 3), an erosional
surface at the base of the Glance Conglomerate truncates the
edges of the calderas from which the extensive ignimbrites
of the pre-Glance Canelo Hills Volcanics were erupted
(Vedder, 1984). This relationship indicates that Glance
depocenters were synrift tectonic features superimposed
across underlying volcanic structures. Inherited topography
along the volcano-tectonic rift belt may nevertheless have
in¯uenced Glance deposition locally (Lipman and
Hagstrum, 1992).
3.2. Basal Glance Conglomerate
In Arizona and Sonora, successions of conglomerate and
sedimentary breccia at the base of the Bisbee Group are
termed the Glance Conglomerate, but analogous strata in
New Mexico are generally treated as basal or inter®ngering
beds of the generally ®ner-grained Hell-to-Finish Formation. Locally in both New Mexico and Chihuahua, ,50 m
of basal `Glance Conglomerate' has been mapped separately from the latter formation (Brown and Dyer, 1987;
Drewes, 1991a). Lateral variations in thickness re¯ect
deposition of the Glance interval over a corrugated rift topography (Bilodeau and Lindberg, 1983; Mack et al., 1986;
Bilodeau et al., 1987; Lawton and Olmstead, 1995), with
thickest Glance accumulations con®ned to downdropped
keels of structurally controlled depressions formed as
grabens or half-grabens. Thinner Glance successions form
more extensive horizons of more consistent thickness where
gravel deposition was a transient phase of pediment-like or
strandline sedimentation as aggradation of sediment overtopped eroding horsts and tilt blocks (Dickinson et al., 1987,
1989).
In favorable instances, abrupt thickness changes de®ne or
closely constrain the positions of bounding paleofaults of
local half-grabens (Bilodeau, 1978, 1979, 1982; Sumpter,
1986; Bayona, 1998; Bayona and Lawton, 2000). In typical
cases, syn-Glance paleofaults strike west-northwest (Fig. 3),
with their south sides downdropped (Bilodeau et al., 1987).
485
This pervasive pattern of deformation suggests that the
Bisbee rift topography was largely an asymmetric array of
tilt blocks and half-grabens rather than symmetric horsts and
grabens. The prevailing structural geometry implies that
segments of the Bisbee basin ¯oor pulled systematically
downward and away from the continental interior, backtilting fault blocks toward the northern ¯ank of the basin along
the edge of the Mogollon paleohighland in the manner of
giant louvers. At the northern limit of Bisbee Group exposures, a basin-margin paleofault of the same character may
be exposed near the northern edge of the mid-Cenozoic
Catalina core complex (Janecke, 1987). Low-angle preLaramide normal faults in the Florida Mountains of New
Mexico may also record syn-Bisbee faulting near the northern edge of the Bisbee core basin (Amato, 2000). The rather
consistent thickness of Glance Conglomerate within the
Bisbee ¯ank basin suggests that the substratum south of
the Bisbee core basin was not disrupted to the same intricate
degree prior to onlap by the Bisbee Group.
The general absence of preserved Bisbee strata between
the Bisbee core basin and the Bisbee ¯ank basin within a
belt passing through Cananea (Fig. 3) has suggested the
presence of a `Cananea high' (Grijalva-Noriega, 1995)
separating the two segments of the Bisbee basin. McKee
and Anderson (1998) have argued that Bisbee sedimentation
in the nearby Sierra Azul (Fig. 3) involved syndepositional
downslope displacement of Mural Limestone blocks and
slabs off the Cananea high into a deep trough to the south.
Jacques-Ayala (1995) noted, however, that the stratigraphy
and lithology of the Bisbee Group in the Sierra Azul is
consistent with adjacent segments of the Bisbee basin, and
cogent reasons were outlined to discount the hypothesis of
local gravity sliding. The apparent continuity of post-Glance
formations of the Bisbee Group across the locus of the
inferred structural barrier (Fig. 4) suggests that the `Cananea
high' was not qualitatively different from other paleotopographic features of the syn-Glance rift morphology.
Glance depositional facies re¯ect aggradation of alluvial
fans and braidplains of varying lateral extent. Some coarse
alluvial-fan deposits enclose thick lenses of debrisavalanche megabreccia (Bilodeau, 1979; Bilodeau et al.,
1987; Dickinson et al., 1987), commonly mapped as `exotic
blocks' from their monolithologic character, but similar in
their internally shattered character to megaclasts in analogous deposits described from the Cenozoic Basin and Range
province (Yarnold and Lombard, 1989). Megabreccia
bodies, which are most common in the Empire Mountains,
Canelo Hills, and Huachuca Mountains (Bilodeau, 1978,
1979; Bilodeau et al., 1987), are either known or inferred
to occur near paleofault scarps.
3.2.1. Gravel clast assemblages
Clast assemblages uniformly re¯ect the nature of subjacent or nearby sub-Glance bedrock, with carbonate-clast
and quartzite-clast conglomerate derived from Paleozoic
sedimentary sequences, schist-clast and granitic-clast
486
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
conglomerate from Precambrian basement, and volcanicclast conglomerate from subjacent Jurassic volcanic assemblages. Varied mixed-clast conglomerates are also present,
re¯ecting the irregular distribution of sub-Bisbee bedrock
types (Bilodeau et al., 1987; GonzaÂlez-LeoÂn and JacquesAyala, 1990). Close study has shown that some Glance
successions display `inverted clast stratigraphy,' documenting progressive unroo®ng of an adjacent horst or tilt block,
with gravel derived from sedimentary cover overlain
conformably by gravel derived from basement beneath the
cover. Given the distribution of rock types in the basin
substratum, carbonate-clast Glance is generally dominant
to the northeast within the Bisbee basin, with volcanicclast Glance dominant to the southwest where Jurassic
volcanic assemblages underpin the Bisbee basin (Fig. 3).
Glance successions containing high proportions of clasts
derived from basement are more irregularly distributed in
areas where especially rugged synrift topography induced
deep syn-Glance erosion.
3.3. Post-Glance facies pattern
Sediment ®ll of the Bisbee core basin can be subdivided
into three lithofacies tracts (Fig. 3): (1) an eastern facies belt
composed entirely of shallow-marine strata, except for thin
basal conglomerate of locally nonmarine character, (2) a
central facies belt composed of mixed marine and nonmarine strata, and (3) a western facies belt composed exclusively of nonmarine strata, except for marginal-marine
deposits intercalated from the east near the eastern margin
of the belt at the Aptian±Albian time horizon represented by
transgressive marine limestone farther east (Klute, 1987).
The lateral transition from marine deposits on the east to
nonmarine deposits on the west re¯ects the regional sedimentary gradient from the marine Chihuahua trough toward
the nonmarine McCoy basin (Fig. 2). The westward facies
transition from mixed marine-nonmarine to nonmarine
deposits continues along trend southward from the Bisbee
core basin across the Bisbee ¯ank basin (Fig. 3).
3.3.1. Medial marine limestone interval
The most widespread marine horizon (Fig. 4), forming
the lithologically continuous Mural Limestone (Arizona±
Sonora) and U-Bar Formation (New Mexico), is well
bracketed biostratigraphically between mid-Aptian and
mid-Albian time (Scott, 1987; Warzeski, 1987; Scott and
Warzeski, 1993; Monreal, 1995; Rosales-DomõÂnguez et al.,
1995; Seager and Mack, 1998; Lucas, 2000; Lucas and
Estep, 2000b; Lucas et al., 2000a). Lateral continuity with
similar Aptian±Albian carbonate strata of the Chihuahua
trough and Lampazos shelf (Monreal-Saavedra, 1997) is
implied by regional paleogeography (Figs. 1 and 2). The
limestone interval is a complex depositional mosaic of
tidal and subtidal carbonate platform, shelf, and lagoonal
facies including prominent coralgal-rudistid patch reefs
(Scott, 1979; Roybal, 1981; Schreiber and Scott, 1987;
Monreal, 1994). Mixed carbonate±siliciclastic intervals
are widely distributed within the lower part of the carbonate
sequence (Mack et al., 1986), providing evidence for intermittent delivery of terrigenous detritus to the carbonate
province prior to the full development of reef-dotted platforms. The limestone interval is a generally progradational
stratal interval, with shelf and lagoonal facies overlain by
biostromal platform and reef facies (Lindberg, 1987; Klute,
1991). Marine ¯ooding of the Bisbee basin thereby
produced a continuous carbonate blanket that completed
the burial of rift topography throughout the marine eastern
and mixed central facies belts. Time-equivalent marginalmarine strata along the eastern edge of the nonmarine
western facies belt are represented by oyster-bearing beds
associated with estuarine or lagoonal deposits within the
Shellenberger Canyon Formation of the Whetstone Mountains (Fig. 3) in Arizona (Archibald, 1987), and by lagoonal
to strandline beds of the Arroyo SaÂsabe Formation in the
westernmost part of the Bisbee ¯ank basin in Sonora
(Jacques-Ayala, 1989, 1995).
3.3.2. Nonmarine and marginal-marine strata
Clastic strata both above and below the Aptian±Albian
limestone interval (Fig. 4) include a wide variety of local
facies representing shelf, lagoonal, tidal-¯at, estuarine, and
¯uvial-plain environments that evolved over time in
complex spatial patterns (Mack, 1987a; Klute, 1991;
Jacques-Ayala, 1992b,c). The overall intrabasinal paleogeography of the Bisbee core basin was a broad and low-lying
coastal plain prograding to the east or southeast toward the
marine environments of the eastern facies belt. Within the
Bisbee ¯ank basin of Sonora, commonly bimodal paleocurrent patterns suggest widespread in¯uence of tidal currents
on sedimentation (Jacques-Ayala and Potter, 1987; JacquesAyala, 1995), but overall northeasterly to southeasterly
sediment dispersal is inferred throughout Sonora (JacquesAyala, 1992a; GonzaÂlez-LeoÂn, 1994). Immediately above
and below the Mural/U-Bar limestones, shallow-marine
shelf or lagoonal deposits transitional to the marine carbonate complex are characteristic (Dickinson et al., 1987), but
farther up and down section nonmarine assemblages are
dominant, except in the exclusively marine eastern facies
belt. More rapid thermotectonic subsidence during Morita
deposition, prior to the transgression that deposited Aptian±
Albian Mural/U-Bar limestones, than during post-Mural/UBar Cintura/Mojado deposition is suggested by the prevalence of single-story ¯uvial channels encased in Morita
mudstones, as opposed to amalgamated multistory channel
complexes in Cintura/Mojado ¯uvial strata (GonzaÂlezLeoÂn, 1994).
Below the Mural/U-Bar limestone interval, paleocurrent
indicators and facies patterns within the Morita Formation
of Arizona-Sonora and the Hell-to-Finish Formation of New
Mexico suggest transport of detritus into the Bisbee basin
from both the northern and southwestern basin margins
(Mack et al., 1986; Klute, 1991; Jacques-Ayala, 1995).
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Above the Mural/U-Bar limestone interval, continuation of
the same sediment delivery systems is indicated for the
Cintura Formation of Arizona±Sonora, but ¯uvial strata of
the Mojado Formation in New Mexico display axial paleocurrents indicating sediment transport from west to east,
parallel to the trend of the Bisbee core basin (Mack et al.,
1986). Within the exclusively nonmarine facies belt farther
west in Arizona, ¯uvial paleocurrents are also dominantly
axial, trending southeasterly parallel to the trend of the
Bisbee core basin (Risley, 1987; Sumpter, 1986; Archibald,
1987).
3.3.3. Lacustrine strata
Fluvial facies tracts of the western nonmarine facies belt
are interrupted locally by intercalated lacustrine facies
deposited in lakes that were apparently ponded within residual half-graben depressions con®ned between tilt blocks
inherited from synrift topography. The Apache Canyon
Formation, representing a mixed carbonate-clastic lacustrine deposystem exposed in the Empire and Whetstone
Mountains (Fig. 3), thins from 600 m within 6 km of a
basin-bounding paleofault to only 200±250 m at distances
of 13±25 km up the paleoslope of the associated tilt block
downfaulted against the ¯ank of the basin (Soreghan, 1999).
The Empire-Whetstone lakebeds (Archibald, 1987) and
analogous lacustrine facies exposed farther west in the
Tucson Mountains (Risley, 1987) were in time buried
beneath ¯uviodeltaic deposits, as accumulating sediment
®lled the ponded lacustrine basins. The western nonmarine
facies belt can be viewed as a segment of the corrugated
¯oor of the Border rift system that never subsided below sea
level but had suf®cient internal tectonic relief to promote
development of local freshwater or saline lakes until sedimentation within the rift belt overwhelmed the synrift topography. West of the Tucson Mountains (Fig. 3), depositional
environments were probably entirely ¯uvial throughout
basin history for the Sand Wells Formation (Haxel et al.,
1980; Beikman et al., 1995), deposited on the reservation of
the Tohono O'Odham Nation near the preserved fringe of
the Bisbee basin.
3.3.4. Onlap of rift shoulder
Along the rift shoulder north of the Bisbee core basin,
mid-Cretaceous (Albian±Cenomanian) marine to marginalmarine deposits resting unconformably on a pre-Bisbee
substratum are locally preserved along the trend of the
Mogollon paleohighland (Dickinson, 1981; Dickinson et
al., 1986, 1989), termed the Burro uplift in New Mexico.
Sandy strata, typically resting on transgressive ravinement
surfaces, include the Pinkard Formation of Arizona (Molenaar, 1983), the Beartooth Formation (Chafetz, 1982) near
Silver City, and the Sarten Formation (Lucas et al., 1988) of
the Cookes Range farther east in New Mexico (Fig. 3). The
onlapping sandstones are overlain by marine shales that are
laterally contiguous with Upper Cretaceous marine strata of
the interior seaway that occupied the Rocky Mountain retro-
487
arc foreland basin (Molenaar, 1983; Mack et al., 1988;
Lucas and Estep, 1998a; Lucas et al., 2000b). Detailed biostratigraphy implies, however, that initial marine transgression of the Mogollon paleohighland was from the Bisbee
basin on the southwest (Lucas and Estep, 2000a).
3.3.5. Uppermost marine interval
Within the Bisbee basin, correlative marine and littoral
(foreshore) deposits of known or inferred Albian±Cenomanian age are present at or near the exposed stratigraphic top
of the Bisbee Group in both New Mexico (Mack et al.,
1986) and Arizona (Inman, 1987), and also near Arizpe in
Sonora (GonzaÂlez-LeoÂn and Jacques-Ayala, 1990; GonzaÂlez-LeoÂn and Lucas, 1995; Baron-Szabo and GonzaÂlezLeÂon, 1999). The marine to marginal-marine strata of the
uppermost Mojado Formation in the Bisbee basin of New
Mexico are laterally equivalent and lithologically comparable to the Beartooth and Sarten Formations onlapping the
Mogollon paleohighland, and the latter units can therefore
be regarded as members of the Mojado Formation (Lucas
and Estep, 1998b, 2000a). The presence of a marine-in¯uenced mid-Cretaceous stratigraphic interval marking the
close of Bisbee deposition suggests that the previously ®lled
Bisbee rift basin was incorporated into the proximal ¯ank of
the Rocky Mountain retroarc foreland basin just prior to
regional regression that carried the shoreline of the interior
seaway off to the northeast (Molenaar, 1983; Cobban and
Hook, 1984; Mack, 1987a; Mack et al., 1988; Lucas and
Lawton, 2000). The transient mid-Cretaceous return of
marine waters to the Bisbee basin probably resulted,
however, from global Albian±Cenomanian eustasy rather
than tectonic downwarping (Mack, 1987b).
