You are on page 1of 17

Variation of Cenozoic extension and volcanism across the southern Sierra Madre Occidental volcanic province, Mexico

ngel F. Nieto-Samaniego* Luca Ferrari Susana A. Alaniz-Alvarez

Unidad de Ciencias de la Tierra, Universidad Nacional Autnoma de Mxico, Campus Juriquilla, Apdo. Postal 1742, Quertaro, Qro., 76001, Mxico

Guillermo Labarthe-Hernndez Instituto de Geologa, Universidad Autnoma de San Luis Potos, Dr. Manuel Nava 5, Zona Universitaria, San Luis Potos, S. L. P., 78240, Mxico Jos Rosas-Elguera Centro de Ciencias de la Tierra, Universidad de Guadalajara, Blvd. M. Garca Barragn y Calz. Olmpica, 44840, Guadalajara, Jalisco, Mxico

ABSTRACT The middle to late Cenozoic tectonic-magmatic evolution of the Sierra Madre Occidental volcanic province south of the Tropic of Cancer is summarized and analyzed for the first time, based on new geologic and structural work and published information. In the eastern part of the study region (Mesa central physiographic province) silicic volcanism occurred in a short-lived episode culminating at ca. 30 Ma and was followed by crustal-scale extension between 30 and 27 Ma. In the western part of the study area (Sierra Madre Occidental physiographic province) a voluminous episode of ignimbrite volcanism at 2421 Ma was succeeded by east-west extension that produced regularly spaced grabens affecting only the upper crust. In the westernmost part of the study region, an andesitic to rhyolitic arc, formed between 17 and 12 Ma, was affected by crustal-scale, north-northwesttrending, extensional faulting, leading to the formation of the Gulf of California. In the Mesa central the maximum extension was oriented approximately east-west and amounted to ~20%. In the eastern Sierra Madre Occidental physiographic province extension was only 8% and oriented approximately east-west. We observe that trenchward shifting of the climax of subduction volcanism and extension occurred during late Oligocene, early Miocene, and late Miocene time. Comparison with the offshore tectonics indicates that the first two tectonic-magmatic pulses coincide with periods of fast spreading at the Pacific-Farallon boundary, south of the Shirley fracture zone. We propose that increases in the spreading rate are related to periods of high subduction rate, which in turn correspond to episodes of retreating subduction. A retreating slab may have generated a flux of hotter asthenospheric material into the mantle wedge, producing widespread melting at the base of the crust as well as intraarc extension in the overriding plate. Boundary conditions (i.e., plate tectonics) ultimately determined timing, magnitude, and orientation of extension, whereas volcanic and tectonic styles are controlled by the internal structure of crustal blocks and by the gravitational and thermal effects of magmatism.
*E-mail: afns@servidor.unam.mx.

INTRODUCTION Three major tectonic events occurred in the northern half of the Mexican Pacific coast during Tertiary time: (1) an episode of subduction that lasted until middle Miocene time (Atwater, 1989); (2) a progressive waning of subduction through various episodes of microplate capture induced by the approach of the East Pacific Rise to the trench (Lonsdale, 1991); and (3) the opening of the Gulf of California (e.g., Stock and Hodges 1989; Lonsdale, 1989). These events shaped the volcanic and structural configuration of the central and western parts of northern Mexico, forming the Sierra Madre Occidental volcanic province and a wide extensional province, referred to as the real southern Basin and Range (Henry and Aranda-Gmez, 1992) (Fig. 1). In the past two decades there have been many contributions to the reconstruction of the volcanic evolution of the northern and central part of the Sierra Madre Occidental volcanic province (McDowell and Keizer, 1977; McDowell and Clabaugh, 1979; Henry and Fredrikson, 1987; Wark et al., 1990; Aguirre and McDowell, 1991). Structural studies in Sinaloa (Henry, 1989), Durango (Aguirre-Diaz and McDowell, 1993; Aranda-Gmez et al., 1997), and Sonora (Gans, 1997; McDowell et al., 1997) (Fig. 2) provide information on the timing and style of the extensional tectonics. In general, these works report a westward migration of extension and volcanism during post-Eocene time, as well as an inception of extension well before the first direct interaction between the Pacific and the North American plates. The Sierra Madre Occidental volcanic province south of the Tropic of Cancer (hereafter referred to as the southern Sierra Madre Occidental volcanic province) is less well-known, at least in the international geologic literature. This region can be divided into the Sierra Madre Occidental and the Mesa central physiographic provinces (Fig. 1). Geologic mapping and structural studies carried out in the past 15 yr have documented the volcanic and tectonic evolution of the Mesa central in good detail (Labarthe-Hernndez et al., 1982; Aranda-Gmez et al., 1989; Henry and Aranda-Gmez, 1992; Nieto-Samaniego, 1990; Martnez-Reyes, 1992; Quintero-Legorreta, 1992; Nieto-Samaniego and Alaniz-Alvarez, 1994; Nieto-Samaniego et al., 1996, 1997), but most of this information is published in university reports or national journals, with limited international distribution. The timing of the volcanic and extensional events in the Sierra Madre

Data Repository item 9919 contains additional material related to this article.
GSA Bulletin; March 1999; v. 111; no. 3; p. 347363; 10 figures; 2 tables.

347

NIETO-SAMANIEGO ET AL.

Figure 1. Geodynamic setting, physiography (A), and Cenozoic volcanic arcs (B) of Mexico. The study region (boxed) comprises the southern Sierra Madre Occidental volcanic province (SMO VP) and extends through the physiographic provinces of the SMOc (Sierra Madre Occidental), Mesa central (MC), Sierra Madre Oriental (SMOr), and Mexican volcanic belt (MVB). Tertiary extension is based on Henry and Aranda-Gomez (1992) and this work. Extension south of the MVB is not shown. OligoceneMiocene volcanic rocks south of the MVB are diagonally ruled.

Occidental is less well-known than for the Mesa central. Good stratigraphic and geochronologic information is available only for a few places (Lyons, 1988; Scheubel et al., 1988; Nieto-Obregn et al., 1985), whereas there are only reconnaissance geologic and structural works (Gastil et al., 1978; Damon et al., 1979; Webber et al., 1994; Ferrari, 1995; Ferrari et al., 1999) and several scattered ages of volcanic rocks for the remaining region (Clark et al., 1981; Nieto-Obregn et al., 1981; Moore et al., 1994). Nevertheless, these data permit reconstructing the spatial evolution of the volcanism. In addition, we have studied the western Sierra Madre Occidental at a regional scale through satellite images, aerial photographs, and reconnaissance studies along roads. In this paper we integrate and summarize our regional stratigraphic and structural information with published geological and geophysical data to provide the first synthesis of the middle to late Cenozoic volcanic and tectonic evolution of the southern Sierra Madre Occidental volcanic province. Our data document an Oligocene three-dimensional strain in the southern Mesa central, an early Miocene two-dimensional strain in the eastern part of the southern Sierra Madre Occidental , and a major middle to late Miocene extensional phase of deformation in the western part of the southern Sierra Madre Occidental. We propose that variations in deformation and volcanism observed since Oligocene time in the southern volcanic province are due to the difference in thickness and composition of the major crustal blocks that form the Sierra Madre Occidental and the Mesa central. When compared to the rest of the Sierra Madre Occidental volcanic province, our results indicate that the westward shift of the zone of maximum volcanic and tectonic activity was a general phenomenon in OligoceneMiocene time, although in the southern volcanic province geologically significant extension affected a much wider region than to the north. The westward migration is tentatively related to the along-trench variation of the FarallonNorth America subduction rate during Oligocene and Miocene time.

110N 30N

105N

SON
CHIH

25N

SIN DGO

ZAC

Tropic of Cancer
0 20N km 500

NAY

SLP GTO

JAL

Figure 2. Boundaries of Mexican states cited in the text. SON Sonora, SINSinaloa, CHIHChihuahua, NAYNayarit, DGO Durango, ZACZacatecas, AAguascalientes, SLPSan Luis Potos, JALJalisco, GTOGuanajuato.

TOPOGRAPHY AND CRUSTAL THICKNESS The spatial variations in the Sierra Madre Occidental volcanic province are illustrated in an east-west topographic profile in Figure 3. This profile distinguishes three physiographic provinces with different elevations: an eastern province with mean elevation of 1400 m corresponding to the Sierra

348

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO

TABLE 1. K/AR AGES FOR THE SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE Sample SLP 7* GL 409 MCU-1 MP-B1 GL 406 CTO-02 GL 137 GL 241 SLP 19* Site Cerro El Bey, SLP N of Comanjilla, Gto. Cerro La Montaa, Gto. Campuzano, Gto. E of Calvillo, Gto. Sierra Nochistln Cerro Cano, Gto. E of Calvillo, Gto. La Salitrera, SLP Rock type Andesite Andesite Ignimbrite Ignimbrite Ignimbrite Ignimbrite Basalt Ignimbrite Bas.- Andesite Material dated Hb WR WR WR WR San WR WR Pl Long (W) 100.68 101.45 101.28 101.14 101.04 102.84 101.07 101.08 100.52 Lat (N) 22.44 21.08 21.18 20.84 21.10 21.59 21.09 21.12 22.93 K (wt%) 0.53 2.23 4.63 6.43 4.56 7.89 1.59 3.71 0.57
40Ar* (mol/gr) 3.50E-11 1.43E-10 2.73E-10 3.02E-10 2.39E-10 3.67E-10 8.05E-11 1.83E-10 1.31E-11 40Ar*

(%) 52.0 64.3 93.1 98.2 95.8 76.2 67.7 96.8 35.0

Age (Ma) 37.6 32.9 29.3 28.6 26.9 26.6 25.9 25.2 13.2

1 (Ma) 1.9 1.6 1.5 1.4 1.3 0.7 1.3 1.3 0.6

Notes: Hbhornblende, Sansanidine, Plplagioclase, gmsgroundmass, WRwhole rock. *Analyses performed at Instituto Mexicano del Petroleo, Mexico City, in 1988. Analyses performed at Teledyne Isotopes, in 1995. Analyses performed at Geochron Laboratories, Cambridge, Massachusetts, in 1994.

