You are on page 1of 19

Downloaded from gsabulletin.gsapubs.

org on February 20, 2012

Geological Society of America Bulletin


Arc-rift transition volcanism in the Puertecitos Volcanic Province, northeastern
Baja California, Mexico
Arturo Martn-Barajas, Joann M. Stock, Paul Layer, Brian Hausback, Paul Renne and Margarita
Lpez-Martnez
Geological Society of America Bulletin 1995;107, no. 4;407-424
doi: 10.1130/0016-7606(1995)107<0407:ARTVIT>2.3.CO;2

Email alerting services

click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new articles


cite this article

Subscribe

click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of


America Bulletin

Permission request

click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA

Copyright not claimed on content prepared wholly by U.S. government employees within scope of their
employment. Individual scientists are hereby granted permission, without fees or further requests to GSA,
to use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make
unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and
science. This file may not be posted to any Web site, but authors may post the abstracts only of their
articles on their own or their organization's Web site providing the posting includes a reference to the
article's full citation. GSA provides this and other forums for the presentation of diverse opinions and
positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or political
viewpoint. Opinions presented in this publication do not reflect official positions of the Society.

Notes

Geological Society of America

Downloaded from gsabulletin.gsapubs.org on February 20, 2012

Arc-rift transition volcanism in the Puertecitos Volcanic Province,


northeastern Baja California, Mexico
Arturo Martn-Barajas CICESE, Ensenada, Mexico, P.O. Box 434843, San Diego, California 92143
Joann M. Stock Caltech, Pasadena, California 91125
Paul Layer University of Alaska, Fairbanks, Alaska 99775
Brian Hausback California State University, Sacramento, California 95819
Paul Renne Institute for Human Origins, Berkeley, California 94709
Margarita Lopez-Martnez CICESE, Ensenada, Mexico, P.O. Box 434843, San Diego, California 92143

ABSTRACT
The Neogene Puertecitos Volcanic Province of northeastern Baja California
records a transition from arc-related volcanic activity to rift volcanism associated
with opening of the Gulf of California. The
eastern Puertecitos Volcanic Province is divided into three volcanic sequences based
on mapping, petrology, and 40Ar/39Ar geochronology. The lowest sequence comprises
early to middle Miocene (20 16 Ma) arcrelated andesitic lava flows, volcanic necks,
and proximal pyroclastic and epiclastic deposits up to 400 m in thickness, with minor
basaltic lava flows. Following the initiation
of crustal extension in the region (11 6
Ma), synrift volcanism produced two rhyolitic sequences that discordantly overlie
the arc-related rocks. The older synrift sequence (6.4 5.8 Ma) is composed of rhyolite
domes and a series of pyroclastic flows up to
300 m thick. The upper sequence (3.22.7
Ma) consists of ash-flow tuffs and pumicelapilli pyroclastic flows, collectively up to
200 m thick. Minor andesite eruptions followed each episode of silicic synrift volcanism. Synvolcanic faults produced topographic relief that controlled deposition of
the pyroclastic flows and caused gentler
dips upsection. Rhyolite domes are aligned
parallel to the predominant north-northwest to north-northeast fault pattern.
All three volcanic sequences are calc-alkaline. However, the synrift andesite is
characterized by lower K2O, lower incompatible element concentrations, and less
fractionation of light rare earth elements
than the arc-related basalt and andesite.

This suggests that the primary melts were


more primitive for synrift andesite than for
the arc-related rocks.
INTRODUCTION
Both the geochemistry and the style of volcanism in the Baja California peninsula
changed in the last 10 15 Ma. This was
partly due to the cessation of subduction
along the west coast of Baja California and
partly due to the development of the Gulf of
California rift system (e.g., Hausback, 1984;
Saunders et al., 1987; Sawlan and Smith,
1984; Sawlan, 1991) (Fig. 1).
Postsubduction volcanic rocks have variable chemical compositions. In the central
and southern parts of the peninsula, calcalkaline lavas with moderate to high K2O
evolved to alkaline lavas rich in MgO during
middle and late Miocene time (Sawlan and
Smith, 1984; Saunders et al., 1987). Volcanism related to continental rifting maintained
its mainly calc-alkaline character along the
margins of the Gulf of California (Sawlan,
1991). In contrast, compositions are typically tholeiitic within the oceanic spreading
centers of the gulf itself (Saunders, 1983;
Saunders et al., 1982). Synrift calc-alkaline
volcanic rocks in the northern gulf region
(Fig. 2a) include ca. 9 Ma andesite and dacite in the Sierra Pinta (Gastil et al., 1979),
a 10 12 Ma subalkaline volcanic field in the
Sierra San Alberto (Sawlan and Smith,
1984), and some 1214 Ma ignimbrite deposits in the Baha de Los AngelesSan
Borja area (Gastil et al., 1979; DelgadoArgote et al., 1992). In coastal Sonora, reported ages for the youngest calc-alkaline

Data Repository item 9514 contains additional material related to this article.
GSA Bulletin; April 1995; v. 107; no. 4; p. 407 424; 14 figures; 1 table.

