Download Ignimbrite flare-up and deformation in the southern Sierra Madre

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Geology wikipedia , lookup

Plate tectonics wikipedia , lookup

Algoman orogeny wikipedia , lookup

Large igneous province wikipedia , lookup

Sierra Madre Occidental wikipedia , lookup

Transcript
TECTONICS, VOL. 21, NO. 4, 10.1029/2001TC001302, 2002
Ignimbrite flare-up and deformation in the southern
Sierra Madre Occidental, western Mexico:
Implications for the late subduction history of the Farallon plate
Luca Ferrari
Centro de Geociencias, Universidad Nacional Autónoma de México, Campus Juriquilla, Queretaro, Mexico
Margarita López-Martı́nez
Departamento de Geologı́a, Centro de Investigación Cientı́fica y Educación Superior de Ensenada, Ensenada, Mexico
José Rosas-Elguera
Centro de Ciencias de la Tierra, Universidad de Guadalajara, Guadalajara, México
Received 17 May 2001; revised 10 February 2002; accepted 27 February 2002; published 24 August 2002.
[1] The Sierra Madre Occidental (SMO) of western
Mexico is one of the largest silicic volcanic provinces
on Earth, but the mechanism for the generation of
such a large volume of ignimbrites has never been
clearly defined. We present new 40Ar/39Ar ages,
geologic mapping, and structural data for the
southern part of the SMO demonstrating that most
of this volcanic province was built in two episodes of
ignimbrite flare-up in Oligocene (31.5–28 Ma) and
early Miocene (23.5 – 20 Ma) time, and that
extensional deformation occurred mostly before the
transfer of Baja California to the Pacific plate.
Extensive ignimbrite successions, with 40Ar/39Ar
ages clustering at 23 and 21 Ma, cover most of
the southern SMO, thus correlating in age with
ignimbrites exposed in southern Baja California and
central Mexico. Grabens with a 020 to N-S
orientation developed in the east almost concurrently
with this volcanic episode. Half grabens and NNW
striking listric normal fault systems formed at the end
of middle Miocene as far as 150 km from the present
coast. A belt of left-lateral transpressional structures
formed along the southern boundary of the SMO
during the same period. We link these magmatic and
tectonic events to the evolution and dynamics of the
Farallon and North America plates during the
Miocene. Particularly, we propose that a first
detachment of the lower part of the Farallon plate in
early Miocene time produced a transient thermal event
and partial melting of the crust via mafic underplating.
Middle Miocene extension would be related to a
second detachment event, resulting from the slowing
subduction that preceded the final capture of the
Magdalena microplate by the Pacific plate at 12.5 Ma.
Transpression at the southernmost end of the SMO
Copyright 2002 by the American Geophysical Union.
0278-7407/02/2001TC001302$12.00
occurred along the inland projection of the MagdalenaCocos plate boundary and may be explained by a
difference in subduction rate and by a temporal
convergence between the two plates in the eve of the
INDEX
end of subduction of the Magdalena plate.
TERMS: 5480 Planetology: Solid Surface Planets: Volcanism
(8450); 8150 Tectonophysics: Evolution of the Earth: Plate
boundary—general (3040); 8109 Tectonophysics: Continental
tectonics—extensional (0905); KEYWORDS: ignimbrite flare-up,
Sierra Madre Occidental, western Mexico, extensional tectonics,
slab detachment
1. Introduction
1.1. Purpose and Objectives
[2] The huge silicic volcanic plateau of the Sierra Madre
Occidental of western Mexico (SMO) is an enigmatic
feature in the geology of the North America plate. The
SMO runs for over 2000 km from the U.S.-Mexico border
to the Trans-Mexican Volcanic Belt (TMVB) (Figure 1). It
mostly consists of silicic ignimbrites and, to a lesser extent,
rhyolitic domes that cover about 300,000 km2 with an
average thickness of 1 km [McDowell and Keizer, 1977;
McDowell and Clabaugh, 1979]. With an estimated volume
of 300,000 km3 the SMO ranks among the major silicic
large igneous provinces on Earth. These kind of provinces
are generally related to continental breakup with variable
interaction of mantle plumes [e.g., Bryant et al., 2000;
Pankhurst et al., 2000], which provide sufficient thermal
energy to thin the continental lithosphere and, eventually, to
melt the crust. By contrast, in the SMO the production of
massive silicic volcanism occurred during subduction of a
young plate with no involvement of mantle plumes and took
place 20 to 10 Ma before the breaking of the continental
crust which led to the formation of the Gulf of California.
The inception of silicic volcanism in the SMO has been
generally considered as a marker of the beginning of
extension following the Laramide orogeny [e.g., Wark et
al., 1990; Luhr et al., 2001]. A plate tectonics mechanism
17 - 1
17 - 2
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Figure 1. Geodynamic map of Mexico showing Tertiary extension and volcanism north of the TransMexican Volcanic Belt (TMVB) and the present configuration of plates. Tertiary extension is from Henry
and Aranda-Gomez [2000] and this work. Regional tilt domains are from Stewart et al. [1998].
Abbreviations are as follows: Nay., Nayarit; Jal., Jalisco.
for the occurrence of massive ignimbritic volcanism, however, has not been provided.
[3] The SMO has been also strongly affected by extensional tectonics during the Tertiary. Extensional structures
form two broad, NNW-SSE trending, belts on both sides of
the SMO [Stewart et al., 1998], which merge in Sonora and
Chihuahua, to the north, and in Nayarit-Jalisco, to the south
(Figure 1). The western belt borders the Gulf of California
and is also known as the Gulf Extensional Province [Gastil
et al., 1975]; the eastern belt extends for several hundreds of
kilometers east of the core of the SMO and has been
considered the southern continuation of the Basin and
Range province of the western United States [Henry and
Aranda-Gomez, 1992]. Henry and Aranda-Gomez [2000]
further suggested that most of the Mexican Basin and Range
was linked to the interaction between the Pacific and the
North America plates during late Miocene time, after the
end of the subduction of the last remnant of the Farallon
plate (12.5 Ma) [Atwater and Stock, 1998]. Several studies,
however, indicate that extension initiated as early as late
Oligocene time both in the north [Nourse et al., 1994; Gans,
1997; McDowell et al., 1997; Stewart et al., 1998] and in
the south [Nieto-Samaniego et al., 1999; Luhr et al., 2001].
Therefore this first part of the extensional history remains to
be explained. In summary, despite a number of studies
published in the last two decades the causes of the ignimbrite flare-up and extension in the SMO are still not
completely understood.
[4] This paper presents new geologic, geochronologic,
and structural data that define the timing of silicic volcanism
and extension as well as the tectonic setting of the SMO
south of the Tropic of Cancer (latitude 23300). This vast
region, which was previously very poorly known, holds an
important piece of information for the geologic reconstruction of the whole province since it faces the last and larger
fragment of the Farallon plate (Magdalena microplate) to be
subducted before the initial rifting of the Baja California
peninsula [Lonsdale, 1991]. Our data fill the last gap in the
geologic reconnaissance of the SMO and enable us to
analyze at a global scale the temporal pattern of ignimbrite
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 3
Figure 2. Regional tectonic map of the southwestern Sierra Madre Occidental with locations of Figures
4 and 5. Shaded area is the late Miocene to Quaternary Trans-Mexican Volcanic Belt. Abbreviations of
tectonic structures are as follows: Co, Concordia fault; SP, San Pedro fault system; JM, Jesus Maria fault
system; SA, San Agustin graben; Ve, La Ventana graben; At, Atengo half graben; Po, Pochotitán fault
system; Hu, Huajimic half graben; Pc, Puente de Camotlán half graben; SMSR, Santa Maria del Oro –
Santa Rosa transpressional corridor.
flare-up and extension. On the basis of this rationale, we
propose a possible mechanism for the occurrence of these
phenomena with implications for the dynamics of the
Farallon slab during its final stage of subduction.
1.2. Location and Methodology
[5] South of the Tropic of Cancer the Sierra Madre
Occidental volcanic province can be divided into the
physiographic provinces of the Mesa Central high plateau
to the east and the Sierra Madre Occidental proper to the
west [Nieto-Samaniego et al., 1999]. The Sierra Madre
Occidental can be further divided into three domains: (1)
an eastern domain affected by several NNE to N-S trending
grabens, (2) a western domain where NNW trending half
grabens dominate the landscape, and (3) a southern domain
characterized by left-lateral transpressional structures (Figure 2). Our fieldwork mostly focused on the latter two
domains, bounded by the Guadalajara-Fresnillo highway to
the east, the Gulf of California to the west, the Mezquital
River to the north, and the Trans-Mexican Volcanic Belt to
the south (Figure 2).
[6] Only sparse geologic and structural works were
available in this region, mainly along its borders [Damon
et al., 1979; Clark et al., 1981; Nieto-Obregon et al., 1981;
Lyons, 1988; Scheubel et al., 1988; Ferrari, 1995]. The
paucity of studies was mostly due to the difficult access to
17 - 4
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Figure 3. Composite stratigraphic columns for the study region indicating typical thickness of units and
existing ages. New ages presented in this work are in bold. Full references of other ages are given in the
text.
this region. The volcanic plateau, with elevations ranging
from 2100 to 2900 m, is dissected by valleys with
elevations as low as 500 m that cut along the main
extensional structures. No paved roads exist, and only
two graded roads were recently completed to traverse this
part of the SMO in a WSW-ENE direction (Figure 2). To
make fieldwork more efficient in this complicated area we
have first studied 1:250,000 satellite images, 1:80,000, and
1:50,000 air photos, and digital terrain models at various
scales to compile a geologic base map. This was subsequently integrated with fieldwork mainly along two transects (Valparaiso-Estación Ruiz; Bolaños-Tepic) and
eventually completed with a few helicopter flights. The
resulting volcanic stratigraphy and geologic mapping are
illustrated in Figures 3, 4, and 5, and will be described in
the next section.
Figure 4. Geologic map of the Valparaiso –Estación Ruiz transect. Question marks indicate contacts between units
inferred from air photos only or unknown extension of units. Curved dashed lines indicate calderas inferred by aerial photos
and topographic features.
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 5
Santa Maria Ocotlán
15
20
?
28
15
San 30
Agustin
2000
10
2300
8
400
10
2200
22° 30'
Santa
Teresa
1300
2200
2000
Rio San
Pedro
o
dr m
Pe te
n ys
Sa ult s
fa
19.9
20.0
1100
30
Santa Fe
500
17.0
El Zopilote
Estación Ruiz
105° 00'
11.9
?
?
?
?
?
?
?
2400
? ?
?
?
2500
?
Mexquitic
?
1350
1800
?
?
21.1
?
2000
2100
?
1700
?
?
?
22° 00'
?
0
104° 30'
Ferrari et al. - Figure 4
25 km
2500
Monte
Escobedo
104° 00’
River
21.1 Ar/Ar date
Sierra
Los Huicholes
1900
2400
2300
2200
1000
22° 00'
50
ntiago
2100
28
50
?
1600
?
2500
700
15
2000
?
Peyotán
Jesus
Maria
Rio Sa
32
San Juan
Bautista
?
?
21.3 23.5
Mesa del
Nayar
1900
Bolaños
graben
?
2600
15
21.2
20.9
2100
2700
Santa Lucia
30
500
100
Huejuquilla
27.9
20 S. Juan
30
500
35
28
?
12
1900
2000
28.6
31.0
?
25
25
26
?
25
18
28 25
21.0
1400
?
?
1100
?
13
2000
?
?
23.5
Jesus Maria
half graben
2100
2400
Valparaiso
1850
28
2450
2400
1100
Francisco
Madero
Huazamota
35
?
26
20
32
28
3000
?
18
20
31.5
Las Canoas
?
1700
?
Sierra de Valparaiso
?
a
1800
?
?
s
su
Je
o ria
Ri Ma
?
20
30
25
20
8
?
15
103° 45’
?
?
ntan
zquital
15
1200
2700
2500
a Ve
10
?
20
?
2450
Rio L
500
21.1
?
22° 45’
?
1500
Rio Me
?