Madre Oriental; a central province with a mean elevation of 2100 m corresponding to the Mesa central; and a western province with a mean elevation of 1800 m corresponding to the Sierra Madre Occidental. The difference in uplift between the Mesa central and the Sierra Madre Occidental is even more pronounced if one looks at the geologic units exposed in the two provinces. In the Mesa central the Mesozoic basement is commonly exposed from 2000 to 2700 m (Fig. 3), whereas in the Sierra Madre Occidental rocks of the same age are not exposed, even at 800 m inside the major depression of the area. The crustal thickness of the physiographic provinces also shows notable differences: based on a gravimetric profile along the Tropic of Cancer, Campos-Enriquez et al. (1994) proposed crustal thickness of 37 km for the Sierra Madre Oriental, 33 km for the Mesa central and 41 km for the Sierra Madre Occidental. Molina-Garza and Urrutia-Fucugauchi (1993) estimated a similar value for the Sierra Madre Occidental from interpretation of Bouguer anomaly along a north-south profile at long 103W. Using group velocity of surface waves, Fix (1975) estimated a crustal thickness of 30 km for central Mexico, whereas Rivera and Ponce (1986) proposed a value of 42 km for the Sierra Madre Occidental. The latter value was also obtained by Meyer et al. (1958) using seismic refraction data. By modeling gravity and seismic refraction data, Couch et al. (1991) calculated a crustal thickness ranging from 40 km at the center of the Sierra Madre Occidental to 25 km along the coast of the Gulf of California. These data indicate that the thinner and topographically elevated Mesa central is bounded by thicker and depressed crustal blocks of the Sierra Madre Oriental and Sierra Madre Occidental (Fig. 3). The exposure of the Mesozoic basement in the Mesa central indicates a larger uplift of this block with respect to the adjacent ones. In the following sections we will show that the Mesa central and the Sierra Madre Occidental also have important differences in the age and style of volcanism and extension. REGIONAL STRATIGRAPHY In this section we present a generalized stratigraphy of the Mesa central and Sierra Madre Occidental that integrates the results of previous works with our reconnaissance investigations of Tertiary rocks in the region. This synthesis is supported by 91 published ages (Table DR1, in GSA Data Repository1) and nine new dates reported in Table 1.
Data Repository item 9919, data table, is available on request from Documents Secretary, GSA, P.O. Box 9140, Boulder, CO 80301. E-mail: editing@geosociety.org.
1GSA

Mesozoic Basement The oldest rocks of the region are exposed in the Mesa central and belong to two major paleogeographic domains: the Guerrero terrane island-arc assemblage (Centeno-Garcia et al., 1993) and the Sierra Madre Oriental marine sedimentary sequence. The Guerrero terrane, exposed in the western part of the southern Mesa central, is a volcanic-sedimentary sequence made of flysch-like sediments interbedded with pillow basalt, which shows a low metamorphic grade and severe contractile deformation. The succession is exposed in the Guanajuato and the Aguascalientes areas (Fig. 4).On the basis of paleontologic evidence (Chiodi et al., 1988) and K-Ar ages (Monod et al., 1990), the succession has been interpreted as Early Cretaceous in age. To the north and east of the southern Mesa central, a Cretaceous marine carbonate and clastic sequence (e.g., Carrillo-Bravo, 1971; Labarthe-Hernndez et al., 1982) underwent thin-skinned contraction during the Laramide orogeny. In the western Sierra Madre Occidental the existence of a prevolcanic Mesozoic basement is substantiated by localized outcrops of argillite and limestone exposed along the lower part of the Rio Grande de Santiago. These rocks occur in small and isolated outcrops on top of late Oligocene to early Miocene intrusive stocks, suggesting that they were roof pendants uplifted by the plutons. Extensive exposures of Mesozoic rocks are not known in the Sierra Madre Occidental south of lat 2230 N and west of the longitude of Zacatecas (Fig. 3). Paleogene Continental Sedimentary Rocks This group includes two continental postorogenic units: the Guanajuato conglomerate (Edwards, 1955) exposed in Guanajuato, and the Cenicera Formation in San Luis Potos (Labarthe-Hernndez et al., 1982). They consist of polymictic conglomerate and sandstone deposited in coalescing alluvial fans after the Laramide orogeny. According to paleontologic determinations (Edwards 1955) and a K-Ar date from an interbedded andesite (Aranda-Gmez and McDowell, 1998), the Guanajuato conglomerate is Eocene in age. Palynologic determinations indicate a Paleocene-Eocene age for the Cenicera Formation (Labarthe-Hernndez et al., 1982). The overlying Casita Blanca Andesite, dated as 44.1 2.2 Ma (LabartheHernndez et al., 1982), gives a minimum age for the Cenicera Formation. Eocene Volcanic Group Eocene volcanic rocks are widespread in the Sierra Madre Occidental volcanic province (Aguirre-Daz and McDowell, 1991, and references therein),

Geological Society of America Bulletin, March 1999

349

104W

102W

100 km

100W

22N

22N

21N

21N

104W

102W

100W

Figure 3. (A) Generalized topographic map of the southern Sierra Madre Occidental volcanic province. Thick lines bound physiographic provinces and correspond to major tectonic structures discussed in the text. (I) Sierra Madre Occidental, (II) Mesa central, and (III) Sierra Madre Oriental physiographic provinces. Stars represent extensive outcrops of the Mesozoic basement. Thick dashed line indicates the trace of topographic profile AA. SMASan Miguel de Allende; NVMNuevo Valle de Moreno. (B) Topographic profile based on 1:1 000 000 scale topographic map filtered for the high frequencies. The mean elevation of the three physiographic provinces has been calculated without considering the coastal plain. (C) Interpretation of the crustal structure along the profile based on surface geology and geophysical modeling (see text for references). Structure of middle and lower crust is not to scale and is based on geologic considerations discussed in the text.

350

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO


TABLE 2. COMPUTED PALEOSTRESS TENSORS FOR THE SIERRA MADRE OCCIDENTAL Eastern Sierra Madre Occidental (Calvillo, Juchipila, Tlaltenengo and Bolaos graben) Trend/Plunge 1 2 3 225/75 007/11 099/08 stress ratio = = Magnitudes relative to 1 100 44 39 Western Sierra Madre Occidental (Aguamilpa Dam, Jess Maria, La Cofrada, Paso de Lozada, Presa Baadero, and Sierra Pajaritos) Trend/Plunge Magnitudes relative to 1 222/81 333/03 066/07 stress ratio = = 100 32 27

2 3 1 3

= 0.08

2 3 1 3

= 0.07

coefficient of friction = = 0.6; n = 58 Note: The method of Reches (1987) assumes that slip obeys the Coulomb criterion. Coefficient of friction, cohesion, the misfit angle between observed and calculated slip axis (V), and the mean misfit angle between general and ideal principal stresses axes () are used as criteria to separate the fault groups. These analyses simultaneously satisfy the following constraints: 0.6 > > 0.4; cohesion near zero; < 25;. < 40.

coefficient of friction = = 0.4; n = 61

although they are commonly obscured by the extensive Oligocene rhyolitic cover. In the southern Mesa central, Eocene rhyolitic ignimbrites and domes and andesite lava flows crop out locally in Zacatecas (Ponce and Clark, 1988; Lang et al., 1988), Guanajuato (Bufa Rhyolite, Gross, 1975), San Luis Potos (Labarthe-Hernndez et al., 1982), and Aguascalientes (Nieto-Samaniego et al., 1996). A sequence of andesitic flows on the Mesozoic basement 80 km northwest of San Luis Potos yields a K-Ar age of 37.6 1.9 Ma (Table 1). In the Sierra Madre Occidental the oldest volcanic rocks are highly altered and pervasively fractured microcrystalline andesitic lavas, which are exposed north of Guadalajara and yield a middle Eocene age (Webber et al., 1994; Ferrari et al., 1999). Their base is not exposed and the top is eroded and locally covered unconformably by red sandstone and conglomerate containing andesitic fragments. The minimum thickness of this succession is 400 m. Based on the geographic distribution of the outcrops mentioned, the Eocene volcanic rocks, although scattered and of lesser volume, were probably more extensive than the Oligocene-Miocene rocks. Oligocene Volcanic Group Rhyolite domes and lava flows with minor intercalated pyroclastic deposits are volumetrically the most important part of the Tertiary volcanism in the Mesa central. These rocks are exposed all over the southern Mesa central, often forming the core of major ranges (Fig. 4). Domes are typically about 2 km in size. They are associated with ore deposits of tin, and with rhyolite lava that commonly contains topaz. The age is well constrained to be Oligocene by five K-Ar dates in San Luis Potos and Guanajuato (30.8 0.8 to 29.2 0.8 Ma, Table DR1, see footnote 1). This group is absent west of Aguascalientes in the Sierra Madre Occidental, as well as to the east in the Sierra Madre Oriental (Fig. 4). Oligocene rhyolitic ignimbrites with subordinate andesitic lava flows are exposed in the southern Mesa central and the southeastern Sierra Madre Occidental (Fig. 4). The aggregate thickness of these rocks is usually less than 200 m. No clear caldera structures have been genetically or spatially associated with these ignimbrites. In San Luis Potos and northern Guanajuato (Fig. 2) the ignimbrites have K-Ar ages ranging from 29.0 to 26.8 Ma (Table DR1, see footnote 1). This time span is confirmed by three new whole-rock K-Ar dates ranging between 29.3 and 26.9 Ma from the Sierra de Guanajuato (Table 1). A second succession of ignimbritic sheets with latest Oligocene ages (Table DR1, see footnote 1) is exposed in the southernmost part of the

Mesa central. The northernmost exposed ignimbrite and one interbedded basaltic flow of this succession give ages of 25.2 1.3 Ma and 25.9 1.3 Ma, respectively (Table 1). In the Sierra Madre Occidental, Oligocene ignimbrites are only reported in the Santa Maria del Oro and Juchipila areas (Fig. 5) (Table DR1, see footnote 1). In addition we obtained a K-Ar age of 26.6 0.7 Ma (Table 1) from one of the uppermost ignimbrites capping the Sierra de Nochistln (Fig. 5). These old ignimbrites are commonly interbedded with and capped by volcaniclastic sequences made of clay, sandstone, and conglomerate, and can be interpreted as the westernmost extension of the Mesa central Oligocene volcanic sequence. Miocene Volcanic Group The Miocene volcanic group consists of three main components: early Miocene silicic volcanic rocks, early Miocene basaltic andesites, and late Miocene basalts. A significant part of the southern Sierra Madre Occidental is covered by rhyolitic ignimbrites and minor domes emplaced in a short time span (24 to 21 Ma, Table DR1, see footnote 1). More than 1200 m of ashflow tuffs with ages between 23.7 and 20.1 Ma are exposed in the Bolaos graben, and the upper part of this succession can be followed up to about 100 km southward (Scheubel et al., 1988; Lyons, 1988; Moore et al., 1994). Ignimbrites with an aggregate thickness of about 1000 m are exposed in the Sierra de Santa Lucia, in the core of the Sierra Madre Occidental (Fig. 5). The uppermost tuffs of this succession have been dated at 23.5 Ma (Clark et al., 1981), whereas the base yielded an age of about 25 Ma (Ferrari et al., 1997a). These silicic ignimbrites resemble the 1000-m-thick and ca. 23-Ma El SaltoEspinazo del Diablo sequence, exposed ~100 km to the north in southwestern Durango (McDowell and Keizer, 1977). Early Miocene ash flows are not found in the Mesa central, which suggests a westward younging of the ignimbrite flare up. Ash flows and rhyolites with similar ages are exposed in southernmost Baja California (Hausback, 1984) and could represent distal facies of the early Miocene Sierra Madre Occidental arc. North of Guadalajara the Sierra Madre Occidental succession is dissected by north- to north-northeasttrending grabens (Fig. 5). Basaltic andesites are found as lavas at the bases of these grabens and as dikes intruded parallel to the normal faults. The age of these rocks is 2119 Ma in the Bolaos graben (Nieto-Obregn et al., 1981) and 21.8 Ma in the Tlaltenango graben (Moore et al., 1994).

Geological Society of America Bulletin, March 1999

351

NIETO-SAMANIEGO ET AL.

