Biochronology, AUTHORS Javier Helenes ~ Departmento de paleoenvironments, and Geolog´ıa, Centro de Investigación Cient´ıfica y de Educación Superior de Ensenada stratigraphic sequences of the late (CICESE), Ensenada, Baja California, México; [email protected] Albian–middle fore-arc Javier Helenes obtained a geological engineering degree from Instituto Politécnico Vizcaino basin, western Baja Nacional, Mexico, and an M.S. degree and Ph.D. from Stanford University, California. He has California, Mexico worked as a biostratigrapher ( and dinoflagellates) in Switzerland, Canada, and Venezuela, and since 1995, he has Javier Helenes, Arturo Martin-Barajas, been a researcher at the Departamento Juan G. Flores-Trujillo, Iraida Paredes, Maritza Canache, de Geolog´ıa at CICESE, where he has studied Ana-Luisa Carreño, and Adriana Miranda –Holocene dinoflagellates from tropical areas. Javier Helenes is the corresponding author of this paper. ABSTRACT Arturo Martin-Barajas ~ Departmento de Geolog´ıa, CICESE, Ensenada, Baja The Vizcaino fore-arc basin accumulated approximately 4 km California, México; [email protected] (~13,000 ft) of upper Albian–middle Eocene siliciclastic marine sedimentary rocks derived from the Peninsular Ranges in Baja Arturo Martin-Barajas obtained a geological engineering degree from Universidad Nacional California. Data from eight exploratory wells document the mi- Autónoma de México (UNAM), and an M.S. cropaleontological content and lithological characteristics of these degree and Doctorate from the Université de – rocks. The strata studied represent mostly neritic upper bathyal Paris-Sud, France. He worked as a geologist marine environments and overlie a basement composed of Creta- in the Consejo de Recursos Minerales, and ceous granitic rocks, or Aptian–Albian volcaniclastic sedimentary since 1990, he has been a researcher at the rocks correlative with the Alisitos Formation. We recognize four Departamento de Geolog´ıa at CICESE, where major depositional sequences within the basin that are related to the he has studied volcanism, tectonics, and regional geology. The basal Albian–Turonian sequence 1 represents stratigraphy of basins around Baja California. the initiation of fore-arc basin sedimentation, contains continental Juan G. Flores-Trujillo ~ Dependencia conglomerates that change to bathyal shales, and correlates with the Académica de Ingenier´ıayTecnolog´ıa, lower part of the Valle Group of the Vizcaino Peninsula. Sequence Universidad Autónoma del Carmen 2 is Coniacian–, includes basal conglomeratic sandstones (UNACAR), Ciudad del Carmen, Campeche, grading into Maastrichtian bathyal shales, and usually overlies México; gfl[email protected] a Coniacian–Santonian unconformity. Sequence 2 is represented at Juan Gabriel Flores-Trujillo obtained a Ph.D. at the surface by the Rosario Group innorthwesternBajaCalifornia CICESE, studying taxonomy, paleoecology of fl and the upper part of the Valle Group in the Vizcaino Peninsula. dino agellates, and their relation to paleo- Sequence 3 is Paleocene–middle Eocene, represents continuity oceanography. Currently, Flores-Trujillo is working at the UNACAR, Campeche, México. of fore-arc sedimentation in neritic–upper bathyal conditions, is Flores-Trujillo is interested in dynamics capped by a major unconformity, and correlates with the Sepultura of coastal processes and environments as and Bateque Formations to the north and south of the basin, well as palynology and climatic change and is a member of the Red Tematica´ sobre Florecimientos Algales Nocivos (Network on Copyright ©2019. The American Association of Petroleum Geologists. All rights reserved. Green Open Harmful Algal Blooms) in México. Access. This paper is published under the terms of the CC-BY license. Manuscript received September 24, 2017; provisional acceptance December 12, 2017; revised manuscript received February 7, 2018; revised manuscript provisional acceptance June 5, 2018; 2nd revised manuscript received June 30, 2018; final acceptance October 2, 2018. DOI:10.1306/10021817349

AAPG Bulletin, v. 103, no. 5 (May 2019), pp. 1045–1069 1045

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Iraida Paredes ~ Stratigraphic respectively. The uppermost Miocene–Pliocene sequence 4 is Consultants, Estado Miranda, Venezuela; composed of marine sandstone–siltstone unconformably overlying [email protected] sequence 3 and is correlative with the Tortugas Formation that Iraida Paredes obtained a B.S. degree and an represents sedimentation after the end of subduction of the M.S. degreeinbiological sciencesinVenezuela. Farallon plate beneath the North America plate. She also received extensive training in palynology at ETH-Zurich¨ in Switzerland, Amsterdam University, Robertson Research, and Petrobras (CENPES) in Rio de Janeiro. INTRODUCTION From 1979 to 2003, she worked as apalynologistandstratigrapheratPetróleosde The western margin of North America was a continuous con- Venezuela-Instituto de Tecnolog´ıa Venezolana vergent zone during subduction of the Farallon plate from the para el Petróleo (PDVSA-INTEVEP), and Early Cretaceous–Paleogene. During this interval, the tectonic she is currently working as a palynological activity in Baja California changed from an extensional arc system consultant in Venezuela. to a contractional arc apron (Busby et al., 1998; Centeno-Garc´ıa Maritza Canache ~ Stratigraphic et al., 2011). In the course of this change, the westernmost Consultants, Estado Miranda, Venezuela; margin developed the Vizcaino fore-arc basin (VB) in central Baja [email protected] California (Figure 1), where a sedimentary infill greater than 3 km > Maritza Canache has a geological engineering thick ( 9,842 ft) was deposited. This important stratigraphic degree from Venezuela, and she has received record provides new information with which to evaluate tec- several training courses at PDVSA-INTEVEP, tonic models and the depositional history of southwestern where she worked from 1996 to 2003 as North America (Gastil et al. 1975, 1978; Rangin, 1978, 1986; a stratigrapher and micropaleontologist Barnes, 1984; Bottjer and Link, 1984; Johnson et al., 2003). (foraminifera) studying all the basins in Several reports have addressed the stratigraphy and sedi- Venezuela. She has also worked material from mentology of strata cropping out along the PacificmarginofBaja Guatemala and México, and she is currently California and southern California (Kilmer, 1977, 1984; Patterson, a consultant in Venezuela. 1978; Beggs, 1984; Boles and Landis, 1984; Kimbrough, 1984, Ana-Luisa Carreño ~ Departamento de 1985; Yeo, 1984; Cunningham and Abbott, 1986; Vazquez-´ Paleontolog´ıa, Institute of Geology, UNAM, Garc´ıa and Schwennicke, 1996; Kimbrough et al., 2001; Ciudad de México, México; anacar@ Gonzalez-Barba,´ 2002; Miranda-Mart´ınez and Carreño, 2008). geologia.unam.mx However, a thicker, larger, and mostly undisrupted part of this Ana-Luisa Carreño obtained a Ph.D. in sedimentary basin lies beneath the continental shelf of central Baja micropaleontology and is working at the California in the VB (Figure 1). Lithostratigraphic, chronostrati- Instituto de Geolog´ıa, UNAM. She is interested graphic, and paleoenvironmental data from this subsurface record in using foraminifera, ostracoda, and calcareous nannoplankton in provide time constraints for regional stratigraphic and tectonic sequences related to the opening of the Gulf of events that marked the evolution of the southern part of the California. She has published over 100 technical fore-arc basin during the Early and Late Cretaceous and Paleogene. and scientific papers, reviews, miscellaneous An exploration program of Pemex® (the Mexican oil company) publications, and book chapters and teaches from 1951 to 1981, drilled several exploration wells in western Baja and directs undergraduate and graduate California, however only limited stratigraphic data from these wells theses. have been published (Lozano-Romen, 1975; Alvarado-de-la-Tejera, Adriana Miranda ~ Departamento de 1976; Cavazos-Prado, 1976; Diaz-Cuevas, 1976; Gonzalez-Garc´ ´ıa, Paleontolog´ıa, Institute of Geology, UNAM, 1976; Ramos-Garc´ıa, 1976; Helenes, 1984). According to these Ciudad de México, México; amiranda@ reports, the VB contains a sedimentary succession that spans the late ciencias.unam.mx Albian–middle Eocene (Lozano-Romen, 1975). Well C penetrated AdrianaMirandaisinterestedinbiostratigraphy 3700 m (12,139 ft) of siliciclastic rocks and did not reach basement of foraminifera and dinoflagellates and is in the deepest part of the VB (Figure 1). currently finishing a Ph.D. on the age of marine This report presents data and interpretations from lithology as deposits related to the evolution of the Gulf of well as micropaleontological and palynological analyses carried out in cuttings and core samples from eight wells in the VB. Based

1046 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 on these results, we construct a detailed chronostratigraphic CaliforniaintheNeogene.Sheisinchargeofthe framework and propose a paleobathymetric evolution of the basin micropaleontology database at the National that help constrain evolutionary models of this continental mar- MuseumofPaleontology,andsheteachesearth gin. In addition, we propose a sequence stratigraphic context sciences and paleobiology at UNAM. regionally correlated with the formal stratigraphic units cropping out onshore from California and south into the Vizcaino Peninsula ACKNOWLEDGMENTS (Figure 1). Petróleos Mexicanos Exploración y Producción financed this study and provided samples and well log data for the analyses, GEOLOGIC FRAMEWORK and the Centro Nacional de Información de Hidrocarburos of the Comisión Nacional de The VB lies on the western margin of the Baja California Peninsula Hidrocarburos (CNH) of México granted (Figure 1) that is south of the continental borderland province permission to publish this manuscript (CNH consent letter no. 271.071 /2016). We thank (Emery, 1960; Krause, 1964, 1965) and northeast of the Vizcaino technician E. Collins (CICESE) for processing Peninsula (Gastil et al., 1975; Lozano-Romen, 1975). The eastern samples and technician V. Frias (CICESE) for margin of the fore-arc basin lies parallel to the western shoreline of assistance with figures. We also thank Baja California south of Punta Canoas (Figure 1). To the west, the reviewers whose comments and suggestions continental margin extends to the Vizcaino Peninsula and Cedros improved our report. Island. To the south, the VB extends into the coastal plain con- fi ned between the Sierra San Andres´ and the Peninsular Ranges DATASHARE 104 batholith (PRB) in Baja California. To the south, a subsurface structural high known as the Lagunitas structural high (Lozano- Supplementary material is available in an electronic version on the AAPG website Romen, 1975) separates the VB from the Pur´ısima Basin to the (www.aapg.org/datashare) as Datashare 104. south (Figure 1). Regional gravimetric anomalies (Lozano-Romen, 1975; Alvarado-de-la Tejera, 1976) show the approximate location and extent of the VB and its southern limit in the Lagunitas high (Lozano-Romen, 1975, figure 2 as “Plano 2”; Alvarado-de-la Tejera, 1976, figures3,4,6,and7).Asoutheast-trending gravimetric minimum offshore of Punta Canoas extends south to offshore of Guerrero Negro, roughly following the northwest–southeast axis of thebasin(Figure1),whereasanalmosteast–west gravimetric maximum shown south of Guerrero Negro indicates the approx- imate position and alignment of the Lagunitas structural high (figure 2 in Lozano-Romen, 1975). Two stages of arc magmatism formed the PRB in Baja California (Silver and Chappell, 1988). The first was an arc apron com- posed of Lower Cretaceous intermediate to mafic volcaniclastic plutonic and sedimentary rocks known as the Alisitos arc (Santillan´ and Barrera, 1930; Allison, 1974; Busby, 2004). The second stage produced an Andean-type continental arc during the Late Cretaceous. Subsequent uplift and erosion have exposed the underlying plutons from southern California to Baja California (Gastil et al., 1975; Silver and Chappell, 1988; Ortega-Rivera et al., 1997). These rocks separate the regional geology into prebatholithic, batholithic, and postbatholithic rock belts. The prebatholithic rocks are represented by a small number of Paleozoic–Lower Cretaceous metasedimentary and volcaniclastic pendants (Gastil, 1983).

HELENES ET AL. 1047

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Figure 1. Location map of the Vizcaino basin (VB) and wells studied (A though H). North of the VB, the San Diego (SD), Ensenada (ENS), and El Rosario areas also contain Upper Cretaceous–Paleogene fore-arc basin strata, but the continental shelf is disrupted by Neogene faulting in the continental borderland. South of VB, the Lagunitas structural high divides the Purisima basin, which also contains Upper Cretaceous–Paleogene fore-arc sedimentary rocks beneath Neogene rocks. GR = the town of Guerrero Negro; SSA = Sierra San Andres in the northern Vizcaino Peninsula; sedim. = sedimentary.

1048 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 The batholithic belt consists of tonalite–diorite plutons of Cretaceous age (140–80 Ma; Ortega-Rivera, 2003). Upper Cretaceous and Cenozoic sedimentary successions unconformably overlie the prebatholithic and batholithic rocks representing the postbatholithic stage (Gastil et al., 1975).