407

rocks are ca. 12 Ma (Gastil and Krummenacher, 1977; Gastil et al., 1979; Mora-Alvarez, 1993). Andesitic to dacitic pyroclastic
flows on Isla Tiburo
n are between 15 and 11
Ma (Neuhaus, 1989). In several places calcalkaline volcanism continued in the gulf depression during the mature stage of the rift
in Pliocene to Recent(?) time. Plio-Pleistocene calc-alkaline volcanic rocks have been
reported in the Tres Vrgenes area, in Isla
San Esteban (Fig. 2a), and possibly in the
Santa Ana and La Reforma calderas (Sawlan, 1981; Desonie, 1992; Demant, 1981).
The Miocene-Pliocene Puertecitos Volcanic Province is situated along the northwestern margin of the Gulf of California
(Figs. 2a and 2b) and was a major locus of
late Neogene calc-alkaline volcanism, dominantly rhyolite ignimbrite and lava. Normal
faulting north and northwest of the Puertecitos Volcanic Province began after 11 Ma
but before 6 Ma and was accompanied by
the emplacement of voluminous rhyolitic
pyroclastic material and lavas (Stock and
Hodges, 1990; Stock, 1989, 1993). In the
eastern and northeastern part of the Puertecitos Volcanic Province, late Miocene to
Pliocene rhyolite lava and ignimbrite overlie
andesitic rocks from the early Miocene volcanic arc (Stock et al., 1991). Near the coast,
Pliocene volcanic and marine rocks are cut
by normal faults, indicating continued extension along the coastal margin of the
Puertecitos Volcanic Province (MartnBarajas and Stock, 1993).
Here we (1) describe the volcanic stratigraphy and geochronology in the northeastern part of the Puertecitos Volcanic Province, (2) document the relationship between

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

Figure 1. Simplified tectonic map of northwest Mexico showing the age of the youngest magnetic anomalies identified on the Pacific
Plate along the California and Baja California margins (modified from Atwater, 1989). Diagonal shading shows possible distribution and
ages of arc-related rocks in southern Baja California (from Sawlan, 1991, and Hausback, 1984). PVP, Puertecitos Volcanic Province; EPR,
East Pacific Rise. Dotted lines show the position of a dead rift, representing the cessation of Guadalupe-Pacific spreading, and coeval
cessation of subduction, at the time of anomaly 5A (12.5 Ma). Note that, since they formed, these sea-floor features have been displaced
relative to Baja California, because of dextral faulting along the western margin of the Peninsula and in the California Continental
Borderland region.
volcanism and extension, (3) show the geochemical and petrologic characteristics of
the volcanic succession, and (4) examine the
evidence for a crustal contribution to the
synrift volcanism.
TECTONIC SETTING
Neogene volcanism in Baja California is
related to the consumption of the FarallonGuadalupe plate beneath North America,
the cessation of subduction beneath the continent, and the transition to Gulf of California rifting and strike-slip motion. The Farallon-Pacific ridge first contacted the North
American plate at some location north of
the Baja California peninsula, not earlier
than 29 Ma (Atwater, 1989; Stock and Molnar, 1988). Subsequently, volcanic activity
decreased from north to south as subduction
stopped (Fig. 1). In northern Baja California, arc-volcanism apparently ended at
1714 Ma, in contrast to southern Baja
California, where it finished at ca. 12 Ma
(Sawlan and Smith, 1984; Hausback, 1984;
Sawlan, 1991). Calc-alkaline volcanism continued, however, in some areas of the gulf
depression during the development of the
Gulf of California rift system (Fig. 2a). A
genetic association has been suggested between extensional and transtensional tectonics and explosive calc-alkaline volcanism,

408

due to melting of the continental crust


(Gastil et al., 1975; Sawlan, 1991).
The Puertecitos Volcanic Province is cut
by north-northwest to north-northeaststriking high-angle normal faults (Fig. 2b) that
are both synthetic and antithetic to the
main gulf escarpment (Dokka and Merriam,
1982). North of the Puertecitos Volcanic
Province, the escarpment is defined by the
north-northweststriking San Pedro Martr
fault, whereas to the south its topographic
expression is less dramatic and consists of a
series of smaller normal faults that strike
subparallel to the escarpment. It is clear locally that these faults offset pre 6 Ma rocks
more than post 6 Ma rocks, so that possibly
about half of the extensional faulting occurred prior to 6 Ma (Stock and Hodges,
1990). Toward the east, volcanic and sedimentary rocks of Plio-Pleistocene age are
commonly tilted, indicating that significant
rift-related extensional faulting was still ongoing in the northeast part of the Puertecitos Volcanic Province into Pliocene and
possibly Quaternary time.
METHODS USED IN THIS STUDY
Our terminology for volcanic and volcaniclastic units follows Fisher (1961), Cas and
Wright (1987), and Bard (1986). We first establish a descriptive classification (Fisher,

1961), and only in those cases when the petrologic characteristics and the field relationships define it is a genetic classification proposed (Cas and Wright, 1987).
Analytical Techniques
Whole-rock major-oxide analyses were
obtained by X-ray fluorescence (XRF) at
San Diego State University (SDSU) and at
the University of Massachusetts, Amherst
(UMass). Standard andesite from Great
Valley (AGV) was used for major-oxide
analyses at SDSU. Most trace-element analyses were made using XRF at UMass.
Selected samples were analyzed for additional trace elements by inductively coupled
plasma mass spectrometry (ICP-MS) at
Washington State University. The detection
limits in the ICP-MS analyses were 0.51.0
times chondrite values, and reported values
are at 10% confidence level. Standards were
BCR-1 and in-house standards (Charles
Knack, 1992, written commun.). Additional
XRF analyses for trace elements were performed by K. Righter (Berkeley, California)
to verify ICP-MS results. Microprobe analysis was done at Harvard University on a
Cameca probe.
40
Ar/39Ar dates were done by step heating
at the Geochronology Laboratory, University of Alaska, or by laser fusion at the In-

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

Figure 2. (a) Index map of northwest Mexico showing the areas with early-rift calc-alkaline volcanism: SP, Sierra Pinta (Gastil et al.,
1979); SRB, Santa Rosa Basin (Bryant, 1984); PVP, Puertecitos Volcanic Province (this study); BA, Bahia de Los Angeles (Delgado-Argote
et al., 1992); SB, San Borja (Sawlan, 1991); IT, Isla Tiburon (Neuhaus, 1989); CS, Central Sonora (Gastil et al., 1979); 3V, Volca
n Tres
Vrgenes (Sawlan, 1981, 1991); SA, Sierra San Alberto (Sawlan, 1991); SU, Sierra Santa Ursula (Mora-Alvarez, 1993); SR, Santa Rosala;
LR, La Reforma and Santa Ana Calderas; SE, Isla San Esteban (Desonie, 1992); CP, Cerro Prieto (tholeiitic; Herzig, 1990). Modified from
Sawlan (1991).
(b) Simplified geologic map of the Puertecitos Volcanic Province in northeast Baja California, showing location of the studied area.
SPMF, San Pedro Martr Fault; MGS, Main Gulf Escarpment. Geologic background modified from Gastil et al. (1975); geologic map at
1:250 000 scale.

stitute for Human Origins (IHO) at


Berkeley (Tables 1a and 1b). Analyses at
IHO generally employed the methods and
facilities of Deino and Potts (1990). Isotopic methods are described in Appendix 1, and data tables are in the GSA Data
Repository.1

1
GSA Data Repository item 9514, Tables 2, 3,
4, and 5 and Appendices 1 and 2, is available on
request from Documents Secretary, GSA, P.O.
Box 9140, Boulder, CO 80301.