Llano Grande
Ameca
La Vieja
Atengo
half
graben
ng o
800
1900
La Ventana
graben
Rio Ate
San Agustín
graben
1700
25
(this work)
Paved road
23.5 Published
Unpaved road
K/Ar date
Extensional
fault
2300
Elevation (m)
Caldera wall
Dip of the ignimbrites
Horizontal strata
Extensional
anticline
Flexure
Miocene subvolcanic bodies
Conglomerate and sandstone
Miocene basalt and andesite
Miocene rhyolitic domes
Nayar ignimbrite
succession (21 Ma)
Oligocene basalt and andesite
Pre-Miocene rhyolitic domes
Canoas ignimbrite
succession (23 Ma)
Atengo ignimbrite
succession (28 Ma)
>28 Ma ignimbrites
Figure 5. Geologic map of the Bolaños-Tepic transect. Symbols as in Figure 4.
17 - 6
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
?
105° 00'
104° 45’
El Zopilote
?
?
104° 30’ 2080
500
?
El Venado
C
C
. La
e
le t
bol
Hu
Rio
11.9
a Huaynamota
c 104° 15’
mi
aji
?
?
1840
?
1920
14.0
2111
1000
10
Huajimic
19.9
2720
21
23.2
amot
lán
Bolaños
2320
22.2
900
2680
Rio C
2200
Sierra de
Pajaritos
18.7
22.4
Sierra Alic
a
nde d
ic
Rio Huajim
n
tá m
ti e
o st
ch sy
o t
P ul
fa
s
ic
an
lc a)
vo 0 M
VB 1 TM (1
Rio Gra
2640
1700
Aguamilpa
ago
e Santi
? ?
Tuxpan
de Bolaños
Guadalupe
Ocotlán
11.5
2400
2040
500
400
?
104° 00’
900
1000
1000
2240
Puente de
Camotlán
21°45’
San Martin
de Bolaños
23.7
2300
19.4
?
35
River
40
25
El
Anticline
21.1
Dip of the
ignimbrites
25
20
20
20
26
15
10.9
New age (this work)
19
25
15
30
?
1100
11.4
19 - 21 Ma ignimbrite succession
23 Ma ignimbrite succession
1640
? ?
2740
2500
Dated mafic dike
Miocene to Quaternary
volcano-sedimentary deposits
Oligocene andesites
and ignimbrites
1500
?
?
?
Scale
Santa Fe
25 km
21°15’
1600
Miocene subvolcanic bodies
?
0
45
?
50
0
Miocene rhyolitic domes
?
2500
te
Santa Maria del Oro
Early-middle Miocene basalts
21°30’
? ?
?
be
Published K/Ar age
be
10
20
20
20
23.5
?
?
2300
5
40
25
15
25
21.3
gra
La Manga
40
15
a
Pin
El
ra
Syncline
25
10
r
Sie
Extensional
fault
700
2140
5
45
17.2
15
1960
45
?
Bo
25
35
b
Ro
s
le
a
he
ños
TMVB volcanics
(11 - 0 Ma)
TEPIC
r z
10
VOLCAN
LAS NAVAJAS
e
on
lañ
500
Bo
la
Pochotitán
Unpaved road
20.7
35
25
Paved road
n
21.3
os
(m)
Rio
2300 Elevation
1700
TMVB volcanics
(11 - 0 Ma)
500
45
20
Ferrari et al. - Figure 5
ntiago
Rio Sa
?
?
TMVB volcanics
(11 - 0 Ma)
103°45’
30.1
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
[7] Our geologic reconnaissance is supported by 17 new
Ar/39Ar ages (Table 1), 2 new K-Ar ages (Table 1), and 16
published K-Ar data (details in the following section). The
40
Ar/39Ar ages were obtained at Centro de Investigación
Cientı́fica y Educación Superior de Ensenada’s (CICESE)
geochronology laboratory using a MS-10 mass spectrometer; details on the methodology are given in Appendix 1,
which is available as electronic supporting material.1 The
samples were step-heated between 700 and 1500C. A
summary of the 40Ar/39Ar is given in Table 1; the plateau
and isochron ages calculated are in Appendix A. In general,
we obtained good agreement between the plateau and
isochron ages. On those samples where duplicate analyses
were performed we obtained consistent results; the isochron
age calculated for the combined fractions of the duplicate
experiments is also reported in Table 1. The majority of the
samples yielded well-defined isochron ages which we take
as the best estimate of the age for the sample; however, we
had cases where the distribution of the data on the correlation diagram did not permit a realistic calculation of the
regression line or only the intercept with one of the axes
could unequivocally be calculated. This is the case for
samples TS-10, TS-21, and TS-46, for which we took the
plateau age as our best estimate of the age. For sample TS28 we analyzed concentrates of plagioclase and biotite, and
because of the scarcity of these minerals, only a few
milligrams could be obtained. Two experiments were conducted on the biotite with one and two fractions collected.
The four fractions collected on the plagioclase yielded a
saddle-shaped age spectrum. Our best estimate for the age
of sample TS-28 is taken from the isochron age calculated
with the combined fractions of the biotite and plagioclase.
[8] Scarce access and dense vegetation prevented a
detailed microtectonic study of the fault systems observed
in the region. However, the geometry and sense of slip of
over 200 striated faults and 57 dikes could be measured at
18 sites. The local paleostress regime responsible for the
observed deformation was computed by fault slip data
inversion with the method of Angelier [1990], and the
relative results are listed in Table 1. Using different
approaches, Pollard et al. [1993], Cashman and Ellis
[1994], Nieto-Samaniego and Alaniz-Alvarez [1997], and
Twiss and Unruh [1998] demonstrated that the assumptions
of the stress inversion methods are not always satisfied and
that caution should be used in using them. Particularly
simplistic assumptions often not fulfilled in nature are (1)
a homogeneous stress field, (2) no dynamic and kinematic
interaction between faults, and (3) the parallelism between
the maximum resolved shear stress on a fault plane and the
slickenline. To minimize this problem, we inverted slip data
of faults affecting previously unfaulted rock units and/or the
fault planes cutting all the other ones, i.e., those with the
higher probability of complying with the assumptions of
40
1
Supporting Appendix A is available via Web browser or via
Anonymous FTP from ftp://kosmos.agu.org, directory ‘‘append’’ (Username = "anonymous," Password = "guest"); subdirectories in the ftp
site are arranged by paper number. Information on searching and
submitting electronic supplements is found at http://www.agu.org/pubs/
esupp_about.html.
17 - 7
the inversion methods. Some faults with incongruous orientations or striations relative to the dominant population
appeared in the data set, but they were discarded in the final
computation. These results should then be representative of
the local paleostress conditions.
2. Geology and Geochronology
[9] In this section we present an overview of the geology
of the study region on the basis of information gathered
mostly along two ENE-WSW transects, 70 to 100 km apart,
that cross the southwestern SMO (Figures 2, 4, and 5).
Reconnaissance field geologic mapping was often extended
for 25 to 50 km to both sides of the transect when dirt roads
or tracks were available. Composite stratigraphic sections
were elaborated for the four areas with more information
and are illustrated in Figure 3.
2.1. Oligocene Volcanism
[10] The oldest rocks observed are exposed in the northeastern part of the study region and, possibly, in the west
along the coast (Figure 4). In the northeast they are silicic
ignimbrites and rhyolitic domes that extend widely to the
east outside the study region at elevations ranging between
2100 and 2400 m. In the Sierra de Valparaiso the ignimbrites attain at least 300 m of thickness. We obtained an age
of 31.5 ± 0.3 Ma for a sanidine concentrate of a welded ash
flow tuff also rich in quartz and plagioclase widely exposed
in the central part of Sierra de Valparaiso (sample TS 56,
Table 1). Similar ignimbrites are widespread toward the
south, in the Huejuquilla area (Figures 3 and 4). We have
dated two ignimbrites exposed in the footwall of the Atengo
half graben (Figure 4). A feldspar separate yielded an age of
31.0 ± 0.7 Ma (sample TS-5, Table 1), indistinguishable
from the Sierra de Valparaiso one. A sanidine concentrate
from an ignimbrite at an upper stratigraphic position produced an age of 28.6 ± 0.3 Ma (sample TS-10, Table 1). In
the western part of the study region, massive andesitic flows
constitute the base of the succession east of Estación Ruiz
(Figures 3 and 4). These rocks proved to be too altered to be
dated. Farther to the south, in the Santa Maria del Oro area
and along the Rio Santiago valley, a succession of ignimbrites and subordinated andesitic lavas underlie the Miocene
ignimbrite succession. We have dated a biotite separate
from a welded ash flow tuff at the southernmost edge of
the SMO, 50 km northwest of Guadalajara, at 30.1 ± 0.8
Ma by K-Ar method (sample ES 1, Table 1; Figure 5).
[11] Andesites and Oligocene ignimbrites are locally
covered by volcanic conglomerate and/or red sandstone.
Outcrops of volcano-sedimentary beds are normally too
small to appear on the geologic maps of Figures 3 and 4
but were observed in the Huejuquilla area (Figure 4), 15 km
NW of El Zopilote (Figure 4), in the Santa Maria del Oro and
Santa Fe areas (Figure 5), and in the Rio Santiago valley
southeast of these towns (Figure 5). Small plateaus of
basaltic to andesitic flows also cover the late Oligocene
ignimbrites in the northwestern part of the study region
(Figure 4). These, in turn, are capped by a distinct pyroclastic succession of poorly to mildly welded and crystal
Latitude, North
Elevation, m
Rock Type
Material Dated
10425.7000
10450.8920
Mexpan, Nay.
Arroyo El Naranjo, Nay.
Rio Chico, Jal.
Cotija area, Mich.
Rock Type
ignimbrite
rhyolite
ignimbrite
gms
feld
bio
bio
gms
feld
san
pl
w.r.
pl
san
feld
san
san
K-Ar Ages Obtained at Geochron Laboratory, Cambridge, Massachussets
2110.4900
ignimbrite
biotite
10346.2520
1953.7050
ignimbrite
feldspar
10302.4760
Elevation, m
1050
460
1050
ignimbrite
granodiorite
ignimbrite
ignimbrite
bas.-and.
ignimbrite
ignimbrite
ignimbrite
ignimbrite
ignimbrite
san
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
0.5
0.3
0.7
0.3
0.4
0.3
0.2
0.5
0.4
0.3
0.4
0.3
0.3
0.7
0.8
5.967
0.946
Average K,
wt %
19.9 ± 0.2
–
19.0 ± 0.6
18.3 ± 0.4
20.7 ± 0.2
19.9 ± 0.3
19.9 ± 0.2
–
21 ± 4
21 ± 1
23.3
21.3
21.1
21.1
21.1
21.1
21.0
20.4
19.7
31.6
31.1
28.6
27.9
27.7
28.1
tp Ma
0.01258
0.001301
Average 40Ar*,
ppm
31.5 ± 0.3
31.0 ± 0.7
–
27.9 ± 0.2
27.8 ± 0.8
27.6 ± 0.8
28 ± 2b
23.5 ± 0.4
–
–
21.1 ± 0.3
20.9 ± 0.4
21.2 ± 0.3
–
20.0 ± 0.6
19.9 ± 0.4
19.9 ± 0.4b
17 ± 9
22 ± 1
20 ± 2
20 ± 2b
20.6 ± 0.2
20.1 ± 0.6
19.8 ± 0.4
20.0 ± 0.3
20.0 ± 0.3
19 ± 2
–
–
17 ± 1b
tc Ma
18
44
12
11
15
64
48
41
24
5
8
7
17
7
10
5
5
2
±2
±
±
±
±
±
±
±
±
±
±
±
±
±
± 21
± 14
±8
±
±
±
±
±
± 27
± 21
Ar*, wt %
42.5
19.5
40
306
291
–
324
276
318
319
289
–
–
294
302
294
–
318
283
308
304
296
298
298
314
295
307
299
299
313
–
–
315
(40Ar/36Ar)i
30.1 ± 0.8
23.5 ± 0.9
Age, Ma
7.6/4
–
–
15.5/7
1.2/3
3.4/4
–
0.1/4
0.4/3
2.6/4
4.0/7
1.1/4
–
–
2.7/3
2.5/4
8.0/4
–
0.5/3
0.2/4
1.7/7
81/4
1.8/4
6.8/4
109/12
1.2/4
10.6/4
17.1/5
34.3/9
SumS/n
All errors are 1s. The abbreviations are exp, experiment(s); tp, plateau age with the error in J included; tc, isochron age calculated from the 36Ar/40Ar versus 39Ar/40Ar correlation diagram; SumS of York
[1969] n, number of points fitted; San, sanidine; bio, biotite; feld, feldspar; gms, groundmass; pl, plagioclase; w.r., whole rock. Preferred ages are shown in bold. Detailed results are given in Appendix A,
which is available as electronic supporting material.
b
Isochron age calculated with the combined fractions of more than one experiment.
a
ES 2
CO 1
2101.8000
2201.9090
10425.7000
Mexpan, Nay.