Regional –Cenozoic Stratigraphy

Subduction-related stratigraphic units exposed in northwestern Baja California include the Lower Cretaceous Alisitos Group (Beggs, 1984), the Upper Figure 2. Mesozoic–Cenozoic sedimentary units cropping out Cretaceous Rosario Group (Kennedy and Moore, in San Diego County (1), the state of Baja California (2, 3), Cedros 1971), the Paleocene Sepultura Formation, and Island (4), and Vizcaino Peninsula (Vizc. Penin.) (5, 6). The units the Eocene Delicias and Buenos Aires Formations. shown are formations, except for the La Jolla, Valle, and Alisitos Groups. Information was taken from the following: 1Gastil and Postsubduction sedimentation includes the Miocene 2 Rosarito Beach and the Pliocene–Pleistocene Cantil Higley, 1977; Yeo, 1984; Herzig and Kimbrough, 2014. Acosta, 1966; Yeo, 1984. 3Kilmer, 1963; Boehlke and Abbott, 1986; Costero Formations (Figure 2). To the southwest, in Helenes and Tellez-Duarte,´ 2002. 4Berry and Miller, 1984; Smith the Vizcaino Peninsula and Cedros Island, the Me- et al., 1993b; Kimbrough et al., 2001. 5Patterson, 1984; Kimbrough sozoic stratigraphic succession is different and related et al., 2001; Busby, 2004. Cret. = Cretaceous; Sta. = Santa; volc. = to an accreted island arc terrane (Sedlock, 2003). volcanics. The Lower Cretaceous Alisitos Group (Santillan´ and Barrera, 1930; Allison, 1974; Beggs, 1984) in- north in San Diego County, the marine Upper Cre- cludes volcanic and volcaniclastic sedimentary strata taceous Point Loma and Cabrillo Formations are of island-arc affinity. This unit has been interpreted as composed of siltstone and sandstone (Kennedy and an extensional volcaniclastic plutonic island arc that Moore, 1971; Gastil and Higley, 1977; Nilsen and fringed the continental margin of North America Abbott, 1979, 1981; Yeo, 1982) and are part of the (Busby, 2004). The Alisitos Group is composed Rosario Group. mainly of andesitic–rhyolitic volcaniclastic rocks The Paleocene Sepultura Formation includes ma- (Beggs, 1984; Wetmore et al., 2003; Busby, 2004) rine rocks that crop out in small areas in westcentral with massive biohermal limestones formed when Baja California (Figure 2; Santillan´ and Barrera, volcanic activity decreased by the end of the Early 1930; Gastil et al., 1975; Zinsmeister and Paredes- Cretaceous (Suarez-Vidal,´ 1987). Mej´ıa, 1988; Abbott et al., 1993; Helenes and After collision of the Alisitos arc with the western Tellez-Duarte,´ 2002; Tellez-Duarte´ et al., 2006). margin of North America, the Upper Cretaceous Eocene marine sedimentary rocks are scarce in Rosario Group (Kennedy and Moore, 1971; Patterson, central Baja California although outcrops of the 1978; Yeo, 1982; Miller and Abbott, 1989) was shallow marine sandstones assigned to the Buenos deposited in the El Rosario fore-arc basin (Figure 1). Aires or Delicias Formations have been described These strata represent deposition on a continental in the northwestern part of the peninsula (Flynn, margin (Gastil et al., 1975; Busby, 2004) as indicated 1970; Carreño and Smith, 2007). Farther north, in by the arkosic–lithic composition with evidence the San Diego Embayment of southern California, that indicates continental–upper bathyal environ- Eocene marine sedimentary rocks are assigned to ments in the type area near El Rosario (Kilmer, 1963; the La Jolla Group (Kennedy and Peterson, 1975) Patterson, 1978; Boehlke and Abbott, 1986). Upper and are partially correlative with the Delicias For- Cretaceous stratigraphic sections with similar lithologic mation. No Oligocene marine deposits have been and depositional settings are described north of El reported in the northern Baja California Peninsula. Rosario, along the western margin of Baja California A clastic wedge of Neogene deposits punctuated (Acosta, 1966; Yeo, 1982; Maestas et al., 2003). Farther with volcanic rocks includes the middle Miocene

HELENES ET AL. 1049

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Rosarito Beach Formation that crops out in the MATERIALS AND METHODS northwestern part of the peninsula, from the United States–Mexico border south to Ensenada (Figure 1). Lithology in wells was interpreted based on gamma- This unit consists of flows interbedded with ray, spontaneous potential, and resistivity logs, which pyroclastic and marine epiclastic deposits (Minch, were verified with mud-log reports from Pemex 1967; Carreño and Smith, 2007; Salinas-Marquez´ (Figure 3A, B). Samples for micropaleontological et al., 2016). In addition, the highly fossiliferous analysis mostly represent fine-grained lithofacies and Pliocene Cantil Costero Formation unconformably were selected based on their distribution along the caps the Maastrichtian Rosario Group north of El well. We analyzed 439 cuttings and 8 core samples Rosario (Carreño and Smith, 2007). from 9 exploratory wells, which were processed As mentioned above, the stratigraphy of Mesozoic and analyzed for palynology and foraminifera. Age rocks exposed on the Vizcaino Peninsula and Cedros and paleobathymetric assignments (Appendix 1, sup- Island is different from that in northern Baja California plementary material available as AAPG Datashare 104 (Figure 2). In the Vizcaino region, at www.aapg.org/datashare) are based on age ranges of ophiolite and related plutonic rocks compose the the planktonic and benthonic foraminifera as well as on basement and are tectonically juxtaposed alongside dinoflagellates and spores recovered from these sam- Early Cretaceous–Cenozoic marine strata (Mina, ples (Appendix 2, supplementary material available as 1957; Lozano-Romen, 1975; Helenes, 1984; Abbott AAPG Datashare 104 at www.aapg.org/datashare). et al., 1995; Busby, 2004). The oldest Mesozoic rocks Individual chronostratigraphic tables from each exposed on Cedros Island and the Vizcaino Peninsula well present the paleontological ages assigned and the include a Triassic–Jurassic ophiolitic complex at the paleobathymetric ranges indicated by the micropale- base of the section (Finch and Abbott, 1977; Rangin, ontological assemblages observed. Taxonomy used 1978; Pessagno et al., 1979; Whalen and Pessagno, follows Stewart and Pearson (2000) (PLANKRANGE) 1984; Moore, 1985; Whalen and Carter, 2002). A for foraminifera, Fensome et al. (2008) (DINOFLAJ2) Jurassic tectonic melange´ unconformably overlies the for dinoflagellates, Perch-Nielsen (1985) for calcareous Jurassic ophiolitic rocks (Boles and Landis, 1984; Kilmer, nannofossils, and the various references mentioned for 1984; Kimbrough, 1984, 1985). On the Vizcaino age determination of pollen and spores. Peninsula, Triassic cherts and Jurassic volcaniclastic rocks overlie the ophiolitic rocks (Hickey, 1984; Kilmer, 1984; Moore, 1984, 1985; Kimbrough, Palynological Processing and Analyses 1985). Cretaceous, shallow- to deep-water marine clastic deposits of the Valle Group rest unconformably Samples for palynology were processed following on basement rocks (Mina, 1957; Barnes, 1984; Berry standard treatment (Wood et al., 1996), which in- and Miller, 1984; Patterson, 1984a; Smith and Busby, cluded treatment with HCl, HF, and sodium poly- 1993; Kimbrough et al., 2001). In the Vizcaino tungstate (specificgravity= 2.0), without oxidation. Peninsula, the Valle Group forms two subbasins in the The residue was then sieved through 125-mmand north–south direction, which indicates the uncon- 15-mm mesh sieves, and strew mounts were prepared formity has significant structural relief. Paleocene– with the residue retained in the 15-mmmeshsieve. lower Eocene marine terrigenous deposits crop out One slide per sample was analyzed under the micro- only in the southern part of the Vizcaino Peninsula scope, and numerical abundances of marine and ter- (Abbott et al., 1995; Carreño and Smith, 2007), restrial palynomorphs were recorded. whereas Miocene–Pliocene strata of the Tortugas The assigned ages are based mainly on the known Formation are exposed mainly in the northwestern stratigraphic range of the species of dinoflagellates and part of Vizcaino Peninsula and in Cedros Island (Mina, planktonic foraminifera reported here. Spores and pollen 1957; Robinson, 1975, 1979; Kilmer, 1977, 1979; grains were also observed, and their known stratigraphic Helenes-Escamilla, 1980, 1984). Our study docu- ranges are noted. The absolute ages of the palynomorph ments the stratigraphic succession in the fore-arc basin ranges are calibrated after Gradstein et al. (2012). and complements the patchy stratigraphic record Stratigraphic ranges used for the dinoflagellates previously studied onshore. were compiled from Drugg and Stover (1975),

1050 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 on 24 September 2021 by guest Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf H LNSE AL. ET ELENES

Figure 3. Ages, paleobathymetry, and proposed correlation of wells A, B, C, and D. Numbers 1–4, which are encircled, indicate the four sequences here recognized. Sequence 4 appears only in wells A and D. Small green arrows indicate the maximum flooding surfaces (mfs) of each secuence. The larger blue arrow indicates the deepest mfs of the Vizcaino basin. Consensus ages are indicated as follows: Albi = Albian; Apti = Aptian; Aqui = Aquitanian; Bart = Bartonian; Burd = Burdigalian; Camp = Campanian; Dani = Danian; Gela = Gelasian; Lute = Lutetian; Maas = Maastrichtian; mid = middle; Pria. = Priabonian; Radiom. = Radiometric; Sela = Selandian; Than = Thanetian; Turo = Turonian; Ypre = Ypresian; Zanc = Zanclean. 1051 Lithology logs interpreted from gamma ray (GR, red), resistivity (ILD, blue) and/or spontaneous potential (SP, red) and mud log reports. The paleobathymetric curves show our interpretation of the sedimentary setting at the time of deposition with the symbols shown in the key. Williams and Bujak (1985), Wrenn et al. (1986), and numerical abundances of the taxa were recorded. Helby et al. (1987), Haq et al. (1988), Matsuoka and Stratigraphic ranges used here for the planktonic fo- Bujak (1988), Powell (1992), and Williams et al. raminifera are those shown in Sliter (1968), Kennett (1993, 2004). Most of the stratigraphic ranges used and Srinivasan (1983), Bolli and Saunders (1985), here for the terrestrial palynomorphs are reported in Caron (1985), Toumarkine and Luterbacher (1985), Germeraad et al. (1968), Pares-Regali et al. (1974a, Premoli-Silva and Sliter (1999), and Premoli-Silva et al. b), Muller¨ et al. (1987), and Lorente (1986). Addi- (2004). Additionally, age assignment was also assigned tionally, stratigraphic ranges of all palynomorphs are after considering information in the computer database complemented with information contained in the PLANKRANGE (Stewart and Pearson, 2000). computer databases: TAXON (R. L. Ravn, 1993, per- sonal communication) and the Palynodata datafile (White, 2008). The absolute ages of the palynomorph Paleobathymetric Interpretations ranges are calibrated with Gradstein et al. (2012). Paleobathymetric interpretations are based primarily on information from benthonic foraminifers follow- Palynological Marine Index ing Sliter and Baker (1972) in combination with The initial weight of the samples was low (mean = 9.4 g), lithologic and palynological data. We also consider and thus recovery was generally low to moderate; that the planktonic/benthonic (P/B) ratio increases for this reason, no statistical analyses were done on the with increasing water depth (van der Zwaan et al., microfossil assemblage. Nevertheless, we use a paly- 1990), with lower P/B ratios commonly occurring nological marine index (PMI) as an auxiliary tool with beneath coastal water and in deep-water environ- the paleoenvironmental interpretations (Helenes ments below the foraminiferal lysocline (Akimoto, et al., 1998). The PMI is calculated using the fol- 1994). Integration of all micropaleontology and li- lowing formula: PMI = (DM/DC+1) · 100, where thology data and considerations allowed the sub- DM = diversity or richness of marine palynomorphs division of the paleoenvironments according to the (dinoflagellates and acritarchs) and DC = diversity or following depth intervals: continental above sea level; richness of continental palynomorphs (pollen and transitional between high and low tides; inner ne- spores). Null values of PMI indicate samples without ritic 0–50 m (0–164 ft); middle neritic 50–100 m marine palynomorphs and are interpreted as repre- (164–328 ft); outer neritic 100–200 m (328–656 ft); senting continental environments. Low values of the upper bathyal 200–500 m (656–1640 ft); and middle PMI indicate deposition in transitional environments, bathyal 500–2000 m (1640–6562 ft). whereas higher values indicate marine deposition. Larger values generally correspond to neritic envi- Age Assignment ronments as indicated by benthonic foraminifera. The PMI is useful when comparing values from Consensus ages for the samples from each well were adjacent samples. Large differences indicate changes assigned on the basis of micropaleontological data in depositional environments and help in the presented here and, in some cases, by integrating the definition of sequence boundaries and maximum information provided by Pemex (Figures 3 and 4). flooding surfaces. The absolute ages of the species reported here have been taken from the references mentioned above Foraminiferal Processing and Analyses with ages determined by the correlation of faunal chronozones with radiometric dates and magnetic Samples for foraminiferal analyses were prepared polarity chronozones (Haq et al., 1988; Gradstein according to the normal oxidation procedure et al., 2012). Appendix 1 (supplementary material (Thomas and Murney, 1985) including washing of available as AAPG Datashare 104 at www.aapg.org/ the samples with detergent and treatment with warm datashare) contains biostratigraphic information for

hydrogen peroxide (H2O2). The residue was sieved each well presented in Tables S1–S8 (wells A–H) (0.063 mm) and the larger fraction dried and ana- (supplementary material available as AAPG Data- lyzed. The entire residue was scanned for specimens share 104 at www.aapg.org/datashare) with the main