VOLCANIC STRATIGRAPHY AND


GEOCHRONOLOGY
Based on petrologic characteristics, stratigraphic position, and isotopic ages, we divide the volcanic rocks into three groups
(Fig. 3). Group 1 comprises rocks of arc origin: early and middle Miocene andesitic
and dacitic rocks with minor basalt (Tma).
Group 2, Miocene rhyolite (Tmr), consists
of rhyolite domes, flows (Tmrl and Tmru),
and ash flows of the Tuff of El Canelo (Tmc)
that lie between rhyolite flows. Group 2 also

includes local andesitic flows (Tba). Group


3 is a series of pyroclastic flows of variable
thickness (Pliocene rhyolite, Tpr) and also
includes an andesitic volcano and young
andesitic lava flows (Tpa).
Group 1: Arc-Related Andesite
Miocene arc rocks of Group 1 are well
exposed in Arroyo Los Heme (Fig. 3), where
outcrops of Tma form a volcanic center ;4
km in diameter. The volcanic necks show a
pronounced paleorelief of up to 400 m. Co-

Geological Society of America Bulletin, April 1995

409

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.
TABLE 1. ISOTOPIC AGES OF SELECTED VOLCANIC UNITS IN THE NORTHEASTERN PUERTECITOS
VOLCANIC PROVINCE
(a)

40

Ar/39Ar ages by step heating

ID number
1
2
3
4
5
6
7
8

Unit

Subunit

ti (Ma)*

tp (Ma)

%39Ar

Tma
Tma
Tmr
Tba
Tba
Tph
Tph
Tph
Tph
Tpr
Tpr
Tpa

LH
LH
Tmru, VC
VC
VC
Tph 28 (?), LH
Tph 28 (?), LH
Tph 28 (?), LH
Tph 28 (?), LH
Tpt-b, VC
Tpt-b, VC
VP

16.1 6 0.3
16.8 6 0.2
6.02 6 0.04
4.7 6 0.6
5.2 6 0.2
2 6 0.1
2.5 6 0.6
3.29 6 0.04
5.6 6 0.1
3.9 6 0.1
3.99 6 0.06
2.6 6 0.1

15.9 6 0.2**
16.3 6 0.1
5.80 6 0.03

62
66
68

5.1 6 0.3

40

2.7 6 0.4
3.36 6 0.03

74
91

3.5 6 0.1**
3.39 6 0.06**
2.6 6 0.1

92
56
81

Sample

Unit

Subunit

Age (Ma)

(40Ar/36Ar)i

EST-55 (san)
12-I-91 (plag)
MT-91-41 (plag)
EST-65-1 (plag)

Tmc
Tpr
Tph
Tpr

Tmc1, LC
Tpvc, VC
Tph 28, LH
Ttg (Tpt-b), SF

6.44 6 0.02
3.27 6 0.04
2.65 6 0.05
3.08 6 0.04

Sample
II-15 (hbl)
21-III-6 (hbl)
22-III-1 (wr)
30-IV-4 (wr)
30-IV-4 (wr)
I-16 (wr)
I-16 (plag)
II-13 (wr)
II-13 (plag)
III-6 (wr)
III-6 (wr)
VP-1 (wr)

(b) 40Ar/39Ar ages by laser fusion


ID number
9
10
11
12

291.6 6 4.9
298.7 6 17.4
299.7 6 5.0
312.8 6 11.2

MSWD

1.01
1.13
0.69
1.77

20
17
20
19

Abbreviations: plag, plagioclase; wr, whole rock; san, sanidine; hbl, hornblende.
Note: Samples 1 8 were analyzed at the Geochronology Laboratory, University of Alaska at Fairbanks. Samples 9 12 were
analyzed at the Geochronology Center of the Institute of Human Origins. See Appendix 1 for details. Complete analytical results
are in the Data Repository.
*Integrated age.

Plateau age.

Percentage of 39Ar released in the plateau segments.


**Preferred ages consisting of only one fraction; see text for details.

Isochron age.

Number of points fitted.

eval pyroclastic flow deposits and epiclastic


deposits crop out near the volcanic necks
and are interstratified with lava flows. The
pyroclastic rocks consist primarily of blocks
and lithic lapilli, and the epiclastic deposits
consist of matrix-supported bedded conglomeratic breccia and gravel- to sand-sized
deposits. Suspended particles are up to 20
cm in diameter, and the matrix is composed
of ash and lithic lapilli. The breccias are interpreted as proximal avalanche and debrisflow deposits.
In Arroyo Los Heme the arc sequence includes a cinder cone deposit interstratified
with the epiclastic deposits. The cinder cone
deposit consists of clast-supported scoria
beds centimeters to decimeters thick. An andesite dike cuts and a sill overlies the cinder
cone deposit, which is clearly coeval with the
orogenic andesite rocks.
Hornblende from two samples of a volcanic neck (sample 2) and a breccia (sample
1) yielded very similar age spectra (Fig. 4a),
but the bulk of the 39Ar was released in one
or two fractions. However, the agreement
between the ages together with the homogeneous 37ArCa/39ArK ratios for fractions
with the bulk of the 39Ar released, support
the 16 Ma age as a good estimate for the
closure age for the hornblendes (Table 1a).