1710
100
1550
650
550
1780
1120
870
2050
2100
ignimbrite
Material Dated
2101.8000
10412.2920
Sierra Los Pajaritos, Nay.
Latitude, North
2135.0610
10505.6710
San Juan Bautista, Nay.
Longitude, West
2208.6100
10444.6970
SW Mesa del Nayar, Nay.
Location
2209.6450
10428.0910
10430.6670
10504.5400
10443.5660
10506.5630
10410.6910
10445.3330
east of Jesus Marı́a, Nay.
Jesus Maria, Nay.
west of Llano grande, Nay.
Arroyo El Frayle, Nay.
north San Juan Bautista, Nay.
south of Santa Lucia, Nay.
west of Mesa del Nayar, Nay.
1330
Sample
2217.4450
2215.1450
2245.6000
2210.3620
2211.0960
2205.6300
2220.6300
10359.195
Rio Atengo, Zac.
2240.089
0
feld+bio
40
Ar/ 39Ar Ages Obtained at CICESE Laboratory, Ensenada, Baja California
2253.2250
2450
ignimbrite
san
10343.3050
2239.5540
1480
ignimbrite
feld
10359.2510
2240.0890
1330
ignimbrite
san
10359.1950
2237.5890
2445
rhyolite
san
10413.1030
Longitude, West
0
Sierra Valparaiso, Zac.
west of Huejuquilla, Jal.
west of Huejuquilla, Jal.
NNE of Sta. Lucia, Zac.
Location
TS 56
TS 5
TS 10
TS 15
TS 11 1st exp
TS 11 2nd exp
TS 11 all exp
TS 22
TS 21
ESC-3
TS 25
ESC-2
ESC-4
TS 46
TS 26 1st exp
TS 26 2nd exp
TS 26 all exp
ESC 1 1st exp
ESC 1 2nd exp
ESC 1 3rd exp
ESC 1 all exp
TSS 1
Gdl 228 1st exp
Gdl 228 2nd exp
Gdl 228 all exp
Gdl 228
TS 28
TS 28 1st exp
TS 28 2nd exp
TS 28 all exp
Sample
Table 1. Summary of New Isotopic Ages for the Southern Sierra Madre Occidentala
17 - 8
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
poor ash flow tuffs, pumice flows, and ash and pumice falls
deposits. The latter was named the Atengo ignimbrite
succession and attains a minimum thickness of 200 m. A
sanidine concentrate for an ash flow tuff from the middle part
of this succession was analyzed twice, yielding an isochron
age of 28 ± 2 Ma (sample TS 11, Table 1). Moderately
welded conglomerate and sandstone cover the Atengo succession inside the half graben. On the western side of the
Atengo half graben a large rhyolitic dome complex was
dated at 27.9 ± 0.2 Ma (sample TS-15, Table 1).
2.2. Miocene Volcanism
[12] The Oligocene succession of the Huejuquilla-Atengo
area is covered by a younger ignimbrite succession 6 km
east of the village of Las Canoas (Figure 4). The Las Canoas
succession is made of several ash and pumice flows and
pumice falls that reach an aggregate thickness of at least 350
m. The most representative unit is a gray to pink, moderately welded ash flow tuff with small phenocrysts of biotite,
plagioclase, alkali-feldspar and hornblende. Clark et al.
[1981] dated this ignimbrite in the vicinity of Las Canoas
by the K-Ar method obtaining an age of 23.5 ± 0.5 Ma for a
feldspar separate. We have followed this unit over 30 km to
the southwest, where it is down-dropped by the large
extensional fault system of the Jesus Maria half graben
(Figure 4). Our new 40Ar/39Ar date of 23.5 ± 0.4 Ma
(sample TS-22, Table 1) for a plagioclase concentrate from
one of the westernmost appearance of the Las Canoas
succession is indistinguishable from the previous K-Ar
age of Clark et al. [1981]. Toward the south the Las Canoas
succession may be correlated with the lower part of the
succession exposed in the Bolaños graben (Figure 5). Here
Scheubel et al. [1988] report K-Ar ages of 23.7 ± 0.6 Ma for
an andesite intercalated in a sequence of ignimbrites and
rhyolites forming the lower part of the graben. They also
report an age of 23.2 ± 0.5 Ma for an ignimbrite of the same
succession (Figure 3). At places, thin mafic flows cap the
Las Canoas succession (Figures 3, 4, and 5). We obtained
whole rock 40Ar/39Ar ages of 21.3 ± 0.3 Ma for one of these
flows exposed close to the town of Jesus Maria (104300W,
22150N; Figure 4; sample TS-21, Table 1). Our 40Ar/39Ar
age is in complete agreement with the 40Ar/39Ar age of 21.6
± 0.3 Ma obtained by Rossotti et al. [2002] for a succession
of basaltic flows at the southern end of the Bolaños graben
and with the K-Ar dates obtained by Scheubel et al. [1988]
and Nieto-Obregon et al. [1981] for basaltic flows and dikes
in the Bolaños mining area (Figures 3 and 5).
[13] Immediately to the west of Jesus Maria (Figure 4),
the Las Canoas succession is covered by another silicic
succession that dominates the southwestern half of the study
area (Figures 4 and 5). This consists of several whitish to
light cream colored welded ash flows and air fall tuffs, and,
to a lesser extent, rhyolitic domes. The ignimbrites thicken
toward the Mesa del Nayar area (Figure 4). In addition,
several subcircular lineaments suggestive of caldera depressions are observed in the satellite images and digital topography model in this area (Figures 2 and 4). We checked in
the field the Nayar and Santa Teresa subcircular lineaments
and found coignimbritic lag fall and breccia deposits,
17 - 9
lacustrine deposits, and/or rhyolite domes associated with
them. These elements allow us to propose the existence of a
number of caldera depressions aligned in a NNW direction,
(the Nayar caldera field), which are the likely source of the
Nayar ignimbrite succession (Figure 4). The latter was
likely the product of many explosive eruptions occurred
over a very short time span. This is confirmed by the
40
Ar/39Ar geochronology. In the valley of Arroyo el Frayle,
which runs along the southern rim of the proposed Nayar
caldera, at least 11 ignimbrite sheets with an aggregate
thickness of 1000 m are observed. In this area we have
dated a number of ignimbrite sheets located at different
stratigraphic levels in the succession. The 40Ar/39Ar experiments performed on sanidine separates yielded ages of
21.1 ± 0.3 Ma for the lowermost ignimbrite and 21.0 ±
0.2 Ma for the uppermost sheet, located 1000 m above the
previous one (samples TS-25 and TS-46, Table 1, respectively). A third, moderately welded ignimbrite, at 3 km to
the west of the Nayar caldera rim, was dated 19.9 ± 0.4 Ma
(sample TS 26, Table 1). The Nayar succession can be
followed for tens of kilometers from the Mesa del Nayar
area along the coastal area at least up to the latitude of
Acaponeta (Figure 2). To the north of Mesa del Nayar the
succession caps the last 600 to 1000 m of the volcanic
plateau. Plagioclase phenocrysts from one of the highest
ignimbrites of the succession in the Llano Grande area
(105030W, 22450N, Figure 4) produced an age of 21.1 ±
0.7 Ma (sample ESC 3, Table 1). West of Mesa del Nayar
the succession is found up to the present coast where we
have dated the uppermost ignimbrite of a tilted block north
of the village of San Juan Bautista at 20.9 ± 0.4 Ma
(Figure 4; sample ESC 2, Table 1). To the east the topmost
ignimbrite in the Sierra Los Huicholes (Figure 4) provided
an age of 21.2 ± 0.3 Ma (sample ESC 4, Table 1). To the
south we have dated again the uppermost ignimbrite of the
succession exposed in the Sierra de Pajaritos (Figure 5).
Here a biotite separate yielded an age of 20.7 ± 0.2 Ma
(sample TSS 1, Table 1). In addition, several K-Ar ages
reported in previous works for ignimbrites in the Aguamilpa
and Santa Maria del Oro area (Figure 5) fall within the
limits of our 40Ar/39Ar ages [Gastil et al., 1979; Damon et
al., 1979; Soto and Ortega, 1982]. Similarly, in the Bolaños
graben, Scheubel et al. [1988] report K-Ar ages of 21.3 ±
0.5 Ma and 20.1 ± 0.4 Ma for ignimbrites of their Huichol
group, made of silicic ash flows and minor basalts (Figures 3
and 5), which dominate the succession exposed in the San
Martin de Bolaños area. As a whole, the Nayar ignimbrite
succession covers at least 5000 km2. Taking into account
erosional removal, we consider an average of 900 m as a
reasonable estimate of the original thickness. This would
imply an original volume of 4500 km3 emplaced in
1.4 Ma, a value comparable to modern large caldera
systems like the Taupo Volcanic Zone and Yellowstone
[Christiansen, 1984; Houghton et al., 1995].
[14] Several rhyolitic domes are intercalated with the
Nayar ignimbrite succession along the ring fault of the
inferred calderas. Toward the coast, rhyolitic domes are
more abundant and in some cases postdate the Nayar
succession (Figure 4). We obtained an isochron age of 17
17 - 10
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
± 1 Ma for a biotite and feldspar separate from a rhyolitic
dome in the El Zopilote area (Figure 4; sample TS-28,
Table 1).
[15] The Nayar ignimbrite succession is capped in places
by basaltic flows with a thickness of 50 to 100 m. In the
Mesa del Nayar area these lavas are undated. However, in
the Bolaños areas the uppermost basalts yielded a K-Ar age
of 19.9 ± 0.4 Ma [Nieto-Obregon et al., 1981]. Subvolcanic
intrusive bodies with granitic to quartz-monzonitic composition crop out on the western and southern margins of the
SMO along the San Pedro and Santiago rivers where they
flow at 150 to 50 m of elevation (Figures 4 and 5). We have
performed three 40Ar/39Ar experiments for a sample of the
San Juan Bautista granodiorite, one of the largest bodies in
the area (Figure 4). The resulting isochron age of 20 ± 2 Ma
(sample ESC 1, Table 1) is comparable to the K-Ar age of
17.2 ± 1.0 Ma reported by Rodriguez-Castañeda and
Rodriguez-Torres [1992] for another granitic body 70 km
to the south, and indicates that these rocks may be the
intrusive equivalent of the Nayar ignimbrite succession.
[16] Mafic volcanism represent the last volcanic event in
the study region. Many mafic dikes are commonly
emplaced parallel to NNW striking extensional faults in
the westernmost part of the area (Figures 4 and 5). Previous
K-Ar determinations for these dikes have consistently
yielded ages between 11.9 Ma and 10.9 Ma [Damon et
al., 1979; Clark et al., 1981]. Henry and Aranda-Gómez
[2000] obtained nearly identical ages for similar mafic dikes
in southern Sinaloa, just to the north of our study area. A
thick succession of basaltic flows that marks the inception
of the TMVB volcanism is emplaced at the boundary
between the SMO and the Jalisco block between Tepic
and Ixtlán del Rio and to the north of Guadalajara (Figure 2)
[Ferrari et al., 2000b]. The emplacement of these lavas
began at 11 Ma [Ferrari et al., 2000b, and references
therein].
2.3. Regional Correlation of the Silicic Volcanism
[17] Our geochronology results for the southwestern
SMO fit well with previous studies in the surrounding
regions. Particularly, part of our stratigraphy and ages
resemble closely those reported by McDowell and Keizer
[1977] along the Durango-Mazatlán transect, located
100 km to the north. Our ages for the Valparaiso-Huejuquilla area (31.5 –28 Ma) overlap with those obtained by
these authors in the Durango area, and our age of 27.9 Ma
for the rhyolitic dome on the western side of the Atengo
valley is identical to their age for a large rhyolitic dome at
Las Adjuntas, 110 km west of Durango city. Similarly, the
ages of the Las Canoas succession are identical to the 1000 m
thick El Salto-Espinazo del Diablo ignimbrite sequence, for
which McDowell and Keizer [1977] obtained eight ages,
indistinguishable within error, which cluster at 23.5 Ma.