1052 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 on 24 September 2021 by guest Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf

H – LNSE AL. ET ELENES Figure 4. Ages, paleobathymetry, and correlation of wells E, F, G, and H. Numbers 1 3, which are encircled, indicate the sequences recognized here. Sequence 4 is absent from these wells. Consensus ages are indicated as follows: Albi = Albian; Apti = Aptian; Bart = Bartonian; Camp = Campanian; Dani = Danian; ear = early; Maas = Maastrichtian; Mess = Messinian; mid = middle; Radiom. = Radiometric; Sant = Santonian; Turo = Turonian; Ypre = Ypresian. Lithology logs interpreted from gamma ray (GR, red) resistivity (ILD, blue) and/or spontaneous potential (SP, red) and mud log reports. The paleobathymetric curves show our interpretation of the sedimentary setting at the time of deposition with the symbols shown in the key. 1053 biostratigraphic markers for each interval and their biostratigraphic data indicative of either Turonian–Albian assigned age ranges in Ma. In addition, Appendix 2 or Cenomanian–Albian age ranges. (supplementary material available as AAPG Datashare In well B, the interval from 2380 to 2535 m 104 at www.aapg.org/datashare)containsthecom- (7808–8317 ft) contains with an age range from plete names and assigned stratigraphic ranges of the Albian–Berriasian (100.5–145 Ma). Because the age taxa used in this study, separated by biological group. range is an uncertainty, and considering the corre- lation of this interval with wells A and C, we in- terpret the age of this interval as Turonian–Albian RESULTS (89.8–113 Ma). This interpretation is supported by the early Albian–Berriasian (126–144 Ma) radio- Terrigenous deposits in the fore-arc basin range from metric age of the underlying igneous rock. conglomerates to mudstones with few calcareous units. In well D, core samples in the interval Fossil recovery ranged from good to poor, depending on 1224–1255 m (4016–4118 ft) contain fossils with the lithology of the sample. Volcaniclastic sedimentary an age range from late Albian–late Pennsylvanian rocks were barren of microfossils, and conglomerate and (101.7–299 Ma). Again, because the age range is an sandstone lithofacies contained continental and some uncertainty and considering the correlation of this marine palynomorphs, whereas siltstone and shale in- interval with well E, we interpret the age of this in- tervals contain well-preserved planktonic and ben- terval as being no older than Albian (<113 Ma). thonic foraminifera, dinoflagellates, pollen, and spores.

Cenomanian Age Biostratigraphy Strata with age ranges that include the Cenomanian werefoundinallwells.WellChasthestrongestevidence The stratigraphic columns drilled in the VB contain for an interval of Cenomanian age—at depths from 2350 microfossils ranging in age from Albian (113–100.5 Ma) to 2505 m (7710–8218 ft). This interval is assigned to Miocene–Pliocene (23–3.6 Ma). A plutonic base- a Cenomanian–latest Albian (93.9–101.9 Ma) age range. ment was drilled and dated (unpublished data from Pemex) in wells B (135–9 Ma K/Ar, feldspar) and H(102–10 Ma K/Ar, feldspar). Six wells drilled Turonian Age Albian–Cenomanian conglomeratic intervals at the Strata with this age are clearly indicated in wells C base, but wells D and E drilled basal Albian volcani- and F. The rest of the wells contain strata with age clastic deposits overlain by bioclastic limestones that are ranges that include the Turonian. in turn overlain by basal conglomerates. On top of these basal strata, a succession of mainly sandstone, Santonian Age siltstone, and mudstone deposits with subordinate Strata of this age are recognized in a short interval conglomeratic horizons constitutes the fore-arc basin (1680–1700 m [5512–5577 ft]) in well A. fill. Most wells contain an unconformity between Turonian and Campanian strata, except wells C and F. A regional unconformity is also evident between Coniacian Age middle Eocene and Miocene–Pliocene strata. The Strata with this age were not clearly defined, and only following summaries highlight important aspects and thin intervals of mudstone–siltstone strata with age occurrences of the biostratigraphic ages that we rec- ranges including the Coniacian were found in some ognized and that were subsequently used to construct wells. The Coniacian is almost certainly missing in a biochronologic framework for the VB. wells A, C, F, G, and H, whereas in wells B, D, and E this age was not identified. Albian Age Strata with this age were clearly recognized in wells A, Campanian Age C, and D, whereas lithological and stratigraphically Strata with this age are clearly identified in wells A, B, correlative levels from the rest of the wells contain C, F, and G. In wells E and H, the Campanian is

1054 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 probably included in intervals with assigned age not being sampled because of the casing of the upper ranges from Maastrichtian–Campanian or Campanian– 200 m of well. late Albian. In addition, the uppermost sample (80–85 m [262–279 ft]) from well D has an age range of Gelasian–Aquitanian (1.81–23 Ma). Because the age Maastrichtian Age range is an uncertainty, and considering the correla- Strata of this age were clearly recognized in all wells, tion of this interval with well A, we interpret the age except D, in which the Maastrichtian is not recog- of this interval as Miocene (23.3–5.33 Ma). The ages nized because of a lack of samples in the interval from of the correlatable surface formations in the region – 455 to 985 m (1493 3232 ft), located between Pa- (Cantil Costero and Rosarito Beach) also support this leocene and Turonian strata. age assignment.

Paleocene Hiatuses Strata of this epoch were clearly recognized in seven The biostratigraphic data described here allow for the of the eight wells studied, but no evidence of recognition of several hiatuses, or unconformities. aCretaceous–Paleocene unconformity was found. However, without information on the structure of the However, wells B and F drilled only parts of the strata, we cannot determine whether these are an- Paleocene. Well B does not contain lower Paleocene gular unconformities. The oldest hiatus is represented strata, and well F contains an upper Paleocene by either the nonconformable contact between upper unconformity. Albian conglomeratic sandstones and the Cretaceous granitic rocks, or the disconformity between these conglomerates and the Albian limestones and vol- Eocene Epoch caniclastic rocks of the Alisitos Group. Subsequently, On top of the Paleocene interval, strata of early to the lack of upper Turonian strata in most wells, ex- middle Eocene age were clearly recognized in wells A, cept in F and G, indicates a paraconformity or dis- B, F, G, and H. Rocks of this age unconformably conformity between middle and upper Turonian – underlie strata of Miocene Pliocene age in wells A strata. In wells F and G, the limit between lower and – and D. In well D, the 115- to 150-m (377 492-ft) upper Turonian coincides with a deepening of the – interval has an age range of Oligocene middle Eo- depositional environments from transitional–neritic cene. However, this mudstone interval is interpreted and an increase of siltstone and mudstone lithofacies here as being middle Eocene because it shares the (Figure 3B). Next, the unconformity between lower characteristic alternating mudstone and siltstone of Paleocene and upper Paleocene strata is indicated – these rocks. A Miocene Pliocene sandy upper in- by a paleobathymetric change from deeper (outer terval unconformably overlies Eocene strata, ex- neritic–upper bathyal) to shallower (transitional–inner pressing an unconformity previously recognized in neritic) environments together with a lithological ı the region (Mart´n et al., 2015). change from silty shale to sandstone. Finally, the lack of upper Eocene–Oligocene strata in all the wells clearly shows a regional unconformity between Miocene Epoch middle Eocene and Miocene–Pliocene strata. Strata representing this epoch unconformably overlie Eocene sedimentary deposits in wells A and D, in which a hiatus of 24.3–24.8 Ma is recognizable. In Paleobathymetry well E, the uppermost interval (0–95 m [0–312 ft]) was deposited in continental to shallow marine The stratigraphic record in the VB represents de- environments sometime from Eocene–Miocene position in continental–neritic marine environments (Ypresian–Messinian; 50.3–5.33 Ma). So, this interval during the Albian and Cenomanian (Figures 3 and 4). could be correlated with the uppermost sequence 4. Subsequently, neritic–bathyal depth environments In other wells, strata of this age are not recognized; prevailed during the Turonian through the Eocene, this is probably because of either erosion or the strata followed by neritic, transitional, and continental

HELENES ET AL. 1055

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 deposits in the Neogene. The depositional environ- environments, and ends again with transitional–neritic ments represented in the studied sections commonly depth sedimentary rocks. The three bathymetric cycles fluctuate from continental to neritic, in which pal- overlie either upper Albian volcaniclastic or calcareous ynological assemblages are varied and contain both deposits, or Cretaceous granitic rocks. Most of the strata continental and marine palynomorphs. Continental filling the VB were deposited in neritic environments strata in the VB consists of conglomerates and along a continental margin with maximum depths in sandstones characterized by the exclusive pres- each cycle commonly reaching upper bathyal depths. ence of terrestrial palynomorphs. Dinoflagellates Green arrows in Figures 3 and 4 show the approximate appear in transitional–inner neritic deposits and locations of the deepest sedimentary environments that become more abundant and varied in middle in turn assist in correlating individual sequences. The neritic–outer neritic environments. The abundance blue arrow in Figure 3 marks a middle bathyal interval in and diversity of the continental fossils diminish with well B that represents the maximum depth within depth, whereas marine taxa tend to reach their highest a cycle of sequence 3. These three T–R cycles underlie abundances in outer neritic environments. a thin Neogene sedimentary unit and a regional Neritic environments in the VB are characterized unconformity that separates middle Eocene from by sandstones and siltstones with common to abundant middle–lower Miocene strata. benthonic foraminifera and scarce planktonic forami- nifera. The most common Cretaceous–Paleogene Basement benthonic foraminiferal genera in inner neritic–middle As mentioned before, the basement in wells B and H neritic environments are Eponides and Cibicides, is a granitic intrusion that was drilled and dated in whereas Dentalina, Nodosaria, Lenticulina, Gyroidina, wells B (135–9 Ma K/Ar, feldspar) and H (102– Uvigerina, and Spiroplectammina arecommoninouter 10 Ma K/Ar, feldspar, unpublished data from neritic environments. Pemex). However, basement rocks in wells D and Almost all wells contain intervals representing E are volcaniclastic deposits topped by calcareous upper bathyal depth environments; however, the and volcaniclastic strata similar to the Alisitos Group. deepest paleodepths recorded in well H are outer These latter rocks were deposited in the late Albian, as neritic (100–200 m [328–656 ft] deep) although well shown in well D, and have intraoceanic arc affinities, H is located near the Lagunitas high. Cretaceous thus we interpret these strata to represent part of the and Paleogene upper bathyal depth environments arc-apron that locally constituted the basement of the commonly contain Gyroidinoides soldanii, Gyroidina fore-arc basin. octocamerata, Cibicides coalingensis, Oridorsalis sp., and Dorothia sp. Middle bathyal deep environ- ments are present only in a relatively short (100 m Bathymetric and Depositional Cycles thick [328 ft]) Maastrichtian interval in well B and contain benthonic foraminifera representative of Cycle 1 deep water, such as the following: Anomalinoides This cycle records the initial filling of the VB during kennae, Osangularia plummerae,aswellasthegenera the Albian. Basal sedimentary rocks of this cycle are Ammodiscoides sp., Osangularia sp., Melonis sp., and mostly nonmarine–shallow marine deposits and are Gyroidinoides sp. chronologically poorly defined. Probably, tectonically induced subsidence caused the basin seafloor to reach upper bathyal depths in the early–middle Turonian. Paleobathymetric Evolution The latter sedimentary rocks were locally eroded Our paleobathymetric analyses allow the identifica- during deposition of the basal conglomerates of tion of three main transgressive–regressive (T–R) cycle 2. Wells D and G do not contain the maximum sedimentary cycles or sequences during the late depths of this transgression. However, the transgression Albian–middle Eocene evolution of the VB (Figures 3 in well H is represented by outer neritic deposits and 4). Each T–R cycle begins with deposition of that reached a maximum depth in the Cenomanian. shallow transitional–inner neritic depth sedimentary The age of the lower boundary of this cycle does not rocks, follows with outer neritic–bathyal depth match the age of any eustatic cycle, but the age of the