410

Group 2: Synrift Rhyolite to Andesite


Lava and Tuffs
Synrift Rhyolite Lavas. Rhyolite domes
crop out over a distance of 10 km in the
northern region (Fig. 3). Along Arroyo La
Cantera, the northernmost rhyolite dome
complex is 3 km in east-west extent. A series
of glassy flows of unknown age is the lowest
unit (Tmrl), which underlies the Tuff of El
Canelo (Tmc), which pinches out against the
southern flank of the dome. Bedded lithiclapilli deposits interpreted as airfall deposits
locally overlie the lower rhyolite flows. An
upper series of rhyolite flows (Tmru) overlies the Tuff of El Canelo east and southeast
of Arroyo La Cantera.
One whole-rock sample of the upper
rhyolite lava (Tmru) in Valle Curbina
(sample 3, Table 1a) was dated by the stepheating technique, yielding a two-step plateau with an age of 5.80 6 0.03 Ma. Additionally, we attempted to determine the
age of the upper rhyolite lava by dating an
andesitic dike that cuts it in Valle Curbina.
A whole-rock, step-heating analysis was
run twice (sample 4, Table 1a). A perturbed age spectrum was obtained on both
experiments (see Fig. 4c). Two fractions,
representing 40% of the 39Ar released,

yielded a 5.1 6 0.3 Ma age, which may be


taken as a lower age estimate for the emplacement of the dike. Due to the large
uncertainty, however, these data do not
provide additional constraints on the age
of the rhyolite sequence. The time of inception of the synrift silicic volcanism is
not yet constrained, but the lower rhyolite
flows may be partially coeval or close in
age to the Tuff of El Canelo discussed below. This is suggested by the air-fall deposit, likely related to the lower rhyolite
dome, that overlies the lowermost unit of
the Tuff of El Canelo (Table 2 in the Data
Repository). This deposit suggests that the
emplacement of the rhyolite dome is
partly coeval with the lower units of the
Tuff of El Canelo.
Rhyolite lavas are volumetrically important elsewhere in the northern Puertecitos
Volcanic Province and, together with outcrops in the study area, may form a latest
Miocene rhyolite dome field that now
reaches more than 40 km in a north-south
direction and 15 km in an east-west direction. Other rhyolitic lavas crop out in the
southern Sierra San Fermin (Fig. 3), north
and northwest of Cerro Canelo, and both
north and south of Arroyo Matom(Fig. 2b).
Southwest of Puertecitos, the upper rhyolite
flows crop out near the base of some arroyos, and southward, along the coastal
zone, both the rhyolite and the Tuff of El
Canelo pinch out.
Small andesite flows and dikes overlie the
rhyolite lava and the Tuff of El Canelo. In
the Arroyo Los Heme area, a small-volume
basaltic andesite lava flow (Tba) overlies
epiclastic deposits derived from the arc
rocks along an erosional unconformity
(Fig. 3). Additionally, the andesite flow in
Arroyo Los Heme underlies a heterolithologic conglomerate that contains beds a few

Figure 3. Simplified geologic map of the


studied area in the northeastern Puertecitos Volcanic Province showing the main
structural grain. Shaded units are andesitic
units of the three main volcanic groups. Site
A, measured section of Tmec (Table 2 in the
Data Repository). Site A*, measured section
of the Tuff of Mesa El Ta
bano (Group 3, in
composite stratigraphic column Mesa El
Ta
banoArroyo La Cantera of Fig. 5). Sites
B and B* measured sections of Tph12
Tph29 and Tph1Tph11, respectively (see
Table 3 in Data Repository).

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012

Geological Society of America Bulletin, April 1995

411

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

Figure 4. Step heating 40Ar/39Ar versus heat diagrams showing plateau ages for selected samples analyzed at the Geochronology Laboratory, University of Alaska. ID
sample number as in Table 4 in the Data Repository, and location in Figure 3. Note that
in Figure 4c the progressive increase of apparent age with temperature does not permit
identification of a stable plateau, although a minimum age for the sample would be 4.5
Ma (the age of the first plateau).

decimeters thick of well-sorted pumiceouslapilli. The conglomerate may represent an


erosional unconformity with the units of
Group 3. This andesite flow is interpreted to

412

be coeval with the ca. 5.0 Ma andesite from


Valle Curbina.
Tuff of El Canelo (Tmc). This sequence of
multiple ash flows crops out between Arroyo

La Cantera and the Sierra San Fermn


(Fig. 3), and for some distance west of Cerro
Canelo (Fig. 2b). On Mesa El Tabano and
Mesa La Angostura, the Tuff of El Canelo is