Henry and Fredrikson [1987] also obtained similar ages for
rhyolite and dacite lavas in Sinaloa. The El Salto succession
can be followed several tens of kilometers to the SSE in the
direction of the Las Canoas area. A possible caldera structure
lies in between, centered on the village of Temoaya [Swanson
and McDowell, 1984] (Figure 2). Inside this area we found a
widespread fluvio-lacustrine sedimentary succession, which
would confirm the existence of a volcano-tectonic depression. Other early Miocene calderas may be located west of
Temoaya and southeast of Santa Lucia (Figure 2), but their
existence was only suggested by remotely sensed features
and helicopter surveys. Southeast of Bolaños, ignimbrites
23 m.y. old are reported in the Teul area [Moore et al., 1994]
(Figure 2) and can be followed to the south up to the Rio
Santiago, where Nieto-Obregon et al. [1985] report a K-Ar
age of 23.6 ± 0.6 for a plagioclase concentrate from an
ignimbrite 15 km to the northwest of the Santa Rosa dam
(Figure 2). South of the river the early Miocene ignimbrites
are covered by the younger volcanism of the Trans Mexican
Volcanic Belt (review by Rosas-Elguera et al. [1997]).
However, isolated blocks of ignimbrites tilted to the NNE
crop out from the northernmost part of the Pliocene succession of the Trans Mexican Volcanic Belt near the village of
Ixtlán del Rio (Figure 2). We have dated a groundmass
concentrate from a welded ash flow tuffs in this area at
20.0 ± 0.3 Ma (sample Gdl 228, Table 1).
[18] Previous workers reported ignimbrites with early
Miocene ages farther east. In the Juchipila area (Figure 2),
Webber et al. [1994] describe a succession of 10 ash flow
tuffs, four of which were dated by the fission track method
and yielded ages between 25.2 ± 2.2 and 24.9 ± 2.7 Ma.
However, the tuffs cover mafic lavas with a K-Ar age of
23.7 ± 1.4 Ma, leaving the possibility that the age of the
ignimbrite succession could still be 23 Ma. In the Los Altos
de Jalisco region, located west of Guadalajara (Figure 2),
ignimbrite and rhyolitic topographic highs emerged from a
late Miocene basaltic plateau. Castillo-Hernandez and
Romero-Rios [1991] obtained K-Ar ages of 24.1 ± 0.8 Ma
and 24.7 ± 1.0 Ma, for sanidine separated from an ignimbrite
and a rhyolite, respectively. The easternmost occurrence of
early Miocene ignimbrites is found in central Guanajuato
(Figure 6). Remnants of a single ignimbrite sheet cover
Oligocene rhyolitic domes in the La Sauceda area. This
ignimbrite, dated by K-Ar at 24.8 ± 0.6 Ma [Nieto-Samaniego et al., 1996], represents a distal facies and probably
came from the Los Altos de Jalisco area.
[19] Ignimbritic successions are also widespread to the
south of the TMVB but are typically Late Cretaceous to early
Paleocene in age within the Jalisco block [Ferrari et al.,
2000a]. However, Tertiary ignimbrite units are exposed more
to the east (Figure 6). Although most of the silicic volcanism
in Michoacán and Guerrero is Oligocene [Morán-Zenteno
et al., 1999], some younger ignimbrites are reported immediately to the south of the TMVB. About 35 km south of
Lake Chapala (Figures 2 and 6), a thick pyroclastic succession overlies Tertiary intrusives and Cretaceous sedimentary
rocks from the Michoacán block. The uppermost ignimbrite
of this succession is a light brownish moderately welded unit
with microphenocrysts of sanidine and white pumice supported by an ash matrix, which we called the Cazos ignimbrite. We have obtained a K-Ar age of 23.5 ± 0.9 Ma for a
feldspar separate from the Cazos ignimbrite (sample CO-1,
Table 1). In the Morelia area (Figure 6), a 200 – 300 m thick
rhyolitic ignimbrite succession forms prominent cliffs
15 km south of the city (Puerto la Sosa area). Pasquarè
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 11
Figure 6. Geographic plot of available ages (circles and boxes) of Oligocene to early Miocene silicic
rocks in central Mexico (data from Ferrari et al. [1999], and this work). Note the broad NNW trending
Oligocene silicic arc and the ESE migration of silicic volcanism between the 23 to the 21 Ma pulse. Cross
pattern indicates the extension of the Los Cabos and Jalisco (Puerto Vallarta) late Cretaceous batholiths.
Boundaries of Mexican states are also shown. Abbreviations are as follows: BC, Baja California; Nay,
Nayarit; Jal, Jalisco; Gto, Guanajuato; Mich, Michoacán; Gro, Guerrero.
et al. [1991] dated a ignimbrite from the upper part of the
succession by K-Ar at 21 ± 1 Ma. These ages suggest a
possible continuation of the 23 and 21 Ma silicic
volcanism up to some tens of kilometers south of the TMVB.
[20] In southern Baja California, Hausback [1984] and
Umhoefer et al. [2001] obtained ages in the range 23– 17
Ma for ignimbrites intercalated in the lower part of the
Comondú Formation. Particularly interesting are the ages
obtained for the ‘‘La Paz tuff’’, a distinct succession of
welded ash flow tuffs exposed in the La Paz-Punta Coyote
area. Hausback [1984] reports K-Ar ages ranging between
21.8 ± 0.2 and 20.6 ± 0.2 Ma for this succession, which
largely overlap our ages for the Nayar succession. Given the
large geographic extent of the Nayar ignimbrite succession,
the possibility exists that the La Paz tuff could be part of the
latter. This correlation poses an additional constraint on
the prerifting position of Baja California and confirms that
the Los Cabos block was located immediately to the northwest of the Jalisco block prior to the opening of the Gulf of
California [Schaaf et al., 2000] (Figure 6).
3. Tectonics of the Southwestern SMO
3.1. Geometry, Kinematics, and Time of Faulting
[21] Several large structures cut the Tertiary succession of
the southwestern SMO (Figures 2, 4, and 5). According to the
geometry and kinematics they can be grouped in an eastern
and a western extensional domains, and in the Santa Maria
del Oro-Santa Rosa transpressional corridor (Figure 2).
3.1.1. Eastern Extensional Domain
[22] This domain comprises six full grabens 30 to 40 km
apart: Aguascalientes, Juchipila, Tlaltenango, Bolaños, La
Ventana, and Mezquital (Figure 2). The last two were not
studied but are grouped with the others because of their
orientation and probable age. The first four grabens trend
010 to N-S and are typically over 100 km long and 20 km
wide. The grabens are bounded by high-angle faults with
dominant dip-slip displacements. Inside the graben, volcanic or sedimentary beds show tilts of 10 to 20 either to
the east and the west. Offset of stratigraphic units can only
be observed in the Bolaños graben, where it may reach
1800 m. Nieto-Samaniego et al. [1999] interpreted the four
grabens as the results of E-W extension during the 23 to
21 Ma interval. This age for the extensional faulting was
mainly inferred from geologic relations in the Tlaltenango
graben where the flows of a basaltic shield volcano dated
21.8 ± 1.0 Ma abut against the fault scarps and appear
unfaulted [Moore et al., 1994]. During our fieldwork,
however, we observed that two normal fault scarps, <50
m high, also affect the shield volcano. Furthermore, we
found that the 21 Ma ignimbrite succession is faulted in
the Bolaños graben.
17 - 12
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
[23] In the Bolaños mining area (103450W, 21490N;
Figure 5) Lyons [1988] estimates 1 to 1.5 km of normal
offset of the mineralized body. The mineralization was
dated at 20.8 ± 1.0 Ma, but it is unequivocally covered by
basaltic flows dated 21.0 ± 0.4 Ma [Nieto-Obregon et al.,
1981]. According to our observations in the Bolaños area
the offset of the 21 Ma succession does not exceed 400–
500 m. In the San Martı́n de Bolaños area (Figure 5),
Scheubel et al. [1988] dated at 21.3 ± 0.5 Ma an ash flow
tuff on the floor of the graben, which is also found outside
the graben over 1300 m higher. However, the geologic cross
section of Scheubel et al. [1988] shows the ash flow tuff
emplaced against the upper part of the western graben wall
with a subvertical contact, suggesting that it could have
flowed into a preexisting depression. Lyons [1988] also
observed three generations of faults in the Bolaños area: A
060 normal fault set is cut by a 030 set which is in turn cut
by N-S normal faults. Our regional observations, however,
seem to indicate that this is not the rule along the Bolaños
graben. In fact, the mining area is located within a broad
accommodation zone that corresponds to a faulted relay
ramp between two right stepping N-S trending normal
faults. In this context the progression of faulting observed
by Lyons [1988] may be interpreted as the progressive
reorientation of secondary extensional structure during the
northward propagation of the southernmost fault.
[24] We interpret the above observations in the following
way. The eastern extensional domain was extended shortly
after the emplacement of the 23 Ma ignimbrite succession.
The extensional pulse probably start to wane during the
emplacement of the shield volcano in the Tlaltenango
graben (21.8 ± 1.0 Ma) and after the deposition of the
Nayar ignimbrite succession (21 Ma) in the Bolaños
graben area. Alternatively, the Bolaños graben could have
been involved in the younger episode of extension that
affected the western domain.
3.1.2. Western Extensional Domain
[25] The region between the Bolaños graben and the
coastal plain is affected by N-S to NNW trending major
structures that cut the grabens of the eastern domain
(Figure 2). They are the Jesus Maria half graben and its
northern continuation, the San Agustin graben, the Sierra
Alica, Atengo, Sierra de Pajaritos, and Puente de Camotlán
half grabens. To the west of these structures lies the
relatively undeformed zone of the Nayar caldera field.
Farther to the west the Pochotitán and San Pedro normal
fault systems bound the Gulf of California (Figures 2, 4,
and 5). With the sole exception of the northern part of the
Jesus Maria half graben and a small area east of Jesus
Maria village (104300W, 22150N; Figure 4), all major
faults in this region systematically dip to the west or WSW
and tilt the hanging blocks up to 30 to the east or ENE.
This broadly extended region is separated from an undeformed zone to the north by the Rio Mezquital accommodation zone (Figure 2). This is also the boundary of two
domains with opposite vergence: To the north, in fact, the
beds systematically dip to the WSW, and major faults dip to
the ENE (Figure 2) as documented by Henry and ArandaGomez [2000]. The Rio Mezquital accommodation zone
was not studied in the field, but it could be a fault zone
with a right-lateral component of motion. At the latitude of
Tepic, the El Roble shear zone is a left-lateral normal
accommodation zone that divides the western extensional
domain from the Santa Maria del Oro transpressional
corridor (Figures 2, 5, and 7). The Jesus Maria and the
Atengo half grabens have vertical offset of over 2.2 km and
1 km, respectively. They both end to the west with flexure
zones that bound flat lying and undeformed zones
(Figure 4). In its northern part the Jesus Maria structure
becomes a full graben (San Agustin graben), and the
flexure becomes a broad extensional anticline (Figures 2
and 4). The Sierra Alica and Sierra de Pajaritos half
grabens have vertical offset of 1.3 km, whereas the
Puente de Camotlán half graben has an offset of 700 m.
The San Pedro and Pochotitán fault systems form a major
breakaway zone bounding the Gulf of California (Figures
2, 4, and 5). They consist of a series of normal faults
cutting the flat lying SMO plateau to the east and have a
cumulative minimum displacement in the order of 1.7
km. The faults are restricted to a 20 km wide and 160 km
long belt, where ignimbrites as young as 19 Ma are found
in a step-like structure with tilts up to 35 to the ENE.
Outcrop-scale faults and accompanying dikes show a
dominant NW to NNW trend (Figure 7). North of the
Mezquital accommodation zone these fault systems continue into the Concordia fault [Henry and Aranda-Gomez,
2000], although their dip and the sense of tilting is reversed
(Figure 2).
[26] On average the direction of extension (sHmin) was
055 ± 17, indicating a consistent SW direction of
extension for the whole western domain and part of the
Bolaños graben (Table 2 and Figure 7). We interpret the
uniform direction of extension as an indication of a common
phase of deformation.