Figure 4. Geologic map of the Mesa central showing the main stratigraphic units discussed in the text. The figure is based on geologic sheets at scales 1:100 000, 1:50 000, and 1:20 000 published by the Geologic Institute of the San Luis Potos University between 1977 and 1995 (see NietoSamaniego et al., 1996; Labarthe-Hernndez et al., 1982, for the complete list of references) and Nieto-Samaniego (unpublished data). PHPalo Hurfano Volcano; SCSierra de Codornices; SGSierra de Guanajuato; NVMNuevo Valle de Moreno; SSMSierra de San Miguelito. The region shown in Figure 6 is outlined.

The southernmost Mesa central is covered by an extensive basaltic plateau and by andesitic stratovolcanoes (Ferrari et al., 1994a). Stratigraphic evidence indicates that the stratovolcanoes are older than the basaltic plateau extending to the south (Valdez-Moreno and Aguirre-Diaz, 1997; Prez-Venzor et al., 1997). Ages range from about12 Ma for one of the stratovolcanoes (Prez-Venzor et al., 1997) to 10.58.1 Ma for the basaltic plateau (Table DR1, see footnote 1). Basaltic to andesitic lava flows also cap the silicic succession in the Mesa central to the north. The only age for these rocks is 13.2 0.6 Ma (Table 1) from one flow of an extensive basaltic-andesitic succession overlying the Oligocene ignimbrites about 80 km east-southeast of San Luis Potos. This

age is considered representative of the postignimbrite mafic volcanism of the region, because it coincides with the age of other basalts in the southernmost Mesa central (Table DR1, see footnote 1), and with the Metates (McDowell and Keizer, 1977) and Los Encinos basalts (Luhr et al., 1995), which are exposed in eastern Durango and northern San Luis Potos, respectively. Along the southern limit of the Sierra Madre Occidental, the early Miocene ignimbrites are covered by 10.28.5 Ma (Table DR1, see footnote 1) basaltic lava flows, well exposed along the Rio Santiago canyon north of Guadalajara (Fig. 5), where they have an aggregate thickness exceeding 600 m. This succession also extends east of Guadalajara for more than 130 km in the Los Altos de Jalisco (Fig. 5), with an average thickness of 200 m and ages of 11 to 8

352

Geological Society of America Bulletin, March 1999

Escuinapa

Ib Ia II
Zacatecas
Sierra St

III
Salinas de Hidalgo

Ate

Jess Maria
Mesa delNayar

ngo

Rio
Bolaos

S. Pajari

nes

lica

de M

och

Rio
Pochotitn

istl

Sa
tos

oro

Villa
n
Sier ra

Sierra A

a nti

go

Pte.de Camotln

Jalpa
de N

de R

Aguamilpa

Tlaltenango

Ojuelos
Calvillo

Sie

rra

Tepic

Juchipila

Lagos deMoreno
S.S.B.

eyes

P San
El Zopilote

Aguascalientes

grab

22N

o edr

S.S.M.

en

San Pedro

Rio

a.Lucia

Acaponeta

Pinos

SanLuis Potos

S.C.

Pacific Ocean
Sa nP ed ro -C eb

Sta. Maria del Oro

Pla nd eB arr an

Rio Sa nti

Dolores Hidalgo

San Luis de la Paz


San Cristobal

21N

ag

Guanajuato
cas -Sa
S ta.Rosa Dam

Geological Society of America Bulletin, March 1999


or CB1 uc og ra be n

nta

Punta Mita

Ro

Jalisco block
0 50 100 km

sa

graGuadalajara ben
PR9

Rio

rd Ve

Len El Ba jo

San Miguel de Allende

Altos de Jalisco

gr ab en
Celaya

Quertaro

106W

104W

Atotonilco

102W

Penjamillo graben

Figure 5. Structural map and generalized stratigraphy of the southern Sierra Madre Occidental volcanic province, based on sources cited in the text for the Mesa central (MC) and on interpretation of satellite image, digital terrain models, aerial photographs and reconnaissance field work (Ferrari and Rosas-Elguera, 1998, personal commun.). Thick gray lines are boundaries of structural domains described in the text. IIISierra Madre Oriental, II

Mesa central, Iaeastern Sierra Madre Occidental, Ibwestern Sierra Madre Occidental. S.C.Sierra del Cubo; S.S.M.Sierra de San Miguelito; S.S.B.Sierra de Santa Barbara; PR9deep geothermal well of La Primavera caldera; CB1deep geothermal well of Ceboruco volcano. Stratigraphic columns not to scale.

353

NIETO-SAMANIEGO ET AL.

Ma (Ferrari et al., 1994a, 1996); these ages allow them to be correlated with the late Miocene basalts and andesites described from the Mesa central. Volcanic Rocks Related to the Gulf of California In the southwestern part of the Sierra Madre Occidental, north of Santa Maria del Oro, eroded andesitic volcanoes commonly cover the early Miocene ash flows. One of these centers has been dated as 14.7 Ma (Table DR1, see footnote 1). Gastil et al. (1979) also reported a succession of andesitic and basaltic flows with subordinate rhyolites with ages between 20.4 and 13.8 Ma south of Tepic (Table DR1, see footnote 1). These rocks suggest the existence of a middle Miocene intermediate-composition volcanic arc in the western part of the Sierra Madre Occidental. This volcanism represents the southern occurrence of the early to middle Miocene andesitic arc, recognized by Gastil et al. (1979) and Fensby and Gastil (1991) around the Gulf of California, that was subsequently separated along its axis by the rifting process. A plateau of alkali basalts with ages of 11 to 8.9 Ma (Table DR1, see footnote 1) is exposed between the Sierra Madre Occidental and the Pacific coast north of Tepic (Fig. 5). These rocks unconformably overlie upon tilted ignimbrites of the Sierra Madre Occidental. Several mafic dikes with ages of 1211 Ma (Damon et al., 1979; Clark et al., 1981; Table DR1, see footnote 1) and with north-south to northwest trends are found in the westernmost part of the Sierra Madre Occidental. This mafic, commonly alkalic volcanism marks the initiation of the rifting process at the mouth of the Gulf of California. Quaternary Volcanic Rocks The youngest volcanic event consists of scattered cinder cones, maars, and lava flows emplaced in the northern part of the Mesa central. These volcanic rocks include nephelinite, basanite, alkaline basalt, and basalt. According to 8 K-Ar age determinations, they are early to middle Pleistocene in age (Aranda-Gomez and Luhr, 1996). Continental Sedimentary Deposits Poorly to mildly consolidated alluvial and lacustrine deposits fill numerous basins cut in the Tertiary succession. They consist of conglomerate, gravel, sand, sandstone, and subordinate marl, limestone, and chert. These deposits probably span the entire late Tertiary time and possibly Quaternary time, as indicated by intercalation of Miocene ignimbrites and basalts and by Pleistocene vertebrate fossils discovered in the upper part of the sequence near San Miguel de Allende (Carranza-Castaeda et al., 1994) and Aguascalientes (Reynoso-Rosales and Montellano-Ballesteros, 1994). The Juchipila graben is floored by a stratified volcanic-sedimentary succession consisting of 10 to 20 m of a coarse volcanic conglomerate, followed by as much as 250 m of lacustrine clays, silts, and limestone. Alluvial deposits (gravel and sandstone) of unknown age also fill the southernmost part of the Tlaltenango graben. Summary of the Volcanic Evolution The variation in the stratigraphic record across the southern parts of the Mesa central and Sierra Madre Occidental is summarized in the stratigraphic columns of Figure 5. Eocene, mostly andesitic, volcanism appears to have occurred across the entire Mesa central and in the eastern part of the Sierra Madre Occidental. In early Miocene time voluminous rhyolitic ignimbrites were emplaced in the Sierra Madre Occidental while erosion and continental sedimentation were underway in the Mesa central. In late Miocene time

basaltic to andesitic volcanism related to the Mexican volcanic belt and to the opening of the Gulf of California occurred at the southern boundaries of the Mesa central and the Sierra Madre Occidental. A few basalts and andesites were also emplaced within the Mesa central but are not known in the Sierra Madre Occidental. In Quaternary time, minor alkaline basalts were emplaced in the Mesa central only. The most striking difference between the Mesa central and the Sierra Madre Occidental concern the age and the style of the silicic volcanism: in the former, silicic volcanism consists largely of rhyolitic domes and occurred mostly in Oligocene time; in the latter, it was essentially represented by caldera-related ignimbrites of early Miocene age. We note a westward (or west-southwestward) shift of the focus of silicic volcanism in Oligocene to Miocene time, similar to that observed in the northern and central Sierra Madre Occidental (Gans, 1997; Aranda-Gomez et al., 1997). MIDDLE TO LATE TERTIARY FAULTING IN THE SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE Geometry and Kinematics The southern volcanic province is bounded to the south and also divided longitudinally by major crustal structures (Figs. 3 and 5). Although the geomorphic expression of these structures is locally modest, their tectonic significance is given by the fact that they bound different types of basement and/or control the style of Cenozoic extension and volcanism. Faulting also affected the interior of these blocks at different times. In this section we describe the geometry, kinematics, and age of the boundary faults and of the other faulting inside the Mesa central and Sierra Madre Occidental blocks. Boundary Fault Systems. The boundaries between the Mesa central and Sierra Madre Occidental and Oriental are major north-south fault systems, along which Cenozoic normal movement formed half grabens. The boundary between Sierra Madre Oriental and the Mesa central is a complex system of north-northwest to north-trending and west-dipping normal faults, here named the San Miguel de AllendeCatorce fault system (Fig. 5). The fault system separates the intensely and complexly faulted Mesa central to the west from the Sierra Madre Oriental, which shows no Cenozoic extension. The southern part of this structure is the TaxcoSan Miguel de Allende fault, first identified by Demant (1978) as a major tectonic boundary. Minimum offsets of the fault system in the San Miguel de Allende area is 450 m (NietoSamaniego and Alaniz-Alvarez, 1994). North of San Miguel de Allende the fault system cuts Quaternary sediments, forming a 50-m-high scarp. The fault trace is interrupted within the Sierra del Cubo and appears again as the northern extension of the Villa de Reyes graben (Fig. 5). Because there are no western bounding faults to form a complete graben, we deduce that the fault system is a half graben, the master fault dipping to the west. We chose the Aguascalientes fault as the western boundary of the Mesa central because it is a large-displacement fault that marks the western limit of Mesozoic basement and separates regions of different structural style. The north-striking, east-dipping fault has a 250-m-high scarp into a basin filled with at least 500 m of sedimentary rocks and intercalated ignimbrites. Water wells indicate a total displacement of 1000 m (Jimnez-Nava, 1993). Subsidiary high-angle, normal faults parallel the main scarp. A west-dipping, antithetic fault that creates a full graben is found only near Aguascalientes, so the basin is mostly a half graben with an east-dipping master fault (Fig. 5). The southern edge of the Mesa central is the dominantly normal Bajo fault (Aranda-Gomez et al., 1989). The eastern half of the Bajo fault trends east and, according to subsurface data, shows an offset about 700 m west of Celaya (Ramos-Salinas, 1996). The western half strikes N50W and bounds Mesozoic basement and Paleogene plutons (Fig. 4). Using displacements reported by Quintero-Legorreta (1992) and Aranda-Gmez et al. (1989) and adding an