3.4. Laramide deformational style
Laramide structures that disrupted the Bisbee basin have
been exposed by erosion at varying crustal levels (Keith and
Wilt, 1985, 1986). In the northeastern part of the basin, only
supracrustal structures are exposed to view, with internally
undeformed basement faulted against Laramide and older
sedimentary successions. In the southwestern part of the
basin, a record of ductile infracrustal deformation, involving
deep-seated plutonism and widespread metamorphism of
both basement and cover, is widely displayed (Haxel et
al., 1984; Tosdal et al., 1990; Jacques-Ayala et al., 1990).
The contrast in thermal history and crustal level of exposure
re¯ects the greater proximity of the southwestern domain to
the thermal core of the Cretaceous magmatic arc along the
continental margin (Reynolds et al., 1988; Barton and
Hanson, 1989; Gastil et al., 1992).
Two contrasting structural styles and patterns of deformational geometry have been proposed for Laramide tectonism
within the area of the Bisbee basin. On the one hand,
Drewes (1976, 1978, 1981, 1982, 1988, 1991b) has long
maintained that Laramide deformation was controlled by
translation of extensive subhorizontal thrust sheets, each
488
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 5. Con®guration of McCoy basin (California-Arizona). Bisbee-equivalent strata include lower McCoy Mountains Formation (black) and Winterhaven
Formation (ruled). Stippled areas denote post-Bisbee upper McCoy Mountains Formation. Outcrop distributions after Harding and Coney (1985), Haxel et al.
(1985), Spencer et al. (1985), Reynolds et al. (1986, 1989b, 1991), Richard et al. (1987, 1993), Stone and Pelka (1989), Sherrod et al. (1990), Richard (1992),
Richard and Spencer (1994) and Tosdal and Stone (1994). Locality key: BR Ð Black Rock Hills; CM Ð Coxcomb Mtns.; DR Ð Dome Rock Mtns.; GW Ð
Granite Wash Mtns.; LHa Ð Little Harquahala Mtns.; LHi Ð Livingston Hills; MM Ð McCoy Mtns.; NW Ð New Water Mtns.; PaM Ð Palen Mtns.; RH
Ð Red Hills; sPM Ð southern Plomosa Mtns.
composed of both basement and cover, which moved on the
order of 100 km from southwest to northeast. The subregional thrusts are viewed as components of a long-lived foreland tectonic system that evolved progressively from Late
Jurassic to Eocene time (Drewes, 1991b). On the other hand,
others argue that observed Laramide thrust faults are
entirely Late Cretaceous to Paleogene structures that ¯ank
basement-cored uplifts, analogous to structures of the
central Rocky Mountains, and root more locally into subjacent basement (Davis, 1979; Seager and Mack, 1986;
Krantz, 1989; Lawton, 2000). We favor the latter interpretation for the following reasons:
1. Contacts inferred to represent important segments of the
Hidalgo and Cochise thrusts (Drewes, 1980), the
supposed master structures of the subregional thrust
system (Drewes, 1976, 1978, 1981), are known instead
to be either unconformities (Keith and Barrett, 1976;
Dickinson et al., 1987; Lipman and Hagstrum, 1992) or
low-angle normal faults (Dickinson, 1984) and detachment faults (Dickinson, 1991) associated with mid-Cenozoic extensional deformation.
2. We can detect no areal repetitions of Bisbee lithofacies or
petrofacies patterns that could re¯ect systematic lateral
displacements of separate compartments of the Bisbee
basin on subregional thrust surfaces.
3. The concept of a progressively evolving foreland system
requires the Bisbee basin to be viewed as one of a linked
series of foreland basins (Drewes, 1991a, Figs. 31 and
36±40), shifting successively eastward as an integrated
Cordilleran orogeny unfolded. However, the following
observations argue against the picture of a foreland
setting for the Bisbee basin: (a) lack of any evidence
for syndepositional thrusting to the west or southwest,
(b) derivation of the basal Glance Conglomerate exclusively from intrabasinal sources rather than from a provenance outside the basin, (c) local stepwise thickening of
the Glance Conglomerate across syndepositional normal
faults, (d) overall thickening of the Glance Conglomerate
into the keel of the basin rather than toward a postulated
basin-¯ank thrust system, and (e) thickening of the
Bisbee Group as a whole toward the northeast from the
Bisbee ¯ank basin into the Bisbee core basin in a direction away from the continental-margin orogen.
4. McCoy basin stratotectonics
Stratal remnants of the McCoy Mountains Formation are
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 6. Internal stratigraphy of Bisbee-equivalent and Laramide-age strata
within McCoy Mountains Formation of McCoy basin (petrofacies abbreviations from Fig. 10). Overall thickness relations after Harding and Coney
(1985). Unconformity between lower and upper McCoy Mountains Formation after Tosdal and Stone (1994). Member correlations discussed in text.
`t' denotes horizon of ,80 Ma tuff bed.
exposed within an elongate east±west belt, pinched structurally between the north-vergent Mule Mountains thrust
system on the south and the south-vergent Maria fold-andthrust belt on the north, with segments of the McCoy basin
®ll incorporated into the latter (Fig. 5). Consequently, the
original width of the basin is unknown, but it probably
extended southward far enough to include the unfossiliferous Winterhaven Formation (Fig. 5), which is lithologically
similar to the lower part of the McCoy Mountains Formation and occurs in the same stratigraphic position (Haxel et
al., 1985). All exposures of the McCoy Mountains Formation are variably metamorphosed; at least weak foliation and
fracture cleavage are ubiquitous.
Basal beds of the McCoy Mountains Formation locally
rest gradationally (Fackler-Adams et al., 1997) and elsewhere disconformably (Harding and Coney, 1985; Richard
et al., 1987; Tosdal and Stone, 1994), but in any case
concordantly, on Jurassic volcanic rocks that have yielded
discordant U±Pb ages projecting to concordia at 165±
155 Ma (Reynolds et al., 1987; Richard et al., 1987;
Asmerom et al., 1991; Fackler-Adams et al., 1997).
Comparable volcanic rocks are mapped as a lower member
of the Winterhaven Formation (Haxel et al., 1985), with
only the overlying quartz arenite and argillitic siltstone
members being correlative with the lower McCoy Mountains Formation. The implied Oxfordian to Kimmeridgian
age for the base of the McCoy Mountains Formation is
closely comparable to the maximum age of the Glance
Conglomerate along the rifted keel of the Bisbee basin.
The underlying volcanic rocks are dominantly rhyodacite
porphyry petrologically similar to the Canelo Hills Volcanics beneath the Bisbee Group of southeastern Arizona, and
they appear to represent an analogous rifted-arc assemblage.
489
Ma®c dikes and sills that intruded the McCoy Mountains
Formation when it was still only partly consolidated are
compatible geochemically with a rift environment (Gleason
et al., 1999), and petrologically similar ma®c lavas are
present locally within the lower McCoy Mountains Formation (Sherrod and Koch, 1987; Sherrod et al., 1990). The
older horizons of the subjacent silicic volcanic assemblage
have yielded ages as old as ,175 Ma (Fackler-Adams et al.,
1997), similar to the oldest ages reported for the Canelo
Hills Volcanics.
In Arizona, an intraformational angular unconformity
(Fig. 6) separates diverse lower members of the McCoy
Mountains Formation from contrasting conglomeratic strata
forming the stratigraphically lowest of the upper members
(Tosdal and Stone, 1994). The stratal discordance at the
unconformity diminishes westward into California, where
a hiatus in deposition may or may not have occurred. The
age of the strata directly below the unconformity is
unknown, but a tuff bed within the conglomeratic succession above the unconformity (Fig. 6) has yielded a U±Pb
age of 79 ^ 2 Ma (Campanian), indicating general correlation of the upper McCoy Mountains Formation with syntectonic strata of Laramide basins in Arizona and Sonora
(Tosdal and Stone, 1994). Fossil wood collected stratigraphically above the tuff represents a genus well known from
Upper Cretaceous strata elsewhere (Stone et al., 1987).
The lower McCoy Mountains Formation is evidently
correlative with the Bisbee Group, representing a distal
northwestern extension of the nonmarine facies belt of the
Bisbee basin, with the intraformational unconformity or
laterally equivalent change in lithology re¯ecting the
onset of Laramide deformation. Lateral variations from
sharply angular unconformities to disconformities or
diastems are also common between Bisbee and Laramide
strata in southeastern Arizona (Dickinson et al., 1989). With
a maximum observed thickness of only ,360 m in the type
section (Haxel et al., 1985), the sedimentary members of the
Winterhaven Formation are almost an order of magnitude
thinner than the lower McCoy Mountains Formation (Figs. 2
and 6), and were probably deposited along a subdued basin
¯ank analogous tectonically to the Bisbee ¯ank basin along
tectonic strike to the southeast (Fig. 1).
4.1. Basin-bounding thrust systems
Displacement on the Mule Mountains thrust, placing
Precambrian basement and Jurassic arc plutons against the
deep keel of the McCoy basin, is bracketed within Campanian time between 79 ^ 2 and 70 ^ 4 Ma (Tosdal, 1990).
The thrusting apparently represents Laramide deformation
analogous to deep-seated, basement-involved Laramide
structures of nearby southwestern Arizona. The Maria
fold-and-thrust belt along the northern ¯ank of the McCoy
basin is at least in part somewhat older, for strands of the
thrust system are crosscut by a pluton that has yielded a
40
Ar/ 39Ar age of 79.3 ^ 0.4 Ma (Richard et al., 1998).
490
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 7. Areal distribution of dominant clast types in Glance Conglomerate of Arizona-Sonora, and equivalent strata in basal conglomerate of Hell-to-Finish
Formation in New Mexico and Sand Wells Formation of Tohono O'Odham Indian Nation west of Tucson. Outline of Bisbee basin (dashed line) from Fig. 3.
Tie lines between symbols denote mixed-clast conglomerate, most polymictic in the Canelo Hills (CH), Sierra Azul (SAz) and Sierra Anibacachi (SAn). Bars
between stacked symbols separate clast assemblages of stratigraphic units de®ning inverted clast stratigraphy of Empire (EM) and Mule Mountains (MM), or
more complex but also successive clast facies of Huachuca (HM) and Dragoon (DM) Mountains. Legend symbols denote ages of source rocks in basin
substratum: J Ð Jurassic; Pz Ð Paleozoic; Pr Ð Proterozoic. Data from Bilodeau (1978, 1979), Haxel et al. (1980), Mack et al. (1986), Archibald (1987),
Bilodeau et al. (1987), Lindberg (1987), Riggs (1987), Segerstrom (1987), GonzaÂlez-LeoÂn and Jacques-Ayala (1990), Jacques-Ayala (1995), Lawton and
Olmstead (1995), Lawton and Harrigan (1998) and Lucas and Lawton (2000).
Peak metamorphism of pre-Mesozoic rocks along the northern ¯ank of the Maria fold-and-thrust belt was approximately 75 Ma (Hoisch et al., 1988; Miller et al., 1992),
and K±Ar cooling ages for the 79 Ma pluton fall in the
range 70±65 Ma (Richard et al., 1998). As it evolved, the
Maria fold-and-thrust belt incorporated strata of the lower
McCoy Mountains Formation into the thrust system (Fig. 5),
gradually reducing the width of the McCoy basin and shedding coarse detritus southward to produce the conglomeratic
basal member of the upper McCoy Mountains Formation
(Tosdal, 1990; Miller et al., 1992; Tosdal and Stone,
1994). Paleocurrent indicators consistently re¯ect southerly
directed ¯ow, and the polymictic conglomerate contains
variable proportions of plutonic and supracrustal detritus
including both volcanic and sedimentary rocks (Harding
and Coney, 1985). Strata comprising much of the upper
McCoy Mountains Formation were deposited after eruption
of the 79 Ma tuff bed, evidently as post-tectonic basin ®ll,
and some ®ner-grained strata are probably lacustrine (Harding and Coney, 1985). The homogeneous structural style
displayed by the entire McCoy Mountains Formation, with
foliation and cleavage uniform in attitude as exposed along
the preserved keel of the McCoy basin, suggests either that
some deformation within the Maria fold-and-thrust belt
continued into Campanian time, or that a uniform tectonic
overprint was associated with movement along the Mule
Mountains thrust system south of the basin.
The relationship of the Maria fold-and-thrust belt to the
coeval Sevier fold-and-thrust belt, with its southern terminus only 175 km to the north-northwest, remains enigmatic.
The trend of the Maria belt is perpendicular to the Sevier
belt, and the Maria belt is south-vergent, away from the
continental interior, whereas the Sevier belt is east-vergent,
toward the continental interior. Their mutual geometric relationship is reminiscent of the trend of the Uinta uplift and its
bounding Laramide thrust systems at right angles to the
Sevier belt farther north in Utah. Just as the trend of the
Uinta uplift was apparently controlled by the structural grain
of a Precambrian aulacogen-like feature in which the Uinta
Mountains Group was deposited, the trend of the Maria
fold-and-thrust belt may have been controlled by the
previously established structural grain of the aulacogenlike Border rift belt (Reynolds et al., 1986).
5. Bisbee±McCoy petrofacies associations
Available for analysis of sandstone petrofacies were point
counts of framework modes for 1076 sandstone samples
from the Bisbee and McCoy basins, including the Bisbee
Group or its equivalents and overlying Laramide strata of
Late Cretaceous and Paleocene age. Both areal and stratigraphic coverage should be adequate to detect any signi®cant compositional trends. Samples from the Bisbee basin
…n ˆ 626† include 339 from Arizona, 147 from New
Mexico, and 140 from Sonora, plus 250 additional samples
collected from overlying strata deposited within superposed
Laramide basins. Samples from the McCoy Mountains
Formation …n ˆ 200† include 124 from the Bisbee-equivalent lower members and 76 from the Upper Cretaceous
upper members (Fig. 6).
As the point counts were performed by multiple operators
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
491
Fig. 8. QmFLt diagram showing compositional ranges of Bisbee basin sandstone petrofacies: Qm Ð quartz grains; F Ð feldspar grains; Lt Ð lithic grains
including chert and detrital limeclasts. No modes plot within ruled areas. Symbols and abbreviations for petrofacies ®elds keyed to Figs. 9±11. Data from
Goodlin (1985), Sumpter (1986), Archibald (1987), Inman (1987), Jamison (1987), Klute (1987, 1991), Mack (1987a), Risley (1987), Olmstead (1992),
GonzaÂlez-LeoÂn (1994), Jacques-Ayala (1995), Mann (1995) and Bayona (1998).
(n , 20) working at six different academic institutions,
results may not be comparable in detail. In particular, we
are unsure how rigorously the different operators followed
the Gazzi-Dickinson or other conventions (Ingersoll et al.,
1984) in summing modal percentages for points falling on
sand-sized quartz and feldspar grains enclosed within polycrystalline lithic fragments. Both contrasts and similarities
in petrofacies appear robust, however, at the level of detail
we treat them here using QmFLt diagrams (Dickinson,
1985). In many of the rocks studied, plagioclase consists
of secondary (low-temperature) albite pseudomorphs of
originally more calcic plagioclase, but no workers have
reported any evidence for albite replacing K feldspar (Dickinson et al., 1982).
In addition to petrofacies analysis, control for provenance
interpretations is provided by variations in gravel clast
composition within the Glance Conglomerate (Fig. 7).