[27] Available ages of faulted units indicate that the
Atengo half graben is younger than 28 ± 2 Ma, the Jesus
Maria and Sierra de Pajaritos half graben are younger than
20.6 Ma and the Pochotitán and San Pedro fault systems are
younger than 19 Ma (Figures 4 and 5). However, no dated
geologic units can constrain the upper age of activity of the
former three structures. In the case of the San PedroPochotitán fault system the mafic dikes intruded along the
normal faults have been dated at 11.9 ± 0.3 Ma at El Zopilote
(Figure 4) [Clark et al., 1981] and 11.5 ± 0.5 Ma at
Aguamilpa [Ferrari et al., 2000a] (Figures 2 and 5). At this
latter location we measured 39 dikes with an average strike
of N238E, some of which are tilted up to 70 (Figure 7 and
Table 2, site 2). Between Aguamilpa and Tepic (Figure 4),
however, basaltic flows dated 8.93 ± 0.11 Ma [Righter et al.,
1995] are horizontal and cover tilted blocks of SMO ignimbrites. In Sinaloa, just to the north of our study region, Henry
and Aranda-Gomez [2000] observed a similar situation.
They obtained 40Ar/39Ar ages of 10.7 ± 0.2 and 11.0 ±
0.2 Ma for dikes moderately tilted and intruded in the
conglomerate filling of the NNW trending half graben
bounded by the Concordia fault. Accordingly with the above
observations we consider that extension along this part of the
eastern margin of the Gulf of California may have initiated a
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 13
Figure 7. Tectonic map and microtectonic data collected in the western extensional domain. Bold arrow
indicates the inferred direction of motion of a block to the west (Los Cabos block?) needed to produce the
observed faulting.
few millions of years before the intrusion of the dikes and
continued for 1 – 2 m.y. after their emplacement. A middle to
late Miocene age for faulting at the southern end of the Gulf
of California is also confirmed by recent thermochronologic
studies on the exhumation of the Los Cabos block, which
would have been located just to the west of our study region
at that time [Fletcher et al., 2000]. Particularly, they indicate
that fast extensional exhumation of the Los Cabos block
commenced by 12 Ma.
3.1.3. Santa Maria del Oro-Santa Rosa Transpressional
Corridor
[28] The southernmost part of the SMO, located close to
the boundary of the Jalisco block, contrasts markedly with
areas to the north. Instead of extensional structures contractile deformation dominates. The most striking structures
are 10 to 40 km long NNW trending open folds arranged in
an en echelon pattern between Santa Maria del Oro and the
Santa Rosa dam (103440W, 20550N; Figures 2 and 5)
[Ferrari, 1995]. Ignimbrites involved in the folds typically
dip 30– 35 with the exception of the western flank of the
Sierra El Pinabete asymmetric fold (Figure 5), where strata
dip up to 75 – 80 on its western limb. In the field the rocks
are affected by many 135 to 150 striking left-lateral
oblique thrust and strike-slip faults (Figure 8). All these
structures are indicative of a left-lateral transpressional
shearing. A second group of folds with N-S trending
and parallel axes lies to the WNW of the en echelon folds in
the La Manga area (Figures 5 and 8). The strata are only
gently tilted with dip of 5 – 10. These folds appear to be
formed by low- intensity E-W compression on the rear of
the northern part of the transpressional shear zone.
[29] Folds and faults cut rocks as young as 19.0 ± 0.4 Ma
east of Santa Maria del Oro [Damon et al., 1979] and 16.9 ±
0.5 Ma [Nieto-Obregon et al., 1985] in the Santa Rosa dam
area. The folds are cut by vertical mafic dikes dated at 11.4 ±
0.3 Ma and 10.9 ± 0.2 Ma [Damon et al., 1979] (Figure 5).
[30] The average direction of horizontal compression
(sHmax) was 100 ± 26 (Table 2). However, the style of
deformation varies from compression in the northwest (sites
11– 13, Table 2; Figure 8) to left-lateral transpression and
oblique thrusting at the center of the area (sites 14– 17), to
almost left-lateral transcurrence in the Santa Rosa area (site
east of Jesus Maria
Aguamilpa Dam
Bolaños Graben
La Cofradı́a
Presa El Bañadero
Paso De Lozada 1
Paso De Lozada 3
Sierra Pajaritos
east Bolaños Graben
Aguamilpa dam
El Cajon 1
El Cajon 2
Santa Fe
Plan de Barrancas Highway km 115
Plan de Barrancas Highway km 115, 3
Paso de La Yesca
Santa Rosa Dam
00
1044500000
1042705500
1042705000
1042600500
1041103000
1041105000
1040403900
10444 30
1043103000
1043101000
1041603000
various sites
SW of site 3
1045701000
0
1042605900
1044500000
1034105100
Longitude, West
Lithology
101/9
114/2
121/15
121/16
275/15
115/15
278/20
216/60
141/62
early Miocene
g
33/81
39/72
323/79
225/69
280/69
84/71
216/60
163/67
s1c
23.2 Ma
11.9 Ma
20.1 Ma
early Miocene
early Miocene
early Miocene
17.2 Ma
20.7 Ma
21.0 Ma
Age of Rocksb
Southern Domain: Left Lateral Transpression
215100000
ignimbrite
22.4 – 18.7 Ma
212502000
ignimbrite
early Miocene
212501500
ignimbrite
early Miocene
0 00
2116 05
ignimbrite
early Miocene?
0 00
2101 00
ignimbrite
early Miocene
210004500
ignimbrite
early Miocene
210804800
ignimbrite
early Miocene
ignimbrite
17 Ma
221802100
215100000
215702800
Western Domain: Extensiónf
ignimbrite
basaltic dikes
ignimbrite
ignimbrite
0 00
2136 30
ignimbrite
213003000
andesite
213003000
granite
213803000
ignimbrite
various sites
basaltic dikes
SW of site 3
220002000
ignimbrite
Latitude, North
9/10
204/11
214/12
331/72
129/73
2/55
81/69
4/26
315/28
129/1
147/6
140/11
125/4
139/17
326/9
4/26
340/23
s2c
230/76
15/79
342/71
214/8
7/9
214/31
186/6
110/14
47/2
219/9
238/16
230/1
34/21
45/13
234/16
100/14
70/1
235/1
s3c
8
8
8
7
9
12
9
13
9
19
39
20
8
8
9
13
8
15
Nd
0.75
0.87
0.85
0.73
0.39
0.53
0.24
0.52
0.53
0.50
0.27
0.20
0.17
0.52
0.14
0.35
Phie
b
Number of sites as in Figures 7 and 8.
Age of rocks according to dated samples (details in the text) or to stratigraphic relations.
c
Trend and plunge of stress tensor axes determined by fault slip data inversion according to the method of Angelier [1990] except at sites 2 and 9, where they are eigenvectors of pole to dikes obtained by
density analysis with the program Orient [Charlesworth et al., 1988].
d
Number of planes used in the computation.
e
Tensor shape is (s2 – s3)/(s1 – s3).
f
Average direction sHmin = N55 ± 17.
g
Average direction of sHmax = N100 ± 26.
a
11,
12,
13,
14,
15,
16,
17,
18,
10, east of El Venado
1,
2,
3,
4,
5,
6,
7,
8,
9,
Site Number and Locationa
Table 2. Kinematics of the Deformation
17 - 14
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 15
Figure 8. Tectonic map and microtectonic data collected in the Santa Maria del Oro-Santa Rosa
transpressional corridor. The geometry of the northern boundary of the Jalisco block is inferred on the
basis of Ferrari et al. [2000a]. Bold arrow indicates the inferred direction of motion of the Jalisco block
needed to produce the observed deformation.
18). This variety of structures is likely related to the original
geometry of the boundary between the Jalisco block and the
SMO, which was reactivated by WNW-ESE compression in
middle Miocene time (Figure 8).
3.2. Magnitude of Extension
[31] Although the normal faults of the study region are
prominent features on satellite images and aerial photos, the
deformation they accommodated was modest. Nieto-Samaniego et al. [1999] estimated an 8% of stretching in a nearly
E-W direction from a profile crossing the Juchipila, Tlaltenango, and Bolaños graben. In the western extensional
domain the relatively steep dipping (70 – 60) of faults
and modest (30) tilting of blocks are also indicative of
a limited amount of extension. The present detail of map-
ping and the lack of a clear stratigraphic marker prevent the
construction of a precise retrodeformable cross section in
this area. However, we have tried to give a rough estimate
of the amount of extension along a profile normal to the
major extensional structure, the Jesus Maria half graben,
using the area balance method described by Groshong
[1994] (Figure 9). In this case we have considered the top
of the 23 Ma ignimbrite as the reference level prior to
extension and erosion (Figure 9). Since this surface outcrops
at 2200 m elevation east of the fault system and at
1400 m west of the deformed zone, an average height
of 1800 m was chosen. We took into account different
depths of detachment and chose 10 km as a realistic, though
conservative, approximation. This value gives a horizontal
extension of 9% (Figure 9). In the less likely case of a
shallower detachment level the extension would remain
17 - 16
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Figure 9. Interpretative cross section through the Jesus Maria half graben with estimation of minimum
amount of extension using the method proposed by Groshong [1994]. The top of the 23 Ma Las Canoas
ignimbrite succession was taken as reference level. Vertical exaggeration 1.65X.
relatively small: For a detachment at 5 km depth, extension
would be 17%.
4. Discussion
4.1. Episodic Nature of Ignimbrite Volcanism in the
SMO
[32] Previous studies recognized that most of the silicic
volcanism in the SMO occurred in early Oligocene time
(32 –28 Ma) throughout the province and in early Miocene
time (23 Ma) in the southern part of the province
[McDowell and Keizer, 1977; McDowell and Clabaugh,
1979; Henry and Fredrikson, 1987; Wark et al., 1990;
Aguirre-Dı́az and McDowell, 1993; McDowell et al.,
1997; Gans, 1997; Nieto-Samaniego et al., 1999]. Our
new geochronologic data confirm these results but also
demonstrate the regional significance of the early Miocene
episode. Ignimbrites in the southwestern SMO actually
cluster in the 32 –28 Ma and 24 –20 Ma time spans, and
the latter episode covers about two thirds of the southwestern SMO. Thus our data confirm that ignimbritic volcanism
in the SMO was concentrated in two pulses. The episodic
nature of this volcanism can be appreciated in the probability density histogram of Figure 10, where all the available
ages of silicic rocks of the southern SMO are plotted. Even
so, the real size and the short duration of the two pulses of
ignimbrite volcanism are not completely evident because
available ages between 28 and 24 Ma often show large
experimental errors and because the volume of volcanism is
only loosely related with the number of available ages.
[33] The distribution of this voluminous silicic volcanism
in space and time define a clear trend (Figure 6). During the
Oligocene ignimbrites were emplaced in a broad NW trending arc with a width of 400 km. The two early Miocene
pulses are located in narrower belts within a progressively
more southeastern position. It is clear that the locus of
ignimbrite flare-up has converged toward the southwest from
early Oligocene to early Miocene (24– 22 Ma) and then
migrated toward the SSE during the 21 to 17.5 Ma time
span. A comparison of the timing and space distribution of
the Oligocene and early Miocene ignimbrite flare-up (Figures
6 and 10) indicates that the Oligocene episode was less
concentrated in time and space, it occurred over a wider area,
and had internal pulses less pronounced than the early
Miocene one.
4.2. Generation of Silicic Volcanism in the SMO
[34] Voluminous episodes of silicic volcanism have been
related to different causes. The most common situation is
continental break-up with variable involvement of mantle
plumes as in the case of the huge Whitsunday Volcanic
Figure 10. Histogram and cumulative probability curve
for ages of silicic volcanic rocks in the Sierra Madre
Occidental south of the Tropic of Cancer plotted constructed
with Isoplot/Ex [Ludwig, 1999]. Note the marked pulse of
silicic volcanism at 23 and 21 Ma.
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Province of northeastern Australia [Bryant et al., 2000] or
the Chon Aike volcanic province of Patagonia and western
Antarctica [Pankhurst et al., 2000]. On the other hand, the
late Miocene-Quaternary central Andean silicic province
has been explained as a consequence of delamination of the
lower lithosphere and, possibly, the lower crust, with subsequent rising of hotter asthenospheric mantle [Kay and
Kay, 1993]. In all the above cases a rapid increase in the
uppermost mantle temperature is thought to have driven
extensive melting of the mantle lithosphere and partial
melting of the lower crust via basaltic underplating.