1056 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 maximum depths in this sequence roughly coincides event. Thus, this transgression likely represents a with the early Turonian maximum transgression of response to tectonically induced subsidence. the Cretaceous at circa 94 Ma (Hardenbol et al., 1998; Haq, 2014). DISCUSSION Cycle 2 Sequence Stratigraphy Initial deposition of this cycle includes conglomerate in wells B, C, D, and E (Figure 3A, B) that changed to We recognize the three main depositional cycles in the fi ne-grained epiclastic deposits in the Maastrichtian. VB through our paleobathymetric and biochronologic The transgression producing this cycle correlates analyses. These cycles accumulated three major se- with the late Turonian and commonly reached upper quences separated by significant depositional hiatuses bathyal depths in the early Maastrichtian. In this that we interpret to be unconformities. Therefore, we cycle, well B reached middle bathyal depths, whereas propose that these deposits represent three distinctive wells H and D only reached neritic environments. sequence stratigraphic units that filled the VB fore-arc The age of the lower boundary of this sequence basin from the late Albian through middle Eocene. matches with the base of the Tu3 cycle at circa Seismic lines in the VB (Ramos-Garc´ıa, 1976; Garc´ıa- 90 Ma, whereas the maximum depths in this Serratos, 2013) indicate the presence of strati- sequence correspond to the late Maastrichtian graphic units correlative throughout the basin transgression at circa 69 Ma (Hardenbol et al., 1998; (see figures 23 and 24 in Garc´ıa-Serratos, 2013) that Haq, 2014). roughly correspond to the three stratigraphic se- quences described here. A poor quality seismic image fi Cycle 3 from the onshore area, north of Guerrero Negro ( gures ı The uppermost bathymetric–depositional cycle in 4, 6, and 7 in Ramos-Garc´a, 1976), shows thicker the VB began in the early Paleocene with deposition stratigraphic sequences near well C along the basin – of neritic sandstones with subsequent deposition of axis approximately connecting wells A Ctowell bathyal shales during the late Paleocene–early Eo- F (Figure 1). cene. In wells A, B, D, E, and F, this cycle 3 reached The chronostratigraphic correlation of unconformity- upper bathyal environments in the middle Eocene, bound units and sedimentary cycles described here – whereas in wells C, G, and H, cycle 3 reached only with Cretaceous Cenozoic regional stratigraphic middle–outer neritic depths. The age of the lower units are the basis for naming these subsurface se- boundary of this cycle matches with the base of the quences as Valle, Rosario, and Sepultura sequences as Da 4 or Sel 1 cycles at circa 62 Ma, whereas the age of parts of the VB (Figure 5). The Alisitos arc interval – the maximum water depths interpreted in this cycle forms part of the basement, and the overlying post coincide with the early Eocene transgression at circa fore-arc Neogene transgressive cycle is here named 53 Ma (Hardenbol et al., 1998; Haq, 2014). the Tortugas sequence in reference to coeval marine deposits cropping out in the Pacific shoreline in the Vizcaino Peninsula (Helenes-Escamilla, 1980). Cycle 4 The VB is underlain by volcanic rocks repre- The youngest bathymetric–depositional cycle drilled senting an island arc setting, or alternatively, the VB comprises a thin (<200 m [<656 ft]), sandstone– overlies granitic rocks on the east, representing ex- siltstone unit representing neritic conditions that humation of arc-related intrusives. The boundary unconformably overlies the Paleogene cycle 3. Se- between the Alisitos rocks and rocks from the sub- quence 4 is a Neogene unit found only in wells A duction complex on the west is not well defined, but and D and was deposited in the early–middle Miocene onshore outcrops suggest a tectonic origin (Sedlock, as indicated in well A, in which it represents middle 1988). Our data indicate that volcaniclastic and neritic environments. Neither the limits of, nor the calcareous strata similar to those described for the transgression represented by this cycle can be accurately Alisitos Formation in the type locality (Allison, 1974) correlated with any recognized Neogene eustatic were deposited in the late Albian in the southern part

HELENES ET AL. 1057

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 on 24 September 2021 by guest Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf 1058 tairpyo aeAlbian Late of Stratigraphy – ideEcn icioBsn aaClfri,Mexico California, Baja Basin, Vizcaino Eocene Middle

Figure 5. Regional chronostratigraphic chart of the fore-arc basin in Baja California and southern California. This includes Cretaceous–Paleogene sedimentary units from San Diego, Baja California, Vizcaino basin, Cedros Island, and the Vizcaino Peninsula. Stratigraphic columns correspond to outcrop sections from literature as noted in Figure 2, or to wells A–H studied here. The age of the Santiago Peak volcanics is from Herzig and Kimbrough (2014). Lithology logs interpreted from gamma ray (GR, red), resistivity (ILD, blue) and/or spontaneous potential (SP, red) and mud log reports. The paleobathymetric curves show our interpretation of the sedimentary setting at the time of deposition with the symbols shown in the key. Alb = Albian; Apt = Aptian; Camp = Campanian; Ceno = Cenomanian; Coni = Coniacian; Eoc = Eocene; Isl. = Island; Maas = Maastrichtian; Mio = Miocene; Oli = Oligocene; Paleo = Paleocene; PRB = Peninsular Ranges batholith; Sant = Santonian; Sn. = San; SubCo = Subduction complex; Turo = Turonian. of the VB (see wells D and E). These basement rocks, axis extending from well A to H. The basin reached as well as the stratigraphic continuity shown by the upper bathyal depths in the Turonian, coincident seismic data, suggest that the strata forming the fill with the widespread Cretaceous eustatic trans- of the VB overlies either Cretaceous granitic rocks of gression (Haq, 2014). The subduction complex in the Peninsular batholith or Lower Cretaceous vol- the Vizcaino Peninsula was already attached to the caniclastic and calcareous strata correlative with the southwestern part of the basin as indicated by the Alisitos Group. Following Noda (2016), we postulate presence of the plutonic clasts in conglomerates in that the basement for the VB in the west is represented the Valle Group at this time (Kimbrough et al., by rocks of the subduction complex that include 2001). ophiolite rocks and rocks from the arc complexes The absence of sequence 1 in outcrops from San exposed in the Vizcaino Peninsula and that, overall, Diego to El Rosario and in wells D and E is indicated represent the accretionary wedge. Sequences 1, 2, by the unconformity of sequence 2 with the un- and 3 contain sedimentary rocks derived from derlying Alisitos Group. This unconformity repre- the batholith and Lower Cretaceous volcanic rocks sents an erosional episode that was probably caused (Kimbrough et al., 2001), and their seismic con- by a compressional event hypothetically related to the tinuity indicates that they constitute the inner influence of the Kula–Farallon spreading ridge as wedge of a convergent margin (Noda, 2016) that it moved northward along western North America developed a fore-arc basin during this time interval in the Late Cretaceous (Engebretson et al., 1985; (Busby, 2004). Umhoefer, 1987). Coincidentally, a paleogeographic reconstruction of western North America locates the Kula–Farallon ridge near the southern part of the VB Sequence Boundaries at 85 Ma (Blakey, 2003). Subsequently, deposition of sequence 2 started The four sequence boundaries represent changes in in the Turonian (ca. 92 Ma), as indicated in wells the sedimentation in the VB and are most likely re- C (Figure 3) and F (Figure 4). However, in most wells lated to regional plate-tectonic events. The charac- in the north, south, and west of the basin, the lower teristics of these boundaries are as follows. Initially, limit of sequence 2 is Santonian (ca. 86 Ma) and the lower boundary of sequence 1 is indicated by the indicates renewed sedimentation after the com- unconformity between the Valle Group and the pressive regime ended. subduction complex that represents the Albian– Afterward, sequence 3 initiated in the early Pa- Cenomanian onset of the fore-arc sedimentation. leocene (ca. 62 Ma) after a regressive event that Subsequently, the lower boundary of sequence 2 is caused a widespread hiatus or allowed deposition of distinguished by an intermediate Turonian–Coniacian sandy intervals in shallow marine environments. hiatus underlying the Rosario Group. Afterward, the Regionally, this hiatus is expressed in outcrops from lower boundary of sequence 3 is distinguished by a San Diego to El Rosario and the southern Vizcaino Maastrichtian–Paleogene local unconformity and shal- Peninsula, where there is a Maastrichtian–Paleogene lowing that likely represents a lowstand sea level that unconformity. Both, the unconformity and the exposed parts of the local fore-arc basin. Finally, the shallowing conditions were likely produced by a lower boundary of sequence 4 overlies a Paleogene– regional response to the Laramide flat subduction Neogene unconformity that represents the inversion beneath southwestern North America (Lovera et al., of the fore-arc basin and partial erosion of sequence 3. 1999; Kimbrough et al., 2001; English and Johnston, The geologic significance of these boundaries is 2010). probably related to the following tectonic conditions. Finally, sequence 4 lies above the regional Initially, sequence 1 represents the beginning of Eocene–Miocene unconformity. The scarcity of the subsidence and sedimentation in the fore-arc basin in upper Paleogene and Neogene sedimentary record is the late Albian (ca. 105 Ma). At this time, subsidence also observed in northwestern Baja California and was took place after the emplacement of the eastern PRB created during a compressive event and basin uplift (Gastil, 1975) and developed into a flat-bottomed related to the consumption of the Farallon plate and basin with neritic depths along a northwest–southeast the first collision of the Pacific–Farallon spreading

HELENES ET AL. 1059

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 ridge against western North America at circa 37 Ma Sequence 2 (Rosario) (Atwater, 1989; Lonsdale, 1991). This sequence includes terrigenous deposits from late Turonian–late Maastrichtian and is correlative with the Rosario Group (Figure 5) exposed along the Stratigraphic Sequences peninsular margin from San Diego County in southern California to the El Rosario area in western Baja Sequence 1 (Valle) California. Sequence 2 is also partially correlative with The oldest fore-arc sequence (Figure 5) comprises the older strata from the Upper Cretaceous–Paleogene upper Albian–lower Turonian coarse to fine-grained ValleGroupfromthesouthernVizcainoPeninsula continental–shallow marine deposits and correlates (Patterson, 1984b; Kimbrough et al., 2001). with the upper part of the Valle Group in Cedros The sedimentary environments represented by Island and the Vizcaino Peninsula. In Cedros, the sequence 2 range from continental and transitional Morro Redondo (Kilmer, 1984) and Pinos (Smith marine in the middle–upper Turonian, upper bathyal et al., 1993b) Formations and the Malarrimo For- in the Maastrichtian of most wells and reach middle mation in the eastern side of the Vizcaino Peninsula bathyal depths in well B. Paleontological data from (Berry and Miller, 1984; Patterson, 1984b) are also sections cropping out at the type locality of the correlative with our sequence 1. The lower part of Rosario Formation (Kilmer, 1963; Acosta, 1966; the Valle Group in Cedros and Vizcaino (Kimbrough Patterson, 1978; Yeo, 1982; Boehlke and Abbott, et al., 2001) could be interpreted as the distal de- 1986; Maestas et al., 2003) and in correlative strata position in the outer wedge of the trench–slope set- in San Diego County (Gastil and Higley, 1977; Yeo, ting of the convergent margin. The sedimentary 1982) also indicate continental–upper bathyal de- environments represented by the sequence 1 in the positional settings. Deep-water structures have been VB indicate a regional transgressive event that interpreted in Maastrichtian strata cropping out near gradually changed from continental and transitional San Fernando, south of El Rosario (Morris and Busby- in the Albian–Cenomanian interval to neritic and Spera, 1988; Busby et al., 1998; Dykstra and Kneller, upper bathyal in the Turonian. Our interpretation 2007). Foraminifera (Dykstra and Kneller, 2007) and of these depositional environments is consistent ichnofossil assemblages (Callow et al., 2013) reported with environments indicated by the benthonic fo- from the San Fernando outcrops indicate middle raminifera reported by Berry and Miller (1984) in bathyal paleoenvironments during deposition for these strata from the Vizcaino Peninsula. These authors rocks. These paleobathymetric interpretations corre- noted the presence of several planktonic species spond to the deepest paleoenvironments interpreted together with the benthonic taxa Globorotalites, for the Maastrichtian interval in well B. However, Gavelinella, Gyroidinioides,andSpiroplectammina sequence 2 biofacies observed in the wells studied are that are indicators of outer neritic–upper bathyal mostly neritic, similar to those reported in strata from environments (Sliter and Baker, 1972). The pres- outcrops of the Rosario Group from Punta Canoas ence of Marssonella oxycona in one of the samples to San Diego County. from the Los Pinos Formation in Cedros Island suggests that the upper strata were deposited at middle bathyal depths of more than 500 m Sequence 3 (Sepultura) (>1640 ft) (Smith et al., 1993a). Sequence 1 in This sequence spans lower Paleocene–middle Eocene wells D and E in the southern part of the VB (Figure 5) and represents the final stage of deposition overlie volcaniclastic and shallow-water calcareous in the fore-arc basin. The lower part of this sequence strata that correlate with the Aptian–Albian Alisitos correlates with the Paleocene Sepultura Formation in Formation. Two samples from the Malarrimo For- El Rosario area where the upper part correlates with mation in the northern Vizcaino Peninsula (Robinson, the Delicias and Buenos Aires Formations described 1975) contain benthonic foraminifera of Coniacian– near Tijuana (Flynn, 1970) and the La Jolla Group of Santonian age. In general, the facies represented the San Diego area (Kennedy and Moore, 1971; in the VB indicate deeper environments toward Gastil and Higley, 1977; Yeo, 1982). To the south, the Vizcaino Peninsula to the southwest. the sequence 3 correlates with the Paleocene–Eocene