Because all Glance successions were deposited close to
source areas, compositional trends within the conglomerates
provide direct information on the nature of intrabasinal
source rocks. In the southwestern part of the Bisbee basin,
the dominant clasts are volcanic rocks derived from the
subjacent Jurassic arc assemblage (Fig. 3), whereas limestone clasts derived from underlying Paleozoic sedimentary
sequences are dominant in the northeastern part of the basin
(Fig. 7). In southeastern Arizona and adjacent Sonora, some
intrabasinal fault blocks were eroded deeply enough to shed
granitic clasts from either Precambrian basement underlying
Paleozoic strata or from Jurassic plutons intrusive into the
arc assemblage. Precambrian schist clasts and quartzite
clasts derived from Cambrian strata in basal horizons of
the limestone-rich Paleozoic cover are also prominent
locally. Inverted clast stratigraphy is re¯ected locally by
stratigraphic progressions upward from limestone-clast
conglomerate to granitic-clast or schist-clast conglomerate
(EM and MM of Fig. 7). Typically, an intervening gradational interval of mixed limestone-clast and quartzite-clast
conglomerate is present as the record of an intermediate
stage in the erosional stripping of cover rocks from underlying basement in the source.
5.1. Bisbee petrofacies
Five empirically delineated petrofacies and one subfacies
are present within the Bisbee basin (Fig. 8), within which
their areal and stratigraphic distributions are indicated by
Fig. 9. Where local stratigraphic correlations allow the
distinction, data are plotted separately for sandstones of
the upper Bisbee Group overlying the Mural/U-Bar limestone interval, and for the lower Bisbee Group below that
horizon. In establishing Bisbee petrofacies boundaries and
distributions, only 11 point counts (,2% of the total available) were ignored as apparently unrepresentative of the
various local sample suites as a whole. Empirical
492
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 9. Distribution of sandstone petrofacies (Fig. 8) within Bisbee basin. Solid and open symbols for pre-mid-Aptian and post-mid-Albian strata lying
respectively below and above the Aptian±Albian Mural/U-Bar limestone interval or its lateral equivalent in nonmarine strata (horizon shown as horizontal
lines breaking symbol clusters). Symbols half-solid where correlation of sample sites uncertain. Tie lines connect interstrati®ed petrofacies (.symbol where
one dominant). Outline of Bisbee basin (dashed line) and limit of marine Mural/U-Bar transgression from Fig. 3.
petrofacies boundaries were selected to be in harmony with
perspectives on compositional variation developed in subregional syntheses for Arizona (Klute, 1987, 1991), New
Mexico (Mack, 1987b), and Sonora (Jacques-Ayala,
1995). Petrofacies are identi®ed by descriptive designations,
relevant only for this study, as follows (percentages cited are
relative to the QmFLt grain population).
5.1.1. Bisbee quartzose petrofacies (Q)
Sand frameworks of the quartzose petrofacies uniformly
contain more than ,80% monocrystalline quartz grains
(Fig. 8). Quartz grains are most typically subrounded, and
some display abraded quartz overgrowths indicative of
reworking from older sedimentary successions (Klute,
1987, 1991). The quartzose petrofacies is dominant in sandstones throughout the Bisbee Group in the central part of the
Bisbee core basin and within the upper Bisbee Group of
New Mexico to the east (Fig. 9).
5.1.2. Bisbee subquartzose petrofacies (S)
The quartz content of the subquartzose petrofacies ranges
downward toward ,45%, and lithic grains are commonly
but not uniformly more abundant than feldspar grains. Chert
is the most common lithic grain type in most sandstones of
the subquartzose petrofacies, and a chert-rich subfacies
(Fig. 8) containing only 15±45% monocrystalline quartz
grains is present sparingly in New Mexico. Some chert
grains display relict or ghost replacement features indicative
of derivation from chert nodules in carbonate successions
(Mack et al., 1986; Klute, 1987, 1991). Minor feldspar
includes both plagioclase and K feldspar in varying proportions (Mack, 1987a; Klute, 1987, 1991). The subquartzose
petrofacies is dominant in the lower Bisbee Group of New
Mexico and is widespread, though subordinate to the quartzose petrofacies, throughout the Bisbee Group in nearby
southeasternmost Arizona and adjacent Sonora (Fig. 9).
5.1.3. Bisbee lithic petrofacies (L)
The quartz content of the lithic petrofacies is less than
45%, with quartz contents of only 10±30% most common.
Many quartz grains display straight extinction suggestive of
derivation from phenocrysts in volcanic rock. Polycrystalline lithic grains, which are chie¯y volcanic rock fragments,
are typically, though not uniformly, more abundant than
feldspar grains (Fig. 8), which are dominantly plagioclase
(Inman, 1987; Klute, 1991; Jacques-Ayala, 1995). The association of plagioclase with volcanic rock fragments is indicative of volcaniclastic detritus (Dickinson, 1985). Both
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
493
frameworks assigned here to a transitional petrofacies occupying a QmFLt compositional ®eld intermediate between
the lithic and the arkosic petrofacies (Figs. 8 and 9). Admixture of arkosic and lithic sands derived, respectively, from
plutonic and volcanic sources is suggested, and some
samples of the transitional petrofacies may be volcanoplutonic sands (Dickinson, 1982) derived from dissection of
Jurassic arc assemblages. Alternate or additional admixture
of either or both kinds of igneous detritus with quartzose or
subquartzose sands may also be represented in the transitional petrofacies.
5.2. Laramide petrofacies
Fig. 10. QmFLt diagram illustrating wide variability of sandstone petrofacies in Laramide synorogenic successions (Arizona±Sonora and New
Mexico) within geographic area of older Bisbee basin. Data from Mark
(1985), Inman (1987), Hayes (1987), Jacques-Ayala and Potter (1987),
James and Russo (1988), Wilson (1991), Mann (1995), GonzaÂlez-LeoÂn
and Lawton (1995) and Basabilvazo (2000).
microlitic grains derived from intermediate (andesitic)
rocks and felsitic grains derived from more felsic rocks
are common (Klute, 1991; GonzaÂlez-LeoÂn, 1994; JacquesAyala, 1995). The lithic petrofacies is dominant throughout
the Bisbee Group in the southwestern part of the Bisbee
basin, particularly over most of the Bisbee ¯ank basin
(Fig. 9). Selected sandstones in ¯uvial strata of the Bisbee
Group contain frameworks typical of the lithic petrofacies
as far northeast as the Huachuca …n ˆ 5† and Mule …n ˆ 3†
mountains in southeastern Arizona and selected nearby
ranges of New Mexico …n ˆ 2†: The lithic detritus is
commonly more angular toward the southwest (JacquesAyala, 1995) and better sorted toward the northeast
(Klute, 1991). Northeastern representatives of the lithic
petrofacies are undiluted by mixing with more quartzose
sands, but are interstrati®ed with typical representatives of
the quartzose and subquartzose petrofacies.
5.1.4. Bisbee arkosic petrofacies (A)
Arkosic (quartz-feldspar) frameworks, containing less
than 14-19% polycrystalline lithic grains (Fig. 8), re¯ect
derivation from subjacent or nearby granitic rocks of
Precambrian basement or Jurassic plutons. Plagioclase and
K feldspar are coequal in abundance, as expected from
plutonic sources. Sand frameworks are most typically
subangular and are more poorly sorted than other petrofacies (Klute, 1991), as expected from short distances of transport and deposition in depositional environments proximal
to source. The arkosic petrofacies is dominant in sandstones
of the lower Bisbee Group in the northwestern part of the
Bisbee basin, but is absent elsewhere and in the upper
Bisbee Group (Fig. 9).
5.1.5. Bisbee transitional petrofacies (T)
Some sample suites from southern and northwestern
segments of the Bisbee basin include sandstones displaying
Upper Cretaceous to Paleogene sandstones of Laramide
basins in Arizona and adjacent parts of Sonora vary widely
and unsystematically in modal composition (Fig. 10) but
have uniformly feldspathic to lithic frameworks re¯ecting
composite derivation from Laramide volcanic ®elds and
pre-Laramide rocks exposed in local fault-bounded uplifts.
Near-vent volcanic or more distal pyroclastic rocks are
locally intercalated within many Laramide sedimentary
successions (Dickinson et al., 1989). Widespread and
broad overlap of the Laramide petrofacies with Bisbee
petrofacies is evident, except that Laramide strata of
Arizona and Sonora contain no frameworks comparable to
the quartzose or subquartzose Bisbee petrofacies (Fig. 10).
Even that distinction is lost, however, in New Mexico where
some Laramide sandstones display subquartzose and even
quartzose frameworks (Fig. 10). Much of the overlap in
framework composition between Bisbee and Laramide
sandstones probably stems from reworking of detritus
from the Bisbee Group, for Bisbee clasts are common in
conglomerates of Laramide successions (Goodlin and
Mark, 1987; Wilson, 1991; GonzaÂlez-LeoÂn and Lawton,
1995; Mann, 1995; Basabilvazo, 2000).
The ambiguity between Bisbee and Laramide petrofacies
is shown by sandstones of the Amole Arkose in the Tucson
Mountains (Fig. 9). Although most Amole exposures
include lithofacies typical of the nonmarine western facies
of the Bisbee Group (Risley, 1987), ignimbrite of Laramide
age is known to be interbedded within strata mapped as
uppermost Amole Arkose intruded by an Upper Cretaceous
pluton near the western edge of the range (Lipman, 1993).
No clear-cut petrofacies distinction can be drawn between
the sandstones of known Laramide age and the older sandstones exposed farther down in the local stratigraphic
sequence as part of typical Bisbee stratal assemblages,
which include laminated algal limestone resembling strata
of the lacustrine Apache Canyon Formation exposed in
ranges farther east (Risley, 1987).
5.3. McCoy petrofacies
Reliable detrital modes are more dif®cult to determine for
sandstones of the McCoy Mountains Formation than for the
Bisbee Group. Typical rocks in most exposures are foliated
494
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Fig. 11. QmFLt diagram (see Fig. 8) showing compositional ranges of sandstone petrofacies of McCoy Mountains Formation. No modes plot within ruled
areas. For comparison, dashed lines indicate boundaries of Bisbee basin petrofacies (abbreviations in brackets from Fig. 8). Data from Robison (1979, 1980),
Harding (1980, 1982), Laubach et al. (1987), Richard et al. (1987), Fackler-Adams et al. (1997) and Dickinson (2000). Plotted points represent individual
control samples (Dickinson, 2000), with ticks indicating mixed (M) petrofacies.
to varying degrees, and neomorphism of metamorphic
minerals has obscured detrital textures in many cases,
with variable development of diagenetic-metamorphic
epimatrix interstitially between relict framework grains.
For this study, we collected a control set of 30 samples
from outcrops in Arizona displaying the least foliation and
metamorphism (Dickinson, 2000). Point counts from this
selective sample suite were given greater weight in delineating petrofacies than the generally comparable results
reported from 170 other samples described by previous
workers. Even so, we dismissed modal data from only
eight of the other available McCoy counts (,5% of the
total sample suite) as unrepresentative. Petrofacies boundaries (Fig. 11) were established entirely on the internal basis
of McCoy point counts without external reference to Bisbee
petrofacies. Petrofacies are identi®ed by the following
descriptive designations, relevant only for this study (the
®rst three, abbreviated with the pre®x `m' to distinguish
them from Bisbee petrofacies, occur in Bisbee-equivalent
lower McCoy strata, but the other two occur only in postBisbee-age upper McCoy strata).
5.3.1. McCoy quartzose petrofacies (mQ)
The McCoy quartzose petrofacies is essentially indistinguishable from the Bisbee quartzose petrofacies (Fig. 11),
and the two are inferred to re¯ect the same type of provenance. The dominant quartz grains are generally well sorted
and subrounded, and feldspar is rare or absent (Dickinson,
2000). The quartzose petrofacies is characteristic of `basal
sandstone member 1' of Harding and Coney (1985), which
equates to the `quartz arenite member' (Sherrod et al., 1990)
and the Crystal Hill `formation' or member (Richard et al.,
1993; Richard and Spencer, 1994).
5.3.2. McCoy subquartzose petrofacies (mS)
The compositional ®eld for sand frameworks of the
McCoy subquartzose petrofacies is contained almost
entirely within the ®eld for the Bisbee subquartzose petrofacies (Fig. 11). In the control samples, the dominant lithic
grains are chert, and an analogous provenance is implied for
the two similar subquartzose petrofacies. Both feldspars are
present, with either the more abundant in different local
collections (Fackler-Adams et al., 1997; Dickinson, 2000).
The position of the QmFLt compositional ®eld for the
McCoy subquartzose petrofacies along a potential mixing
path between compositional ®elds for the stratally associated McCoy quartzose and McCoy lithic petrofacies
suggests that sedimentological mingling of quartzose and
volcaniclastic sands may have produced the McCoy
subquartzose sandstones in which plagioclase is more abundant than K feldspar. Samples of the McCoy subquartzose
petrofacies in which K feldspar is more abundant than plagioclase were probably derived, however, from the same types
of sources as the Bisbee subquartzose petrofacies. The
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
495
near Ramsey Mine (Sherrod and Koch, 1987; Sherrod et al.,
1990), and also occurs within the `mudstone member' of
Harding and Coney (1985).
Fig. 12. Provenance relations of Bisbee and McCoy basins during Early
Cretaceous time. Intrabasinal sources from residual synrift paleotopography not shown. Gulf of California closed by reversal of slip on San Andreas
transform system (SAF, Neogene San Andreas fault trend); note offset of
international boundary (SA offset). Arrows denote dispersal of volcaniclastic detritus from Alisitos arc, quartzose cratonic detritus from continental
surface (across Mogollon paleohighland ¯anking Bisbee basin), and
subquartzose detritus from Burro uplift (eastern segment of Mogollon
paleohighland). Cities: A Ð Albuquerque; Ch Ð Ciudad Chihuahua; EP
Ð El Paso; H Ð Hermosillo; LA Ð Los Angeles; LC Ð Las Cruces; LV
Ð Las Vegas, P Ð Phoenix; SD Ð San Diego; Tu Ð Tucson. Modern
rivers: C.R. Ð Colorado; R.G. Ð Rio Grande.
McCoy subquartzose petrofacies is characteristic of the
`Ranegras member' (Spencer et al., 1985; Richard et al.,
1987), but also occurs within the `mudstone member' of
Harding and Coney (1985).
5.3.3. McCoy lithic petrofacies (mL)
The compositional ®elds for the McCoy and Bisbee lithic
petrofacies are virtually coextensive (Fig. 11), except that
the former overlaps somewhat with the related transitional
facies of the Bisbee basin, and both lithic petrofacies are
dominantly volcaniclastic. Most volcanic rock fragments
are felsite similar to the quartz porphyry bodies that underlie
the McCoy Mountains Formation, and the feldspar is exclusively or predominantly plagioclase (Harding, 1982; Dickinson, 2000). The McCoy lithic petrofacies is characteristic
of `basal sandstone member 2' of Harding and Coney
(1985), forms the `Harquar member' (Richard et al., 1987;
Spencer et al., 1985) and the related sedimentary succession
5.3.4. McCoy quartzo-feldspathic petrofacies (QF)
The McCoy quartzo-feldspathic petrofacies is similar in
overall grain composition to Bisbee arkosic and transitional
petrofacies (Fig. 11) but is generically more closely related
to feldspathic representatives of analogous Laramide petrofacies in Arizona and Sonora (Fig. 10). Both feldspars are
present in signi®cant proportions, although plagioclase is
more abundant (,2:1) than K feldspar (Dickinson, 2000).