[35] Isotope studies indicated contrasting scenarios for
the generation of the Oligocene silicic volcanism in the
SMO. On one hand, K. L. Cameron and coworkers [e.g.,
Cameron et al., 1980; Smith et al., 1996, and reference
therein] have suggested that fractional crystallization of
mafic magmas with relatively small crustal interaction
generated the Oligocene ignimbrites in Chihuahua, while
Ruiz et al. [1988, 1990] advocate a major involvement of
the crust through anatexis. From binary mixing calculations,
Verma [1984] also estimated values of involvement as high
as 80% for Oligocene ignimbrites in the Zacatecas area.
More recently, Albrecht and Goldstein [2000] modeled Sr,
Nd, and Pb isotopic data from the central SMO in Chihuahua and obtained values of crustal involvement of up to
70%. Particularly, they propose that the silicic rocks were
the result of assimilation/fractional crystallization or melting, assimilation, storage, and homogenization processes
affecting first the lowermost crust and later the middle crust,
as the thermal anomaly would rise into higher sections of
the crust.
[36] Although no isotope and geochemical studies are
available for the early Miocene ignimbrite flare-up of the
southern SMO, several lines of reasoning also suggest that
melting of the crust played an important role in generating
this massive silicic volcanism. First, the rate of silicic
magma production and emplacement in the SMO during
the early Miocene episodes is very high. The two silicic
successions (23 and 21 Ma) have a thickness of 1000 m
close to the source, which gradually decreases to <100 m in
distal areas. They form a 700 km long and 120 km wide belt
from central Sinaloa to western Guanajuato (Figure 6),
covering an area of 84,000 km2. Thus, considering a mean
thickness of 500 m, a volume of 42,000 km3 appears to have
been emplaced in 2 m.y. These values approximate those
of plume-related silicic large igneous provinces and suggest
a similar mechanism. Second, partial melting of the crust
needs much less amount of primary mafic magmas to
produce a rhyolite. Several studies [Huppert and Sparks,
1988; Harry and Leeman, 1995] have shown that one
volume unit of basalt may be able to produce from two
thirds to one equivalent volume of rhyolite by crustal
melting over timescale of 102 to 103 years. By contrast,
fractional crystallization needs much more basalt volume,
takes longer time periods, and implies the formation of rocks
of intermediate composition. Our study of the southern SMO
shows that volcanism occurred in fast and short pulses and
that rocks of intermediate composition are absent. We thus
conclude that the early Miocene ignimbrite flare-up was
17 - 17
primarily produced by a massive and fast generation of
subcrustal mafic melts, which provided sufficient thermal
energy to melt the crust.
4.3. Plate Tectonic Control of Ignimbrite Flare-Up
[37] The rapid generation of mafic melts in the subcrustal
lithosphere over several hundreds of kilometers asks for a
plate tectonic mechanism. In the western United States, the
Tertiary silicic flare-up has been related to the removal of
the Farallon slab from beneath the North America plate
[e.g., Coney, 1978; Humphreys, 1995] as the slab, which
was subhorizontal during the Laramide orogeny, progressively rolled back and foundered in the mantle, exposing the
upper plate to hotter asthenospheric mantle. This mechanism may have worked as well in Mexico during the
Oligocene episode of ignimbrite flare-up that was coeval
over a large area from the U.S.-Mexico border to the south
of the Trans-Mexican Volcanic Belt. Indeed, it could have
been triggered by a relatively sudden rollback of the
subducted plate [Nieto-Samaniego et al., 1999] following
the slowing in the Farallon-North America relative convergence that should have occurred some millions of years
before the first contact of the East Pacific Rise with the
continent [Atwater, 1989].
[38] For the early Miocene episode, however, we propose
that the detachment of the Farallon slab (also called ‘‘slab
breakoff’’) was the ultimate control over the timing and
localization of silicic volcanism and extension in the southern SMO. Seismic tomography studies [Van der Lee and
Nolet, 1997] have shown that the upper mantle beneath the
SMO and north central Mexico is characterized by two
subparallel high velocity anomalies (Figure 11). The most
obvious interpretation of this double anomaly is that it
represents two fragments of the subducted Farallon slab
that detached at different times. Detachment of the subducted slab is well known in the Mediterranean-Carpathian
region as a consequence of the arrival of the more buoyant
continental crust in the subduction zone (see Wortel and
Sparkman [2000] for a review). In a similar way, in western
Mexico the initiation of detachment was the natural consequence of the approach of the East Pacific Rise to the
paleotrench west of Baja California. The arrival of very
young (<10 m.y. old) and buoyant oceanic crust in the
subduction zone eventually results in the waning and
cessation of subduction of the last remnant of the Farallon
plate and their capture by the Pacific plate [Lonsdale, 1991].
The captured portion of the subducted slab began to move
in an opposite direction (with the absolute motion of the
Pacific plate), whereas the rest of the slab kept on sinking in
the mantle. Therefore a tear in the subducted slab should
have initiated at the time of the stalling of subduction and
should have propagated laterally following the progressive
termination of subduction off North America. According to
the reconstruction of Atwater and Stock [1998] the first
breakoff of the subducted part of the Farallon slab initiated
at 28 Ma at the latitude of southern California, after the
first interaction between the Pacific and the North America
plates. At this time a trench parallel tear started to separate
the short slab attached to the young Monterey and Arguello
17 - 18
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Figure 11. Map and cross section showing the main upper mantle seismic anomaly detected by the
tomographic model of Van der Lee and Nolet [1997]. A double-high velocity anomaly corresponds
broadly to the upper mantle region below the SMO in the early Miocene. Light shaded area in the cross
section represents a positive anomaly exceeding 50 m/s, which is interpreted as portions of subducted
slabs. Bold dashed line shows a possible geometry of the slab fragments (proposed in this work). Dark
shading indicates a low-velocity anomaly below 50 m/s, corresponding to the shallow warm region
which underlain the Gulf of California rift.
microplates from the deeper and sinking part of the Farallon
slab (Figure 12a). We propose that the tear in the subducted
slab propagated laterally toward the SSE until reaching the
southern SMO at the beginning of early Miocene time
(Figures 12a and 13a). The seismic velocity contrast
between the slab and the surrounding mantle observed in
tomography studies may correspond to a temperature difference of several hundreds of degrees [Van der Lee and Nolet,
1997; Schmid et al., 2001]. Thermomechanical models of
slab detachment also confirm that the temperature may
increase over 500C above the lithospheric gap filled with
upwelling hot asthenosphere [van de Zedde and Wortel,
2001]. Thus the abrupt removal of the slab via its detachment may have increased the temperature at the base of the
crust of an amount analogous to that of a mantle plume.
Unlike a mantle plume, however, this advective-type source
of heat has a transient nature and should produce only a
temporal episode of melting [van de Zedde and Wortel,
2001], which is what we observed in the southern SMO.
4.4. What Caused Deformation in the Southern SMO?
[39] As outlined in section 3.1, two episodes of extension
may be recognized in the southern SMO. The first one
produced regularly spaced grabens between 23 and
20 Ma, whereas the second affected the western part of the
study region approximately between 15 and 11 Ma. In both
cases the magnitude of extension was moderate and below
10– 15%. In addition, left-lateral transpression affected the
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 19
Figure 12. Generalized reconstruction of Pacific-North America plate interaction and volcanic and
tectonic events in the southern SMO during the Miocene. (a) Reconstruction of plate boundaries and slab
windows at 24 Ma (adapted from Nicholson et al. [1994] and Atwater and Stock [1998]) showing the
proposed zone of slab breakoff, which is considered to have caused the 23 and 21 Ma episodes of
silicic volcanism. Abbreviations are MP, Morro microplate; AP, Arguello microplate. (b) Plate
configuration at 16 Ma [from Nicholson et al., 1994]. A second slab detachment event (which possible
location is indicated) is postulated to have occurred between 15 and 11 Ma concurrently with the
waning of the subduction of the Guadalupe and Magdalena plates. This, in turn, is considered to have
produced widespread extension in the southern SMO and at the site of the future Gulf of California. (c)
Plate configuration at the end of subduction of the Magdalena plate (12.5 Ma) (partly based on Stock and
Lee [1994]). Recognizable magnetic anomaly pattern on the Pacific plate (PAC) are also shown [from
Atwater, 1989]. The Jalisco block (JB) is thought to have moved toward the ESE shortly before this time
producing a left-lateral tranpressional deformation along the southern margin of the SMO. Note the
clockwise reorientation of magnetic anomaly on the Pacific plate north of the 22300N fracture zone
during the last period of spreading.
southern end of the SMO between 17 and 11 Ma.
Although more detailed structural and geochronologic studies are needed to resolve the fine time and space evolution
of these episodes, the general geometry, kinematics, and age
data summarized in this work may be used to make a
preliminary assessment of the causes that generated them.
[40] The first episode is bracketed by the two peaks of
ignimbrite flare-up at 23 and 21 Ma. The orientation of
the grabens produced during this episode defines a sort of
fan from the N15E trend of the easternmost graben
(Juchipila) to the N-S trend of the westernmost one (La
Ventana graben) (Figures 2 and 14). The area affected by
17 - 20
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Figure 13. Schematic cross section of the two slab detachment event postulated in this work that
produced the early Miocene episode of ignimbrite flare up (SMO) and the late Miocene mafic volcanism
at the eastern side of the Gulf of California (GCMV). Abbreviations are Farallon, Farallon plate; NOAM,
North America plate; M, Magdalena microplate; BC, Baja California.
this episode has also the highest topography within the
southern SMO, with elevation ranging from 2400 to 3000 m
(Figures 4 and 14). Moreover, elevation contours are roughly
perpendicular to the graben orientation (Figure 14). Considering these facts, we speculate that the first episode was
driven by magmatic intrusion into the crust and/or removal of
the mantle lithosphere. The volume of granitic magma that
remains in the crust is often 4 – 5 times the volume of silicic
rocks emplaced at surface [Crisp, 1984]. However, part of
this volume was already in place since we propose that a
considerable amount of the crust was melted. It is reasonable
to think, therefore, that material addition due to intrusion in
the crust was about equivalent to the extruded material.
Recent studies [Cruden, 1998; Améglio and Vigneresse,
1999; Grocott et al., 1999] showed that granitic batholiths
frequently grow by lateral intrusion at the ductile/brittle
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
17 - 21
Figure 14. Topography and structures of the southern SMO. The NW trending elliptical uplift and the
fanning graben broadly coincide with the early Miocene ignimbrite successions and are thought to be due
to magmatic intrusion at the ductile/brittle transition in the crust.
transition zone by depressing the floor and uplifting the roof.
Considering this mechanism one possibility is that the
regional topography of part of the southern SMO and the
fanning grabens of the eastern extensional domain are
related to the intrusion of a tabular granitic batholith that
fed the 23 and, particularly, the 21 Ma ignimbrite pulses.
In addition, another mechanism capable of producing the
observed topography and extension is the thermal removal
of the mantle lithosphere beneath the southwestern SMO. In
fact the material rising at the base of the plate may have
replaced the cold mantle lithosphere with hotter asthenosphere, resulting in a doming and stretching of the crust
(Figure 13a). The two mechanisms, however, may have
worked concurrently.
[41] The second extensional episode overlaps in time
with the last subduction of the Guadalupe and Magdalena
microplates and the initial transfer of Baja California to the
Pacific plate (Figure 12b). Indeed, plate tectonic reconstructions indicate that subduction of the Farallon plate ended
along the southern half of Baja California at 12.5 Ma,
when the Peninsula and a short slab fragment stalled
beneath it began to be captured by the Pacific plate
[Atwater, 1989; Lonsdale, 1991; Atwater and Stock,
1998]. We agree with Henry and Aranda-Gomez [2000]
in considering this episode of extension as a general
manifestation of the ‘‘proto-Gulf extension,’’ driven by
the capture of Baja California by the Pacific plate. However,
Henry and Fredrikson [1987] and the data presented in this
work suggest the possibility that extension may have
initiated some time before the end of spreading between
the Magdalena and Pacific plates. Indeed the component of
active subduction of the Magdalena microplate beneath
North America in the few million years before the end of
spreading was likely small or close to zero. That is because
after 15 Ma the ridge started to rotate clockwise tending to
be perpendicular to the absolute motion of the Pacific plate
and oblique to the trench. This implies that the spreading
was driven by the Pacific motion rather than by slab pull.
We propose that trench-normal extension observed in the
western extensional domain may have initiated in response
to the waning of subduction of the Magdalena microplate in
middle Miocene time that, in turn, triggered a second slab
detachment event (Figures 12b and 13b). This mechanism
would have produced also a partial melting of the mantle
below the present Gulf of California. In this case, however,
the extensional regime induced by plate boundary forces
allowed rapid upraise of mafic melts (12 –10 Ma mafic
dikes and flows) rather than the formation of silicic crustal
melts (Figure 13b).