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

overlain in angular unconformity by the upper rhyolite flows and by Pliocene ignimbrites. The Tuff of El Canelo is cut by
north-northweststriking high- and lowangle normal faults that produced strongly
tilted blocks; in the northern part of the
study area, the Tuff of El Canelo is tilted
508708WSW, but south and east, the dip
decreases. At site A of Figure 3, the dip decreases upward, from 258308 at the base to
58108 at the top. No paleosols or abrupt
angular unconformities are observed within
the sequence. In Arroyo La Cantera, the
Tuff of El Canelo rests upon, and pinches
out against, the older rhyolite flows (Tmrl)
and related beds of stratified lithic and pumice lapilli, interpreted as airfall deposits.
In Arroyo La Cantera (site A of Fig. 3),
the Tuff of El Canelo comprises six cooling
units (Tmc1 to Tmc6) with a cumulative
thickness of nearly 300 m (Table 2 in the
Data Repository). The basal unit (Tmc1), an
unwelded crystal-rich pyroclastic flow with a
maximum outcrop thickness of 5 m, directly
overlies and pinches out against the lower
rhyolite flows and bedded airfall deposits
and is locally overlain by up to 1 m of bedded
rhyolite lithic lapilli interpreted as airfall deposits. Units 2 and 4 are thick compound
cooling units (each ;100 m), separated by a
thin (1.5 m thick) bedded ash and lithic-lapilli deposit (unit 3). Unit 5 is a 10- to 15m-thick unwelded lithic tuff, and unit 6 is a
thick (up to 70 m) densely welded, ash-flow
tuff. Unit 6 may correlate northwestward
into the lower reaches of Arroyo El Canelo,
with a 300-m-thick compound cooling unit
of densely welded tuffs (unit t12 of the Tuffs
of El Canelo of Stock et al., 1991). In the
thickest units of Tmc in Arroyo La Cantera,
the lithic content is ,25% and glassy juvenile material predominates. Thickness and
welding variations in the Tuff of El Canelo
indicate that the source may be located to
the west and close to our study area.
Laser fusion of sanidine from the basal
ash-flow tuff (Tmc1, Table 2 in the Data Repository) yielded a well-defined isochron
with an intercept age of 6.44 6 0.02 Ma (Table 1b). This result constrains the maximum
age for Tmc and is consistent with the 5.8 6
0.4 Ma age of the overlying rhyolite flows,
discussed above. Thus, most of the rhyolitic
volcanism of Group 2 occurred in latest Miocene time.
Group 3
Rhyolite Ignimbrites (Tpr). This sequence comprises the Tuff of Los Heme

(Tph) and the Tuff of Mesa El Tabano


(Tpt), which are two approximately coeval
sequences of pyroclastic flows collectively
labeled Tpr (Pliocene rhyolites). Tpr forms
the mesas and tilted blocks visible from the
coastal highway from Arroyo Matom to
south of Puertecitos. The most complete sequence of Tuff of Los Heme is observed in
the Arroyo Los Heme area, where it directly
overlies arc-related andesites, and locally
overlies the rocks of Group 2. Here, the Tuff
of Los Heme contains .25 cooling units,
including ash flows, pyroclastic flows with
pumice and minor lithic lapilli, and pumiceous airfall tuffs (Table 3 in the Data Repository). High within the sequence is a local unit of agglutinated mafic lava spatter.
North and west from the mouth of Arroyo
Los Heme, the number of individual units
decreases, and around Valle Curbina only
the lower part of the sequence (the Tuff of
Valle Curbina and the Tuff of Mesa El Tabano) are present (see Martn-Barajas and
Stock, 1993).
The basal unit of the Tuff of Los Heme
(Tph1) correlates with the Tuff of Valle
Curbina from the Valle Curbina area
(Fig. 3), on the basis of thickness and mineralogical and textural characteristics. It is
consistent with the isotopic data discussed
below and with magnetic polarity data
(T. Melbourne, unpub. data). In Valle Curbina, this unit is .60 m thick in some of the
paleotopographic depressions formed by
graben and half-graben structures, some of
which were invaded by marine waters during
late Miocene and Pliocene time. The Tuff of
Valle Curbina is thus interstratified with
shallow marine deposits (Stock et al., 1991;
Martn-Barajas et al., 1993). The distal part
of the Tuff of Valle Curbina reached as far
north as the southern Sierra San Fermn
(Fig. 3), where it is 35 m thick. Toward the
south, where it lies at the base of the Tuff of
Los Heme (Tph1), it is .70 m thick with a
middle welded zone. The unit extends .30
km in a north-south direction and .5 km in
an east-west direction, with an estimated
minimum volume of 5 km3 for the Tuff of
Valle Curbina.
The increase in thickness and number of
units toward the south and east within the
Tpr sequence suggests that their source is
east-southeast of the study area. A collapse
structure west of Volcan Prieto cuts most of
the Pliocene ignimbrites (Fig. 3) and consists of a normal fault dipping 608 458E with
a semicircular, concave east pattern in map
view. The uppermost epiclastic and pyroclastic units of the Tuff of Los Heme pinch

out against the footwall block. Proximal deposits in the downthrown block include airfall deposits with ballistic blocks of rhyolite
glass .1 m in diameter, agglutinated spatter
deposits of mafic lava (unit Tph 26), and
coarse epiclastic deposits. The vent or vents
that produced the ignimbrite deposits do
not crop out and may be buried by the pyroclastic deposits or may be under water farther east.
Isotopic ages for four pyroclastic rocks
from the young synrift sequence (Tpr) range
from 3.27 to 2.65 Ma. Plagioclase grains separated from three samples of Tpr were laser
fused (see Table 1b). The Tuff of Valle Curbina (sample 11) yielded a reliable isochron
age of 3.27 6 0.04 Ma. For this unit, the
fusion step analyses yielded dates ranging
from 3.06 6 0.13 to 3.44 6 0.04 Ma. Analyses yielding lower percentages of radiogenic 40Ar were relatively imprecise, so the
isochron age is much better defined than the
initial 40Ar/36Ar age.
The mean of 19 fusion analyses of single
plagioclase grains from subunit Tph28 is
2.65 6 0.05 Ma (sample 11, Table 1b). This
ash-flow tuff was sampled in site B of Figure
3 and is one of the uppermost units of Tph
in the coastal zone of Arroyo Los Heme
(Table 3 in the Data Repository). The isotopic age of this subunit constrains well the
minimum age of the Pliocene ignimbrite sequence. Plagioclase from a pumice-lapilli
tuff interstratified in the marine sequence in
the southern Sierra San Fermn (Fig. 3) was
dated at 3.08 6 0.04 Ma (sample 12, Table 1b). This unit is considered to be the
submarine equivalent of unit b of the Tuff of
Mesa El Tabano, and the age obtained is
consistent with a 3.1 6 0.5 Ma age (K/Ar,
plagioclase) for the uppermost unit of the
Tuff of Mesa El Tabano (Gastil, 1975).
The 40Ar/39Ar ages obtained by laser fusion analysis are in good agreement with the
step-heating experiments. Three samples
from unit Tpr were analyzed by the stepwise heating technique (Figs. 4d, 4e, and 4f).
Samples 5 and 6 in Table 1a are basal vitrophyres from the uppermost unit of welded
ash-flow tuff, 8 km west of the coastline
along Arroyo Los Heme (Fig. 3). This unit
correlates with unit Tph28 as discussed
above. On samples 5 and 6, whole-rock and
plagioclase separates were analyzed. Sample
5 displays a saddle-shaped spectrum for the
whole-rock experiment, and the plagioclase
separate yielded a segment consisting of two
fractions (74% of the 39Ar released) with a
2.7 6 0.4 Ma age. This age is consistent with
the isochron age of Tph28. Sample 6 yielded