The quartzo-feldspathic petrofacies is restricted to the upper
McCoy Mountains Formation (Harding and Coney, 1985)
and is characteristic of all its members. Samples from the
uppermost (siltstone) member fall entirely within the
compositional ®eld for the Bisbee transitional petrofacies,
indicating dilution of arkosic debris dominant in both the
underlying members (conglomerate and sandstone) with
detritus from more supracrustal sources. Sedimentary±
metasedimentary rock fragments are more abundant
(,4:1) than volcanic rock fragments (Dickinson, 2000).
By inference, the arkosic detritus was derived principally
from basement rocks uplifted along the Maria fold-andthrust belt north of the McCoy basin, and the supracrustal
detritus was derived from cover strata as the deformed belt
was eroded.
5.3.5. McCoy mixed petrofacies (M)
Samples from the Apache Wash formation or member
(Harding and Coney, 1985; Richard, 1992; Richard et al.,
1987, 1993) contain a mixed petrofacies that overlaps
compositional ®elds for lower and upper McCoy sandstones
(Fig. 11). The ratio of plagioclase to K feldspar (,2:1) is
characteristic, however, of upper rather than lower McCoy
petrofacies (Dickinson, 2000). Within the McCoy Mountains Formation, the mixed petrofacies of the Apache
Wash member is also notable for its relatively high content
(5±10%) of detrital limeclasts reworked from older Paleozoic limestone bedrock. The stratigraphic position of the
Apache Wash member was long uncertain (Harding and
Coney, 1985), but detailed mapping (Richard, 1992;
Richard et al., 1993) has now shown that it overlies a
thick lens of debris-avalanche megabreccia previously misidenti®ed as bedrock. The megabreccia in turn overlies the
lower McCoy Mountains Formation, and the Apache Wash
member thus correlates broadly with the basal `conglomerate member' of the upper McCoy Mountains Formation. As
inferred with regard to the compositional overlap of Bisbee
and Laramide petrofacies in Arizona, compositional overlap
of the mixed petrofacies of the upper McCoy Apache Wash
member with lower McCoy petrofacies may stem in part
from reworking of lower McCoy detritus, in this case
from erosion of strata deformed and uplifted within the
Maria fold-and-thrust belt (Fig. 5).
496
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
6. Bisbee±McCoy provenance interpretations
The sources and provenance of Upper Jurassic to Lower
Cretaceous Bisbee±McCoy petrofacies can be inferred from
their geographic and stratigraphic distributions in the two
related basins. Derivation of the arkosic petrofacies in the
lower Bisbee Group of the nonmarine western facies from
intrabasinal basement sources, exposed in local tilt blocks of
the northwestern Bisbee basin, is indicated by gradational to
inter®ngering contacts with granitic-clast Glance Conglomerate. Derivation of the lithic petrofacies from subjacent
Jurassic volcanic rocks exposed nearby is also feasible for
lower parts of the Bisbee Group, and is speci®cally inferred
for sequences of the lower McCoy Mountains Formation
closely associated stratigraphically with directly underlying
Jurassic volcanic rocks in both Arizona and California.
Later widespread deposition of the lithic petrofacies in the
upper Bisbee Group cannot be attributed, however, to
erosion of local intrabasinal sources because the basin
substratum was by then buried under thick sequences of
older Bisbee strata. The extent of the lithic petrofacies
across a broad expanse of the southwestern Bisbee basin,
coupled with generally easterly paleocurrent indicators,
suggests derivation from the Aptian±Albian Alisitos
magmatic arc of modern Baja California, which was located
not far southwest of the Bisbee basin before sea¯oor spreading that opened the Gulf of California in Neogene time (Fig.
12).
Although stripping of sedimentary cover from intrabasinal tilt blocks may have locally contributed minor volumes
of subquartzose sand to the Bisbee basin, the widespread
quartzose and subquartzose petrofacies of the northeastern
Bisbee basin record the entry of extrabasinal detritus into
the rift belt. Southerly directed paleocurrents suggest that
the subquartzose petrofacies dominant in the lower Bisbee
Group of New Mexico was derived from the northern rift
shoulder of the basin where Paleozoic sediment cover at
least 1 km thick (Mack, 1987a) was stripped by erosion
from basement of the Burro uplift (Mack et al., 1986).
Farther west in the central Bisbee basin, where southerly
directed paleocurrents also prevail, the dominance of the
quartzose petrofacies throughout Bisbee deposition (Fig.
9) implies delivery of quartzose sediment from the segment
of the rift shoulder forming the Mogollon paleohighland of
Arizona (Fig. 12). The quartz-rich sand was probably
recycled from Jurassic erg deposits, which are widespread
on the Colorado Plateau in the region directly north of the
Mogollon paleohighland of Arizona but never extended as
far east as the Burro uplift of New Mexico (Riggs and
Blakey, 1993). The absence of the erg deposits overlying
Paleozoic and older rocks along the trend of the Mogollon
paleohighland is interpreted to re¯ect erosional stripping of
early Mesozoic cover during Bisbee sedimentation. A similar sediment supply reached the McCoy basin still farther
west (Fig. 12) where relict erg deposits are also present
directly to the north beyond the rift shoulder. As sediment
®ll built up, quartzose detritus initially ponded in the central
part of the Bisbee basin, but later spilled eastward along the
axis of the basin within ¯uvial systems ¯owing toward the
northern end of the Chihuahua trough. Derivation of the
quartzose petrofacies from any direction other than the
northern rift shoulder is precluded by the dominance of
arkosic and lithic petrofacies to the west and southwest,
respectively, and by easterly-directed paleocurrents in the
quartzose petrofacies of the upper Bisbee Group in New
Mexico.
Transitional and intermingled petrofacies within the
Bisbee basin are inferred to re¯ect mixing and interstrati®cation of intrabasinal arkosic detritus, intrabasinal and
basin-¯ank subquartzose detritus, volcaniclastic lithic detritus derived from both subjacent and extrabasinal volcanic
sources (the latter lying toward the southwest), and cratonderived quartzose detritus delivered to the basin from the
north. The axis of most prominent petrofacies mixing and
intertonguing was oriented northwest±southeast at right
angles to the boundary between the nonmarine western
and the mixed (nonmarine-marine) central facies belts
(Figs. 3 and 9). Lithic to arkosic petrofacies are dominant
to the southwest of the mixed petrofacies belt, and quartzose
to subquartzose petrofacies are dominant to the northeast.
The transitional petrofacies re¯ects intermixing mainly of
arkosic and volcaniclastic sands, whereas some of the
subquartzose petrofacies may re¯ect intermixing of quartzose and volcaniclastic sands, especially within the McCoy
basin. The delivery of subordinate quantities of unmixed
lithic sands from volcanic sources toward the southwest
into the central Bisbee basin indicates, however, that the
¯uvial transport systems leading across the basin ¯oor
from different source terranes could at times maintain
quite separate dispersal paths for long distances.
7. Summary Conclusions
The Bisbee core basin was one of a string of linked Late
Jurassic to mid-Cretaceous depocenters extending to the
northwest from the Gulf of Mexico as far as the McCoy
basin of Arizona±California and de®ning the trend of the
Border rift belt controlled by slab rollback beneath the
Cordilleran magmatic arc (Dickinson and Lawton, 1999,
2001). The Bisbee ¯ank basin, which received less sediment
cover, extended across one of the paleotectonic sills intervening between higher standing platforms that delineated
the southwestern ¯ank of the Border rift belt north of the
Mar Mexicano. The latter oceanic domain occupied central
Mexico prior to arc collision and accretion of the Guerrero
superterrane in Aptian±Albian time (Dickinson and
Lawton, 1999, 2001). Retroarc and intra-arc deformation
related to migratory Laramide magmatism disrupted the
Bisbee and McCoy basins in Late Cretaceous time.
The association within a single rift basin of quartzose,
volcaniclastic, and arkosic petrofacies, together with
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
transitional and subquartzose petrofacies of intermediate
character, re¯ects the unusual geotectonic setting of the
Bisbee basin as a segment of the Border rift belt. The
close geographic and stratigraphic juxtaposition of petrofacies re¯ecting derivation of detritus from the coeval Alisitos
magmatic arc, locally uplifted continental basement, and the
craton surface are understandable in the paleogeographic
context of the Border rift belt, but are otherwise puzzling.
The geodynamic patterns that controlled Bisbee basin
evolution and sediment provenance were apparently unique
in the Phanerozoic history of North America, which led to
the delivery into the same depocenter of multiple petrofacies normally indicative of disparate geotectonic settings.
Acknowledgements
The graduate work of W.L. Bilodeau at Stanford University introduced the ®rst author to problems of the Bisbee
basin. Liaison across the international border with Claudio
Bartolini, Samuel Eguiluz de AntunÄano, Carlos M. GonzaÂlez-LeoÂn, CeÂsar Jacques-Ayala, and Rogelio Monreal was
especially helpful during our investigations. Discussions of
Bisbee-McCoy geology over the years with colleagues and
coworkers R.F. Butler, R.E. Clemons, P.J. Coney, P.E.
Damon, G.H. Davis, K.W. Flessa, E.R. Force, G.B. Haxel,
S.B. Keith, C.F. Kluth, P.W. Lipman, S.G. Lucas, G.H.
Mack, N.J. McMillan, S.J. Reynolds, S.M. Richard, J.
Ruiz, W.R. Seager, J.F. Schreiber, D.R. Sherrod, J.E. Spencer, S.R. Titley, and R.M. Tosdal improved our understanding of various key relationships. We gratefully acknowledge
the collaboration and stimulus of the following students at
the University of Arizona and New Mexico State University
at various stages of our study: L.E. Archibald, Y. Asmerom,
C. Bartolini, G.T. Basabilvazo, G. Bayona, S. Buffum, R.C.
Ferguson, A.R. Fiorillo, J.D. Gleason, C.M. GonzaÂlez-LeoÂn,
T.C. Goodlin, D.L. Hall, L.E. Harding, P.J. Harrigan, M.J.
Hayes, K.F. Inman, K. Jamison, S.U. Janecke, M.A. Klute,
C.K. Krebs, F.A. Lindberg, J.D. Mann, R.A. Mark, S.R.
May, R. Monreal, G.A. Olmstead, A.R. Potochnik, S.M.
Richard, N.R. Riggs, R. Risley, S. Sindlinger, L.T. Sumpter,
P.N. Swift, L.K. Vedder, and D.A. Wilson. An oral version
of this paper was presented at the Cuarta Reunion Sobre la
 reas Adjacentes in
GeologõÂa del Noroeste de MeÂxico y A
Hermosillo in March 2000. Reviews by S.G. Lucas, G.A.
Smith, and J.E. Spencer improved the manuscript. Jim
Abbott of SciGraphics prepared the ®gures. Our research
was funded in part by National Science Foundation grants
EAR-8417106 (to Dickinson) and EAR-9304759 (to
Lawton).
References
Amato, J.M., 2000. Structural relationships in the Florida Mountains, southwestern New Mexico Ð a review. Southwest Passage Ð A Trip
Through the Phanerozoic, Lawton, T.F., McMillan, N.J., McLemore,
497
V.T. (Eds.). New Mexico Geological Society Annual Field Conference
Guidebook 51, 102±108.
Amsbury, D.L., Reaser, D.F., 1988. Lower Cretaceous of western transPecos Texas. South-Central Section of the Geological Society of America, Hayward, O.T. (Ed.). Geological Society of America Centennial
Field Guide 4, 401±406.
Araujo Mendieta, J., Casar GonzaÂlez, R., 1987. EstratigrafõÂa y sedimentologõÂa del JuraÂsico Superior en la cuenca de Chihuahua, norte de MeÂxico.
Revista del Instituto Mexicano del PetroÂleo 19, 6±29.
Araujo Mendieta, J., Estavillo GonzaÂlez, C.F., 1987. EvolucioÂn tectoÂnica
sedimentaria del JuraÂsico Superior y CretaÂcico Inferior en el NE de
Sonora, MeÂxico. Revista del Instituto Mexicano del PetroÂleo 19, 4±37.
Archibald, L.E., 1987. Stratigraphy and sedimentology of the Bisbee Group
in the Whetstone Mountains, southeastern Arizona. Mesozoic Rocks of
Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A.
(Eds.). Arizona Geological Society Digest 18, 273±282.
Asmerom, Y., Zartman, R.E., Damon, P.E., Sha®qullah, M., 1990. Zircon
U±Th±Pb and whole-rock Rb±Sr patterns of lower Mesozoic igneous
rocks in the Santa Rita Mountains, southeast Arizona: implications for
Mesozoic magmatism and tectonics in the southern Cordillera. Geological Society of America Bulletin 102, 961±968.
Asmerom, Y., Damon, P., Sha®qullah, S., Dickinson, W.R., Zartman, R.E.,
1991. Resetting of Rb±Sr ages of volcanic rocks by low-grade burial
metamorphism. Chemical Geology (Isotope Geoscience Section) 87,
167±173.
Baron-Szabo, R.C., GonzaÂlez-LeÂon, C.M., 1999. Lower Cretaceous corals
and stratigraphy of the Bisbee Group (Cerro de Oro and Lampazos
areas), Sonora, Mexico. Cretaceous Research 20, 465±497.
Bartolini, C., Herrera U, S., 1986. EstratigrafõÂa y estructura de la regioÂn de
Lampazos, Sonora, MeÂxico. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 3 (2), 13±22.
Barton, M.D., Hanson, R.B., 1989. Magmatism and the development of
low-pressure metamorphic belts: implications from the western United
States and thermal modeling. Geological Society of America Bulletin
101, 1051±1065.
Basabilvazo, G., 2000. The Upper Cretaceous Ringbone Formation, Little
Hatchet Mountains, southwestern New Mexico. Southwest Passage Ð
A Trip through the Phanerozoic, Lawton, T.F., McMillan, N.J., McLemore, V.T. (Eds.). New Mexico Geological Society Annual Field
Conference Guidebook 51, 203±210.
Bayona, G., 1998. Stratigraphic relationships of Cretaceous and Paleogene
deposits, and their tectonic implications; central Peloncillo Mountains,
Hidalgo County, New Mexico. Unpublished master's thesis, New
Mexico State University, Las Cruces, 151 pp.
Bayona, G., Lawton, T.F., 2000. Aptian±Albian extensional faulting and
subsequent Laramide inversion, central Peloncillo Mountains, southwestern New Mexico. Southwest Passage Ð A Trip Through the
Phanerozoic, Lawton, T.F., McMillan, N.J., McLemore, V.T. (Eds.).
Southwest New Mexico Geological Society Annual Field Conference
Guidebook 51, 85±94.
Beikman, H.M., Haxel, G.B., Miller, R.J., 1995. Geologic map of the
Tohono O'Odham Indian Nation, southern Arizona. US Geological
Survey Miscellaneous Investigations Series Map I-2017, scale
1:125,000.
Berggren, W.A., Kent, D.V., Swisher III, C.C., Aubry, M.-P., 1995. A
revised Cenozoic geochronology and chronostratigraphy. Geochronology Time Scales and Global Stratigraphic Correlation, Berggren, W.A.,
Kent, D.V., Aubry, M.-P., Hardenbol, J. (Eds.). Society for Sedimentary
Geology (SEPM) Special Publication 54, 129±212.