[42] The transpressional deformation observed at the
southern boundary of the SMO has no trivial explanation.
The plate configuration west of Baja California for that time
is not precisely defined because the amount of right-lateral
motion along the Tosco-Abreojos fault zone (Figure 12c) is
uncertain [e.g., Fletcher, 2001]. However, it is reasonable to
think that the middle Miocene plate boundary between the
17 - 22
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Magdalena and Cocos plates was located in front of the
mouth of the Gulf [Lyle and Ness, 1991; Stock and Lee,
1994]. In this case the transpressional deformation zone
would be located along the inland projection of the Magdalena-Cocos plate boundary; it also corresponds to the boundary between the SMO and the Jalisco block (Figure 12c).
These observations led Ferrari [1995] to relate this event to
an east-southeastward motion of the Jalisco block shortly
before the end of subduction of the Magdalena microplate. In
addition, we note that the magnetic anomaly pattern preserved on the western side of the East Pacific Rise indicates a
progressive clockwise reorientation of the spreading ridge
between the Magdalena and the Pacific plates between 14.1
and 12.9 Ma (Figure 12c). The contrast between the waning
subduction of the Magdalena microplate and the steady
subduction of the Cocos plate added to the small convergent direction between them at 14– 12.9 Ma may have
produced the left-lateral transpressional deformation
observed along the inland continuation of their subducted
boundary.
5. Concluding Remarks
[43] Massive generation of silicic volcanism occurred in
the southern SMO during two episodes at 31.5 –28 Ma and
23.5 – 20 Ma. Widespread but moderate extension affected
the upper crust during two periods: 23– 20 Ma in the eastern
part and 15– 10 Ma in the western part. The southern end
of the SMO underwent left-lateral transpression roughly at
the same time. We have linked these magmatic and tectonic
events to the evolution and dynamics of the plate motion and
the mantle in western North America during the Miocene.
Particularly, we propose that two consecutive detachments
of the subducted Farallon slab in early and middle Miocene
time played an important role in controlling the locus and
timing of volcanism and extension. Clearly, the interaction
among tectonomagmatic events and plate tectonic history
postulated in this paper is somewhat speculative but has a
number of implications that could be tested by further
studies. The underplating of basaltic magmas must have
produced a mafic lower crust that may be detected by
seismic and gravimetric studies. The involvement of the
asthenosphere after the slab detachment must have left a
geochemical and isotope imprint in the early Miocene
volcanism that may be quantified by petrologic studies.
We hope that this work may stimulate further studies that
may unravel the enigma of the Sierra Madre Occidental.
[44] Acknowledgments. This research was supported by grant CONACyT P-0152 T and Instituto de Geologı́a, UNAM, to L. Ferrari. We are
indebted to Consejo de Recursos Minerales and particularly to J. C. Salinas,
Head of Geology and Geochemistry Department, for allowing the use of a
helicopter. Suzan van der Lee kindly provided sections of her tomographic
model NA97 used to sketch Figure 11. Formal reviews by Joann Stock and
Wanda Taylor provided valuable suggestions and improved the clarity of
the original manuscript. We also thank A. Nieto-Samaniego, D. MoránZenteno, S. Alaniz-Alvarez, Fred McDowell, Chris Henry, J. ArandaGómez, J. Urrutia Fucugauchi, and G. Aguirre-Diaz for several useful
discussions at different stages of the research. G. Aguirre-Diaz also helped
in an initial phase of fieldwork. Technical assistance by V. Moreno, A. S.
Rosas, and V. Pérez is acknowledged.
References
Aguirre-Dı́az, G. J., and F. W. McDowell, Nature and
timing of faulting and synextensional magmatism in
the southern Basin and Range, central-eastern Durango, Mexico, Geol. Soc. Am. Bull., 105, 1435 –
1444, 1993.
Albrecht, A., and S. L. Goldstein, Effects of basement
composition and age on silicic magmas across an
accreted terrane-Precambrian crust boundary, Sierra
Madre Occidental, Mexico, J. S. Am. Earth Sci., 13,
255 – 273, 2000.
Améglio, L., and J. L. Vigneresse, in Understanding
Granites: Integrating New and Classical Techniques,
edited by A. Castro, C. Fernandez, and J. L. Vigneresse, Geol. Soc. Spec. Publ., 158, 39 – 54, 1999.
Angelier, J., Inversion of field data in fault tectonics to
obtain the regional stress, III, A new rapid direction
inversion method by analytical means, J. Geophys.,
103, 363 – 376, 1990.
Atwater, T., Plate tectonic history of northeast Pacific
and western North America, in The Geology of
North America, vol. N, The Eastern Pacific Ocean
and Hawaii, edited by E. L. Winterer, D. M. Hussong, and R. W. Decker, pp. 21 – 71, Geol. Soc. of
Am., Boulder, Colo., 1989.
Atwater, T., and J. M. Stock, Pacific-North America plate
tectonics of the Neogene southwestern United States:
An update, Int. Geol. Rev., 40, 375 – 402, 1998.
Bryant, S. E., A. Ewart, C. J. Stephens, J. Parianos, and
P. J. Downes, The Witsunday volcanic province,
central Queensland, Australia: Lithological and stratigraphic investigations of a silicic-dominated large
igneous province, J. Volcanol. Geotherm. Res., 99,
55 – 78, 2000.
Cameron, M., W. C. Bagby, and K. L. Cameron, Petrogenesis of voluminous mid-Tertiary ignimbrites
of the Sierra Madre Occidental, Chihuahua, Mexico, Contrib. Mineral. Petrol., 74, 271 – 284, 1980.
Cashman, P. H., and M. A. Ellis, Fault interaction may
generate multiple slip vectors on a single fault surface, Geology, 22, 1123 – 1126, 1994.
Castillo-Hernandez, D., and F. Romero-Rios, Estudio
geológico-regional de Los Altos, Jalisco y El Bajio,
Open File Rep., 02-91, 35 pp., Com. Fed. de Electr.,
Gerencia de Proy. Geotermoelectr., Dep. Explor.,
Morelia, Mich., Mexico, 1991.
Christiansen, R. L., Yellowstone magmatic evolution:
Its bearing on understanding large-volume explosive volcanism, in Explosive Volcanism: Inception,
Evolution, and Hazards, pp. 84 – 95, Nat. Acad.
Press, Washington, D. C., 1984.
Clark, K. F., P. E. Damon, M. Shafiquillah, B. F. Ponce,
and D. Cardenas, Sección geológica-estructural a
través de la parte sur de la Sierra Madre Occidental,
entre Fresnillo y la costa de Nayarit, Mem. Téc., XIV,
pp. 69 – 99, Asoc. Ing. Mineros, Metal. y Geól. de
México, Mexico City, 1981.
Coney, P. J., Mesozoic-Cenozoic Cordilleran plate tectonics, in Cenozoic Tectonics and Regional Geophysics of the Western Cordillera, edited by R. B.
Smith and G. P. Eaton, Mem. Geol. Soc. Am., 152,
33 – 50, 1978.
Crisp, J. A., Rate of magma emplacement and volcanic
output, J. Volcanol. Geotherm. Res., 20, 177 – 211,
1984.
Cruden, A. R., On the emplacement of tabular granites,
J. Geol. Soc. London, 155, 853 – 862, 1998.
Damon, P. E., J. Nieto-Obregon, and L. Delgado-Argote, Un plegamiento neogénico en Nayarit y Jalisco y evolución geomórfica del Rio Grande de
Santiago, Mem. Téc., XIII, pp. 156 – 191, Asoc.
Ing. Mineros, Metal. y Geól. de México, 1979.
Ferrari, L., Miocene shearing along the northern boundary of the Jalisco block and the opening of the
southern Gulf of California, Geology, 23, 751 –
754, 1995.
Ferrari, L., S. A. Nelson, J. Rosas-Elguera, and G. J.
Aguirre-Dı́az, Tectonics and volcanism of the western Mexican Volcanic Belt, in Magmatism and Tectonics in Central and Northwestern Mexico.A
selection of the 1997 IAVCEI General Assembly
excursions, edited by G. J. Aguirre-Diaz, pp. 4 –
25, Inst. de Geol., Univ. Nac. Autón. de Méx., Mexico City, 1997.
Ferrari, L., M. Lopez-Martinez, G. Aguirre-Diaz, and
G. Carrasco-Nuñez, Space-time patterns of Cenozoic arc volcanism in central Mexico: From the
Sierra Madre Occidental to the Mexican Volcanic
Belt, Geology, 27, 303 – 307, 1999.
Ferrari, L., G. Pasquaré, S. Venegas-Salgado, and F.
Romero-Rios, Geology of the western Mexican Volcanic Belt and adjacent Sierra Madre Occidental
and Jalisco block, Geol. Soc. Am. Spec. Pap., 334,
65 – 84, 2000a.
Ferrari, L., S. Conticelli, G. Vaggelli, C. Petrone, and P.
Manetti, Late Miocene mafic volcanism and intraarc tectonics during the early development of the
Trans-Mexican Volcanic Belt, Tectonophysics,
318, 161 – 185, 2000b.
Fletcher, J. M., The Baja California Borderland and the
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Neogene evolution of the Pacific-North American
plate boundary, Eos Trans. American Geophysical
Union, 82(47), Fall Meet. Suppl., Abstract T52C0966, 2001.
Fletcher, J. M., B. P. Kohn, D. A. Foster, and A. J. W.
Gleadow, Heterogeneous Neogene cooling and exhumation of the Los Cabos block, southern Baja
California: Evidence from fission-track thermochronology, Geology, 28, 107 – 110, 2000.
Francis, P. W., and C. J. Hawkesworth, Late Cenozoic
rates of magmatic activity in the Central Andes and
their relationships to continental crust formation and
thickening, J. Geol. Soc. London, 151, 845 – 854,
1994.
Gans, P. B., Large-magnitude Oligo-Miocene extension
in southern Sonora: Implications for the tectonic
evolution of northwest Mexico, Tectonics, 16,
388 – 408, 1997.
Gastil, R. G., R. P. Phillips, and E. C. Allison, Reconnaissance geology of the state of Baja California,
Mem. Geol. Soc. Am., 140, 170 pp., 1975.
Gastil, R. G., D. Krummenacher, and J. Minch, The
record of Cenozoic volcanism around the Gulf of
California, Geol. Soc. Am. Bull., 90, 839 – 857,
1979.
Grocott, J., A. Garde, B. Chadwick, A. R. Cruden, and
C. Swager, Emplacement of Rapakiwi granite and
syenite by floor depression and roof uplift in the
Paleoproterozoic Ketilidian orogen, South Greenland, J. Geol. Soc. London, 156, 15 – 24, 1999.
Groshong, R. H., Area balance, depth of detachment,
and strain in extension, Tectonics, 13, 1488 – 1497,
1994.
Harry, D. L., and W. P. Leeman, Partial melting of melt
metasomatized subcontinental mantle and the magma source potential of the lower lithosphere, J.
Geophys. Res., 100, 10,255 – 10,269, 1995.
Hausback, B. P., Cenozoic volcanic and tectonic evolution of Baja California Sur, México, in Geology of
Baja California Peninsula, vol. 199, edited by V. A.
Frizzell, pp. 219 – 236, Soc. for Sediment., Pac.
Sec., Geol., Tulsa, Okla., 1984.
Henry, C. D., and J. J. Aranda-Gomez, The real southern Basin and Range: Mid- to late-Cenozoic extension in Mexico, Geology, 20, 701 – 704, 1992.
Henry, C., and J. J. Aranda-Gomez, Plate interactions
control middle-late Miocene, proto-Gulf and Basin
and Range extension in the southern Basin and
Range, Tectonophysics, 218, 1 – 26, 2000.
Henry, C. D., and G. Fredrikson, Geology of part of
southern Sinaloa, Mexico, adjacent to the Gulf of
California, Geol. Soc. Am. Maps Chart Ser., MCH
063, 1 sheet, 14 pp., Geol. Soc. of Am, Boulder,
Colo., 1987.
Houghton, B. F., C. J. N. Wilson, M. O. McWilliams,
M. A. Lanphere, D. Weaver, R. M. Briggs, and M.
S. Pringle, Chronology and dynamics of a large
silicic magmatic system: Central Taupo Volcanic
Zone, New Zealand, Geology, 23, 13 – 16, 1995.