354

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO

average 500 m of basin-fill deposits estimated from Hernndez-Laloth (1991), the Bajo fault reaches a vertical offset of 1350 m near Len and 1100 m near Guanajuato. However, if we consider the vertical offset of the parallel Veta Madre fault in the city of Guanajuato (Gross, 1975), the cumulative offset can be as much as 2850 m. The Bajo fault vanishes west of Len, where it intersects the north-trending Penjamillo graben (Fig. 5) (Ferrari et al., 1994b). Between the Penjamillo graben and Guadalajara, the southern boundary of the Mesa central and of the Sierra Madre Occidental is poorly defined because most of the area is covered by late Tertiary continental sediments and by extensive late Miocene basalts. These rocks are cut by two approximately eastwest normal fault systems with 200400 m of vertical offset (Fig. 5). South of the Los Altos de Jalisco plateau is the east-westtrending Chapala graben (Campos-Enriquez et al., 1990; Urrutia-Fucugauchi and Rosas-Elguera, 1994). None of these structures has the large vertical displacement shown by the Bajo fault, and this suggests that the Penjamillo-Guadalajara region is a transitional zone with a diffuse deformation. West of Guadalajara the Sierra Madre Occidental is separated from the Jalisco block by two N55W trending en echelon grabens (Ferrari and Rosas-Elguera, 1999) (Fig. 5). Bounding faults on both grabens display mainly dip-slip motions. However, older right-lateral transtensional motions are observed at some places (Ferrari et al., 1997b). In addition a middle Miocene left-lateral transpression was documented in the southernmost part of the Sierra Madre Occidental and some of this deformation was probably accommodated by strike-slip movement along major faults of the graben (Ferrari, 1995). Mesa central. Within the Mesa central there are three major fault systems. The N30E trending Villa de Reyes graben cuts the eastern part of the Mesa central. This is a symmetric graben but it is segmented by secondary orthogonal graben and faults (Fig. 5). In the city of San Luis Potos the sedimentary deposits that fill the basin are 480 m thick and contain interbedded Oligocene ignimbrite sheets. A vertical offset of ~680 m was calculated using subsurface stratigraphic sections based on water exploration wells and geologic mapping (Tristn-Gonzlez, 1986; Martnez-Ruiz and CuellarGonzlez, 1978). The San Luis de la PazSalinas de Hidalgo fault system is a well-defined group of parallel faults that strike N50W and dip southwest (Fig. 5). These faults have mainly normal movements (Labarthe-Hernandez and JimnezLpez, 1993; Silva-Romo, 1996) with limited offset, and there are no basins associated with these faults. The third major fault system forms the northwest-trending Dolores HidalgoOjuelos depression and cuts the Villa de Reyes graben near San Felipe (Figs. 4 and 5). The basin is poorly defined and only few bounding faults were mapped. Other minor faults in the southern Mesa central have dominantly normal displacement. Sierra Madre Occidental Physiographic Province. The structural style of this province is dramatically different from that in the Mesa central. Instead of two sets of orthogonal faults, several major parallel faults are found (Fig. 5). In the eastern part of the province three nearly symmetric grabens strike north-northeast to north-south. In the western part of the province, north-northwesttrending half grabens and major listric faulting are dominant (Fig. 5). The easternmost structure in the province is the Juchipila graben and its eastward splay, the Calvillo graben. The Juchipila graben is a 80-km-long and 15-km-wide depression with a N10E strike and a relief of as much as 1000 m. The Calvillo graben, is a narrow, N15E trending depression, with a maximum relief of 900 m (Fig. 5). Its total length is 40 km and its average width is 12 km. West of Juchipila, the Tlaltenango graben trends N10E and has a length of 120 km and a width of 15 km, forming a depression of 400 m. The largest structure is the north-south Bolaos graben, which is 140 km long and 1520 km wide. Here offsets of stratigraphic units commonly

range from 1500 to 2000 m (Lyons, 1988). The graben is slightly asymmetric; the maximum relief is 1200 m on the western scarp (Scheubel et al., 1988). The three grabens are bounded by high-angle faults with dominant dip-slip displacements. Inside the grabens, volcanic or sedimentary beds show tilts of 10 to 20 either to the east and the west. The region between the Bolaos graben and the coastal plane is affected by four north-south to north-northwesttrending major structures: the Puente de Camotln, the Sierra de Pajaritos-Atengo, and the Sierra AlicaJesus Maria half grabens and the PochotitnSan Pedro fault system. These structures systematically dip to the west and tilt the hanging blocks as much as 35 to the east and east-northeast. The major structure is the Sierra AlicaJesus Maria half graben, which has a maximum vertical offset exceeding 2.2 km. The region west of this structure is interpreted as a series of parallel listric faults, whose style, age, and orientation strongly indicate a genetic relation with the opening of the Gulf of California (Ferrari et al., 1997b; Ferrari and Rosas-Elguera, 1999). These structures are limited to the south by a left-lateral normal accommodation zone that separates extension to the north from transpressional folding at the boundary with the Jalisco block (Fig. 5) (Ferrari, 1995). Age of Faulting Southern Mesa central. The age of the main depressions in the Mesa central is constrained by the volcanic units interbedded in the sedimentary filling. The filling of the San Miguel de Allende basin contains a ca. 11 Ma basalt the counterpart of which was identified in the range. However, the basalt shows less than 20% of the total displacement of the fault (NietoSamaniego and Alaniz-Alvarez, 1994). Both the Aguascalientes and the El Bajo basins contain late Miocene basalts and undated ignimbrites. In the Villa de Reyes graben, the Panalillo rhyolite (Fig. 6), with age of 26.8 Ma, was identified within the basin fill. According to these data, we consider that major faulting in the southern Mesa central began in late Oligocene time. The oldest Cenozoic unit exposed in Sierra de Guanajuato, the Eocene Guanajuato conglomerate, presents dramatic changes in thickness, varying from 1500 m (Gross, 1975) to less than 100 m in adjacent areas. These variations, together with the presence of dikes, the intercalation of lava flows, and the tilting of sandstone horizons, have been interpreted as an evidence of the inception of extension (Nieto-Samaniego, 1990; Randall-Roberts et al., 1994). Andesitic dikes (Echegoyn-Snchez et al., 1970), rhyolitic domes aligned along fractures, and rhyolitic lava flows that cover normal faults record extension prior to 32 1 Ma (Gross, 1975; Nieto-Samaniego, 1990). The tectonic episode with the higher deformation rate occurred in middleOligocene. The Veta Madre fault offsets 30 Ma rhyolite domes about 1500 m (Nieto-Samaniego et al., 1996) and is filled by vein minerals dated as 29.2 2 Ma (Gross, 1975). Post-Miocene extensional phases with lower intensity are recorded by displacements of 200 to 600 m of an ignimbrite dated as 24.8 0.6 Ma (Nieto-Samaniego et al., 1996) and basaltic lavas dated as 13 to 11 Ma (Table DR1, see footnote 1). In the San Luis Potos region (Fig. 6) the first clear evidence of extension is the emplacement of the San Miguelito rhyolitic domes and dikes along faults, which implies an extensional phase ca. 30 Ma. The main phase of extension is documented by the displacement of ~500 m of the Cantera ignimbrite (29.0 1.5 Ma) (Tristn Gonzlez 1986) and the emplacement of the El Zapote rhyolite (27.0 0.7 Ma) and Panalillo rhyolite (26.8 1.3 Ma) through the faults cutting the Cantera ignimbrite. Minor offsets of the Panalillo rhyolite (Labarthe-Hernndez and Jimenez-Lpez 1994) record a minor phase of extension in post-Oligocene time. The outlined stratigraphic relationships indicate that extension could have begun in Eocene time, at least in Guanajuato and San Luis Potos, and that three subsequent phases of extension can be identified: (1) a mild ex-

Geological Society of America Bulletin, March 1999

355

NIETO-SAMANIEGO ET AL.

10 km

101

10050'

Continental sedimentary deposits (Oligocene-Holocene)


Late Miocene basalt
Ignimbrites and andesites

Cabras Basalt
Panalillo Rhyolite (upper member) Andesite Panalillo Rhyolite (lower member) El Zapote Rhyolite Cantera Ignimbrite San Miguelito Rhyolite Portezuelo Latite Ojo Caliente Traquite Santa Mara Ignimbrite Casita Blanca Andesite Cenicera Formation Caracol Formation

SAN LUIS POTOS


Oligocene volcanic group
A'

2200'

Eocene volcanic group

Paleogene continental sediments Mesozoic basement

SL

Rhyolite domes

A
m a.s.l.
2450 2150 1900

101

10050'

Vill a

de

Re

yes

2145'

2145'

gra

ben

SH

A'

1 km
Figure 6. Geologic map of Sierra de San Miguelito. Geologic mapping is modified after Tristn-Gonzlez (1986), and faults are simplified after Labarthe-Hernndez and Jimnez-Lpez (1992, 1993, 1994). Formal units defined by Labarthe-Hernndez et al. (1982) are in italics. Faults inside the boxed area are reproduced in the lower right corner to show the rhombohedral array at the intersection between Villa de Reyes graben and San Luis de la PazSalinas de Hidalgo fault system (SLSH). Section is after Labarthe-Hernndez and Jimnez-Lpez (1994); the marker is the lower member of the Cantera Ignimbrite.

tensional phase before 30 Ma; (2) a well constrained phase constituting the peak of extension between 30 and 27 Ma; and (3) minor phases after about 24 and after about 11 Ma. Sierra Madre Occidental Physiographic Province. The age of deformation in this province is less well-constrained than in the Mesa central but can be reasonably estimated in some places. In the Bolaos graben an important

mineralization event took place prior to 21 Ma and fills N30E trending faults that cut a 23.7 Ma ignimbritic succession with offsets of 100200 m (Lyons, 1988). The peak of deformation is associated with the north-south bounding faults of the graben, which displace a 21.3 Ma ignimbrite by more than 1 km and expose the mineralization (Lyons, 1988; Scheubel et al., 1988). Several mafic dikes intrude the N30E faults, in some cases feeding basaltic flows that

356

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO

Figure 7. Schmidt projection, lower hemisphere of faults measured in the southern Sierra Madre Occidental volcanic province. Contouring was performed with the Kamb method.