Geological Society of America Bulletin, April 1995

413

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

a whole-rock plateau age of 3.36 6 0.03 Ma,


and the plagioclase separate yielded a perturbed, uninterpretable age spectrum.
Step-wise heating in duplicate analysis on
whole-rock sample 7 yielded very similar age
spectra, with two fractions in agreement at
3.5 Ma (see Fig. 4f). This sample was collected from unit Tpt-b in the Valle Curbina
area. However, this age is too old because it
is stratigraphically inconsistent with the
well-dated Tuff of Valle Curbina, 3.27 6
0.04 Ma, which underlies Tpt.
Volca
n Prieto and Other Young Volcanic
Flows (Tpa). Volcan Prieto (Fig. 3) is a monogenetic, 280-m-high cone that is coeval
with the latest ignimbrite-forming eruptions. Blocky andesite flows are interstratified with scoriaceous deposits from this volcano and cover an elliptical area, 3 4 km
across, east of the semicircular collapse
structure described above. Most lavas from
Volcan Prieto overlie the Tuff of Los Heme,
but some flows to the south underlie one of
the uppermost ash-flow tuffs of the Los
Heme sequence (probably unit Tph29) and
a relatively well sorted lithic sandstone. This
sandstone may represent a beach deposit;
however, no evidence of lava interaction
with water was observed. Andesitic agglutinated spatter (unit Tph26, Table 3 in the
Data Repository), probably from a nearby
vent, is interstratified in the uppermost part
of the Tuff of Los Heme, 5 km north of
Volcan Prieto (site B, Fig. 3). These stratigraphic relationships establish that andesitic
volcanoes and lava flows of Group 3 first
erupted during the waning stages of Pliocene rhyolitic volcanism and continued after
the youngest Los Heme ignimbrites.
Whole-rock ages from Volcan Prieto
(sample 8, Table 1a) and from an andesite
flow west of Cerro Canelo (sample 20,
Fig. 2) are 2.6 6 0.1 Ma and 0.91 6 0.67 Ma,
respectively. Step-wise heating of a sample
from Volcan Prieto (Fig. 4g) yielded a clear
plateau age with 75% of the 39Ar released,
in good agreement with the age of the
youngest Los Heme ignimbrite. The age of
the andesite flow from west of Cerro Canelo
is poorly constrained but is consistent with
the youthful morphology of the flow. Additionally, this flow is related to a northeast-striking fault that cuts Quaternary(?)
alluvium.
Sparse Pliocene andesite in the study area
probably indicates dispersed Pliocene
andesitic volcanism in the Puertecitos Volcanic Province (Stock and Comenetz, 1992).
Andesite from the valley west of Cerro
Canelo (Fig. 2) is similar to the Volcan Prie-

414

to andesite and has a well-preserved morphology. We infer that andesites from both
localities (Volcan Prieto and near Cerro
Canelo) are broadly equivalent in age (late
PliocenePleistocene [?]). At both localities,
the andesites are cut by normal faults (of
strike north-northwest and north-northeast
at Volcan Prieto, and northeast west of
Cerro Canelo).
Summary and Regional Stratigraphic
Correlations
The eastern Puertecitos Volcanic Province includes a sequence of arc-related
andesitic rocks (Group 1) discordantly overlain by two sequences of dominantly silicic
synrift volcanic rocks (Group 2 and Group 3,
Fig. 5).
The ca. 16 Ma ages of the arc-related andesite are consistent with the 20 14.5 Ma
ages obtained from similar units to the
north, in southern Valle Chico (Stock,
1989), where the sequence attributed to the
volcanic arc reaches 300 m in thickness and
includes pyroclastic flows, andesitic breccias, basaltic lavas, reworked tuffs, and epiclastic deposits (Fig. 5). There, and northeast of the Puertecitos Volcanic Province in
Sierra San Fermn (Fig. 3), the andesitic
rocks overlie the Mesozoic granitic and metamorphic basement as well as Tertiary basement-derived sandstone (Lewis, 1994). We
infer similar relationships beneath our study
area. In Arroyo Los Heme, the arc-related
volcanic deposits correspond to proximal facies, whereas in Valle Chico and the Sierra
San Fermn they vary from proximal to distal. Along the Main Gulf Escarpment at this
latitude, and to the north in the Sierra
Juarez (north of the Sierra San Pedro Ma
rtir), the epiclastic deposits dominate the arcsequence (Gastil et al., 1975; Dorsey and
Burns, 1994).
The two synrift sequences (Group 2 and
Group 3) erupted in relatively short periods
of time around 6 and 3 Ma, respectively. The
late Miocene silicic volcanism began no
later than 6.4 Ma, but likely after 7 Ma as
suggested by the lack of reported 40Ar/39Ar
ages older than 7 Ma on rhyolite rocks
within the Puertecitos Volcanic Province
and the Sierra San Fermn (Stock, 1989;
Lewis, 1994). The age of the Pliocene ignimbrite sequence is well constrained between 3.2 and 2.7 Ma. Both sequences end
with andesite flows and scoria deposits. The
6-m.y.-old rocks crop out principally toward
the north and northwest within the study
area. Many local rhyolite domes and flows