1060 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Bah´ıa Ballenas Formation from the southern part terranes to emphasize their distinct origin (Coney of the Sebastian Vizcaino Peninsula (Abbott et al., et al., 1980; Campa and Coney, 1983; Gastil, 1985; 1995) and the Bateque Formation reported near San Sedlock, 1988; Gonzalez-Le´ on,´ 1989; Sedlock, Ignacio in southern Baja California (Mina, 1957; 2003). Alternatively, they have been related to an arc Smith et al., 1993a). The facies described here are system (Busby-Spera, 1988; Morris and Busby, 1996; similar to those outcrops of correlative strata in the El Busby et al., 1998; Busby, 2004) or even to al- Rosario area where a local angular unconformity is lochthonous terranes coming from southern Mexico documented (Abbott et al., 1993). and Central America (Hagstrum et al., 1985; Howell et al., 1985; Beck, 1986; Morris et al., 1986; Lund et al., 1991; Sedlock, 2003). However, stratigraphic Sequence 4 (Tortugas) data from the Rosario Formation (Helenes and Tellez-´ Finally, the youngest sequence is Neogene (Figure 5) Duarte, 2002; Tellez-Duarte´ et al., 2006) as well as and unconformably overlies the Paleogene Sepultura new paleomagnetic data from the PRB plutons sequence. Sequence 4 is widely represented by (Vaughn et al., 2005) and the Valle Group (Kimbrough the Tortugas (Mina, 1957; Kilmer, 1979, Helenes- et al., 2006) indicate that both the Peninsular Ranges Escamilla, 1980) and Almejas Formations (Mina, and the Vizcaino terrane have remained in place 1957; Smith, 1984) in Cedros Island and the northern with respect to cratons in North America since the Vizcaino Peninsula. However, sequence 4 is scarce in Cretaceous. northwestern Baja California, where it is represented The units described here reflect different tectonic by the Rosarito Beach Formation reported from San regimes. The Alisitos sequence is the oldest unit Diego to Ensenada (Minch, 1967; Gastil and Higley, recognized here and represents deposition of volcanic 1977; Yeo, 1982) and by the Cantil Costero For- rocks with oceanic arc affinities in an extensional mation (Santillan´ and Barrera, 1930), which crops out fringing arc (Busby, 2004). The uppermost part of near El Rosario. South of the VB, the San Ignacio and this unit is characterized by the biohermal limestones San Isidro Formations (Mina, 1957) also represent that were exposed along the western side of the PRB Neogene marine sedimentation. Thus, sequence 4 (Suarez-Vidal,´ 1987) and that were drilled in wells D presents shallow marine facies on the continental shelf and E. Subsequently, sedimentation in the fore arc that correlate with reported shallow marine and vol- initiated in mostly transitional–neritic environments canic deposits cropping out in northwestern Baja over a continental margin formed by the fringes of California. Outcrop facies in the south represent ne- the PRB and the Alisitos arc strata as indicated by ritic environments, whereas correlative deposits in the the basement units sampled by Pemex drilling. Vizcaino Peninsula reach bathyal depths (Helenes- Regional volcanic units with oceanic affinity in- Escamilla, 1980). Subsidence of the continental dicate a tectonic regime that represents an earlier shelf in the Neogene is probably related to the phase of an extensional island arc (Busby et al., 1998; westward tilting of the Baja Peninsula because of Busby, 2004), so they are excluded from the fore-arc continental rupture in the Gulf of California (Mark basin stage. These volcanic units, which underlie et al., 2014). Rift-related asymmetric flexure of the the Valle and Rosario Groups as used here, are in- crust probably promoted subsidence of the conti- cluded in the Lower Cretaceous Alisitos volcaniclastic nental shelf to the west and marine flooding. However, sedimentary group. We also exclude most rock units isostatic sea-level changes may also have been a factor exposed in the Cedros Island–northern Vizcaino in promoting shelf sedimentation during highstand Peninsula area. These latter units are interpreted to be sea-level conditions; however, the lack of complete part of the subduction complex developed at the stratigraphic record for sequence 4 prevents accurate trench slope and progressively exhumed during the correlation with the Neogene sea-level curve. termination of the regional subduction. These rocks include the Triassic–Jurassic ophiolitic complex, Evolution of the Fore-Arc Basin a Jurassic tectonic melange,´ the Jurassic volcaniclastic rocks, and the Albian lower part of the Valle Group The stratigraphic record of western Baja California (Kilmer, 1984), including the Valle and Vargas For- has been described in terms of tectonostratigraphic mations in Cedros Island (Mina, 1957; Smith et al.,

HELENES ET AL. 1061

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 1993b) and Los Chapunes Formation in northern the subduction complex and inversion of the fore- Vizcaino Peninsula (Patterson, 1984b). arc basin. The regional unconformity separating se- The boundaries of the sequences described here quences 3 and 4 and the scarcity of the Neogene represent major changes in the sedimentation of sedimentary record in northwestern Baja California the VB basin. These boundaries represent the onset of collectively support our interpretation. The con- the fore-arc sedimentation (sequence 1), two fore-arc sumption of the Farallon plate and the first collision of events—first, a Turonian–Coniacian unconformity the Pacific–Farallon spreading ridge against western marking the base of sequence 2 and later a North America at circa 37 Ma (Atwater, 1989; Maastrichtian–Paleogene unconformity separating Lonsdale, 1991) likely caused this compression and sequences 2 and 3—and finally, the post fore-arc inversion of the Vizcaino basin. sedimentation (sequence 4).

CONCLUSIONS Onset of Fore-Arc Sedimentation Onset of fore-arc sedimentation began after the Our micropaleontologic and lithologic data from synbatholithic crustal shortening noted in the Sierra eight exploratory wells in the VB provide a chrono- San Pedro Martir region (Johnson et al., 1999). A stratigraphic and paleoenvironmental framework to rapid exhumation of cretaceous intrusives and analyze the evolution of the continental margin in the prominent influx of coarse-grained clastic deposits. context of the regional geologic history of south- This requires steep topographic gradients in the western North America. region as already noted by Kimbrough et al. (2001) Our analyses document that the subsiding fore-arc in the Vizca´ıno Peninsula. basin experienced three major transgressive–regressive depositional cycles from the Albian through middle Intra Fore-Arc Events Eocene forming three first-order unconformity-bound stratigraphic sequences. Each sequence displays evi- The older event caused a widespread Coniacian gap dence of different environments and correlates with in the stratigraphic record in the VB that corresponds stratigraphic units cropping out along coastal southern to the upper boundary of sequence 1 (Figure 5). California and Baja California. Basal sequence 1 (Valle) This unconformity spans the entire sequence 1 marks initial sedimentation in the basin and corre- (Cenomanian–Santonian) in outcrops from San lates with the Cretaceous Valle Group, sequence 2 Diego to El Rosario and in wells D and E. This hiatus (Rosario) correlates with the Upper Cretaceous parts of is hypothetically linked to the compression of the the Rosario Group, and sequence 3 (Sepultura) corre- Kula–Farallon ridge against the North American lates with the Paleogene Sepultura and Bateque For- plate. The younger Maastrichtian–Paleogene hiatus mations. Post fore-arc Neogene marine sedimentary constitutes the base of sequence 3 (Figure 5) and rocks unconformably overlie the fore-arc basin fill, represents a significant regressive event. We interpret comprise sequence 4 (Tortugas), and correlate with the this regression as a response to Laramide flat sub- Tortugas Formation and other wedge-shaped marine duction with subsidence and sedimentation in the and volcanic deposits on the continental shelf of Baja fore-arc basin following the Laramide orogenic event, California. as indicated by upper bathyal depths recorded in Basement rocks that underlie the basin-fill in the wells A, B, E, and F (Figures 3 and 4). Finally, an exploratory wells include volcaniclastic deposits cap angular unconformity separates the depositional re- by bioclastic limestones representing the late Early cord of the fore arc and the development of the Cretaceous (Albian) Alisitos arc-apron on the south. transtensional plate boundary in both margins of Cretaceous granitic rocks of the PRB underlie the the Baja California continental block (Mart´ınetal., basin along the eastern margin. Triassic–Jurassic 2015). The youngest sequence 4 (Figure 5) is related ophiolitic rocks and tectonic melange´ exposed on the to margin subsidence after late Oligocene collision Vizcaino Peninsula represent an additional basement of the Pacific–Farallon ridge (Atwater, 1989). This unit to the west but only reported in the Lagunitas collision produced a compressive event and uplift of structural high (Mart´ın et al., 2015).

1062 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Basin sedimentation began in the late Albian paleoceanography of the New Hebrides Island arc and because of continental margin subsidence, with se- north D’entrecasteaux ridge area, in H. G. Greene, J.-Y. quences 1, 2, and 3 subsequently deposited in Collot, L. B. Stokking, K. Akimoto, M. V. S. Ask, P. E. Baker, L. Briqueu, et al., Proceedings of the Ocean a convergent continental margin setting. Initial de- Drilling Program, Scientific Results: College Station, position of shallow marine sedimentary rocks of se- Texas, Ocean Drilling Program, v. 134, p. 265–291, doi: quence 3 signals the end of a regressive event probably 10.2973/odp.proc.sr.134.011.1994. linked to local influence of the Laramide orogeny Allison, E. C., 1974, The type Alisitos Formation (Cretaceous, Aptian–Albian) of Baja California and its bivalve fauna, in followed by renewed Paleogene basin subsidence. R. G. Gastil and J. Lillegraven, eds., A guidebook to the Deepest paleobathymetries in the evolving fore- geology of peninsular California: 49th Annual Meeting, arc basin were reached during the early Turonian, Bakersfield, California, April 24–28, 1974, AAPG Pacific Maastrichtian, and late Paleocene–early Eocene and Section Guidebook 37, p. 20–59. were correlative with peaks in global maximum Alvarado-de-la-Tejera, S., 1976, Integracion´ y evaluacion´ en las cuencas sedimentarias de la Pen´ınsuladeBajaCalifornia: flooding surfaces recorded at 98.3 Ma, 72 Ma, and in III simposio de geolog´ıa del subsuelo zonas noreste: 54 Ma by Hardenbol et al. (1998). Reynosa, Tamaulipas, Mexico, Petroleos´ Mexicanos, A significant Coniacian unconformity indicates Superintendencia General de Exploracion,´ p. 82–92. basin inversion because of compression and is hypo- Atwater, T. M., 1989, Plate tectonic history of the northeast fi thetically related to the collision of the Kula–Farallon Paci c and western North America, in E. L. Winterer, ed., The geology of North America: Boulder, Colorado, ridge against North America. The youngest major The Geological Society of America, p. 21–72. unconformity separates sequences 3 and 4 encom- Barnes, D. A., 1984, Volcanic arc derived, Mesozoic sedi- passing the late Eocene through Oligocene. This an- mentary rocks, Vizcaino Peninsula, Baja California Sur, gular unconformity records inversion of the VB and Mexico, in V. A. Frizzell, ed., Geology of the Baja California peninsula: Tulsa, Oklahoma, SEPM Pacific subsequent partial erosion of sequence 3 strata as – fi – Section 39, p. 119 133. a function of (1) the collision of the Paci c Farallon Beck, M. E. Jr., 1986, Model for late Mesozoic–early Tertiary ridge with the continental margin at 37 Ma, (2) ces- tectonics of coastal California and western Mexico sation of subduction of the Farallon plate beneath and speculations on the origin of the San Andreas the continental margin, and (3) the initiation of Fault:Tectonics,v.5,no.1,p.49–64, doi:10.1029 atranslational–transtensional tectonic regime in this /TC005i001p00049. Beggs, J. M., 1984, Volcaniclastic rocks of the Alisitos Group, region that is related to progressive capture of the Baja California, Mexico, in V. A. Frizzell, ed., Geology of Baja California microplate by the Pacificplate. the Baja California peninsula: Tulsa, Oklahoma, SEPM The Vizcaino basin records an important event in Pacific Section 39, p. 43–52. the evolution and growth of the continental margin Berry, K. D., and P. L. Miller, 1984, Mesozoic biostratigraphy, of southwestern North America. Vizcaino Peninsula and Cedros Island, Baja California Sur, Mexico, in V. A. Frizzell, ed., Geology of the Baja California peninsula: Tulsa, Oklahoma, SEPM Pacific Section 39, p. 67–88. Blakey,R.C.,2003,DeeptimeMaps,2sheets,http:// REFERENCES CITED deeptimemaps.com/wp-content/uploads/2016/05 /SWNA_85Ma.png. Abbott, P. L., A. D. Hanson, C. N. Thomson, D. L. Logue, Boehlke, J. E., and P. L. Abbott, 1986, Punta Baja Formation, K. D. Bradshaw, W. J. Pollard, and T. E. Seeliger, 1993, a Campanian submarine canyon fill, Baja California, Geology of the Paleocene Sepultura Formation, Mesa de Mexico, in P. L. Abbott, ed., Cretaceous stratigraphy la Sepultura, Baja California, Mexico: Ciencias Marinas, western North America: Tulsa, Oklahoma, SEPM Pacific v. 19, no. 1, p. 75–93, doi:10.7773/cm.v19i1.919. Section 46, p. 92–102. Abbott, P. L., D. P. Smith, W. V. Sliter, and L. R. Saul, 1995, Boles, J. R., and C. A. Landis, 1984, Jurassic sedimentary Paleogeography of three Paleocene limestones in Baja melangé and associated facies, Baja California, Mexico: California, in A. E. Fritsche, ed., Cenozoic paleogeog- Geological Society of America Bulletin, v. 95, no. 5, raphy of the western United States II: Tulsa, Oklahoma, p. 513–521, doi:10.1130/0016-7606(1984)95<513: SEPM Pacific Section 75, p. 1–8. JSMAAF>2.0.CO;2. Acosta, M. G., 1966, Geology of the Bahia Soledad Embay- Bolli, H. M., and J. B. Saunders, 1985, Oligocene to Holocene ment, Baja California, Mexico, M.S. thesis, San Diego low latitude planktic foraminifera, in H. M. Bolli, State University, San Diego, California, 93 p. J. B. Saunders, and K. Perch-Nielsen, eds., Plankton stra- Akimoto, K., 1994, Cenozoic benthonic foraminiferal bio- tigraphy, Cambridge Earth Sciences Series: Cambridge, stratigraphy, paleobathymetry, paleoenvironments and United Kingdom, Cambridge University Press, p. 155–262.