have been dated as 19.9 Ma at Veta Rica (Table DR1, see footnote 1). Field evidence suggests that the dikes are synextensional and thus their age probably represents the peak of extension. The minimum age of faulting is not constrained by any geologic unit. An early Miocene age is deduced for the Tlaltenango graben because the bounding faults cut an ignimbritic succession of about 23 Ma but do not affect a 21.8 Ma shield volcano inside the depression (Moore et al., 1994). The uppermost ignimbrite cut by the Sierra AlicaJesus Maria half graben, dated as 23.5 Ma (Clark et al., 1981), provides a maximum age for this structure, but no ages are available for the ending of faulting. In the westernmost part of the province, the PochotitnSan Pedro fault system cuts ignimbrites as young as 19 Ma (Table DR1, see footnote 1), and tilted hanging-wall blocks are covered by late Miocene basalts north of Tepic. Several mafic dikes with ages between 11.9 and 10.9 Ma (Table DR1, see footnote 1) are intruded parallel with the fault system and are inferred to mark the major phase of extension (Ferrari and Rosas-Elguera, 1999). In summary, these stratigraphic relations indicate two main phases of deformation: the first between ca. 21 and 18 Ma in the eastern part of the province and the second during middle to late Miocene in the western part of the province, adjacent to the Gulf of California. Amount of Extension Southern Mesa central. The Mesa central represents a good example of triaxial deformation style. Strong evidence of triaxial strain in the southern Mesa central includes the following: (1) the rhombohedral arrangement of fault traces both at large scale (Fig. 6), and in the intersection of the Villa de Reyes graben with the San Luis de la PazSalinas de Hidalgo fault system (Fig. 6); (2) both fault systems show dip-slip displacement; (3) there is no systematic crosscutting relation between the faults, indicating that the activity was synchronous (Nieto-Samaniego, 1990; Labarthe-Hernndez and Jimnez-Lpez, 1992, 1994). The orientations of the principal strains were obtained, using the odd-axis

model proposed by Krantz (1988), in the Sierra de San Miguelito area, located southwest of San Luis Potos (Fig. 6). Using these results we calculated the direction and the amount of extension for the entire Mesa central (see Appendix 1 for a complete description). Principal extension was oriented 258/12 with a magnitude of 20%. An extension of 11% also occurred oriented 162/02, showing the three-dimensional character of the deformation. Sierra Madre Occidental Physiographic Province. In this province it is clear that deformation was two dimensional and that it affected an almost horizontal and undeformed ignimbritic plateau. Thus we considered it adequate to use fault-slip inversion methods to calculate the paleostress tensor. In the eastern part of the province we measured a total of 137 faults, 108 of which have good kinematic indicators. Fault trends show a peak near N10E (Fig. 7), which is also the trend of the faults at larger scale (Fig. 5). There is a minor concentration with an east-west direction. We interpret these faults as secondary structures that accommodate local strain incompatibilities because they are nearly vertical and show oblique- or strike-slip displacements. Furthermore, there are no major faults belonging to this group, and the grabens are not interrupted by major transverse depressions. The minimum principal stress (3) obtained with the inversion method proposed by Reches (1987) is oriented 099/08, nearly horizontal and orthogonal to the trend of the grabens (Table 2). In the western part of the province we measured a total of 54 faults with good kinematic indicators. Using the same method we obtain a N66E trending 3 (Table 2), which is orthogonal to the dominant fault trend (Fig. 5). To estimate the amount of extension we use the area-balance method described by Groshong (1994) (Fig. 8). Although there is no good stratigraphic control, the method may work because there is a well-defined regional reference level not disturbed by major faulting within the horsts (Figs. 5 and 8). Some minor fault zones between the major grabens are negligible because they do not tilt the sequence and produced displacements

Geological Society of America Bulletin, March 1999

357

NIETO-SAMANIEGO ET AL.

Figure 8. Estimation of amount of extension along a profile perpendicular to the graben of the eastern Sierra Madre Occidental. The area vertically ruled is equal to the gray one. The figure shows the most consistent solution according to the geology of the region (see text for details). Vertical exaggeration is 10x. Dashed lines represent inferred faults at depth. Lo = initial length; Lf = final length.

less than 100 m. We carried out many experiments varying the depth of detachment and obtained 5 km as a realistic approximation. This corresponds to a horizontal extension of 8% (Fig. 8). Extension in the western part of the province is obviously greater than to the east because the crust has been thinned into the oceanic rift of the Gulf of California. We have not estimated the amount of extension in this region because a reference level is lacking. DISCUSSION AND CONCLUSIONS Tectonic-Magmatic Evolution of the Southern Sierra Madre Occidental volcanic province On the basis of stratigraphy, structure, topography, and crustal thickness, we propose that the southern volcanic province contains three major crustal blocks that behaved as roughly independent units. These blocks correspond to the physiographic provinces of the Sierra Madre Oriental, Mesa central and Sierra Madre Occidental. The Sierra Madre Oriental is a relatively stable block, without volcanism and extension during the Cenozoic. The Mesa central is a thinned and elevated block bounded by the thicker and relatively lower Sierra Madre Oriental and Occidental blocks (Fig. 3). Within these blocks volcanism and deformation show marked differences in style and timing. Volcanism in the Mesa central occurred between about 30 and 25 Ma, and is mainly represented by rhyolitic lava domes capped by a thin ignimbritic cover. Deformation was three dimensional, with horizontal, approximately east-west extension of as much as ~20%. Extension was in part concurrent with volcanism, but the peak of the deformation postdates the principal volcanic event. Volcanism and deformation occurred later toward the west. By early Miocene time a thick ignimbritic cover was emplaced in the eastern Sierra Madre Occidental, with rare rhyolitic lava flows and domes and subordinate mafic lava. Deformation was two dimensional, and the principal stretching of ~8% was oriented approximately east-west. Extension was partly synchronous with volcanism, but its climax postdates the ignimbrite flare up. During middle to late Miocene time, volcanism and extension continued to shift toward the west-southwest. Arc volcanism of this

age is found in the western part of the Sierra Madre Occidental, and extension appears to slightly postdate it. Extension direction changed orientation counterclockwise by ~30, from NW81SE to NE66SW, culminating with the opening of the Gulf of California, which now bounds the western edge of Sierra Madre Occidental. The trenchward pattern of tectonicmagmatic migration is clearly shown in Figure 9D, where dated volcanic rocks and extensional events are plotted against their distance from the paleotrench. A similar pattern of migration has been observed in the central and northern Sierra Madre Occidental volcanic province (e.g., Aranda-Gomez et al., 1997; Gans, 1997). Comparison with the Central and Northern Sierra Madre Occidental volcanic province There are some peculiarities in the geometry and kinematics of extension in the southern Sierra Madre Occidental volcanic province. The first difference concerns the size of the extended area. It has been long suggested that the inner core of the Sierra Madre Occidental is a relatively unfaulted horst, separating the highly extended areas of the southern Basin and Range and the Gulf of California (Henry, 1989; Henry and Aranda-Gomez, 1992). Although this appears to be true in Durango and southern Chihuahua, our data indicate that this is not the case for the southern Sierra Madre Occidental; here the crust appears pervasively faulted, and distinguishing Basin and Range from Gulf of California faulting is difficult both in space and in time (Fig. 1). Furthermore, the area affected by extension and volcanism in the southern Sierra Madre Occidental volcanic province is nearly twice as wide as in the central volcanic province, directly north of the study region. Volcanism and significant OligoceneMiocene extension in the former extend as far as eastern San Luis Potos and Quertaro, about 500 km from the present coast, whereas in the central Sierra Madre Occidental volcanic province they are only observed in Durango and Sinaloa. A second difference regards the direction of extension. In the northern Sierra Madre Occidental, Gans (1997) reported a minimum extension of 90% in the northeast-southwest direction from 26 to 20 Ma followed by approximately east-west extension of 10%15% between 20 and 17 Ma. In the central Sierra Madre Occidental (Durango and Sinaloa) dominant ex-

358

Geological Society of America Bulletin, March 1999

24 18 16 14
Escuinapa
Zacatecas

22 14 16

20

Possible accommodation zone with vergence reversal (Axen, 1995)

Possible accommodation zone with vergence reversal (this work)

Northern limit of Oligo-Miocene extensional faulting and volcanism in the MC

34 mm/yr

F Shirley
17-9
43 mm/yr 16 18 22 47 mm/yr
0

racture
23-19
15
Tepic

Zone

San Luis Potos

Aguascalientes

29-27

28

26 20

24

Len

Jalisco block
100

Guadalajara

S.M. de Allende

200 km

35 30 25 20 15 10 5

Half-spreading rate (mm/yr)

Geological Society of America Bulletin, March 1999


North of Shirley

Age (Ma)

South of Shirley

Dated volcanic rock with error bar (Table 1 & 2) Main extensional episode (this work)
300 400 500 600 700

1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 20

30

25

Age (Ma)

Distance from the paleotrench (km)

Figure 9. Comparison between continental and offshore tectonics. The map in (A) shows the main continental structure with time of faulting and average direction of extension (from this work) and the corresponding sea-floor isochron pattern deduced from the magnetic anomaly map of Severinghaus and Atwater (1989) and Lonsdale (1991). Baja California was restored to its prerift position considering that no embayment existed at the mouth of the Gulf of California. Sea-floor age (circled) and extinct ridges (thick double lines) were re-

stored to their pre 12.5 Ma position for the amount proposed in Lonsdale (1991). Lines perpendicular to sea-floor isochrons show examples of computed half rate of sea-floor spreading. Note the systematic higher average half-spreading rate south of the Shirley fracture zone (at 1 m.y. intervals) during Oligocene and early Miocene time (B) and the close match between the timing of high spreading pulses and of the main extensional and volcanic episodes on the mainland (B and C).

359

NIETO-SAMANIEGO ET AL.

tension in late Oligocene to middle Miocene and in PlioceneQuaternary time was east-northeast trending, with a minor episode of west-northwest extension in between (Aranda-Gomez et al., 1997). By contrast, two thirds of the southern volcanic province (Mesa central and eastern Sierra Madre Occidental) were affected by approximately east-west extension in Oligocene and early Miocene time, and only in middle to late Miocene time did east-northeast extension affect the western Sierra Madre Occidental. A final difference concerns the tilt vergence in the western Sierra Madre Occidental. South of lat 23 N blocks are ubiquitously tilted toward the eastnortheast, whereas to the north, in Sinaloa, they are systematically tilted in the opposite direction (Henry, 1989; Aranda-Gomez et al., 1997). This vergence reversal implies that a reverse accommodation zone must exist somewhere in the Escuinapa area (Fig. 9). Axen (1995) suggested a possible matching structure on the other side of the Gulf of California, separating hanging wall and footwall segments along the Main Gulf escarpment in southern Baja California (Fig. 9A). Causes of Extension in the Southern Sierra Madre Occidental volcanic province In general, extension in continental volcanic arcs is caused by the competition among lithospheric body forces, such as those related to the gravitational and thermal effects of magmatism (e.g., Gans et al., 1989), and plate boundary forces, related to the relative motion vectors between subducting and overriding plates or to direct interaction between diverging oceanic and continental plates (e.g., Bohannon and Parsons, 1995). Magmatism can induce extension in two ways: 1) injection of mafic magmas at the base of the crust may induce a buckling of the lithosphere because of the accumulation of low-density hot material; 2) rapid piling up of volcanic rocks at the surface can build a topographically elevated region, where the vertical lithostatic stress may exceed the regional horizontal stress. At a regional scale, deformation induced by high topography is likely to be bidimensional, with extension orthogonal to the axis of the volcanic arc, whereas deformation related to magmatic doming may be three dimensional and close to radial extension. In both cases, however, extension is likely to affect only the upper part of the crust. However, extension in the overriding plate can be induced by a retreating subduction boundary, such as that where the subduction rate exceeds the convergence rate (Jarrad, 1986; Royden, 1993), and this deformation is usually accommodated along the volcanic arc (Hamilton, 1995). Extension driven by a retreating subduction boundary is expected to involve the entire lithosphere, and deformation is expected to be bidimensional if the lithosphere is homogeneous. The extension direction, however, will depend upon the obliquity of the convergence and the occurrence or not of strain partitioning. In addition, a retreating slab will generate a flux of hotter asthenospheric material into the opening mantle wedge, producing widespread melting and magma injection at the base of the crust. Thus, episodes of retreating should match pulses of magmatism and control the location of volcanism in the overriding plate. A detailed discussion of the relative role of body forces and plate boundary forces in determining the tectonic-magmatic evolution of the southern Sierra Madre Occidental volcanic province is beyond the scope of this paper. However, the data presented herein permits at least a qualitative analysis of the issue. The late Miocene episode of extension (proto-Gulf of California) was clearly related to the capture of the Magdalena microplate and Baja California by the Pacific plate (e.g., Stock and Hodges, 1989). Thus, plate tectonics was the dominant factor governing the latest period of extension. However, the first two episodes of extension occurred when Farallon plate remnants were still subducting, magmatism was still active, and both kinds of forces were active. To evaluate the role of plate boundary forces, we reconstructed the pre-