Bilodeau, W.L., 1978. The Glance Conglomerate: a Lower Cretaceous
syntectonic deposit in southeastern Arizona. Land of Cochise Ð Southeastern Arizona, Callender, J.F., Wilt, J.C., Clemons, R.E. (Eds.). New
Mexico Geological Society Annual Field Conference Guidebook 29,
209±214.
Bilodeau, W.L., 1979. Early Cretaceous tectonics and deposition of the
Glance Conglomerate, southeastern Arizona. Unpublished PhD dissertation, Stanford University, 145 pp.
498
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Bilodeau, W.L., 1982. Tectonic models for Early Cretaceous rifting in
southeastern Arizona. Geology 10, 466±470.
Bilodeau, W.L., 1986. The Mesozoic Mogollon highlands, Arizona: an
Early Cretaceous rift shoulder. Journal of Geology 94, 724±735.
Bilodeau, W.L., Keith, S.B., 1986. Lower Jurassic Navajo-Aztec-equivalent sandstones in southern Arizona and their paleogeographic signi®cance. American Association of Petroleum Geologists Bulletin 70,
690±701.
Bilodeau, W.L., Lindberg, F.A., 1983. Early Cretaceous tectonics and sedimentation in southern Arizona, southwestern New Mexico, and northern Sonora, Mexico. Mesozoic Paleogeography of the West-Central
United States, Reynolds, M.W., Dolly, E.D. (Eds.). Rocky Mountain
Section (SEPM) Rocky Mountain Paleogeography Symposium 2, 173±
188.
Bilodeau, W.L., Kluth, C.F., Vedder, L.K., 1987. Regional stratigraphic,
sedimentologic, and tectonic relationships of the Glance Conglomerate
in southern Arizona. Mesozoic Rocks of Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological
Society Digest 18, 229±256.
Brown, M.L., Dyer, R., 1987. Mesozoic geology of northwestern Chihuahua, Mexico. Mesozoic Rocks of Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological Society Digest 18,
381±394.
Busby-Spera, C.J., 1988. Speculative tectonic model for the lower Mesozoic arc of the southwest Cordillera. Geology 16, 1121±1125.
Busby-Spera, C.J., Mattinson, J.M., Riggs, N.R., Schermer, E.R., 1990. The
Triassic±Jurassic magmatic arc in the Mojave±Sonoran Deserts and the
Sierran±Klamath region; similarities and differences in paleogeographic evolution. Paleozoic and Early Mesozoic Paleogeographic
Relations; Sierra Nevada, Klamath Mountains, and Related Terranes,
Harwood, D.S., Miller, M.M. (Eds.). Geological Society of America
Special Paper 255, 325±337.
Campbell, D.H., 1980. The Yucca Formation Ð Early Cretaceous continental and transitional environments, southern Quitman Mountains,
Hudspeth County, Texas. Trans-Pecos Region, Dickerson, P.W.,
Hoffer, J.M. (Eds.). New Mexico Geological Society Annual Field
Conference Guidebook 31, 159±168.
Cantu Chapa, C.M., 1993. Sedimentation and tectonic subsidence during
the Albian±Cenomanian in the Chihuahua basin, Mexico. Cretaceous
Carbonate Platforms, Simo, J.A.T., Scott, R.W., Masse, J.-P. (Eds.).
American Association of Petroleum Geologists Memoir 56, 61±70.
Cantu Chapa, A., 1998. Las transgresiones JuraÂsicas en MeÂxico. Revista
Mexicana de Ciencias GeoloÂgicas 15, 25±37.
Cantu Chapa, C.M., Sandoval Silva, R., Arenas Partida, R., 1985. EvolucioÂn sedimentaria del CretaÂcico Inferior en el norte de MeÂxico. Revista
Instituto Mexicano del PetroÂleo 17, 14±37.
Chafetz, H.S., 1982. The Upper Cretaceous Beartooth Sandstone of southwestern New Mexico. Journal of Sedimentary Petrology 52, 157±169.
Cobban, W.A., Hook, S.C., 1984. Mid-Cretaceous molluscan biostratigraphy and paleogeography of southwestern part of western interior,
United States. Jurassic±Cretaceous Biochronology and Paleogeography
of North America, Westermann, G.E.G. (Ed.). Geological Association
of Canada Special Paper 27, 257±271.
Contreras Montero, B., MartõÂnez Cortes, A., GoÂmez Luna, M.E., 1988.
BiostratigrafõÂa y sedimentologõÂa del JuraÂsico Superior en San Pedro
del Gallo, Durango, MeÂxico. Revista del Instituto Mexicano del PetroÂleo 20, 5±29.
CoÂrdoba, D.A., 1969. Mesozoic stratigraphy of northeastern Chihuahua,
Mexico. The Border Region, CoÂrdoba, D.A., Wengerd, S.A., Shomaker,
J. (Eds.). New Mexico Geological Society Annual Field Conference
Guidebook 20, 91±96.
CoÂrdoba, D.A., RodrõÂguez-Torres, R., Guerrero-GarcõÂa, J., 1970. Mesozoic
stratigraphy of the northern portion of the Chihuahua trough. In:
Seewald, K., Sundeen, D. (Eds.). The Geologic Framework of the
Chihuahua Tectonic Belt. West Texas Geological Society, Midland,
pp. 83±97.
Cuevas Leree, J.A., 1985. Analysis of subsidence and thermal history in the
Sabinas basin, northeastern Mexico. Unpublished master's thesis,
University of Arizona, 81 pp.
Davis, G.H., 1979. Laramide folding and faulting in southeastern Arizona.
American Journal of Science 279, 543±569.
DeFord, R.K., 1969. Some keys to the geology of northern Chihuahua. The
Border Region, CoÂrdoba, D.A., Wengerd, S.A., Shomaker, J. (Eds.).
New Mexico Geological Society Annual Field Conference Guidebook
20, 61±65.
DeFord, R.K., Haenggi, W.T., 1970. Stratigraphic nomenclature of Cretaceous rocks in northeastern Chihuahua. In: Seewald, K., Sundeen, D.
(Eds.). The Geologic Framework of the Chihuahua Tectonic Belt. West
Texas Geological Society, Midland, pp. 175±196.
Dickinson, W.R., 1981. Plate tectonic evolution of the southern Cordillera.
Relations of Tectonics to Ore Deposits in the Southern Cordillera,
Dickinson, W.R., Payne, W.D. (Eds.). Arizona Geological Society
Digest 14, 113±135.
Dickinson, W.R., 1982. Compositions of sandstones in circum-Paci®c
subduction complexes and fore-arc basins. American Association of
Petroleum Geologists Bulletin 66, 121±137.
Dickinson, W.R., 1984. Reinterpretation of Lime Peak thrust as a low-angle
normal fault: implications for the tectonics of southeastern Arizona.
Geology 12, 610±613.
Dickinson, W.R., 1985. Interpreting provenance relations from detrital
modes of sandstones. In: Zuffa, G.G. (Ed.). Provenance of Arenites.
Reidel, Dordrecht, pp. 333±361.
Dickinson, W.R., 1991. Tectonic setting of faulted Tertiary strata associated with the Catalina core complex in southern Arizona. Geological
Society of America Special Paper 264, 106 pp.
Dickinson, W.R., 2000. Detrital modes of selected sandstone samples from
the McCoy Mountains Formation and correlative units in southwestern
Arizona. Arizona Geological Survey Contributed Report CR-00-A, 18
pp.
Dickinson, W.R., Lawton, T.F., 1999. Post-mid-Paleozoic tectonic assembly of Mexico. Geological Society of America Abstracts with Programs
31 (7), A-293.
Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to Cretaceous assembly and fragmentation of Mexico. Geological Society of America Bulletin 113, 1142±1160.
Dickinson, W.R., Ingersoll, R.V., Cowan, D.S., Helmold, K.P., Suczek,
C.A., 1982. Provenance of Franciscan graywackes in coastal California.
Geological Society of America Bulletin 93, 105±107 and 1306 (Errata).
Dickinson, W.R., Klute, M.A., Swift, P.A., 1986. The Bisbee basin and its
bearing on late Mesozoic paleogeographic and paleotectonic relations
between the Cordilleran and Caribbean regions. Cretaceous Stratigraphy, Western North America, Abbott, P.L. (Ed.). Paci®c Section
(SEPM) Book 46, 51±62.
Dickinson, W.R., Klute, M.A., Bilodeau, W.L., 1987. Tectonic setting and
sedimentological features of upper Mesozoic strata in southeastern
Arizona. Geologic Diversity of Arizona and its Margins: Excursions
to Choice Areas, Davis, G.H., VandenDolder, E.M. (Eds.). Arizona
Geological Survey Special Paper 5, 266±279.
Dickinson, W.R., Fiorillo, A.R., Hall, D.L., Monreal, R., Potochnik, A.R.,
Swift, P.N., 1989. Cretaceous strata of southern Arizona. Geologic
Evolution of Arizona, Jenney, J.P., Reynolds, S.R. (Eds.). Arizona
Geological Society Digest 17, 397±434.
Drewes, H., 1971. Mesozoic stratigraphy of the Santa Rita Mountains
southeast of Tucson, Arizona. US Geological Survey Professional
Paper 658-C, 81 pp..
Drewes, H., 1976. Laramide tectonics from Paradise to Hells Gate, southeastern Arizona. Arizona Geological Society Digest 10, 151±167.
Drewes, H., 1978. The Cordilleran orogenic belt between Nevada and
Chihuahua. Geological Society of America Bulletin 89, 641±657.
Drewes, H, 1980. Tectonic map of southeast Arizona. US Geological
Survey Miscellaneous Investigations Series Map I-1109, scale
1:125,000
Drewes, H., 1981. Tectonics of southeastern Arizona. US Geological
Society Professional Paper 1144, 96 pp.
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Drewes, H., 1982. Some general features of the El Paso Ð Wickenburg
transect of the Cordilleran orogenic belt, Texas to Arizona. In: Drewes,
H. (Ed.). Cordilleran overthrust belt, Texas to Arizona. Rocky Mountain Association of Geologists, Denver, pp. 88±96.
Drewes, H., 1986. Geologic map of the northern part of the Animas Mountains, Hidalgo County, New Mexico. US Geological Survey Miscellaneous Investigations Series Map I-1686, scale 1:24,000.
Drewes, H., 1988. Development of the foreland zone and adjacent terranes
of the Cordilleran orogenic belt near the US±Mexican border. Interaction of the Rocky Mountain Foreland and the Cordilleran Thrust Belt,
Schmidt, C.J., Perry Jr, W.T. (Eds.). Development of the foreland zone
and adjacent terranes of the Geological Society of America Memoir
171, 447±463.
Drewes, H., 1991a. Geologic map of the Big Hatchet Mountains, Hidalgo
County, New Mexico. US Geological Survey Miscellaneous Investigations Series Map I-2144, scale 1:24,000.
Drewes, H., 1991b. Description and development of the Cordilleran
orogenic belt in the southwestern United States and Mexico. US Geological Survey Professional Paper 1572, 92 pp.
Drewes, H., Dyer, R., 1993. Geologic map and structure sections of the
Sierra JuaÂrez, Chihuahua, Mexico. US Geological Survey Miscellaneous Investigations Series Map I-2287, scale 1:12,500
Drewes, H., Thorman, C.H., 1980a. Geologic map of the Cotton City Quadrangle and the adjacent part of the Vanar Quadrangle, Hidalgo County,
New Mexico. US Geological Survey Miscellaneous Investigations
Series Map I-1121, scale 1:24,000.
Drewes, H., Thorman, C.H., 1980b. Geologic map of the Steins Quadrangle
and the adjacent part of the Vanar Quadrangle. Hidalgo County, New
Mexico. US Geological Survey Miscellaneous Investigations Series
Map I-1220, scale 1:24,000.
Eaton, J., Dyer, R., Goodell, P.C., 1983. Preliminary report on the stratigraphy of Lower Cretaceous northern Sierra Las Vacas, Chihuahua,
Mexico. Geology and Mineral Resources of North-Central Chihuahua,
Clark, K.F., Goodell, P.C. (Eds.). El Paso Geological Society Publication 15, 269±274.
Eguiluz de AntunÄano, S., 1985. Posibilidades petroleras en el Mar Mexicano. IngenierõÂa Petrolera 25, 39±62.
Eguiluz de AntunÄano, S., Campa Uranga, M.F., 1982. Problemas tectoÂnicas
del sector San Pedro del Gallo en los estados de Chihuahua y Durango.
AsociacioÂn Mexicana de GeoÂlogos Petroleros BoletõÂn 34, 5±42.
Fackler-Adams, B.N., Busby, C.J., Mattinson, J.M., 1997. Jurassic magmatism and sedimentation in the Palen Mountains, southeastern California:
implications for regional tectonic controls on the Mesozoic continental
arc. Geological Society of America Bulletin 109, 1464±1484.
Ferguson, R.C., 1987. Bisbee Group (Lower Cretaceous) carbonates,
Guadalupe Canyon area, southeastern Arizona. Mesozoic Rocks of
Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A.
(Eds.). Arizona Geological Society Digest 18, 365±370.
Gastil, G., Wracher, M., Strand, G., Kear, L.L., Eley, D., Chapman, D.,
Anderson, C., 1992. The tectonic history of the southwestern United
States and Sonora, Mexico, during the past 100 M.Y. Tectonics 11,
990±997.
Gleason, J.D., Spencer, J.E., Richard, S.M., 1999. Geochemistry of ma®c
dikes and sills from the lower McCoy Mountains Formation, La Paz
County, western Arizona. Arizona Geological Survey Open-File Report
99-1, 24 pp.
GonzaÂlez-LeoÂn, C., 1988. EstratigrafõÂa y geologõÂa estructural de las rocas
sedimentarias CretaÂcicas del aÂrea de Lampazos, Sonora. Instituto de
GeologõÂa, Universidad Nacional AutoÂnoma de MeÂxico, Revista 7,
148±162.
GonzaÂlez-LeoÂn, C., 1994. Early Cretaceous tectono-sedimentary evolution
of the southwestern margin of the Sonora basin. Revista Mexicana de
Ciencias GeoloÂgicas 11, 139±146.
GonzaÂlez-LeoÂn, C., Jacques-Ayala, C., 1988. EstratigrafõÂa de las rocas
CretaÂcicas del aÂrea de Cerro de Oro, Sonora central. Universidad de
Sonora BoletõÂn del Departamento de GeologõÂa 5, 1±23.
GonzaÂlez-LeoÂn, C., Jacques-Ayala, C., 1990. PaleogeografõÂa del CretaÂcico
499
Inferior de Sonora. Universidad AutoÂnoma de Nuevo LeoÂn (Linares)
Facultad de Ciencias del Tierra Actas, 125±152.
GonzaÂlez-LeoÂn, C.M., Lawton, T.F., 1995. Stratigraphy, depositional
environments, and origin of the Cabullona basin, northeastern Sonora.
Studies on the Mesozoic of Sonora and Adjacent Areas, Jacques-Ayala,
C., GonzaÂlez-LeoÂn, C.M., RoldaÂn-Quintana, J. (Eds.). Geological
Society of America Special Paper 301, 121±142.
GonzaÂlez-LeoÂn, C., Lucas, S.G., 1995. Stratigraphy and paleontology of the
Early Cretaceous Cerro de Oro Formation, central Sonora. Studies on
the Mesozoic of Sonora and Adjacent Areas, Jacques-Ayala, C., GonzaÂlez-LeoÂn, C., RoldaÂn-Quintana, J. (Eds.). Geological Society of America Special Paper 301, 41±47.