HELENES ET AL. 1063

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Bottjer, D. J., and M. H. Link, 1984, A synthesis of Late Cre- Diaz-Cuevas, A., 1976, III simposio de geolog´ıa del subsuelo taceous southern California and northern Baja California zonas noreste: Reynosa, Tamaulipas, Mexico, Petroleos´ paleogeography, in J. K. Crouch and S. B. Bachman, Mexicanos, Superintendencia General de Exploracion,´ eds., Tectonics and sedimentation along the California p. 73–81. margin: Tulsa, Oklahoma, SEPM Pacific Section 38, Drugg, W. S., and L. E. Stover, 1975, Stratigraphic range p. 79–90. charts of selected Cenozoic dinoflagellates, in W. R. Evitt, Busby, C., 2004, Continental growth at convergent margins ed., Proceedings of a forum on dinoflagellates: Dallas, facing large ocean basins: A case study from Mesozoic Texas, American Association of Stratigraphic Palynolo- convergent-margin basins of Baja California, Mexico: gists Contribution Series 4, p. 73–90. Tectonophysics,v.392,no.1–4, p. 241–277, doi:10.1016 Dykstra, M., and B. Kneller, 2007, Canyon San Fernando, /j.tecto.2004.04.017. Mexico: A deep-water, channel-levee complex exhibiting Busby, C. J., D. P. Smith, W. R. Morris, and B. Fackler-Adams, evolution from submarine canyon—confined to uncon- 1998, Evolutionary model for convergent margins facing fined, in T. Nilsen, R. Shew, G. Steffens, and J. Studlick, large ocean basins: Mesozoic Baja California (Mexico): eds., Atlas of deepwater outcrops: AAPG Studies in Ge- Geology, v. 26, no. 3, p. 227–230, doi:10.1130/0091- ology 56, p. 226–230. 7613(1998)026<0227:EMFCMF>2.3.CO;2. Emery, K. O., 1960, The sea off southern California. A Busby-Spera, C. J., 1988, Evolution of a middle Jurassic back- modern habitat of petroleum: New York, John Wiley & arc basin, Cedros Island, Baja California: Evidence from Sons, Inc., 366 p. a marine volcaniclastic apron: Geological Society of Engebretson, D. C., A. Cox, and R. G. Gordon, 1985, Relative America Bulletin, v. 100, no. 2, p. 218–233, doi:10.1130 motions between oceanic and continental plates in the /0016-7606(1988)100<0218:EOAMJB>2.3.CO;2. Pacific Basin: Boulder, Colorado, Geological Society of Callow, R. H. T., D. McIlroy, B. Kneller, and M. Dykstra, America Special Paper 206, 60 p., doi:10.1130/SPE206. 2013, Integrated ichnological and sedimentological English, J. M., and S. T. Johnston, 2010, The Laramide analysis of a Late Cretaceous submarine channel-levee Orogeny: What were the driving forces: International system: The Rosario Formation, Baja California, Mexico: Geology Review, v. 46, no. 9, p. 833–838, doi:10.2747 Marine and Petroleum Geology, v. 41, p. 277–294, doi: /0020-6814.46.9.833. 10.1016/j.marpetgeo.2012.02.001. Fensome, R. A., R. A. MacRae, and G. L. Williams, 2008, Campa, M. F., and P. J. Coney, 1983, Tectono-stratigraphic DINOFLAJ2, version 1, American Association of Strati- terranes and mineral resource distributions in Mexico: graphic Palynologists, data series no. 1, accessed October Canadian Journal of Earth Sciences, v. 20, no. 6, 29, 2018, http://dinoflaj.smu.ca/wiki/Main_Page. p. 1040–1051, doi:doi.org/10.1139/e83-094. Finch, J. W., and P. L. Abbott, 1977, Petrology of a Triassic Caron, M., 1985, Cretaceous planktic foraminifera, in marine section, Vizcaino Peninsula, Baja California Sur, H. M. Bolli, J. B. Saunders, and K. Perch-Nielsen, eds., Mexico: Sedimentary Geology, v. 19, p. 253–273, doi: Plankton stratigraphy, Cambridge Earth Sciences Series: 10.1016/0037-0738(77)90039-2. Cambridge, United Kingdom, Cambridge University Flynn, C. J., 1970, Post-batholithic geology of the La Gloria- Press, p. 17–86. Presa Rodriguez area, Baja California, Mexico: Geological Carreño, A. L., and J. T. Smith, 2007, Stratigraphy and Society of America Bulletin, v. 81, no. 6, p. 1789–1806, Correlation for the Ancient Gulf of California and Baja doi:10.1130/0016-7606(1970)81[1789:PGOTLG]2.0.CO;2. California Peninsula, Mexico: Ithaca, New York, Pale- Garc´ıa-Serratos, N. E., 2013, Modelo de facies de la cuenca ontological Research Institution, 146 p. cretacico-terciaria´ de Vizca´ıno, Baja California, a partir de Cavazos-Prado, S., 1976, Cronoestratigraf´ıa y paleobatimetr´ıa la interpretacion´ de registros de pozos y s´ısmica de re- de las cuencas Sebast´ıan V´ızcaino y Purisima-Iray,´ in III flexion,´ M.S. thesis, Earth Sciences Graduate Program, simposio de geolog´ıa del subsuelo zonas noreste: Reynosa, Centro de Investigacion´ Cient´ıfica y de Educacion´ Su- Tamaulipas, Mexico, Petroleos´ Mexicanos, Super- perior de Ensenada (CICESE), Baja California, Mexico,´ intendencia General de Exploracion,´ p. 50–60. 78 p. Centeno-Garc´ıa, E., C. Busby, M. Busby, and G. Gehrels, Gastil, R. G., 1975, Plutonic zones in the Peninsular Ranges of 2011, Evolution of the Guerrero composite terrane along southern California and northern Baja California: Geol- the Mexican margin, from extensional fringing arc to con- ogy, v. 3, no. 7, p. 361–363, doi:10.1130/0091-7613 tractional continental arc: Geological Society of America (1975)3<361:PZITPR>2.0.CO;2. Bulletin, v. 123, no. 9–10, doi:10.1130/B30057.1. Gastil, R. G., 1983, Mesozoic and Cenozoic granitic rocks of Coney, P. J., D. L. Jones, and J. W. H. Monger, 1980, Cor- southern California and western Mexico: Boulder, Col- dilleran suspect terranes: Nature, v. 288, p. 329–333, doi: orado, Geological Society of America 159, p. 265–276, 10.1038/288329a0. doi:10.1130/MEM159-p265. Cunningham,A.B.,andP.L.Abbott,1986,Sedimentologyand Gastil, R. G., 1985, Terranes of peninsular California and provenance of the Upper Cretaceous Rosario Formation adjacent Sonora, in D. G. Howell, ed., Tectonos- south of Ensenada, Baja California, Mexico, in P. L. tratiographic terranes of the Circum–Pacificregion: Abbott, ed., Cretaceous stratigraphy western North Moss Landing, California, Circum–Pacific Council for America: Tulsa, Oklahoma, SEPM Pacific Section 46, Energy and Mineral Resources Earth Science Series 1, pp. 103–118. p. 273–283.

1064 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Gastil, R. G., and R. Higley, 1977, Guide to the San Diego area Helenes, J., C. De Guerra, and J. Vasquez,´ 1998, Palynology stratigraphy: San Diego, California, San Diego State and of the Upper Cretaceous in the University, 62 p. subsurface of the Barinas area, western Venezuela: AAPG Gastil, R. G., G. J. Morgan, and D. Krummenacher, 1978, Bulletin, v. 82, no. 7, p. 1308–1328. Mesozoic history of peninsular California and related Helenes, J., and M. A. Tellez-Duarte,´ 2002, Paleontological ev- areas east of the Gulf of California, in D. G. Howell and idence of the Campanian to Early Paleocene paleogeography K. A. McDougall, eds., Mesozoic paleogeography of the of Baja California: Palaeogeography, Palaeoclimatology, western United States: Tulsa, Oklahoma, SEPM Pacific Palaeoecology, v. 186 no. 1–2, p. 61–80, doi:10.1016 Section, p. 107–115. /S0031-0182(02)00444-3. Gastil, R. G., R. P. Phillips, and E. C. Allison, 1975, Reconnaissance Helenes-Escamilla, J., 1980, Stratigraphy, depositional en- geology of the state of Baja California: Boulder, Colorado, viroments and foraminifera of the Miocene Tortugas Geological Society of America Memoir 140, 169 p. Formation, Baja California Sur, Mexico: Bolet´ındela Germeraad, J. H., C. A. Hopping, and J. Muller, 1968, Pal- Sociedad Geologica´ Mexicana, v. 61, no. 1–2, p. 47–67, ynology of Tertiary sediments from tropical areas: Review doi:10.18268/BSGM1980v41n1a3. of Palaeobotany and Palynology, v. 6, no. 3–4, p. 189–348, Herzig, C. T., and D. L. Kimbrough, 2014, Santiago Peak doi:10.1016/0034-6667(68)90051-1. volcanics: Early Cretaceous arc volcanism of the western Gonzalez-Garc´ ´ıa, R., 1976, Bosquejo geologico´ de la zona Peninsular Ranges batholith, southern California, in noreste, in III simposio de geolog´ıa del subsuelo zonas D. M. Morton and F. K. Miller, eds., Peninsular Ranges noreste: Reynosa, Tamaulipas, Mexico, Petroleos´ Mex- batholith, Baja California and southern California: icanos, Superintendencia General de Exploracion,p.5´ –29. Boulder, Colorado, Geological Society of America Gonzalez-Le´ on,´ C. M., 1989, Estratigraf´ıa y Paleogeograf´ıa del Memoir 211, p. 1–19., doi:10.1130/2014.1211(09) Cretacico´ del area´ de Cerro de Oro, Mpio. de Rayon,´ Hickey, J. J., 1984, Stratigraphy and composition of a Jura– Sonora, M.S. thesis, Universidad Nacional Autonoma´ de Cretaceous volcanic arc apron, in V. A. Frizzell, ed., Mexico,´ Mexico City, Mexico, 109 p. Geology of the Baja California peninsula: Tulsa, Gradstein, F. M., J. G. Ogg, M. Schmitz, and G. Ogg, 2012, The Oklahoma, SEPM Pacific Section 39, p. 149–160. geologic time scale 2012: Amsterdam, Elsevier, 1176 p. Howell, D. G., D. L. Jones, and E. R. Schermer, 1985, Hagstrum, J. T., M. McWilliams, D. G. Howell, and Tectonostratigraphic terranes of the Circum–Pacificre- S. Gromme,´ 1985, Mesozoic paleomagnetism and gion, in D. C Howell, ed., Tectonostratigraphic terranes northward translation of the Baja California peninsula: of the Circum–Pacific Region: Moss Landing, California, Boulder, Colorado, Geological Society of America Bulle- Circum–Pacific Council for Energy and Mineral Re- tin, v. 96, no. 8, p. 1077–1090, doi:10.1130/0016-7606 sources Earth Sciences Series 1, p. 3–30. (1985)96<1077:MPANTO>2.0.CO;2. Johnson, S., M. C. Tate, and C. M. Fanning, 1999, New Haq, B. U., 2014, Cretaceous eustasy revisited: Global geologic mapping and SHRIMP U–Pb zircon data in the and Planetary Change, v. 113, p. 44–58, doi:10.1016 Peninsular Ranges batholith, Baja California, Mexico: /j.gloplacha.2013.12.007. Evidence for a suture?: Geology, v. 27, no. 8, p. 743–746, Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic and doi:10.1130/0091-7613(1999)027<0743:NGMASU>2.3 Cenozoic chronostratigraphy and Eustatic cycles, in .CO;2. C. K. Wilgus, B. S. Hastings, C. G. St. C. Kendall, Johnson, S. E., S. R. Paterson, J. M. Fletcher, G. H. Girty, H. W. Posamentier, C. A. Ross, and J. C. Van Wagoner, D. L. Kimbrough, and A. Mart´ın-Barajas, eds., 2003, eds., Sea-level changes: An integrated approach: Boulder, Tectonic evolution of northwestern Mexico and the Colorado, SEPM Pacific Section 42, p. 71–108, doi: southwestern USA: Boulder, Colorado, Geological So- 10.2110/pec.88.01.0071 ciety of America Special Paper 374, 478 p. Hardenbol, J., J. Thierry, M. B. Farley, Th. Jacquin, P.-C. de Kennedy, M. P., and G. W. Moore, 1971, Stratigraphic re- Graciansky, and P. R. Vail, 1998, Mesozoic and Cenozoic lations of Upper Cretaceous and Eocene formations, San sequence chronostratigraphic framework of European Diego coastal region, California: AAPG Bulletin, v. 55, Basins, in P.-C de Graciansky, J. J. Hardenbol, Th. Jacquin, no. 5, p. 709–722. and P. R. Vail, eds., Mesozoic and Cenozoic sequence Kennedy, M. P., and G. L. Peterson, 1975, Geology of the stratigraphy of European Basins: Boulder, Colorado, SEPM San Diego metropolitan area, California: Sacramento, Special Publication 60, p. 3–13. California, California Division of Mines and Geology Helby, R., R. Morgan, and A. D. Partridge, 1987, A paly- Bulletin 200, p. 43–56. nological zonation of the Australian Mesozoic, in Kennett, J. P., and M. S. Srinivasan, 1983, Neogene planktonic P. A. Jell, ed., Studies in Australian Mesozoic palynology: foraminifera: A phylogenetic atlas: Stroudsburg, Penn- Hornsby, New South Wales, Association of Australasian sylvania, Hutchinson Ross, 265 p. Palaeontologists Memoir 4, p. 1–85. Kilmer, F. H., 1963, Cretaceous and Cenozoic stratigraphy Helenes, J., 1984, Dinoflagellates from Cretaceous to and , El Rosario area, Ph.D. thesis, Uni- Early Tertiary rocks of the Sebastian´ Vizcaino Basin, in versity of California, Berkeley, California, 149 p. V. A. Frizzell, ed., Geology of the Baja California Kilmer, F. H., 1977, Reconnaissance geology of Cedros Island, peninsula: Boulder, Colorado, SEPM PacificSection Baja California, Mexico: Bulletin of the Southern Cal- 39,p.89–106. ifornia Academy of Sciences, v. 76, p. 91–98.