also occur there, and the source of the 6.4


Ma Tuff of El Canelo is also probably in that
general vicinity based on the presence of the
lava flows and domes along normal faults,
the increasing thickness of some units
toward the west, and the welding variations
within the compound cooling units. In general, the Group 2 sequence, which is the
older synrift sequence, pinches out toward
the south in the study area.
In southern Valle Chico, a sequence of
rhyolite ignimbrites and lava flows with ages
between 5.8 6 0.5 Ma (alkali feldspar) and
6.1 6 0.16 Ma (alkali feldspar) constitutes
the youngest regionally extensive volcanic
sequence near the escarpment (Tmr3 and
Tmr4, Fig. 5). These rhyolitic rocks overlie
intermediate to silicic lavas dated at 6.47 6
0.2 Ma (whole rock). The rhyolite flows and
the Tuffs of El Canelo in the northern
Puertecitos Volcanic Province are possibly
coeval with the 6.4 6.1 Ma rhyolite lavas
and tuffs reported by Stock (1989) from
Valle Chico. A regionally extensive, densely
welded 11 Ma tuff also crops out in southern
Valle Chico. The 11 Ma unit and the
younger andesite cones and lavas dated at
6.47 Ma by Stock (1989) have no known coeval units in the area included in this study.
The late Miocene rhyolite volcanism extended as far northward as the Sierra San
Fermn (Fig. 2b; Lewis, 1994) and apparently ceased around 5.5 Ma in our study
area. This intense volcanic activity is likely
associated with the extension of the crust as
suggested by the normal faults that produced steplike topography controlling the
rhyolite flows and that produced shallow
dips upsection in the Tuff of El Canelo. Normal faults produced graben structures that
were invaded by marine waters in late Miocene(?) and early Pliocene time. Shallow
marine beds discordantly overlie the late
Miocene volcanic rocks in the northeastern
Puertecitos Volcanic Province and some of
the rhyolite domes may have interacted with
sea water as suggested by the rhyolite tephra
that surrounds the lower rhyolite dome in
Arroyo La Cantera.
A second period of explosive volcanism
within the study area occurred during the
early late Pliocene time. Its locus was east of
the locus of volcanism during late Miocene
time, and the Pliocene ignimbrites pinch out
toward north and west (Fig. 5). The source
of the Pliocene ignimbrites was east of the
present coastline, and some units (such as
the Tuff of Valle Curbina) were partly deposited in a shallow marine environment.
Both periods of synrift volcanism ended with

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

Figure 5. West to east stratigraphic correlation and compilation of isotopic ages of the main volcanic groups in the northern Puertecitos
Volcanic Province. Data from this study in Table 1. The Group 3 in the Mesa El Ta
banoArroyo La Cantera is a composite column of
the Tuff of Mesa El Ta
bano (site A* in Fig. 3) and Tpvc from Valle Curbina. Plag, plagioclase; wr, whole rock; san, sanidine; ksp,
anorthoclase; hbl, hornblende. Note: The age of the uppermost unit in Mesa El Ta
bano is from Somer and Garca (1970).
sparse andesite flows, including the eruption
of Volcan Prieto at the end of the late Pliocene sequence.

contain augite in optically zoned, but compositionally homogeneous crystals (Fig. 7).
Group 2: Rhyolite Flows and Tuffs
of El Canelo

PETROGRAPHY AND MINERAL


COMPOSITION
Group 1: Arc-Related Andesite
Petrographically, these rocks show porphyritic to microporphyritic textures, with
a glassy to microcrystalline matrix. Hornblende is the most abundant mafic mineral
and occurs in phenocrysts (mostly 12
mm) randomly distributed or aligned parallel to the flow direction, defining a
pseudotrachytic texture. Distinctive, compositionally zoned plagioclase phenocrysts
have labradorite (An50 70) cores and andesine
(An30 50) rims (Fig. 6a). The andesitic necks

In Arroyo La Cantera, the lower rhyolite


lavas (Tmrl) are massive, devitrified, rhyolite glass flows, with some zones of autoclastic flow breccia and some zones of low-temperature hydrothermal alteration. These
lavas contain 5%10% phenocrysts, mainly
oligoclase, augite, and ferrohypersthene
(Fig. 7). The hydrothermal alteration is pervasive and produced pyrophyllite, smectite,
and zeolite.
The mineralogy of the Tuff of El Canelo
varies from base to top (Table 2 in the Data
Repository). The basal unit (Tmc1) contains
up to 30% phenocrysts, mainly plagioclase

and quartz, with subordinate opaque minerals and alkali feldspar. Units 2, 3, and 4
probably are a single compound cooling unit
with very consistent phenocryst compositions; all three contain ,10% phenocrysts,
which consist of plagioclase, orthopyroxene,
clinopyroxene, opaques, and hornblende.
The latter is generally altered to iron oxides.
Unit 6 is a densely welded ash-flow tuff containing plagioclase, orthopyroxene (ferrohypersthene), clinopyroxene (augite), and
opaques. In general, hornblende phenocrysts are not present in Tmc and are rare
within the Group 2 rocks.
The upper rhyolite lavas (Tmru) are
glassy and sparsely porphyritic or aphanitic,
with ,5% phenocrysts, principally sodic plagioclase (oligoclase to andesine), clinopyroxene (ferroaugite) (Figs. 6b and 7), and
opaque minerals in a partially recrystallized

Geological Society of America Bulletin, April 1995

415

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

Figure 6. Plagioclase normative composition in representative rocks. (a) Group 1 pre11


Ma hornblende andesite. (b) Group 2 late Miocene andesite and rhyolite lavas (Tba-Tmru).
(c) Pliocene rhyolite and dacite ignimbrites (Tpr). Data from Martn-Barajas and Stock
(1993).