GonzaÂlez-LeoÂn, C.M., RoldaÂn-Quintana, J., RodrõÂguez, E.P., 1992. Deformaciones Sevier y Laramide: Su presencia en Sonora: Universidad de
Sonora BoletõÂn del Departamento de GeologõÂa 9 (1), 1±18.
GonzaÂlez-LeoÂn, C.M., McIntosh, W.C., Lozano-Santacruz, R., ValenciaMoreno, M., Amaya-MartõÂnez, R., RodrõÂguez-CastanÄeda, J., 2000.
Cretaceous and Tertiary sedimentary, magmatic, and tectonic evolution
of north-central Sonora (Arizpe and Bacanuchi Quadrangles), northwest Mexico. Geological Society of America Bulletin 112, 600±610.
Goodlin, T.C., 1985. Stratigraphic and structural relations of the area south
of Hot Springs Canyon, southern Galiuro Mountains, Arizona. Unpublished master's thesis, University of Arizona, 101 pp.
Goodlin, T.C., Mark, R.A., 1987. Tectonic implications of stratigraphy and
structure of Cretaceous rocks and overlying mid-Tertiary cover in Hot
Springs Canyon, Galiuro Mountains, southeastern Arizona. Mesozoic
Rocks of Southern Arizona and Adjacent Areas, Dickinson, W.R.,
Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 177±188.
GoÈtte, M., Michalzik, D., 1992. Stratigraphic relations and facies sequences
of an Upper Jurassic evaporitic ramp in the Sierra Madre Oriental
(Mexico). Zentralblatt fuÈr Geologie und PalaÈontologie (Teil I) 1991,
1445±1466.
Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., Van Veer, P.,
Thierry, J., Huang, Z., 1994. A Mesozoic time scale. Journal of Geophysical Research 99, 24051±24074.
Gries, J.C., Haenggi, W.T., 1970. Structural evolution of the eastern
Chihuahua tectonic belt. In: Seewald, K., Sundeen, D. (Eds.). The
Geologic Framework of the Chihuahua Tectonic Belt. West Texas
Geological Society, Midland, pp. 119±137.
Grijalva-Noriega, J., 1991. Sobre el CretaÂcico Temprano en Sonora y aÂreas
adyacentes. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 8 (2), 1±18.
Grijalva-Noriega, F.J., 1995. El alto Cananea: Un elemento paleogeogra®co del JuraÂsico TardioÂ. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 12 (2), 1±16.
Grijalva-Noriega, F.J., 1996. The Glance episode: a new proposed diachronic unit. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa
13 (1), 75±88.
Harding, L.E., 1980. Petrology and tectonic setting of the Livingston Hills
Formation. Studies in Western Arizona, Jenney, J.P., Stone, C. (Eds.).
Arizona Geological Society Digest 12, 135±145.
Harding, L.E., 1982. Tectonic signi®cance of the McCoy Mountains
Formation, southeastern California and southwestern Arizona. Unpublished PhD dissertation, University of Arizona, 197 pp.
Harding, L.E., Coney, P.J., 1985. The geology of the McCoy Mountains
Formation, southeastern California and southwestern Arizona. Geological Society of America Bulletin 96, 755±769.
Haxel, G., Wright, J.E., May, D.J., Tosdal, R.M., 1980. Reconnaissance
geology of the Mesozoic and lower Cenozoic rocks of the southern
Papago Indian Reservation, Arizona: a preliminary report. Studies in
Western Arizona, Jenney, J.P., Stone, C. (Eds.). Arizona Geological
Society Digest 12, 17±29.
Haxel, G.B., Tosdal, R.M., May, D.J., Wright, J.E., 1984. Latest Cretaceous
and early Tertiary orogenesis in south-central Arizona: Thrust faulting,
regional metamorphism, and granitic plutonism. Geological Society of
America Bulletin 95, 631±653.
Haxel, G.B., Tosdal, R.M., Dillon, J.T., 1985. Tectonic setting and
500
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
lithology of the Winterhaven Formation: a new Mesozoic stratigraphic
unit in southeasternmost California and southwestern Arizona. US
Geological Survey Bulletin 1599, 19.
Hayes, M.J., 1987. Depositional history of the Upper Cretaceous Fort Crittenden Formation in southeastern Arizona. Mesozoic Rocks of Southern
Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.).
Arizona Geological Society Digest 18, 315±325.
Hayes, P.T., 1970. Cretaceous paleogeography of southeastern Arizona and
adjacent areas. US Geological Survey Professional Paper 658-B, B1±
B42.
Hennings, P.H., 1994. Structural transect of the southern Chihuahua fold
belt between Ojinaga and Aldama, Chihuahua, Mexico. Tectonics 13,
1445±1460.
Hoisch, T.D., Heizler, M.T., Harrison, T.M., Miller, C.F., Stoddard, E.F.,
1988. Late Cretaceous regional metamorphism in southeastern California. In: Ernst, W.G. (Ed.). Metamorphism and Crustal Evolution of the
Western United States. Prentice-Hall, Englewood Cliffs, NJ, pp. 538±
571.
Ingersoll, R.V., Bullard, T.F., Ford, R.I., Grimm, J.P., Pickle, J.D., Sares,
S.W., 1984. The effect of grain size on detrital modes: A test of the
Gazzi±Dickinson point-counting method. Journal of Sedimentary
Petrology 54, 103±116.
Inman, K.F., 1987. Depositional environments and sandstone petrography
of Cretaceous sedimentary rocks, Adobe Canyon, Santa Rita Mountains, southeastern Arizona. Mesozoic Rocks of Southern Arizona and
Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 301±314.
Jacques-Ayala, C., 1989. Arroyo SaÂsabe Formation (Aptian±Albian),
northwestern Sonora, Mexico Ð marginal marine sedimentation in
the Sonora back-arc basin. Instituto de GeologõÂa, Universidad Nacional
AutoÂnoma de MeÂxico, Revista 8, 171±178.
Jacques-Ayala, C., 1992a. The Lower Cretaceous Bisbee Group in the
surroundings of Santa Ana. Geology and Mineral Resources of the
Sierra Madre Occidental, Mexico, Clark, K.F., RoldaÂn-Quintana, J.,
Schmitt, H.R. (Eds.). El Paso Geological Society Guidebook, El Paso,
349±357.
Jacques-Ayala, C., 1992b. The Lower Cretaceous Glance Conglomerate
and Morita Formation of the Sierra El Chanate, northwestern Sonora.
Instituto de GeologõÂa, Universidad Nacional AutoÂnoma de MeÂxico,
Revista 10, 37±46.
Jacques-Ayala, C., 1992c. Stratigraphy of the Lower Cretaceous Cintura
Formation. Sierra El Chanate, northwestern Sonora, Mexico. Instituto
de GeologõÂa, Universidad Nacional AutoÂnoma de MeÂxico, Revista 10,
129±136.
Jacques-Ayala, C., 1995. Paleogeography and provenance of the Lower
Cretaceous Bisbee Group in the Caborca Ð Santa Ana area, northwestern Sonora. Studies on the Mesozoic of Sonora and Adjacent
Areas, Jacques-Ayala, C., GonzaÂlez-LeoÂn, C., RoldaÂn-Quintana, J.
(Eds.). Geological Society of America Special Paper 301, 79±109.
Jacques-Ayala, C., Potter, P.E., 1987. Stratigraphy and paleogeography of
Lower Cretaceous rocks, Sierra El Chanate, northwest Sonora, Mexico.
Mesozoic Rocks of Southern Arizona and Adjacent Areas, Dickinson,
W.R., Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 203±
214.
Jacques-Ayala, C., GarcõÂa y BorragaÂn, J.C., De Jong, K.A., 1990. CaborcaAltar geology: Cretaceous sedimentation and compression, Tertiary
uplift and extension. Geological Excursions Through the Sonoran
Desert Region, Arizona and Sonora, Gehrels, G.E., Spencer, J.E.
(Eds.). Arizona Geological Survey Special Paper 7, 165±182.
James, W.C., Russo, J.F., 1988. Diagenesis of a mixed arenite petrofacies,
Lobo Formation, southwestern New Mexico. Cretaceous and Laramide
Tectonic Evolution of Southwestern New Mexico, Mack, G.H.,
Lawton, T.F., Lucas, S.G. (Eds.). New Mexico Geological Society
Annual Field Conference Guidebook 39, 160±173.
Jamison, K., 1987. Petrofacies of Morita Formation (Bisbee Group), southeastern Arizona and northern Sonora, Mexico. Mesozoic Rocks of
Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A.
(Eds.). Arizona Geological Society Digest 18, 257±262.
Janecke, S.U., 1987. Mesozoic and early Tertiary structural history of the
Geesaman Wash area, northeastern Catalina-Rincon metamorphic core
complex. Mesozoic Rocks of Southern Arizona and Adjacent Areas,
Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological Society
Digest 18, 189±201.
Jones, N.W., McKee, J.W., MaÂrquez, D.B., Tovar, J., Long, L.E., Laudon,
T.M., 1984. The Mesozoic La Mula island, Coahuila, Mexico. Geological Society of America Bulletin 95, 1226±1241.
Keith, S.B., Barrett, L.F., 1976. Tectonics of the central Dragoon Mountains: a new look. Arizona Geological Society Digest 10, 169±204.
Keith, S.B., Wilt, J.C., 1985. Late Cretaceous and Cenozoic orogenesis of
Arizona and adjacent regions: a strato-tectonic approach. Cenozoic
Paleogeography of the West-Central United States, Flores, R.M.,
Kaplan, S.S. (Eds.). Rocky Mountain Section (SEPM) Rocky Mountain
Paleogeography Symposium 3, 403±437.
Keith, S.B., Wilt, J.C., 1986. Laramide orogeny in Arizona and adjacent
regions: a strato-tectonic synthesis. Frontiers in Geology and Ore
Deposits of Arizona and the Southwest, Beatty, B., Wilkinson, P.A.K.
(Eds.). Arizona Geological Society Digest 16, 502±554.
Klute, M.A., 1987. Tectonic signi®cance of sandstone petrofacies within
the Bisbee basin of southeastern Arizona. Mesozoic Rocks of Southern
Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.).
Arizona Geological Society Digest 18, 263±272.
Klute, M.A., 1991. Sedimentology, sandstone petrofacies, and tectonic
setting of the late Mesozoic Bisbee basin, southeastern Arizona. Unpublished PhD dissertation, University of Arizona, 268 pp.
Kluth, C.F., Butler, R.F., Harding, L.E., Sha®qullah, M., Damon, P.E.,
1982. Paleomagnetism of Late Jurassic rocks in the northern Canelo
Hills, southeastern Arizona. Journal of Geophysical Research 87,
7079±7086.
Krantz, R.W., 1989. Laramide structures of Arizona. In: Jenney, J.P.,
Reynolds, S.J. (Eds.). Geologic Evolution of Arizona. Arizona Geological Society Digest, 17. , pp. 463±483.
Krebs, C.K., Ruiz, J., 1987. Geochemistry of the Canelo Hills Volcanics
and implications for the Jurassic tectonic setting of southeastern
Arizona. Mesozoic Rocks of Southern Arizona and Adjacent Areas,
Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological Society
Digest 18, 139±151.
Laubach, S.E., Reynolds, S.J., Spencer, J.E., 1987. Mesozoic stratigraphy
of the Granite Wash Mountains, west-central Arizona. Mesozoic Rocks
of Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A.
(Eds.). Arizona Geological Society Digest 18, 91±100.
Lawton, T.F., 2000. Inversion of Late Jurassic±Early Cretaceous extensional faults of the Bisbee basin, southeastern Arizona and southwestern
New Mexico. Southwest Passage Ð A Trip Through the Phanerozoic,
Lawton, T.F., McMillan, N.J., McLemore, V.T. (Eds.). New Mexico
Geological Society Annual Field Conference Guidebook 51, 95±102.
Lawton, T.F., Clemons, R.E., 1992. Klondike basin Ð late Laramide depocenter in southern New Mexico. New Mexico Geology 14, 1±7.
Lawton, T.F., Dickinson, W.R., 1999. Late Paleozoic to Paleogene evolution of the southwest border region (USA±Mexico): implications for
regional tectonics. Geological Society of America Abstracts with
Programs 31 (7), A-294.
Lawton, T.F., Harrigan, P.J., 1997. Jurassic strata of the Little Hatchet
Mountains, Hidalgo County, New Mexico. New Mexico Geology 19,
57.
Lawton, T.F., Harrigan, P.J., 1998. Broken Jug Formation Ð rede®nition of
lower part of Bisbee Group, Little Hatchet Mountains, Hidalgo County,
New Mexico. New Mexico Geology 20, 69±77.
Lawton, T.F., McMillan, N.J., 1999. Arc abandonment as a cause for
passive continental rifting: comparison of the Jurassic Borderland rift
and the Cenozoic Rio Grande rift. Geology 27, 779±782.
Lawton, T.F., Olmstead, G.A., 1995. Stratigraphy and structure of part of
the Bisbee Group, northeastern Chiricahua Mountains, Arizona. Studies
on the Mesozoic of Sonora and Adjacent Areas, Jacques-Ayala,
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
C., GonzaÂlez-LeÂon, C.M., RoldaÂn-Quintana, J. (Eds.). Geological
Society of America Special Paper 301, 21±39.
Lawton, T.F., Basabilvaso, G.T., Hodgson, S.A., Wilson, D.A., Mack,
G.H., McIntosh, W.C., Lucas, S.G., Kietzke, K.K., 1993. Laramide
stratigraphy of the Little Hatchet Mountains, southwestern New
Mexico. New Mexico Geology 15, 9±15.
Lehman, T.M., 1986. Late Cretaceous sedimentation in trans-Pecos Texas.
West Texas Geological Society Bulletin 25, 4±9.
Lehman, T.M., 1991. Sedimentation and tectonism in the Laramide
Tornillo basin of west Texas. Sedimentary Geology 75, 9±28.
Lehmann, C., Osleger, D.A., MontanÄez, I.P., Sliter, W., Arnaud-Vanneau,
A., Banner, J., 1999. Evolution of Cupido and Coahuila carbonate platforms, Early Cretaceous, northeastern Mexico. Geological Society of
America Bulletin 111, 1010±1029.
Lehmann, C., Osleger, D.A., MontanÄez, I., 2000. Sequence stratigraphy of
Lower Cretaceous (Barremian±Albian) carbonate platforms of northeastern Mexico: regional and global correlations. Journal of Sedimentary Research 70, 373±391.
Lindberg, F.A., 1987. Cretaceous sedimentary geology of the Rucker
Canyon area, Cochise County, Arizona. Mesozoic Rocks of Southern
Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.).
Arizona Geological Society Digest 18, 283±299.
Lipman, P.W., 1993. Geologic map of the Tucson Mountains caldera,
southern Arizona. US Geological Survey Miscellaneous Investigations
Series Map I-2205, scale 1:24,000.
Lipman, P.W., Hagstrum, J.T., 1992. Jurassic ash-¯ow sheets, calderas, and
related intrusions of the Cordilleran volcanic arc in southeastern
Arizona: implications for regional tectonics and ore deposits. Geological Society of America Bulletin 104, 32±39.
Lucas, S.G., 2000. Lower Cretaceous ammonites from southwestern New
Mexico. Southwest Passage Ð A Trip Through the Phanerozoic,
Lawton, T.F., McMillan, N.J., McLemore, V.T. (Eds.). New Mexico
Geological Society Annual Field Conference Guidebook 51, 195±201.