Gulf of California plates configuration by closing the Gulf of California and restoring the right-lateral motion along the Tosco-Abreojos fault system (Fig. 9).This restoration places the Shirley Fracture Zone at the same latitude as the northern limit of the southern Sierra Madre Occidental volcanic province, and the vergence reversal in the Gulf of California area. In Figure 9B we calculated the western half rate of spreading at the East Pacific Rise between 30 and 20 Ma along two flow lines to the north and to the south of the Shirley Fracture Zone. Considering a nearly symmetric spreading at the rise (Lonsdale, 1991), one can take these values as representative of the eastern (Farallon) half-spreading rate. These rates show two rapid increases, at 3026 Ma and 2420 Ma that match the two main episodes of volcanism and extension in the Sierra Madre Occidental volcanic province (Fig. 9B and C). Given that the trench has actually moved oceanward since Mesozoic time (Lightgow-Bertelloni and Richards, 1998), we propose that increases of the spreading rate at the Eastern Pacific Rise were a consequence of increases in the subduction rate at the trench and, therefore, that the FarallonNorth America boundary west of the southern volcanic province was a retreating subduction boundary. Spreading south of the Shirley Fracture Zone was always higher than to the north by as much as 20% during the peaks of spreading rate. This explains why extension in the southern volcanic province occurred much farther inland than in the central volcanic province. Along this part of the FarallonNorth America plate boundary, the difference between subduction and convergence rate probably exceeded a critical threshold needed to induce extension in the overriding plate. These observations led us to conclude that, at a continental scale (over 100 km of length), timing, magnitude, and orientation of extension as well as age of volcanic episodes in the Sierra Madre Occidental volcanic province were ultimately controlled by plate tectonic forces, namely by the latitudinal variation of the retreating subduction boundary between the Farallon and North America plates. On a regional to local scale (less than 100 km of length), we believe that the variations in the tectonic and volcanic style in the southern Sierra Madre Occidental volcanic province were controlled by the difference in the structure of the Mesa central and Sierra Madre Occidental crustal blocks and by lithospheric body forces. In the Mesa central, the existence of a prefractured crust and a moderate magmatic rate induced a three-dimensional strain, a complicated fault geometry, and the emplacement of silicic magmas as rhyolitic domes. We believe that, in this case, plate boundary forces and body forces are equally responsible in controlling the dynamics of extension and magmatism. In the Sierra Madre Occidental, a stronger and relatively homogeneous crust and a high magmatic rate caused a bidimensional strain, regularly spaced faults, and the formation of calderas. In this case magmatism was probably the dominant cause of extension. We consider that the silicic volcanism flare-up was triggered by the combined effect of extensive mafic underplating and the beginning of upper plate extension induced by the slab retreat. CONCLUSIONS Our main conclusions can be summarized as follows: 1. In the southern Sierra Madre Occidental volcanic province we observed a clear pattern of trenchward shifting of the climax of subduction volcanism and extension during three distinct episodes in middle Oligocene, early Miocene, and late Miocene time. In all the episodes, the peak of extension slightly postdates volcanism. 2. Tectonic and volcanic style, as well as fault pattern, vary strongly from east to west. 3. Extension and volcanism in the southern volcanic province took place much farther inland than in the central volcanic province. 4. There is a very good correlation between the timing of extensional and

360

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO A

Unrotated
SSM faults 227/48
odd=e
3

k = -0.52

VRG faults 120/65

odd, e
intermediate, e
2

similar=e

=071/77

1=258/12

a =36

e ria st 7 M 4 / SS 16 2

VR 09 G s 7 tri /63 ae

similar, e

a
B

90

intermediate=e
2

=168/02

Corrected for rotation


SSM faults 228/70

k = -0.49
similar=e
1=259/11

VRG faults 120/65

odd=e
3

=064/78

a =35

VR e 09 G s 7 tri ria /63 ae st M /69 S S 13 2

90

intermediate=e
2

=169/02

Figure A1. Application of the odd-axis model (Krantz, 1988) to the Sierra de San Miguelito area. The model is based on the Rechess slip-model and considers four fault planes with orthorhombic symmetry. Simultaneous motion of the faults accommodates a triaxial nonrotational deformation. The odd axis is the one with opposite sign (shortening of stretching) with respect to the other principal axes, assuming no volume change. When the odd-axis is 1 then the similar-axis is 3 and vice-versa. We assumed the stretching to be positive and 1 > 2 > 3. 2 is the angle between the fault traces on the intermediate-similar plane. The Krantz model holds for nonrotational faulting. However, we consider that it can be applicable to the San Miguelito area. In any case, correcting the data for the rotation produces nearly identical results in the similar axis orientation (cf. A and B). Data in B were rotated 22 counterclockwise about an axis 311/00.
0.4 and a coefficient of friction = 0.8. In order to estimate the post-Oligocene strain we need to assume the principal strain orientations obtained for Sierra de San Miguelito as valid for all the Mesa central, based on the similarity in strike of the principal fault systems (cf. Figs. 5 and 7). The crust of the southern Mesa central is thinner and more elevated relative to the adjacent Sierra Madre Oriental and Occidental crustal blocks. A reasonable assumption is to consider that Mesa central had a thickness similar to that of the adjacent crustal blocks prior to the deformation. The symmetry of crustal blocks shown in Figure 3 suggests that deformation within the Mesa central was produced by the relative movement of the Sierra Madre Oriental and Occidental crustal blocks, as supported by the orientation of the bounding structures (Aguascalientes graben and San Miguel de AllendeCatorce fault system), approximately north-south (Fig. 5), and by the approximately east-west orientation of the maximum principal strain (1) obtained in the Sierra de San Miguelito. The crust in the Sierra Madre Occidental is 40 km thick, and we estimated 8% stretching during Cenozoic time. Considering the crustal thickness before deformation of the Mesa central to be equal to the original thickness of the Sierra Madre Occidental, l0 = 43 km, and a final thickness lf = 32 km, the vertical principal elongation (3) is 3 = (lfl0)/l0 = 0.25, where the negative sign indicates that shortening occurred in the vertical direction. To calculate the horizontal principal elongations we need to assume no volume change. Considering an undeformed cube with unit edge and applying to it a strain tensor with principal strains without volume change, we obtain

magmatic pulses and periods of rapid spreading at the East Pacific Rise in the offshore region. 5. The good match among continental and oceanic tectonic events suggests that boundary conditions (i.e., plate tectonics) ultimately determine the timing, magnitude, and orientation of extension. 6. The style of tectonics and volcanism and the distribution of faulting are controlled by the internal structure of the crustal blocks and by the gravitational and thermal effects of magmatism.
APPENDIX I

ON THE ESTIMATION OF THE AMOUNT OF EXTENSION IN THE MESA CENTRAL The method of Krantz (1988) was used to determine the principal directions and the strain ratio = 2/3 = tan2 , where i are the principal elongations, and 1 > 2 > 3 (Fig. A1). We obtained a value of = 0.52. (the negative sign indicates that 3 is a shortening axis), an odd-axis (shortening axis) oriented 071/77, and a similar axis (maximum stretching) oriented 248/12 (Fig. A1). Thus, considering the strain ellipsoid, the principal plane that contains the maximum and intermediate strain axes is nearly horizontal. We also used the graphics proposed by Reches (1983), obtaining

(1 + )(1 + )(1 + ) = 1 .
1 2 3

(1)