HELENES ET AL. 1065

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Kilmer, F. H., 1979, A geological sketch of Cedros Island, Baja Lund, S. P., D. J. Bottjer, K. J. Whidden, J. E. Powers, and California, Mexico, in P. L. Abbott and G. R. Gastil, eds., M. C. Steele, 1991, Paleomagnetic evidence for dis- Baja California geology: San Diego, California, San Diego placements and accretion in southern California, in State University, p. 11–28. P. L. Abbott and J. A. May, eds., Eocene Geologic Kilmer, F. H., 1984, Geology of Cedros Island, Baja California, History, San Diego Region: Tulsa, Oklahoma, SEPM Mexico: Arcata, California, Humboldt State University, Pacific Section 68, p. 99–106. 69 p. Maestas, Y., K. G. MacLeod, R. Douglas, J. Self-Trail, and Kimbrough, D. L., 1984, Paleogeographic significance of the P. D. Ward, 2003, Late Cretaceous foraminifera, pale- middle Jurassic Gran Canon Formation, Cedros Island, oenvironments, and paleoceanography of the Rosario Baha California Sur, in V. A. Frizzell, ed., Geology of the Formation, San Antonio del Mar, Baja California, Mexico: Baja California peninsula: Tulsa, Oklahoma, SEPM Pa- Journal of Foraminiferal Research, v. 33, no. 3, p. 179–191, cific Section 39: p. 107–118. doi:10.2113/33.3.179. Kimbrough, D. L., 1985, Tectonstratigraphic terranes of the Mark, C., S. Gupta, A. Carter, D. F. Mark, C. Gautheron, and Vizcaino Peninsula and Cedros and San Benito Islands, A. Mart´ın, 2014, Rift flank uplift at the Gulf of California: Baja California, Mexico: Moss Landing, California, No requirement for asthenospheric upwelling: Geology, Circum–Pacific Council for Energy and Mineral Re- v. 42, no. 3, p. 259–262, doi:10.1130/G35073.1. sources Earth Sciences Series 1, p. 285–298. Mart´ın, A., J. Contreras, J. Garc´ıa-Abdeslem, M. Gonzalez- Kimbrough, D. L., P. L. Abbott, M. Grove, D. P. Smith, Escobar, J. Helenes, and F. Suarez-Vidal, 2015, El J. B. Mohoney, T. E. Moore, and G. E. Gehrels, 2006, margen continental del noroeste de Mexico: De sub- Contrasting cratonal provenances for Upper Cretaceous duccion a ruptura continental (abs.): Union´ Geof´ısica Valle Group quarzite clasts, Baja California, in Mexicana Reunion´ Anual 2015, Puerto Vallarta, Mexico,´ G. H. Girty and J. D. Cooper, eds., Using stratigraphy, November 2–7, 2015, 1 p. sedimentology, and geochemistry to unravel the geologic Matsuoka, K., and J. P. Bujak, 1988, Cenozoic dinoflagellate history of the southwestern Cordillera: A volume in cysts from the Navarin Basin, Norton Sound and St. honor of Patrick L. Abbott: Tulsa, Oklahoma, SEPM George Basin, Bering Sea: Bulletin of the Faculty of Pacific Section 101, p. 97–110. Liberal Arts, Nagasaki University, Nature and Science, Kimbrough, D. L., D. P. Smith, J. B. Mahoney, T. E. Moore, v. 29, p. 1–147. M. Grove, R. G. Gastil, A. Ortega-Rivera, and Miller, V. V., and P. L. Abbott, 1989, Sedimentology of the C. M. Fanning, 2001, Forearc-basin sedimentary response Upper Cretaceous Rosario Formation near Colonet and to rapid Late Cretaceous batholith emplacement in the Camalu, in P. L. Abbott, ed., Geologic studies in Baja Peninsular Ranges of southern and Baja California: Ge- California: Tulsa, Oklahoma, SEPM Pacific Section 63, ology, v. 29, no. 6, p. 491–494, doi:10.1130/0091-7613 p. 63–74. (2001)029<0491:FBSRTR>2.0.CO;2 Mina, U. F., 1957, Bosquejo geologico´ del Territorio Sur de la Krause, D. C., 1964, Lithology and sedimentation in the Baja California: Bolet´ın de la Asociacion´ Mexicana de southern continental borderland (California), in R. L. Miller, Geologos´ Petroleros, v. 9, no. 3–4, pp. 139–269. ed., Papers in marine geology, Shepard commemorative Minch, J. A., 1967, Stratigraphy and structure of the volume: New York, MacMillan Co., p. 274–318. Tijuana–Rosarito Beach area, northwestern Baja Cal- Krause, D. C., 1965, Tectonics, bathymetry, and geo- ifornia, Mexico: Boulder, Colorado, Geological Society magnetism of the southern continental borderland west of America Bulletin, v. 78, no. 9, p. 1155–1178, doi: of Baja California, Mexico: Boulder, Colorado, Geological 10.1130/0016-7606(1967)78[1155:SASOTT]2.0.CO;2 Society of America Bulletin, v. 76, no. 6, p. 617–650, doi: Miranda-Mart´ınez, A. Y., and A. L. Carreño, 2008, Forami- 10.1130/0016-7606(1965)76[617:TBAGOT]2.0.CO;2. nifera from the Tepetate Formation (Heim) at Las Po- Lonsdale, P., 1991, Structural patterns of the Pacific floor citas, Baja California Sur, Mexico: Ciencias Marinas, offshore of peninsular California, in J. P. Dauphin and v. 34, no. 2, p. 179–195. B. R. T. Simoneit, eds., The Gulf and the peninsular Moore, T. E., 1984, Sedimentary facies and composition of province of the Californias: AAPG Memoir 47, Jurassic volcaniclastic turbidites at Cerro El Calvario, in p. 87–125. V. A. Frizzell, ed., Geology of the Baja California pen- Lorente, M. A., 1986, Palynology and palynofacies of the insula: Tulsa, Oklahoma, SEPM Pacific Section 39, upper Tertiary in Venezuela, Ph.D. thesis, University of p. 131–148. Amsterdam, Amsterdam, 222 p. Moore, T. E., 1985, Stratigraphy and tectonic significance of Lovera, O. M., M. Grove, D. L. Kimbrough, and P. L. Abbott, the Mesozoic tectonstratigraphic terranes of the Viz- 1999, A method for evaluating basement exhumation caino Peninsula, Baja California Sur, Mexico: Moss histories from closure age distributions of detrital minerals: Landing, California, Circum–Pacific Council for En- Journal of Geophysical Research: Solid Earth, v. 104, ergy and Mineral Resources Earth Sciences Series 1, no. 12, p. 29419–29438, doi:10.1029/1999JB900082. p. 315–329. Lozano-Romen, F., 1975, Evaluacion´ petrol´ıfera de la Morris, L. K., S. P. Lund, and D. J. Bottjer, 1986, Paleolatitude Pen´ınsula de Baja California, Mexico:´ Bolet´ındela drift history of displaced terranes in southern and Baja Asociacion´ Mexicana de Geologos´ Petroleros, v. 27 California: Nature, v. 321, p. 844–847, doi:10.1038 no. 4–6, p. 104–329. /321844a0.