Lucas, S.G., Estep, J.W., 1998a. Cretaceous stratigraphy and biostratigraphy in the southern San Andres Mountains, DonÄa Ana County, New
Mexico. Las Cruces Country II, Mack, G.H., Austin, G.S., Baker, J.M.
(Eds.). New Mexico Geological Society Annual Field Conference
Guidebook 49, 187±196.
Lucas, S.G., Estep, J.W., 1998b. Lithostratigraphy and biostratigraphy of
the Lower-middle Cretaceous Bisbee Group, southwestern New
Mexico, USA. New Mexico Museum of Natural History and Science
Bulletin 14, 39±55.
Lucas, S.G., Estep, J.W., 2000a. Lower Cretaceous stratigraphy, correlation, and paleogeography of New Mexico. New Mexico Museum of
Natural History and Science Bulletin 16, 45±61.
Lucas, S.G., Estep, J.W., 2000b. Stratigraphy, paleontology and correlation
of the Lower Cretaceous Carbonate Hill Member of the U-Bar Formation, East Potrillo Mountains, New Mexico. New Mexico Museum of
Natural History and Science Bulletin 16, 81±86.
Lucas, S.G., Lawton, T.F., 2000. Stratigraphy of the Bisbee Group (Jurassic±Cretaceous), Little Hatchet Mountains, New Mexico. Southwest
Passage Ð A Trip Through the Phanerozoic, Lawton, T.F., McMillan,
N.J., McLemore, V.T. (Eds.). New Mexico Geological Society Annual
Field Conference Guidebook 51, 175±194.
Lucas, S.G., Kues, B.S., Hayden, S.N., Allen, B.D., Kietzke, K.K.,
Williamson, T.E., Sealey, P., Pence, R., 1988. Cretaceous stratigraphy
and biostratigraphy, Cooke's Range, Luna County, New Mexico.
Cretaceous and Laramide Tectonic Evolution of Southwestern New
Mexico, Mack, G.H., Lawton, T.F., Lucas, S.G. (Eds.). New Mexico
Geological Society Annual Field Conference Guidebook 39, 143±167.
Lucas, S.G., Estep, J.W., Anderson, B.G., 2000a. Lower Cretaceous stratigraphy and biostratigraphy in the Victorio Mountains, Luna County,
New Mexico. New Mexico Museum of Natural History and Science
Bulletin 16, 63±72.
Lucas, S.G., Estep, J.W., Boucher, L.D., Anderson, B.G., 2000b. Cretaceous stratigraphy, biostratigraphy and depositional environments near
501
Virden, Hidalgo County, New Mexico. New Mexico Museum of
Natural History and Science Bulletin 16, 107±119.
Lucas, S.G., Zeigler, K.E., Lawton, T.F., Filkorn, H.F., 2001. Late Jurassic
invertebrate fossils from the Little Hatchet Mountains, southwestern
New Mexico. New Mexico Geology 23, 16±20.
Mack, G.H., 1987a. Mid-Cretaceous (late Albian) change from rift to retroarc foreland basin in southwestern New Mexico. Geological Society of
America Bulletin 98, 507±514.
Mack, G.H., 1987b. Tectonic, eustatic, and climatic control on Cretaceous
(Aptian to Turonian) sedimentation in southwestern New Mexico.
Mesozoic Rocks of Southern Arizona and Adjacent Areas, Dickinson,
W.R., Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 371±
379.
Mack, G.H., Clemons, R.E., 1988. Structural and stratigraphic evidence for
the Laramide (early Tertiary) Burro uplift in southwestern New Mexico.
Cretaceous and Laramide Tectonic Evolution of Southwestern New
Mexico, Mack, G.H., Lawton, T.F., Lucas, S.G. (Eds.). New Mexico
Geological Society Annual Field Conference Guidebook 39, 59±66.
Mack, G.H., Kolins, W.B., Galemore, J.A., 1986. Lower Cretaceous stratigraphy, depositional environments, and sediment dispersal in southwestern New Mexico. American Journal of Science 286, 309±331.
Mack, G.H., Galemore, J.A., Kaczmarek, E.L., 1988. The Cretaceous foreland basin in southwestern New Mexico. Cretaceous and Laramide
Tectonic Evolution of Southwestern New Mexico, Mack, G.H.,
Lawton, T.F., Lucas, S.G. (Eds.). New Mexico Geological Society
Annual Field Conference Guidebook 39, 135±141.
Mack, G.H., Kottlowski, F.E., Seager, W.R., 1998. The stratigraphy of
south-central New Mexico. Las Cruces Country II, Mack, G.H., Austin,
G.S., Barker, J.M. (Eds.). New Mexico Geological Society Annual
Field Conference Guidebook 49, 135±154.
Mann, J.D., 1995. Stratigraphy, petrology, and depositional environments
of the Upper Cretaceous Rucker Formation, southeastern Arizona.
Unpublished master's thesis, New Mexico State University, Las Cruces,
114 pp.
Mark, R.A., 1985. Structural and sedimentary geology of the area north of
Hot Springs Canyon, southern Galiuro Mountains, Cochise County,
Arizona. Unpublished master's thesis, University of Arizona, Tucson,
96 pp.
Marton, G., Buf¯er, R.T., 1994. Jurassic reconstruction of the Gulf of
Mexico basin. International Geology Review 36, 545±586.
Marvin, R.F., Naeser, C.W., Mehnert, H.H., 1978. Tabulation of radiometric ages Ð including unpublished K±Ar and ®ssion-track ages Ð
for rocks in southeastern Arizona and southwestern New Mexico. Land
of Cochise Ð southeastern Arizona, New Mexico Geological Society
Annual Field Conference Guidebook 29, 243±252.
McBride, E.F., Weidie, A.E., Wolleben, J.A., Laudon, R.C., 1974. Stratigraphy and structure of the Parras and La Popa basins, northeastern
Mexico. Geological Society of America Bulletin 84, 1603±1622.
McFarlan Jr, E., Menes, L.S., 1991. Lower Cretaceous. The Gulf of Mexico
Basin, Salvador, A. (Ed.). Geological Society of America Geology of
North America J, 181±204.
McKee, M.B., Anderson, T.H., 1998. Mass±gravity deposits in the Lower
Cretaceous of Sonora, Mexico. Geological Society of America Bulletin
110, 1516±1529.
McKee, J.W., Jones, N.W., Long, L.E., 1984. History of recurrent activity
along a major fault in northeastern Mexico. Geology 12, 103±107.
McKee, J.W., Jones, N.W., Long, L.E., 1990. Stratigraphy and provenance
of strata along the San Marcos fault, central Coahuila, Mexico. Geological Society of America Bulletin 102, 593±614.
Michalzik, D., Schumann, D., 1994. Lithofacies relations and palaeoecology of a Late Jurassic to Early Cretaceous fan delta to slope depositional system in the Sierra Madre Oriental of north-east Mexico.
Sedimentology 41, 463±477.
Miller, D.M., Haxel, G.B., Tosdal, R.M., 1992. Sonoran and Mojave
Deserts region. The Cordilleran Orogen, Burch®el, B.C., Lipman,
P.W., Zoback, M.L. (Eds.). Conterminous US Geological Society of
America Geology of North America G-3, 230±237.
502
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Minjarez Sosa, I., 1991. La orogenõÂa mesocretaÂcica en Sonora: Universidad
de Sonora BoletõÂn del Departamento de GeologõÂa 8 (2), 19±36.
Minjarez Sosa, I., Palafox, J.J., MartõÂnez, J.A., RodrõÂguez, B., Villalobos,
R., 1985. Consideraciones respecto a la estratigrafõÂa y estructura del
aÂrea de Sahuaripa-Arivechi. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 2, 90±105.
Molenaar, C.M., 1983. Major depositional cycles and regional correlations
of Upper Cretaceous rocks, southern Colorado Plateau and adjacent
areas. Mesozoic Paleogeography of the West-Central United States,
Reynolds, M.W., Dolly, E.D. (Eds.). Rocky Mountain Section
(SEPM) Rocky Mountain Paleogeography Symposium 2, 201±224.
Monreal, R., 1994. Facies changes, unconformities, stratigraphic juxtapositions and their tectonic implications of the Cretaceous of Cerro de Oro,
central Sonora, Mexico. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 11 (1), 1±30.
Monreal, R., 1995. Las facies marinas (Aptiano±Albiano) del Grupo
Bisbee y cronocorrelativas en Sonora. Universidad de Sonora BoletõÂn
del Departamento de GeologõÂa 12 (1), 65±78.
Monreal, R., 1996. A Cretaceous marine section in east-central Sonora.
Universidad de Sonora BoletõÂn del Departamento de GeologõÂa 13 (2),
11±18.
Monreal-Saavedra, R., 1993. PaleogeografõÂa y tectoÂnica de la cuenca de
Chihuahua durante el CretaÂcico Inferior. Universidad de Sonora BoletõÂn
del Departamento de GeologõÂa 10 (1), 1±20.
Monreal-Saavedra, R., 1997. Microfacies of a Lower Cretaceous marine
succession in Cerro Las Conchas, Sonora, Mexico. Revista Mexicana
de Ciencias GeoloÂgicas 14, 38±49.
Monreal, R., Longoria, J.F., 1999. A revision of the Upper Jurassic and
Lower Cretaceous stratigraphic nomenclature for the Chihuahua trough,
north-central Mexico: implications for lithocorrelations. Mesozoic
Sedimentary and Tectonic History of North-Central Mexico, Bartolini,
C., Wilson, J.L., Lawton, T.F. (Eds.). Geological Society of America
Special Paper 340, 69±92.
Monreal, R., Longoria, J.F., 2000a. Stratigraphy and structure of the Lower
Cretaceous of Lampazos, Sonora (northwest Mexico), and its relationship to the Gulf Coast succession. American Association of Petroleum
Geologists Bulletin 84, 1811±1831.
Monreal, R., Longoria, J.F., 2000b. Geology of the Lower Cretaceous rocks
of Sierra Los Chinos, east-central Sonora, Mexico. GeofõÂsica Internacional 39, 309±322.
Monreal, R., Valenzuela, M., GonzaÂlez-LeoÂn, C., 1994. A revision of the
stratigraphic nomenclature for the Cretaceous of northern Sonora, and
some paleogeographic implications. Universidad de Sonora BoletõÂn del
Departamento de GeologõÂa 11 (1), 171±190.
Mora Villalobos, M. de C., 1997. Estudio sistemaÂtico de amonitas CretaÂÂ reas de Lampazos y Cerro de las Concha. Universicicas de Sonora: A
dad de Sonora BoletõÂn del Departamento de GeologõÂa 14 (2), 41±66.
MoraÂn-Zenteno, D., 1994. The geology of the Mexican republic. American
Association of Petroleum Geologists Studies in Geology, 160.
Nourse, J.A., 2001. Tectonic insights from a stretched Upper JurassicLower Cretaceous conglomerate, Caborca±Altar region, Sonora,
Mexico. Journal of South American Earth Sciences 14, 453±474.
Olmstead, G.A., 1992. Depositional environments, lithofacies, and petrology of Lower Cretaceous (Lower to Middle Albian) and Tertiary (Rupelian) strata within the northeastern Chiricahua Mountains, southeastern
Arizona. Unpublished master's thesis, New Mexico State University,
Las Cruces, 101 pp.
Olmstead, G.A., Young, K., 2000. Late Jurassic ammonites from the northeastern Chiricahua Mountains, southeast Arizona. New Mexico Geology 22, 1±7.
Padilla y SaÂnchez, R.J., 1986. Post-Paleozoic tectonics of northeast Mexico
and its role in the evolution of the Gulf of Mexico. GeofõÂsica Internacional 25, 157±206.
Palafox, J.J., MartõÂnez, J.A., 1985. EstratigrafõÂa del aÂrea de Arivechi,
Sonora. Universidad de Sonora BoletõÂn del Departamento de GeologõÂa
2, 30±56.
Palafox, J.J., Mendoza, A., Minjarez, V.A., 1992. GeologõÂa de la regioÂn de
la Sierra Caracahui (Sonora, Mexico). Universidad de Sonora BoletõÂn
del Departamento de GeologõÂa 9 (1), 19±34.
Pingitore Jr, N.E., Stege, B.M., Goodell, P.C., LeMone, D.V., 1983. Limestone stratigraphy, central Sierra PenÄa Blanca, Chihuahua, Mexico.
Geology and Mineral Resources of North-Central Chihuahua, Clark,
K.F., Goodell, P.C. (Eds.). El Paso Geological Society Publication
15, 239±246.
Rangin, C., 1977. Sobre la presencõÂa del JuraÂsico Superior con amonitas en
Sonora septentrional. Universidad Nacional AutoÂnoma de MeÂxico,
Instituto de GeologõÂa, Revista 1, 1±4.
Reynolds, S.J., Spencer, J.E., Richard, S.M., Laubach, S.E., 1986. Mesozoic structures in west-central Arizona. Frontiers in Geology and Ore
Deposits of Arizona and the Southwest, Beatty, B., Wilkinson, P.A.K.
(Eds.). Arizona Geological Society Digest 16, 35±51.
Reynolds, S.J., Spencer, J.E., DeWitt, E., 1987. Stratigraphy and U±Th±Pb
geochronology of Triassic and Jurassic rocks in west-central Arizona.
Mesozoic Rocks of Southern Arizona and Adjacent Areas, Dickinson,
W.R., Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 65±
80.
Reynolds, S.J., Richard, S.M., Haxel, G.B., Tosdal, R.M., Laubach, S.E.,
1988. Geologic setting of Mesozoic and Cenozoic metamorphism in
Arizona. In: Ernst, W.G. (Ed.). Metamorphism and Crustal Evolution of
the Western United States. Prentice-Hall, Englewood Cliffs, NJ, pp.
466±501.
Reynolds, S.J., Spencer, J.E., Asmerom, Y., DeWitt, E., Laubach, S.E.,
1989a. Early Mesozoic uplift in west-central Arizona and southeastern
California. Geology 17, 207±211.
Reynolds, S.J., Spencer, J.E., Laubach, S.E., Cunningham, D., Richard,
S.M., 1989. Geologic map, geologic evolution, and mineral deposits
of the Granite Wash Mountains, west-central Arizona. Arizona Geological Survey Open-File Report 89-04, 51 pp.
Reynolds, S.J., Spencer, J.E., Laubach, S.E., Cunningham, D., Richard,
S.M., 1991. Geologic map and sections of the Granite Wash Mountains,
west-central Arizona. Arizona Geological Survey Map, 30, scale
1:24,000.
Richard, S.M., 1992. Detailed geologic map of the upper Apache Wash
area, central southern Plomosa Mountains, west-central Arizona.
Arizona Geological Survey Open-File Report 92-2, scale 1:12,000
with 12 pp. text.
Richard, S.M., Spencer, J.S., 1994. Detailed geologic map and cross
sections of the Ramsey Mine area, southeastern Plomosa Mountains,
west-central Arizona. Arizona Geological Survey Open-File Report 9414, scale 1:12,000 with 16 pp. text.
Richard, S.M., Reynolds, S.J., Spencer, J.E., 1987. Mesozoic stratigraphy
of the Little Harquahala and Harquahala Mountains, west-central
Arizona. Mesozoic Rocks of Southern Arizona and Adjacent Areas,
Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological Society
Digest 18, 101±119.
Richard, S.M., Spencer, J.E., Stone, P., Tosdal, R.M., 1993. Preliminary
geologic map of the southern Plomosa mountains, La Paz County,
Arizona. Arizona Geological Survey Open-File Report 93-9, scale
1:24,000 with 27 pp. text.