Humphreys, E. D., Post-Laramide removal of the Farallon slab, western United States, Geology, 23,
987 – 990, 1995.
Huppert, H. E., and R. S. J. Sparks, The generation of
granitic magmas by intrusion of basalt into continental crust, J. Petrol., 29, 599 – 642, 1988.
Kay, R. W., and S. Kay, Delamination and delamination
magmatism, Tectonophysics, 219, 177 – 189, 1993.
Lonsdale, P., Structural pattern of the pacific floor offshore peninsular California, in The Gulf and the
Peninsular Province of the Californias, edited by
J. P. Dauphin and B. R. T. Simoneit, AAPG Mem.,
47, 87 – 125, 1991.
Ludwig, K. R., Isoplot/Ex 2.01: A geochronological
toolkit for Microsoft Excel, Berkeley Geochronological Cent. Spec. Publ. 1a, 50 pp., 1999.
Luhr, J., C. D. Henry, T. B. Housh, J. Aranda-Gómez,
and W. C. McIntosh, Early extension and associated
mafic alkalic volcanism from the southern Basin
and Range province: Geology and petrology of
the Rodeo and Nazas volcanic fields, Durango,
Mexico, Geol. Soc. Am. Bull., 113, 760 – 773, 2001.
Lyle, M., and G. Ness, The opening of the Southern
Gulf of California, in The Gulf and Peninsular Provinces of the Californias, edited by J. P. Dauphin
and B. R. T. Simoneit, AAPG Mem., 47, 403 – 423,
1991.
Lyons, J. I., Geology and ores deposits of the Bolaños
silver district, Jalisco, Mexico, Econ. Geol., 83,
1560 – 1582, 1988.
McDowell, F. W., and S. E. Clabaugh, Ignimbrites of
the Sierra Madre Occidental and their relation to the
tectonic history of western Mexico, Geol. Soc. Am.
Spec. Pap., 180, 113 – 124, 1979.
McDowell, F. W., and R. P. Keizer, Timing of midTertiary volcanism in the Sierra Madre Occidental
between Durango city and Mazatlan, Mexico, Geol.
Soc. Am. Bull., 88, 1479 – 1487, 1977.
McDowell, F. W., J. Roldán-Quintana, and R. AmayaMartı́nez, Interrelationship of sedimentary and volcanic deposits associated with Tertiary extension in
Sonora, Mexico, Geol. Soc. Am. Bull., 109, 1349 –
1360, 1997.
Moore, G., C. Marone, I. S. E. Carmichael, and P. Renne, Basaltic volcanism and extension near the intersection of the Sierra Madre volcanic province and
the Mexican Volcanic Belt, Geol. Soc. Am. Bull.,
106, 383 – 394, 1994.
Morán-Zenteno, D. J., et al., Tertiary arc-magmatism of
the Sierra Madre del Sur, Mexico, and its transition
to the volcanic activity of the Trans-Mexican Volcanic Belt, J. S. Am. Earth Sci., 12, 513 – 535, 1999.
Nicholson, C., C. C. Sorlien, T. Atwater, J. C. Crowell,
and B. P. Luyendyk, Microplate capture, rotation of
the western Transverse Ranges, and initiation of the
San Andreas transform as a low-angle fault system,
Geology, 22, 491 – 495, 1994.
Nieto-Obregon, J., L. Delgado-Argote, and P. E. Damon, Relaciones petrológicas y geocronológicas
del magmatismo de la Sierra Madre Occidental y
el Eje Neovolcánico en Nayarit, Jalisco y Zacatecas,
Mem. Téc., XIV, pp. 327 – 361, Asoc. Ing. Mineros,
Metal. y Geól. de México, CITY, 1981.
Nieto-Obregon, J., L. Delgado-Argote, and P. E. Damon, Geochronologic, petrologic and structural data
related to large morphologic features between the
Sierra Madre Occidental and the Mexican Volcanic
Belt, Geofis. Int., 24, 623 – 663, 1985.
Nieto-Samaniego, A. F., and S. A. Alaniz-Alvarez, Origin and tectonic interpretation of multiple fault patterns, Tectonophysics, 270, 197 – 206, 1997.
Nieto-Samaniego, A. F., C. Macı́as-Romo, and S. A.
Alaniz-Alvarez, Nuevas edades isotópicas de la cubierta volcánica cenozoica de la parte meridional de
la Mesa Central, México, Rev. Mex. Cienc. Geol.,
13, 117 – 122, 1996.
Nieto-Samaniego, A. F., L. Ferrari, S. A. Alaniz-Alvarez, G. Labarthe-Hernandez, and J. Rosas-Elguera,
Variation in Cenozoic extension and volcanism
across the southern Sierra Madre Occidental volcanic province, western Mexico, Geol. Soc. Am. Bull.,
111, 347 – 363, 1999.
Nourse, J. A., T. H. Anderson, and L. T. Silver, Tertiary
metamorphic core complexes in Sonora, northwestern Mexico, Tectonics, 13, 359 – 366, 1994.
Pankhurst, R. J., T. R. Riley, C. M. Fanning, and S. P.
Kelley, Episodic silicic volcanism in Patagonia and
the Antarctic Peninsula: Chronology of magmatism
associated with the break-up of Gondwana, J. Petrol., 41, 605 – 625, 2000.
Pasquarè, G., L. Ferrari, V. H. Garduño, A. Tibaldi, and
L. Vezzoli, Geological map of the central sector of
Mexican Volcanic Belt, States of Guanajuato and
Michoacan, Geol. Soc. Am. Map Chart Ser., MCH
072, 1 sheet, 20 pp., Geol. Soc. of Am., Boulder,
Colo., 1991.
Pollard, D. D., S. D. Saltzer, and A. Rubin, Stress inversion methods: are they based on faulty assumptions?, J. Struct. Geol., 15, 1045 – 1054, 1993.
Righter, K., I. S. E. Carmichael, and T. Becker, PlioceneQuaternary faulting and volcanism at the intersection
of the Gulf of California and the Mexican Volcanic
Belt, Geol. Soc. Am. Bull., 107, 612 – 626, 1995.
Rodrı́guez-Castañeda, J. L., and R. Rodrı́guez-Torres,
Geologı́a estructural y estratigrafı́a del área entre
17 - 23
Guadalajara y Tepic, Estados de Jalisco y Nayarit,
México, Rev. Inst. Geol. UNAM, 10, 99 – 110, 1992.
Rosas-Elguera, J., L. Ferrari, M. López-Martinez, and J.
Urrutia-Fucugauchi, Stratigraphy and tectonics of
the Guadalajara region and triple-junction area,
western Mexico, Int. Geol. Rev., 39, 125 – 140,
1997.
Rossotti, A., L. Ferrari, M. Martı́nez-López, and J. Rosas-Elguera, Geology of the boundary between the
Sierra Madre Occidental and the Trans-Mexican
Volcanic Belt in the Guadalajara region, western
Mexico, Rev. Mex. Cienc. Geol., 19, 1 – 15, 2002.
Ruiz, J., P. J. Patchett, and R. J. Arculus, Nd-Sr isotopic
composition of lower crustal xenoliths.Evidence for
the origin of mid-Tertiary felsic volcanics in Mexico,
Contrib. Mineral. Petrol., 99, 36 – 43, 1988.
Ruiz, J., P. J. Patchett, and R. J. Arculus, Reply to
"Comments on Nd-Sr isotopic compositions of lower crustal xenoliths: Evidence for the origin of midTertiary felsic volcanics in Mexico," by K. L. Cameron and J. V. Robinson, Contrib. Mineral. Petrol., 104, 615 – 618, 1990.
Schaaf, P., H. Bohnel, and J. A. Perez-Venzor, Pre-Miocene palaeogeography of the Los Cabos Block, Baja
California Sur: Geochronological and palaeomagnetic constraints, Tectonophysics, 318, 53 – 69,
2000.
Scheubel, F. R., K. F. Clark, and E. W. Porter, Geology,
tectonic environment and structural controls in the
San Martin de Bolaños district, Jalisco, Mexico,
Econ. Geol., 83, 1703 – 1720, 1988.
Schmid, C., S. Goes, S. van der Lee, and D. Giardini,
Thermal and seismic signature of the trailing fragments of the Farallon slab, Eos Trans. American
Geophysical Union, 82(47), Fall Meet. Suppl., Abstract F1197-1998, 2001.
Smith, R. D., K. L. Cameron, F. W. McDowell, S.
Nyemeyer, and D. E. Sampson, Generation of voluminous silicic magmas and formation of mid Cenozoic crust beneath north-central Mexico: Evidence
from ignimbrites, associated lavas, deep crustal
granulites, and mantle pyroxenites, Contrib. Mineral. Petrol., 123, 375 – 389, 1996.
Soto, M. A., and J. G. Ortega, Geologia del Rio Santiago en los estados de Jalisco y Nayarit, México,
paper presented at XII Congress of the Sociedad
Geológico de México, Mexico City, 1982.
Stewart, J. H., et al., Map showing Cenozoic tilt domains and associated structural features, western
North America, in Accommodation Zones and
Transfer Zones: The Regional Segmentation of the
Basin and Range Province, edited by J. E. Faulds
and J. H. Stewart, Geol. Soc. Am. Spec. Pap., 323,
1998.
Stock, J. M., and J. Lee, Do microplates in subduction
zones leave a geologic record?, Tectonics, 13,
1472 – 1487, 1994.
Swanson, E. R., and F. W. McDowell, Calderas of the
Sierra Madre Occidental volcanic field, western
Mexico, J. Geophys. Res., 89, 8787 – 8799, 1984.
Twiss, R. J., and J. R. Unruh, Analysis of fault slip
inversions: Do they constrain stress or strain rate?,
J. Geophys. Res., 103, 12,205 – 12,222, 1998.
Umhoefer, P. J., R. Dorsey, S. Willsey, L. Mayer, and P.
Renne, Stratigraphy and geochronology of the Comondú Group near Loreto, Baja California sur,
Mexico, Sediment. Geol., 144, 125 – 147, 2001.
Van der Lee, S., and G. Nolet, Seismic image of the
subducted trailing fragments of the Farallon plate,
Nature, 386, 266 – 269, 1997.
van de Zedde, D. M. A., and M. J. R. Wortel, Shallow
slab detachment as a transient source of heat at
midlithospheric depth, Tectonics, 20, 868 – 882,
2001.
Verma, S. P., Sr and Nd isotopic evidence for petrogenesis of mid-Tertiary felsic volcanism in the mineral
district of Zacatecas, Zac. (Sierra Madre Occidental), Mexico, Isot. Geosci., 2, 37 – 53, 1984.
Wark, D. A., K. A. Kempter, and F. W. McDowell,
Evolution of waning subduction-related magmatism, northern Sierra Madre Occidental, Mexico,
Geol. Soc. Am. Bull., 102, 1555 – 1564, 1990.
17 - 24
FERRARI ET AL.: IGNIMBRITE FLARE-UP AND DEFORMATION, WESTERN MEXICO
Webber, K., L. A. Fernández, and B. Simmons, Geochemistry and mineralogy of the Eocene-Oligocene
volcanic sequence, southern Sierra Madre Occidental, Juchipila, Zacatecas, México, Geofı́s. Int., 33,
77 – 89, 1994.
Wortel, M. J. R., and W. Sparkman, Subduction and
slab detachment in the Mediterranean-Carpathian
region, Science, 290, 1910 – 1917, 2000.
York, D., Least squares fitting of a straight line with
correlated errors, Earth Planet. Sci. Lett., 5, 320 –
324, 1969.
L. Ferrari, Centro de Geociencias, Universidad
Nacional Autónoma de México, Campus Juriquilla,
Qro. Apdo. Postal 1-742 Centro, Querétaro, 76001
Qro., Mexico. ([email protected])
M. López-Martı́nez, Departamento de Geologı́a,
Centro de Investigación Cientı́fica y Educación Superior de Ensenada, Km. 107 Carretera Tijuana-Ensenada,
C. P. 22860, Ensenada, Baja California, México.
([email protected])
J. Rosas-Elguera, Centro de Ciencias de la Tierra,
Universidad de Guadalajara, Calzada Olı́mpica y Blvd.
Marcelino Garcı́a Barragán, 44840 Guadalajara, Jal.,
México. ( [email protected])