Geological Society of America Bulletin, March 1999

361

NIETO-SAMANIEGO ET AL. Using the mean value for strain ratio = 0.45 in equation 1 and introducing the calculated value 3 = 0.25, we obtained 1 = 0.2, 2 = 0.11, and 3 = 0.25. These values represent the bulk deformation of southern Mesa central, without considering the material transfer due to the magmatism. Some comments are needed regarding the use of the odd-axis model for Sierra de San Miguelito. The San Luis de la PazSalinas de Hidalgo fault system shows a domino style with beds dipping 20 to the northeast (Labarthe-Hernndez and Jimnez-Lpez, 1992, 1994). This bed inclination implies a rotational component in the deformation tensor, whereas the odd-axis model supposes an irrotational strain. Nevertheless, we use the odd-axis model for the following reasons: 1. We try to estimate the strain for all the southern Mesa central, whereas tilting is local. The rotation of 20 observed in the beds of Sierra de San Miguelito is not widespread. The bulk of Sierra de San Miguelito is only tilted ~5. Adjacent to the Sierra de San Miguelito there are other faults and grabens parallel to the San Luis de la PazSalinas de Hidalgo fault system (Bledos and Enramadas grabens, Fig. 7), that do not produce significant tilting. At a regional scale tilting produced by the major grabens is <10. 2. Rotation introduced error in the odd-axis model only in the orientation of the slip vector of the San Luis de la PazSalinas de Hidalgo fault system. It is possible to correct this error (Fig. A1), but the variation produced by rotation does not affect the result significantly because it is less than the variation due to the quality of the geologic data. We conclude that the orientations of the principal strains estimated using the oddaxis model are sufficiently good, in relation to the quality of our geologic data.
Carrillo-Bravo, J., 1971, La Plataforma VallesSan Luis Potos: Boletn de la Asociacin Mexicana Gelogos Petroleros, v. 23, 102 p. Centeno-Garcia, E., Ruiz, J., Coney, P. J., Patchett, P. J., and Ortega-Gutirrez, F., 1993, Guerrero terrane of Mexico: Its role in the Southern Cordillera from new geochemical data: Geology, v. 21, p. 419422. Chiodi, M., Monod, O., Busnardo, R., Gaspar, D., Sanchez, A., and Yta, M., 1988, Une discordance ante Albienne date par une faune dmmonites et de braquiopodes de type Tthysien au Mexique Central: Geobios, v. 21, p. 125135. Clark, K. F., Damon, P. E., Shafiquillah, M., Ponce, B. F., and Cardenas, D., 1981, Seccin geolgica-estructural a travs de la parte sur de la Sierra Madre Occidental, entre Fresnillo y la costa de Nayarit: Asociacin Ingenieros Mineros, Metalurgicos y Gelogos de Mxico, Memoria Tcnica, v. XIV, p. 6999. Couch, R. W., Ness, G. E., Snchez-Zamora, O., Calderon-Riveroll, O., Doguin, P., Plawman, T., Coperude, S., Huehn, B., and Gumma, P., 1991, Gravity anomalies and crustal structure of the Gulf and Peninsular Province of the Californias, in Dauphin, J. P., and Simoneit, B. R. T., eds., The Gulf and the Peninsular Province of the Californias: American Association of Petroleum Geologists Memoir 47, p. 4770. Damon, P. E., Nieto-Obregon, J., and Delgado-Argote, L., 1979, Un plegamiento neognico en Nayarit y Jalisco y evolucin geomrfica del Rio Grande de Santiago: Asociacin Ingenieros Mineros, Metalurgicos y Gelogos de Mxico, Memoria Tcnica, v. XIII, p. 156191. Demant, A., 1978, Caractersticas del Eje Neovolcnico Transmexicano y sus problemas de interpretacin: Universidad Nacional Autnoma de Mxico, Instituto de Geologa, Revista, v. 2, p. 172187. Echegoyn-Snchez, J., Romero-Martnez S., and Velzquez-Silva, S., 1970, Geologa y yacimientos minerales de la parte central del Distrito Minero de Guanajuato: Consejo de Recursos Naturales No Renovables (Mxico), Boletin, v. 75, 36 p. Edwards, J. D., 1955, Studies of some early Tertiary red conglomerates of central Mexico: U.S. Geological Survey Professional Paper 264-H, 183 p. Fensby, S., and Gastil, G., 1991, Geologic-tectonic map of the Gulf of California and surrounding areas, in: Dauphin, J. P., and Simoneit, B. R. T., eds., The Gulf and the Peninsular Province of the Californias: American Association of Petroleum Geologists Memoir 47, p. 7983. Ferrari, L., 1995, Miocene shearing along the northern boundary of the Jalisco block and the opening of the southern Gulf of California: Geology, v. 23, p. 751754. Ferrari, L., and Rosas-Elguera, J., 1999, Late Miocene to Quaternary extension at the northern boundary of the Jalisco block, western Mexico: The Tepic-Zacoalco rift revised: Geological Society of America Special Paper 334 (in press). Ferrari, L., Garduo, V. H., Innocenti, F., Manetti, P., Pasquar, G., and Vaggelli, G., 1994a, A widespread mafic volcanic unit at the base of the Mexican Volcanic Belt between Guadalajara and Queretaro: Geofsica Internacional, v. 33, p. 107124. Ferrari, L., Garduo, V. H., Pasquar, G., and Tibaldi, A., 1994b, Volcanic and tectonic evolution of central Mexico: Oligocene to present: Geofsica Internacional, v. 33, p. 91105. Ferrari, L., Vaggelli, G., and Manetti, P., 1996, Late Miocene mafic volcanism in central Mexico: The continental record of the opening of the Gulf of California [abs.]: Eos (Transactions, American Geophysical Union), v. 77, p. F757. Ferrari, L., Urrutia-Fucugauchi, J., Aguirre-Diaz, G., and Rosas-Elguera, J., 1997a, Stratigraphy and structure of the southern Sierra Madre Occidental along the Zacatecas-Nayarit transect: Unin Geofsica Mexicana, Boletn Informativo, GEOS, v. 17, p. 286. Ferrari, L., Nelson, S. A., Rosas-Elguera, J., and Aguirre-Daz, G. J., 1997b, Tectonics and volcanism of the western Mexican Volcanic Belt, in International Association of Volcanology and Chemistry of the Earth Interior, General Assembly 1997, Field Trip 12 Guidebook: Puerto Vallarta, Mxico, 61 p. Ferrari, L., Pasquar, G., Venegas-Salgado, S., and Romero-Rios, F., 1999, Geology of the western Mexican Volcanic Belt and adjacent Sierra Madre Occidental and Jalisco block: Geological Society of America Special Paper 334, in press. Fix, J. E., 1975, The crust and upper mantle of central Mexico: Royal Astronomical Society Geophysical Journal , v. 43, p. 453499. Gans, P. B., 1997, Large-magnitude Oligo-Miocene extension in southern Sonora: Implications for the tectonic evolution of northwest Mexico: Tectonics, v. 16, p. 388408. Gans, P. B., Mahood, G. A., and Schermer E., 1989, Synextensional magmatism in the Basin and Range province; a case study from the eastern Great Basin: Geological Society of America Special Paper 233, 53 p. Gastil, R. G., Krummenacher, D., and Jensky, W. E., 1978, Reconnaissance geologic map of the west-central part of the state of Nayarit, Mxico: Geological Society of America Map and Chart Series Map MC-24, 1 sheet, scale 1:200 000. Gastil, R. G., Krummenacher, D., and Minch, J., 1979, The record of Cenozoic volcanism around the Gulf of California: Geological Society of America Bulletin, v. 90, p. 839857. Groshong, R. H., 1994, Area balance, depth of detachment, and strain in extension: Tectonics, v. 13, p. 14881497. Gross, W. H., 1975, New ore discovery and source of silver-gold veins, Guanajuato, Mexico: Economic Geology, v. 70, p. 11751189. Hamilton, W. B., 1995, Subduction systems and magmatism, in Smeille, J. L., ed., Volcanism associated with extension at consuming plate margins: Geological Society [London] Special Publication 81, p. 328. Hausback, B. P., 1984, Cenozoic volcanic and tectonic evolution of Baja California Sur, Mxico, in Frizzell, V. A., ed., Geology of Baja California Peninsula: Pacific Section, Society of Economic Paleontologists and Mineralogists Publication 199, p. 219236. Henry, C. D., 1989, Late Cenozoic Basin and Range structure in western Mexico adjacent to the Gulf of California: Geological Society of America Bulletin, v. 101, p. 11471156. Henry, C. D., and Aranda-Gmez, J. J., 1992, The real southern Basin and Range: Mid- to late Cenozoic extension in Mexico: Geology, v. 20, p. 701704.

ACKNOWLEDGMENTS This work was supported by grants CONACYT P152T (to Ferrari) and CONACYT 3159PT (to Nieto-Samaniego). P. Gans kindly made available his paper in advance of publication. Cia. Minera Las Torres is acknowledged for permission to publish K/Ar ages of the Guanajuato area. D. Moran-Zenteno provided useful suggestions to an earlier version of the manuscript. Bulletin editor S. Reynolds and reviewers C. Henry and F. McDowell contributed to significantly improve the manuscript.
REFERENCES CITED Aguirre-Daz, G. J., and McDowell, F. W., 1991, The volcanic section at Nazas, Durango, Mxico, and the possibility of widespread Eocene volcanism within the Sierra Madre Occidental: Journal of Geophysical Research, v. 96, p. 1337313388. Aguirre-Daz, G. J., and McDowell, F. W., 1993, Nature and timing of faulting and synextensional magmatism in the southern Basin and Range, central-eastern Durango, Mexico: Geological Society of America Bulletin, v. 105, p. 14351444. Aranda-Gmez, J. J., and Luhr, J. F., 1996, Origin of the Joya Honda maar, San Luis Potos, Mxico: Journal of Volcanology and Geothermal Research, v. 74, p. 118. Aranda-Gmez, J. J., and McDowell, F. W. , 1998, Paleogene extension in the southern basin and range province of Mexico: Syndepositional tilting of Eocene red beds and Oligocene volcanic rocks in the Guanajuato mining district: International Geology Review, v. 40, p. 116134. Aranda-Gmez, J. J., Aranda-Gmez, J. M., and Nieto-Samaniego, A. F., 1989, Consideraciones acerca de la evolucin tectnica durante el Cenozoico de la Sierra de Guanajuato y la parte meridional de la Mesa central: Universidad Nacional Autnoma de Mxico, Instituto de Geologa, Revista, v. 8, p. 3346. Aranda-Gmez, J. J., Henry, C. D., Luhr, J. F., and McDowell, F. W., 1997, Cenozoic volcanism and tectonics in NW Mexico, in International Association of Volcanology and Chemistry of the Earth Interior, General Assembly 1997, Field Trip #11 Guidebook: Puerto Vallarta, Mxico, Universidad Nacional Autnoma de Mxico, Instituto de Geologa, 94 p. Atwater, T., 1989, Plate tectonic history of northeast Pacific and western North America, in Winterer, E. L., Hussong, D. M., and Decker, R. W., eds., The eastern Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America, Geology of North America, v. N, p. 2171. Axen, G., 1995, Extensional segmentation of the Main Gulf Escarpment, Mexico and United States: Geology, v. 23, p. 515518. Bohannon, R. G., and Parsons, T., 1995, Tectonic implications of post30 Ma Pacific and North American relative plate motions: Geological Society of America Bulletin, v. 107, p. 937959. Campos-Enriquez, J. O., Arroyo-Esquivel, M. A., and Urrutia-Fucugauchi, J., 1990, Basement, Curie isotherm and shallow crustal structure of the structure of the Trans-Mexican volcanic belt from aeromagnetic data: Tectonophysics, v. 172, p. 7790. Campos-Enriquez, J. O., Kerdan, T., Morn-Zenteno, D. J., Urrutia-Fucugauchi, J., SnchezCastellanos, E., and Alday-Cruz, R., 1994, Estructura de la litsfera superior a lo largo del Trpico de Cncer[abs]: Unin Geofsica Mexicana, Boletin Informativo, GEOS, v. 12, p. 7576. Carranza-Castaeda, O., Petersen, M. S., and Miller, W. E., 1994, Preliminary investigation of the geology of northern San Miguel de Allende area, northeastern Guanajuato, Mexico: Brigham Young University Geological Studies, v. 40, p. 19.