Richard, S.M., McWilliams, M.O., Gans, P.B., 1998. 40Ar/ 39Ar dates from
the Harquahala and Little Harquahala Mountains, west-central Arizona;
Part I. Arizona Geological Survey Open-File Report 98-25, 25 pp.
Riggs, N., 1987. Stratigraphy, structure, and geochemistry of Mesozoic
rocks in the Pajarito Mountains, Santa Cruz County, Arizona. Mesozoic
Rocks of Southern Arizona and Adjacent Areas, Dickinson, W.R.,
Klute, M.A. (Eds.). Arizona Geological Society Digest 18, 165±175.
Riggs, N.R., Blakey, R.C., 1993. Early and Middle Jurassic paleogeography
and volcanology of Arizona and adjacent areas. Mesozoic paleogeography of the western United States-II, Dunne, G.C., McDougall, K.A.
(Eds.). Paci®c Section, SEPM (Society for Sedimentary Geology),
Book 71, 347±373.
Riggs, N.R., Busby-Spera, C.J., 1990. Evolution of a multi-vent volcanic
complex within a subsiding arc graben depression: Mount Wrightson
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
Formation, Arizona. Geological Society of America Bulletin 102,
1114±1135.
Riggs, N.R., Busby-Spera, C.J., 1991. Facies analysis of an ancient,
dismembered, large caldera complex and implications for intra-arc
subsidence: Middle Jurassic strata of Cobre Ridge, southern Arizona,
USA. Sedimentary Geology 74, 39±68.
Riggs, N.R., Haxel, G.B., 1990. The Early to Middle Jurassic magmatic arc
in southern Arizona: Plutons to sand dunes. Geologic Excursions
Through the Sonoran Desert Region, Arizona and Sonora, Gehrels,
G.E., Spencer, J.E. (Eds.). Arizona Geological Survey Special Paper
7, 90±103.
Riggs, N.R., Mattinson, J.M., Busby, C.J., 1993. Correlation of Jurassic
eolian strata between the magmatic arc and the Colorado Plateau: new
U±Pb geochronologic data from southern Arizona. Geological Society
of America Bulletin 105, 1231±1246.
Risley, R., 1987. Sedimentation and stratigraphy of the Lower Cretaceous
Amole Arkose, Tucson Mountains, Arizona. Mesozoic Rocks of Southern Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.).
Arizona Geological Society Digest 18, 215±228.
Robison, B.A., 1979. Stratigraphy and petrology of some Mesozoic rocks in
western Arizona. Unpublished master's thesis, University of Arizona,
138 pp.
Robison, B.A., 1980. Description and analysis of Mesozoic Red Beds,
western Arizona and southeastern California. Studies in Western
Arizona, Jenney, J.P., Stone, C. (Eds.). Arizona Geological Society
Digest 12, 147±154.
RodrõÂguez-CastanÄeda, J.L., 1988. EstratigrafõÂa de la regioÂn de Tuape. Instituto de GeologõÂa, Universidad Nacional AutoÂnoma de MeÂxico, Revista
7, 52±66.
RodrõÂguez-CastanÄeda, J.L., 1990. Relaciones estructurales en la parte
centroseptentrional del Estado de Sonora. Instituto de GeologõÂa,
Universidad Nacional AutoÂnoma de MeÂxico, Revista 9, 51±61.
RodrõÂguez-CastanÄeda, J.L., 1991. Mesozoic stratigraphy in north-central
Sonora, Mexico. Universidad de Sonora BoletõÂn del Departamento de
GeologõÂa 8 (1), 13±27.
Rosales-DomõÂnguez, M., del, C., Grajales-Nishimura, J.M., SaÂnchez-Rios,
M.A., GoÂmez-Luna, M.E., DuenÄas, M.A., 1995. Biostratigraphy of the
Lower Cretaceous Bisbee Group, Rancho Culantrillo area, northeastern
Sonora. Studies on the Mesozoic of Sonora and Adjacent Areas,
Jacques-Ayala, C., GonzaÂlez-LeoÂn, C., RoldaÂn-Quintana, J. (Eds.).
Geological Society of America Special Paper 301, 49±57.
Roybal, G.H., 1981. Facies development in a Lower Cretaceous coral-rudist
patch reef (Mural Limestone, southeast Arizona). Mountain Geologist
18, 46±56.
Salvador, A., 1987. Late Triassic-Jurassic paleogeography and origin of
Gulf of Mexico basin. American Association of Petroleum Geologists
Bulletin 71, 419±451.
Salvador, A., 1991. Triassic±Jurassic. The Gulf of Mexico Basin, Salvador,
A. (Ed.). Geological Society of America Geology of North America J,
131±180.
Schreiber Jr, J.F., Scott, R.W., 1987. Lower Cretaceous coral-algal-rudist
patch reefs in southeastern Arizona. Geologic Diversity of Arizona and
its Margins: Excursions to Choice Areas, Davis, G.H., VandenDolder,
E.M. (Eds.). Arizona Geological Survey Special Paper 5, 280±292.
Scott, R.W., 1979. Depositional model of Early Cretaceous coral-algalrudist reefs, Arizona. American Association of Petroleum Geologists
Bulletin 63, 1108±1127.
Scott, R.W., 1987. Stratigraphy and correlation of the Cretaceous Mural
Limestone, Arizona and Sonora. Mesozoic Rocks of Southern Arizona
and Adjacent Areas, Dickinson, W.R., Klute, M.A. (Eds.). Arizona
Geological Society Digest 18, 327±334.
Scott, R.W., GonzaÂlez-LeoÂn, C., 1991. Paleontology and biostratigraphy of
Cretaceous rocks, Lampazos area, Sonora, Mexico. Studies of Sonoran
Geology, PeÂrez-Segura, E., Jacques-Ayala, C. (Eds.). Geological
Society of America Special Paper 254, 51±67.
Scott, R.W., Warzeski, E.R., 1993. An Aptian±Albian shelf ramp, Arizona
and Sonora. Cretaceous Carbonate Platforms, Simo, J.A.T., Scott,
503
R.W., Masse, J.-P. (Eds.). American Association of Petroleum Geologists Memoir 56, 71±79.
Seager, W.R., Mack, G.H., 1986. Laramide paleotectonics of southern New
Mexico. Paleotectonics and Sedimentation in the Rocky Mountain
Region, United States, Peterson, J.A. (Ed.). American Association of
Petroleum Geologists Memoir 41, 669±685.
Seager, W.R., Mack, G.H., 1998. Geology of East Potrillo Mountains and
vicinity, DonÄa Ana County, New Mexico. New Mexico Bureau of
Mines and Mineral Resources Bulletin, 27.
Segerstrom, L., 1987. Geology of the Planchas de Plata area, northern
Sonora, Mexico. In: Dickinson, W.R., Klute, M.A. (Eds.). Mesozoic
Rocks of Southern Arizona and Adjacent Areas. Arizona Geological
Society Digest, 18. , pp. 153±164.
Sherrod, D.R., Koch, R.D., 1987. Structure and stratigraphy of Mesozoic
sedimentary and volcanic rocks in the New Water Mountains, southwestern Arizona. Mesozoic Rocks of Southern Arizona and Adjacent
Areas, Dickinson, W.R., Klute, M.A. (Eds.). Arizona Geological
Society Digest 18, 81±90.
Sherrod, D.R., Koch, R.D., Grubensky, M.J., 1990. Geologic map of the
Vicksburg quadrangle, La Paz County, Arizona. US Geological Survey
Geologic Quadrangle Map GQ-1684, scale 1:62,500.
Smith, C.I., 1970. Lower Cretaceous sedimentation and tectonics of the
Coahuila and West Texas platforms. In: Seewald, K., Sundeen, D.
(Eds.). The Geologic Framework of the Chihuahua Tectonic Belt.
West Texas Geological Society, Midland, pp. 75±82.
Smith, C.I., 1981. Review of the geologic setting, stratigraphy, and facies of
the Lower Cretaceous in northern Mexico. In: Santiago, C.A., Smith,
C.I., Brown, J.B. (Eds.). Lower Cretaceous stratigraphy and structure,
northern Mexico. West Texas Geological Society Publication, 81-74. ,
pp. 1±27.
Sohl, N.F., MartõÂnez, R.E., SalmeroÂn-UrenÄa, P., Soto-Jaramillo, F., 1991.
Upper Cretaceous. The Gulf of Mexico Basin, Salvador, A. (Ed.).
Geological Society of America Geology of North America J, 205±244.
Soreghan, M.J., 1999. Facies distributions within an ancient asymmetric
lake basin: the Apache Canyon Formation, Bisbee basin, Arizona.
Modern and Ancient Lake Systems, Pitman, J.K., Carroll, A.R.
(Eds.). Utah Geological Association Guidebook 26, 163±190.
Spencer, J.E., Richard, S.M., Reynolds, S.J., 1985. Geologic map of the
Little Harquahala Mountains, west-central Arizona. Arizona Geological
Survey Open-File Report 85-9, 18 pp.
Stewart, J.H., Poole, F.G., Ketner, K.B., Madrid, R.J., RoldaÂn-Quintana, J.,
Amaya-MartõÂnez, R., 1990. Tectonics and stratigraphy of the Paleozoic
and Triassic southern margin of North America. Geologic Excursions
Through the Sonoran Desert Region, Arizona and Sonora, Gehrels,
G.E., Spencer, J.E. (Eds.). Arizona Geological Survey Special Paper
7, 183±202.
Stewart, J.H., Amaya-MartõÂnez, R., Stamm, R.G., Wardlaw, B.R., Stanley
Jr, G.D., Stevens, C.H., 1997. Stratigraphy and regional signi®cance of
Mississippian to Jurassic rocks in Sierra Santa Teresa, Sonora, Mexico.
Revista Mexicana de Ciencias GeoloÂgicas 14, 115±135.
Stone, P., Pelka, G.J., 1989. Geologic map of the Palen-McCoy Wilderness
Study Area and vicinity, Riverside County, California. US Geological
Survey Miscellaneous Field Studies Map MF-2092, scale 1:62,500.
Stone, P., Page, V.M., Hamilton, W., Howard, K.A., 1987. Cretaceous age
of the upper part of the McCoy Mountains Formation, southeastern
California and southwestern Arizona, and its tectonic signi®cance:
reconciliation of paleobotanical and paleomagnetic evidence. Geology
15, 561±564.
Sumpter, L.T., 1986. Stratigraphy and sedimentology of the Willow
Canyon Formation, southeastern Arizona. Unpublished master's thesis,
University of Arizona, Tucson, 127 pp.
Tosdal, R.M., 1990. Constraints on the tectonics of the Mule Mountains
thrust system, southeast California and southwest Arizona. Journal of
Geophysical Research 95, 20,025±20,048.
Tosdal, R.M., Stone, P., 1994. Stratigraphic relations and U-Pb geochronology of the Upper Cretaceous upper McCoy Mountains Formation,
504
W.R. Dickinson, T.F. Lawton / Journal of South American Earth Sciences 14 (2001) 475±504
southwestern Arizona. Geological Society of America Bulletin 106,
476±491.
Tosdal, R.M., Haxel, G.B., Wright, J.E., 1989. Jurassic geology of the
Sonoran Desert region, southern Arizona, southeastern California,
northernmost Sonora: construction of a continental-margin magmatic
arc. Geologic Evolution of Arizona, Jenney, J.P., Reynolds, S.R. (Eds.).
Arizona Geological Society Digest 17, 397±434.
Tosdal, R.M., Haxel, G.B., Anderson, T.M., Connors, C.D., May, D.J.,
Wright, J.E., 1990. Highlights of Jurassic, Late Cretaceous to early
Tertiary and middle Tertiary tectonics, south-central Arizona and
north-central Sonora. Geologic Excursions Through the Sonoran Desert
Region, Arizona and Sonora, Gehrels, G.E., Spencer, J.E. (Eds.).
Arizona Geological Survey Special Paper 7, 76±88.
Underwood Jr, J.R., 1980. Geology of the Eagle Mountains, Hudspeth
County, Texas. Trans-Pecos Region, Dickerson, P.W., Hoffer, J.M.
(Eds.). New Mexico Geological Society Annual Field Conference
Guidebook 31, 183±193.
Valencia GoÂmez, V.A., 1994. EstratigrafõÂa y microfacies del CretaÂcico
Inferior de Lampazos, Sonora, Mexico. Universidad de Sonora BoletõÂn
del Departamento de GeologõÂa 11 (2), 59±86.
Vedder, L.K., 1984. Stratigraphic relationship between the Late Jurassic
Canelo Hills Volcanics and the Glance Conglomerate, southeastern
Arizona. Unpublished master's thesis, University of Arizona, Tucson,
129 pp.
Warzeski, E.R., 1987. Revised stratigraphy of the Mural Limestone: a
Lower Cretaceous carbonate shelf in Arizona and Sonora. Mesozoic
Rocks of Arizona and Adjacent Areas, Dickinson, W.R., Klute, M.A.
(Eds.). Arizona Geological Society Digest 18, 335±363.
Wilson, D.A., 1991. Depositional environments, provenance, and age of the
Skunk Ranch Formation, Little Hatchet Mountains, southwestern New
Mexico. Unpublished master's thesis, New Mexico State University,
Las Cruces, 87 pp.
Wilson, J.L., 1990. Basement structural controls on Mesozoic carbonate
facies in northeastern Mexico Ð a review. Carbonate Platforms;
Facies, Sequences, and Evolution, Tucker, M.E., Wilson, J.L., Crevello,
P.D., Sarg, J.R., Read, J.F. (Eds.). International Association of Sedimentologists Special Publication 9, 235±255.
Wilson, J.L., 1999. Controls on the wandering path of the Cupido reef trend
in northeastern Mexico. Mesozoic Sedimentary and Tectonic History of
North-Central Mexico, Bartolini, C., Wilson, J.L., Lawton, T.F. (Eds.).
Geological Society of America Special Paper 340, 135±143.
Wilson, J.L., Ward, W.C., 1993. Early Cretaceous carbonate platforms of
northeastern and east-central Mexico. Cretaceous Carbonate Platforms,
Simo, J.A.T., Scott, R.W., Masse, J.-P. (Eds.). American Association of
Petroleum Geologists Memoir 56, 35±49.
Wilson, J.L., Ward, W.C., Finneran, J., 1984. Introduction. In: Wilson, J.L.,
Ward, W.C., Finneran, J. (Eds.). A Field Guide to Upper Jurassic and
Lower Cretaceous Carbonate Platform and Basin Systems, MonterreySaltillo Area, Northeast Mexico. Gulf Coast Section (SEPM) Foundation, Houston, pp. 1±27.
Winker, C.D., Buf¯er, R.T., 1988. Paleogeographic evolution of early deepwater Gulf of Mexico and margins, Jurassic to middle Cretaceous
(Comanchean). American Association of Petroleum Geologists Bulletin
72, 318±346.
Yarnold, J.C., Lombard, J.P., 1989. Facies model for large rock avalanche
deposits formed in dry climates. Conglomerates in Basin Analysis,
Colburn, I.P., Abbott, P.L., Minch, J. (Eds.). Paci®c Section (SEPM)
Book 62, 9±31.
Zwanziger, J.A., 1992. New concepts in Mesozoic stratigraphy of Chihuahua. In: Goodell, P.C., GutieÂrrez, C., Reyes-Cortes, I. (Eds.). Energy
Resources of the Chihuahuan Desert Region. El Paso Geological
Society, El Paso, pp. 77±125.