1066 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Morris, W. R., and C. J. Busby, 1996, The effects of tectonism peninsula: Tulsa, Oklahoma, SEPM Pacific Section 39, on the high-resolution sequence stratigraphic framework p. 173–182. of non-marine to deep-marine deposits in the Peninsular Perch-Nielsen, K., 1985, Mesozoic calcareous nannofossils, in Ranges forearc basin complex, in P. L. Abbott and H. M. Bolli, J. B. Saunders, and K. Perch-Nielsen, eds., J. D. Cooper, eds., AAPG field conference guide 1996: Plankton stratigraphy: Cambridge, United Kingdom, AAPG Guidebook 73, p. 381–408. Cambridge University Press, p. 329–426. Morris, W. R., and C. J. Busby-Spera, 1988, Sedimentological Pessagno, E. A. Jr., J. W. Finch, and P. L. Abbott, 1979, Upper evolution of a submarine canyon in a forearc basin, Upper Triassic Radiolaria from the San Hipolito Formation, Baja Cretaceous Rosario Formation, San Carlos, Mexico: California: Micropaleontology, v. 25, no. 2, p. 160–197, AAPG Bulletin, v. 72, no. 6, p. 717–737. doi:10.2307/1485265. Muller,¨ J., E. Di Giacomo, and A. W. van Erve, 1987, A Powell, A. J., ed., 1992, A stratigraphic index of dinoflagellate palynological zonation for the Cretaceous, Tertiary, and cysts: London, Chapman and Hall, 290 p. Quaternary of northern South America: Dallas, Texas, Premoli Silva, I., and D. Verga, 2004, Practical manual of American Association of Stratigraphic Palynologists Cretaceous planktonic foraminifera, course 3, in D. Verga Contribution Series 19, p. 7–76. and R. Rettori, eds., International school on planktonic Nilsen, T. H., and P. L. Abbott, 1979, Turbidite sedimen- foraminifera: TipografiadidiPontefelcino,Perugia, tology of the Upper Cretaceous, Point Loma and Cabrillo Universities of Perugia and Milano, 283 p. Formations, San Diego, California, in P. L. Abbott, ed., Premoli-Silva, I., and W. V. Sliter, 1999, Cretaceous paleo- Geological excursions in the southern California area: San ceanography: Evidence from planktonic foraminiferal Diego, California, San Diego State University, p. 139–166. evolution, in E. Barrera and C. C. Johnson, eds., Evo- Nilsen, T. H., and P. L. Abbott, 1981, Paleogeography and lution of the Cretaceous ocean-climate system: Boulder, sedimentology of Upper Cretaceous turbidites, San Diego, Colorado, Geological Society of America Special Paper California: AAPG Bulletin, v. 65, no. 7, p. 1256–1284. 332, p. 301–328, doi:10.1130/0-8137-2332-9.301 Noda, A., 2016, Forearc basins: Types, geometries, and re- Ramos-Garc´ıa, F., 1976, Prospeccion´ geof´ısica en la lationships to subduction zone dynamics: Boulder, Col- pen´ınsula de Baja California, in III simposio de geolog´ıa orado, Geological Society of America Bulletin, v. 128, del subsuelo zonas noreste: Reynosa, Tamaulipas, no. 5–6, p. 879–895, doi:10.1130/B31345.1 Mexico, Petroleos´ Mexicanos, Superintendencia Gen- Ortega-Rivera, A., 2003, Geochronological constraints on the eral de Exploracion,p.61´ –72. tectonic history of the Peninsular Ranges batholith: Rangin, C., 1978, Speculative model of Mesozoic geo- Boulder, Colorado, Geological Society of America Spe- dynamics, central Baja California to northeastern Sonora cial Paper 374, p. 297–335. (Mexico), in D. G Howell and K. A MacDougal, eds., Ortega-Rivera, A., E. Farrar, J. A. Hanes, D. A. Archibald, Mesozoic paleogeography of the western U.S.: Tulsa, R. G. Gastil, D. L. Kimbrough, M. Zentilli, M. Lopez-´ Oklahoma, SEPM Pacific Section 2, p. 85–105. Mart´ınez, G. Feraud,´ and G. Ruffet, 1997, Chronological Rangin, C., 1986, Contribution al` ’etude´ geologique´ du constraints on the thermal and tilting history of the Sierra System Cordillerain´ Mesozoique´ du nord-ouest du San Pedro Martir´ pluton, Baja California, Mexico,´ from Mexique: Paris, France, Societ´ eG´ eologique´ de France U/Pb, 40Ar/39Ar, and fission-track : Memoir 148, 136 p. Boulder, Colorado, Geological Society of America Bul- Robinson, J., 1979, Structure and stratigraphy of the northern letin, v. 109, no. 6, p. 728–745, doi:10.1130/0016-7606 Vizcaino Peninsula with a note in a Miocene recon- (1997)109<0728:CCOTTA>2.3.CO;2 struction of the Peninsula, in P. L. Abbott and G. Gastil, Pares-Regali, M. S., N. Uesugui, and A. S. Santos, 1974a, eds., Baja California geology: Field guides and papers: San Palinologia dos sedimentos Meso–Cenozoicos do Brasil I: Diego, California, San Diego State University, p. 77–89. Boletim Tecnico da Petrobras, v. 17, no. 4, p. 177–190. Robinson, J. W., 1975, Reconnaissance geology of the Pares-Regali, M. S., N. Uesugui, and A. S. Santos, 1974b, northern Vizcaino Peninsula, Baja California Sur, Mex- Palinologia dos sedimentos Meso–Cenozoicos do Brasil II: ico, M.S. thesis, San Diego State University, San Diego, Boletim Tecnico da Petrobras, v. 17, no. 4, p. 263–301. California, 114 p. Patterson, D. L., 1978, The foraminiferal biostratigraphy and Salinas-Marquez,´ F. M., J. G. Flores-Trujillo, J. Helenes, paleoecology of the type Rosario Formation, El Rosario, M. A. Tellez-Duarte,´ and F. J. Aranda-Manteca, 2016, Baja California del Norte, Mexico, M.S. thesis, University Paleoecolog´ıa y cronoestratigraf´ıa de las diatomeas del of California, Santa Barbara, California, 150 p. Miembro Los Indios en la Mesa La Mision,´ del Mioceno Patterson, D. L., 1984a, The Valle Formation, an integrated de Baja California, Mexico:´ Bolet´ındelaSociedad stratigraphy of turbidite deposits and the reconstruc- Geologica´ Mexicana, v. 68, no. 3, p. 537–552, doi: tion of the Late Cretaceous paleogeography of the 10.18268/BSGM2016v68n3a9. Vizcaino Basin, Baja California, Mexico, Ph.D. thesis, Santillan,´ M., and T. Barrera, 1930, Las posibilidades pet- University of California, Santa Barbara, California, rol´ıferas en la costa occidental de Baja California entre los 549 p. paralelos 30° y 32° de latitud norte: Anales del Instituto Patterson, D. L., 1984b, Los Chapunes and Valle sandstones: de Geolog´ıadeMexico,´ vol. 5, p. l–37, map. Cretaceous petrofacies of the Vizcaino Basin, in Sedlock, R. L., 1988, Tectonic setting of blueschist and island- V. A Frizzell Jr., ed., Geology of the Baja California arc terranes of west-central Baja California, Mexico:

HELENES ET AL. 1067

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 Geology, v. 16, no. 7, p. 623–626, doi:10.1130/0091- Toumarkine, M., and H. Luterbacher, 1985, Paleocene 7613(1988)016<0623:TSOBAI>2.3.CO;2. and Eocene planktic foraminifera, in H. M. Bolli, Sedlock, R.L., 2003, Geology and tectonics of the Baja J. B. Saunders, and K. Perch-Nielsen, eds., Plankton California peninsula and adjacent areas, in S. E. Johnson, stratigraphy, Cambridge Earth Sciences Series: S.R.Paterson,J.M.Fletcher,G.H.Girty,D.L.Kimbrough, Cambridge, United Kingdom, Cambridge University and A. Mart´ın-Barajas, eds., Tectonic evolution of north- Press, p. 87–154. western Mexico´ and the southwestern USA: Boulder, Umhoefer, P. J., 1987, Northward translation of “BAJA Colorado, Geological Society of America Special Paper 374, British Columbia” along the Late Cretaceous to Paleo- p. 1–42, doi:10.1130/0-8137-2374-4.1. cene margin of western North America: Tectonics, Silver, L. T., and B. W. Chappell, 1988, The Peninsular v. 6, no. 4, p. 377–394, doi:10.1029/TC006i004p00377. Ranges batholith: An insight into the evolution of the vanderZwaan,G.J.,F.J.Jorissen,andH.C.deStigter, Cordilleran batholiths of southwestern North America: 1990, The depth dependency of planktonic/benthic Earth and Environmental Transactions of the Royal So- foraminiferal ratios: Constraints and applications: Ma- ciety of Edinburgh: The Origin of Granites, v. 79, no. 2–3, rine Geology, v. 95, no. 1, p. 1–16, doi:10.1016/0025- p. 105–121, doi:10.1017/S0263593300014152. 3227(90)90016-D. Sliter, W. V., 1968, Upper Cretaceous foraminifera from Vaughn, J., K. P. Kodama, and D. P. Smith, 2005, Correction southern California and northwestern Baja California, of inclination shallowing and its tectonic implications: Mexico: Lawrence, Kansas, University of Kansas Pale- The Cretaceous Perforada Formation, Baja California: ontological Contributions 49, 141 p. Earth and Planetary Science Letters, v. 232, no. 1–2, Sliter, W. V., and R. A. Baker, 1972, Cretaceous bathymetric p. 71–82, doi:10.1016/j.epsl.2004.11.026. distribution of benthonic foraminifers: Journal of Fora- Vazquez-Garc´ ´ıa, A., and T. Schwennicke, 1996, Paleo- miniferal Research, v. 2, no. 4, p. 167–183, doi:10.2113 channels formed by bypass tempesites of the Tepetate /gsjfr.2.4.167. Formation at Arroyo El Conejo and Aguajito, Baja Smith, D. P., and C. J. Busby, 1993a, Mid-Cretaceous crust California Sur, Mexico, in Resumenes.´ XIII Con- extension recorded in deep-marine half-graben fill, Ce- vencion´ Geologica´ Nacional, Mexico:´ La Paz, dros Island, Mexico: Geological Society of America Bulletin, Mexico, Universidad AutonomadeBajaCalifornia´ v. 105, no. 4, p. 547–562, doi:10.1130/0016-7606(1993) Sur, p. 59. 105<0547:MCCERI>2.3.CO;2. Wetmore, P. H., C. Herzig, H. Asleben, M. Sutherland, Smith, D. P., C. J. Busby, and K. D. Berry, 1993b, Litho- K. L. Schmidt, P. W. Schultz, and S. R. Patterson, 2003, stratigraphy of Cretaceous strata, Cedros Island, Baja Mesozoic tectonic evolution of the Peninsular Ranges California, Mexico: Vargas and Pinos formations defined: of southern and Baja California, in S. Johnson, S. R. Paterson, Cretaceous Research, v. 14, no. 3, p. 337–364, doi: J.M.Fletcher,G.H.Girty,D.L.Kimbrough,andA.Mart´ın- 10.1006/cres.1993.1024. Barajas, eds., Tectonic of northwestern Mexico and south- Smith, J. T., 1984, Miocene and Pliocene marine mollusks and western USA: Boulder, Colorado, Geological Society of preliminary correlations, Vizcaino Pen´ınsula to Arroyo La America Special Paper 347, p. 93–116. doi:10.1130 Pur´ısima, northwestern Baja California Sur, in V. A. Frizzell /0-8137-2374-4.93 Jr., ed., Geology of the Baja California peninsula: Tulsa, Whalen, P. A., and E. S. Carter, 2002, Pliensbachian (lower Oklahoma, SEPM Pacific Section 39, p. 197–217. Jurassic) Radiolaria from Baja California Sur, Mexico: Stewart, D. R. M., and P. N. Pearson, 2000, Plankrange: A Micropaleontology, v. 48, no. 2, p. 97–151, doi:10.2113 database of planktonic foraminiferal ranges, accessed /48.2.97. October 29, 2018, http://palaeo.gly.bris.ac.uk/Data Whalen, P. A., and E. A. Pessagno Jr., 1984, Lower Jurassic /plankrange.html. Radiolaria, San Hipolito Formation, Vizcaino Peninsula, Suarez-Vidal,´ F., 1987, The calcareous facies of the Alisitos Baja California Sur, in V. A. Frizzell Jr., ed., Geology of Formation, evidence for an Early Cretaceous tectonic the Baja California peninsula: Tulsa, Oklahoma, SEPM calm: Ciencias Marinas, v. 13, no. 3, p. 131–154, doi: Pacific Section 39, p. 53–65. 10.7773/cm.v13i3.537. White, J. M., 2008, Palynodata datafile: 2006 version: Ottawa, Tellez-Duarte,´ M. A., J. Helenes-Escamilla, and J. A. Mart´ın- Ontaria, Canada, Geological Survey of Canada Open File Barajas, 2006, Stratigraphy and depositional history of 5793, accessed October 29, 2018, https://geoscan.nrcan the Sepultura Formation (Paleocene) in northwestern .gc.ca/starweb/geoscan/servlet.starweb?path=geoscan/fulle Baja California, Mexico,´ in G. HGirtyand J.D Cooper, .web&search1=R=225704. eds., Using stratigraphy, sedimentology, and geochemis- Williams, G. L., H. Brinkhuis, M. A. Pearce, R. A. Fensome, try to unravel the geologic history of the southwestern and J. W. Weegink, 2004, Southern ocean and global Cordillera: A volume in honor of Patrick L. Abbott: Tulsa, dinoflagellate cyst events compared: Index events for the Oklahoma, SEPM Pacific Section 101, p. 123–144. Late Cretaceous–Neogene, in N. F. Exon, J. P. Kennett, Thomas, F. C., and M. G. Murney, 1985, Techniques for and M. J. Malone, eds., Proceedings of the Ocean Drilling extraction of foraminifers and ostracodes from sediment Program, Scientific Results: College Station, Texas, Ocean samples, in Canadian Technical Report of Hydrography Drilling Program, v. 189, p. 1–98. and Ocean Sciences No. 54: Ottawa, Ontario, Canada, Williams, G. L., and J. P. Bujak, 1985, Mesozoic and Ceno- Department of Fisheries and Oceans, p. 1–24. zoic dinoflagellates, in H. M. Bolli, J. B. Saunders, and

1068 Stratigraphy of Late Albian–Middle Eocene Vizcaino Basin, Baja California, Mexico

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021 K. Perch-Nielsen, eds., Plankton stratigraphy: Cam- Yeo, R. K., 1982, The stratigraphy and sedimentology of the bridge, United Kingdom, Cambridge University Press, Upper Cretaceous sediments of southwestern California Cambridge Earth Sciences Series p. 847–964. and Baja California, Mexico, Ph.D. thesis, Rice Univer- Williams, G. L., L. E. Stover, and E. J. Kidson, 1993, Morphology sity, Houston, Texas, 604 p. and stratigraphic ranges of selected Mesozoic–Cenozoic di- Yeo, R. K., 1984, Sedimentology of Upper Cretaceous strata, noflagellate taxa in the northern hemisphere: Ottawa, On- northern Baja California, Mexico, in P. L. Abbott, ed., tario, Canada, Geological Survey of Canada 92, 137 p. UpperCretaceousdepositionalsystems,southern Wood, G. D., A. M. Gabriel, and J. C. Lawson, 1996, Paly- California–northern Baja California: Tulsa, Oklahoma, nological techniques—Processing and microscopy, in J. SEPM Pacific Section 36, p. 109–120. Jansonius and D. C. McGregor, eds., Palynology: Chapter Zinsmeister, W. J., and L. M. Paredes-Mej´ıa, 1988, Paleocene 3, principles and applications: Dallas, Texas, American biogeography of the west coast of North America: A look Association of Stratigraphic Palynologists, p. 29–50. at the molluscan fauna from the Sepultura Formation, Wrenn, J. H., S. L. Duffield, and J. A. Stein, eds., 1986, Papers Mesa San Carlos, Baja California, in M. V. Filewicz and from the first symposium on Neogene dinoflagellate cyst R. L. Squires, eds., Paleogene stratigraphy, West Coast biostratigraphy: Dallas, Texas, American Association of of North America: Tulsa, Oklahoma, SEPM Pacific Stratigraphic Palynologists, Contribution Series 17, 225 p. Section 58, p. 9–22.

HELENES ET AL. 1069

Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/103/5/1045/4698244/bltn17349.pdf by guest on 24 September 2021