362

Geological Society of America Bulletin, March 1999

SOUTHERN SIERRA MADRE OCCIDENTAL VOLCANIC PROVINCE, MEXICO


Henry, C. D., and Fredrikson, G., 1987, Geology of part of southern Sinaloa, Mexico, adjacent to the Gulf of California: Geological Society of America Maps and Chart series, MCH 063, 1 sheet, 14 p., scale 1:250 000. Hernndez-Laloth, N., 1991, Modelo conceptual de funcionamiento hidrodinmico del sistema acufero del valle de Len, Guanajuato [Bachelors thesis]: Mxico, D. F., Universidad Nacional Autnoma de Mxico, Facultad de Ingeniera, 129 p. Jarrad, R. D., 1986, Causes of compression and extension behind trenches: Tectonophysics, v. 132, p. 89102. Jimnez-Nava, F. J., 1993, Aportes a la estratigrafa de Aguascalientes mediante la exploracin geohidrolgica a profundidad, in Simposio sobre la geologa del Centro de Mxico, Extensin Minera 93, Universidad de Guanajuato, Facultad de Minas, Metalurgia y Geologa, Abstracts and field-trip guidebook, p. 1. Krantz, R. W., 1988, Multiple fault sets and three-dimensional strain: Theory and application: Journal of Structural Geology, v. 10, p. 225237. Labarthe-Hernndez, G., and Jimnez-Lpez, L. S., 1992, Caractersticas fsicas y estructura de lavas e ignimbritas riolticas en la Sierra de San Miguelito, S. L. P.: Universidad Autnoma de San Luis Potos, Instituto Geologa, Folleto tcnico (Open-File Report) 114, 31 p. Labarthe-Hernndez, G., and Jimnez-Lpez, L. S., 1993, Geologa del domo Cerro Grande, Sierra de San Miguelito, S. L. P.: Universidad Autnoma de San Luis Potos, Instituto Geologa, Folleto tcnico (Open-File Report) 117, 22 p. Labarthe-Hernndez, G., and Jimnez-Lpez, L. S., 1994, Geologa de la porcin sureste de la Sierra de San Miguelito, S. L. P.: Universidad Autnoma de San Luis Potos, Instituto Geologa, Folleto tcnico (Open-File Report) 120, 34 p. Labarthe-Hernndez, G., Tristn-Gonzlez, M., and Aranda-Gmez, J. J., 1982, Revisin estratigrfica del Cenozoico de la parte central del Estado de San Luis Potos: Universidad Autnoma de San Luis Potos, Instituto Geologa, Folleto tcnico (Open-File Report) 85, 208 p. Lang, B., Steinitz, G., Sawkins, F. J., and Simmons, S. F., 1988, K-Ar Age studies in the Fresnillo Silver District, Zacatecas, Mexico: Economic Geology, v. 83, p. 16421646. Lightgow-Bertelloni, C., and Richards, M., 1998, The dynamics of global plate motion: Reviews of Geophysics, v. 36, p. 2778. Lonsdale P., 1989, Geology and tectonic history of the Gulf of California, in Winterer, E. L., Hussong, D. M., and Decker, R. W., eds., The eastern Pacific Ocean and Hawaii: Boulder, Colorado, Geological Society of America, Geology of North America, v. N, p. 499521. Lonsdale, P., 1991, Structural pattern of the pacific floor offshore peninsular California, in Dauphin, J. P., and Simoneit, B. R. T., eds., The Gulf and the Peninsular Province of the Californias: American Association of Petroleum Geologists Memoir 47, p. 87125. Luhr, J. F., Pier, J. G., Aranda-Gomez, J. J., and Podosek, F. A., 1995, Crustal contamination in early Basin and Range hawaiites of the Los Encinos volcanic field, central Mexico: Contributions to Mineralogy and Petrology, v. 118, p. 321339. Lyons, J. I., 1988, Geology and ores deposits of the Bolaos silver district, Jalisco, Mexico: Economic Geology, v. 83, p. 15601582. Martnez-Reyes, J., 1992, Mapa geolgico de la Sierra de Guanajuato: Universidad Nacional Autnoma de Mxico, Instituto de Geologa, Cartas Geolgicas y Mineras 8, 1 sheet, scale 1:100 000. Martinez-Ruiz, V. J., and Cuellar-Gonzlez, G., 1978, Correlacin de superficie y subsuelo de la cuenca geohidrolgica de San Luis Postos, S.L.P.: Universidad Autnonoma de San Luis Potos, Instituto de Geologa, Miscellaneous Chart Series, 1 sheet, scale 1:50 000.. McDowell, F. W., and Clabaugh, S. E., 1979, Ignimbrites of the Sierra Madre Occidental and their relation to the tectonic history of western Mexico, in Chapin, C. E., and Elston, W. E., eds., Ash Flows: Geological Society of America Special Paper 180, p. 113124. McDowell, F. W., and Keizer, R. P., 1977, Timing of mid-Tertiary volcanism in the Sierra Madre Occidental between Durango city and Mazatlan, Mexico: Geological Society of America Bulletin, v. 88, p. 14791487. McDowell, F. W., Roldan-Quintana, J., and Amaya-Martinez, R., 1997, The interrelationship of sedimentary and volcanic deposits associated with Tertiary extension in Sonora, Mexico: Geological Society of America Bulletin, v. 109, p. 13491360. Meyer, R. P., Steinhart, J. S., and Woolard, G. P., 1958, Seismic determination of crustal structure in the central plateau of Mexico [abs]: Eos (Transactions, American Geophysical Union), v. 39, p. 525. Molina-Garza, R., and Urrutia-Fucugauchi, J., 1993, Deep crustal structure of central Mexico derived from interpretation of Bouguer gravity anomaly data: Journal of Geodynamics, v. 17, p. 181201. Monod, O., Lapierre, H., Chiodi, M., Martinez-Reyes, J., Calvet, P., Ortiz, E., and Zimmermann, J. L., 1990, Reconstitution dn arc insulaire intra-ocanique au Mexique central: La squence volcano-plutonique de Guanajuato (Crtac infrieur): Paris, Academie des Sciencies Comptes Rendues , v. 310, p. 4551. Moore, G., Marone, C., Carmichael, I. S. E., and Renne, P., 1994, Basaltic volcanism and extension near the intersection of the Sierra Madre volcanic province and the Mexican Volcanic Belt: Geological Society of America Bulletin, v. 106, p. 383394. Nieto-Obregon, J., Delgado-Argote, L., and Damon, P. E., 1981, Relaciones petrolgicas y geocronolgicas del magmatismo de la Sierra Madre Occidental y el Eje Neovolcnico en Nayarit, Jalisco y Zacatecas: Asociacin Ingenieros Mineros, Metalurgicos y Gelogos de Mxico, Memoria Tcnica, v. XIV, p. 327361. Nieto-Obregon, J., Delgado-Argote, L., and Damon, P. E., 1985, Geochronologic, petrologic and structural data related to large morphologic features between the Sierra Madre Occidental and the Mexican Volcanic Belt: Geofsica Internacional, v. 24, p. 623663. Nieto-Samaniego, A. F., 1990, Fallamiento y estratigrafa cenozoicos en la parte sudoriental de la Sierra de Guanajuato: Universidad Nacional Autnoma de Mxico, Instituto de Geologa, Revista, v. 9, p. 146155. Nieto-Samaniego, A. F., and Alaniz-Alvarez, S. A., 1994, La Falla de San Miguel de Allende: Caractersticas y evidencias de su actividad cenozoica: Tercera Reunin Nacional de Geomorfologa, Guadalajara, Jalisco, Mxico, Sociedad Mexiacana de Geomorfologa, Abstracts: p. 139142. Nieto-Samaniego, A. F., Macas-Romo, C., and Alaniz-Alvarez, S. A., 1996, Nuevas edades isotpicas de la cubierta volcnica cenozoica de la parte meridional de la Mesa central, Mxico: Revista Mexicana de Ciencias Geolgicas, v. 13, p. 117122. Nieto-Samaniego, A. F., Alaniz-Alvarez, S. A., and Labarthe-Hernndez, G., 1997, La deformacin cenozoica poslaramdica en la parte meridional de la Mesa Central, Mxico: Revista Mexicana de Ciencias Geolgicas, v. 14, p. 1325. Perez-Venzor, J. A., Aranda-Gmez, J. J., McDowell, F. W., and Solorio-Munguia, J. G., 1997, Bosquejo de la evolucin geolgica del volcn Palo Hurfano, Guanajuato: Revista Mexicana de Ciencias Geolgicas, v. 13, p. 174183. Ponce, S. B. F., and Clark, K. F., 1988, The Zacatecas mining district: A Tertiary caldera complex associated with precious and base metal mineralization: Economic Geology, v. 83, p. 16681682. Quintero-Legorreta, O., 1992, Geologa de la regin de Comanja, estados de Guanajuato y Jalisco: Universidad Nacional Autnoma de Mxico, Instituto de Geologa, Revista, v. 10, p. 625. Ramos-Salinas, J. A., 1996, Edades y composicin quimica del vulcanismo temprano en la porcin septentrional del eje Neovolcanico, regin Guanajuato: 5th International Meeting volcan Colima, Gobierno del Estado de Colima, Colima, Mxico, Abstracts (PC diskette). Randall-Roberts, J. A., Saldaa, E., and Clark, K. F., 1994, Exploration in a volcano-plutonic centre at Guanajuato, Mexico: Economic Geology, v. 89, p. 17221751. Reches, Z., 1983, Faulting of rocks in three-dimensional strain fields. II. Theoretical analysis: Tectonophysics, v. 95, p. 133156. Reches, Z., 1987, Determination of the tectonic stress tensor from slip along faults that obey the Coulomb yield condition: Tectonics, v. 6, p. 849861. Reynoso-Rosales, V. H., and Montellano-Ballesteros, M., 1994, Revisin de los Equidos de la fauna Cedazo del Pleistoceno de Aguascalientes, Mexico: Revista Mexicana de Ciencias Geologicas, v. 11, p. 87105. Rivera, J., and Ponce, L., 1986, Estructura de la corteza al oriente de la Sierra Madre Occidental, Mxico, basada en la velocidad del grupo de las ondas Rayleigh: Geofsica Internacional, v. 25, p. 383402. Royden, L., 1993, The tectonic expression of slab pull at continental convergent boundaries: Tectonics, v. 12, p. 303325. Scheubel, F. R., Clark, K. F., and Porter, E. W., 1988, Geology, tectonic environment and structural controls in the San Martin de Bolaos district, Jalisco, Mexico: Economic Geology, v. 83, p. 17031720. Severinghaus, J., and Atwater, T., 1989, Cenozoic geometry and thermal state of the subducting slabs beneath western North America, in Wernicke, B. P., ed., Basin and Range extensional tectonics near the latitude of Las Vegas, Nevada: Geological Society of America Memoir 176, p. 122. Silva-Romo, G., 1996, Estudio de la estratigrafa y estructuras tectnicas de la Sierra de Salinas, Estados de S. L. P. y Zacatecas [Masters Thesis]: Universidad Nacional Autnoma de Mxico, Facultad de Ciencias, 139 p. Stock, J. M., and Hodges, K. V., 1989, Pre-Pliocene extension around the Gulf of California and the transfer of Baja California to the Pacific plate: Tectonics, v. 8, p. 99115. Tristn-Gonzlez, M., 1986, Estratigrafa y tectnica del Graben de Villa de Reyes en los estados de San Luis Potos y Guanajuato, Mxico.: Universidad Autnoma de San Luis Potos, Instituto Geologa, Folleto tcnico (Open-File Report) 107, 91 p. Urrutia-Fucugauchi, J., and Rosas-Elguera, J., 1994, Paleomagnetic study of the eastern sector of Chapala Lake and implications for the tectonics of west-central Mexico: Tectonophysics, v. 239, p. 6171. Valdez-Moreno, G., and Aguirre-Diaz, G. J., 1997, La Joya stratovolcano (Quertaro). An example of the earlier volcanism in the MVB: International Association of Volcanology and Chemistry of the Earth Interior, General Assembly 1997 (Abstracts), Puerto Vallarta, Universidad Nacional Autnoma de Mxico, p. 58. Wark, D. A., Kempter, K. A., and McDowell, F. W., 1990, Evolution of waning subductionrelated magmatism, northern Sierra Madre Occidental, Mexico: Geological Society of America Bulletin, v. 102, p. 15551564. Webber, K. L., Fernndez, L. A., and Simmons, B., 1994, Geochemistry and mineralogy of the Eocene-Oligocene volcanic sequence, southern Sierra Madre Occidental, Juchipila, Zacatecas, Mxico: Geofsica Internacional, v. 33, p. 7789. MANUSCRIPT RECEIVED BY THE SOCIETY JUNE 11, 1997 REVISED MANUSCRIPT RECEIVED DECEMBER 2, 1997 MANUSCRIPT ACCEPTED MARCH 24, 1998

Printed in U.S.A.

Geological Society of America Bulletin, March 1999

363

You might also like