matrix. Devitrification textures are spherulitic, composed of intergrowths of alkali


feldspar (Or40 50) and cristobalite. Microcrystalline calcite occurs in microfractures
and is also disseminated in the matrix.
Group 2: Synrift Andesite Lava (Tba)
Andesite lava from a dike in Valle Curbina and from a flow in Arroyo Los Heme
are aphanitic to microporphyritic. These
lava show microlitic flow texture and have
10%50% microphenocrystsmainly plagioclase, clinopyroxene (augite), and orthopyroxene (bronzite) (Fig. 7). Some plagioclase
phenocrysts show normal zoning composed
of labradorite cores and andesine rims
(Fig. 6b). Mafic minerals in some breccia
zones within the andesite dike from Valle
Curbina are selectively altered to chlorite
and limonite, probably by hydrothermal alteration. Ground mass is mainly plagioclase,
pyroxene, opaque minerals and ,10% glass.
No olivine was observed in these lavas.
Group 3: Synrift Pliocene Ignimbrites
(Tpr) and Andesites from Volca
n Prieto
and West of Cerro Canelo (Tpa)
The pyroclastic units at the base of the
Tuff of Los Heme (Tph1) and the Tuff of
Valle Curbina have similar mineralogic
composition. Both have plagioclase (An2550),
biotite, clinopyroxene (diopsidic augite-ferro-

416

augite), and rare hornblende. Quartz is extremely rare. Lithic fragments are principally rhyolite, with minor andesite porphyry,
basaltic andesite, and granite. The Tuff of
Valle Curbina has two types of glass: one
type is medium brown in thin section and
the other is colorless. Both are rhyolitic in
composition, but the brown variety is
1%2% less silicic and contains 0.5%1%

more iron. Two compositions of glass shards


are also present in the basal pyroclastic flow
deposit of the Tuff of Los Heme (Tph1).
Upsection, thin, densely welded to slightly
welded ash flows predominate. These units
have phenocrysts of orthopyroxene and clinopyroxene (both 1 mm) and lack hydrous
phases such as biotite or hornblende. Olivine phenocrysts are fairly common in rhyolite and dacite rocks. Plagioclase compositions vary from andesine to oligoclase
(An20 30) (Fig. 6c). Clinopyroxenes are augite and ferroaugite; orthopyroxene is hypersthene (En50 60, Fig. 7); and sparse
fayalite is observed with varying degrees of
iddingsite alteration. This phenocryst assemblage is characteristic of most of the tuff
of the Pliocene ignimbrite sequence (Tpr).
In andesite lavas from Volcan Prieto
(Tpa) and the andesite flow west of Cerro
Canelo, textures vary from aphanitic to porphyritic, and some flows have up to 20%
phenocrysts. In thin section, the texture is
microlitic to felsitic, with orthopyroxene, clinopyroxene, and olivine phenocrysts in a microcrystalline groundmass composed principally of acicular plagioclase, pyroxene, and
opaque minerals.
GEOCHEMISTRY
Our discussion of chemical compositions
focuses on the differences between the basaltic and andesitic rocks attributed to the
old volcanic arc (pre11 Ma) and the

Figure 7. Pyroxene quadrilateral composition in selected samples from the main volcanic
sequences in the northeast Puertecitos Volcanic Province. Pre11 Ma hornblende andesite
from Arroyo Los Heme (Tma, sample II-1), synrift andesite from Arroyo Los Heme (filled
square) (Tba, sample 2-V-2), and rhyolite flow from the Valle Curbina area (empty square)
(Tmru, sample III-11). Tpt is Tuff of Mesa El Ta
bano (samples III-6, III-2); Tpvc is the Tuff
of Valle Curbina (empty diamonds). The Fo-Fa diagram shows composition of olivine in
unit Tpt-c (sample III-2). Data are in Martn-Barajas and Stock (1993) and available on
request.

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

andesitic rocks contemporaneous with the


rift. Because of the scarcity of mafic rocks
that might represent the primary magmas of
synrift late Miocene and Pliocene rocks, we
analyzed andesite rocks of those sequences.
For the trace-element analysis of representative samples, care was taken to analyze
only the freshest rocks.
Major-Oxide Geochemistry

Figure 8. Total alkali versus silica diagram. Dotted lines delimit fields from Le Bas et
al. (1986). Dividing lines between alkalic and subalkalic fields from Irvine and Baragar
(1971) and McDonald and Katsura (1964). Data from Tables 4 and 5 and Martn-Barajas
and Stock (1993). Rock fields: TB, trachybasalt; BTA, basaltic trachyandesite; B, basalt;
BA, basaltic andesite; TA, trachyandesite; A, andesite; D, dacite; TD, trachydacite; R,
rhyolite.

Figure 9. Ternary diagram for volcanic rocks from the Puertecitos Volcanic Province.
Tholeiitic divide from Irvine and Baragar (1971).

Arc-Related Rocks. The arc-related sequence is composed mainly of subalkaline


andesite and basalt (Fig. 8). We include two
analyses of alkalic basalts (20 Ma) and one
of a subalkalic basalt (17 Ma) from southern
Valle Chico to show that pre11 Ma rocks
also have geochemical variations apparently
related to the subduction tectonic setting.
Subalkalic andesite and basalt plot in the
calc-alkaline field in the ternary diagram of
Figure 9 (Group 1). The alkalic basalt is
characterized by high K2O (2.3% and 2.7%),
MgO, and P2O3 content and very low Al2O3
content (Fig. 10). This basalt directly overlies granitic basement and/or basementderived sediments (Stock, 1989). For arcrelated rocks, a clear decreasing trend is
observed in the FeO*, TiO2, MgO, and CaO
content with increasing silica, whereas
Al2O3 and P2O5 values scatter (Fig. 10).
With the exception of the alkalic basalt, K2O
and Na2O in andesitic to dacitic arc-rocks
increase with increasing silica.
Synrift Rocks. Synrift volcanic rocks are
predominantly rhyolite to dacite, with some
andesite. All synrift andesitic to rhyolitic
rocks are subalkaline (Fig. 8) and form a
broad calc-alkaline trend in the ternary diagram (Fig. 9). Na2O and Al2O3 are the oxides that show the most scatter on Harker
variation diagrams (Fig. 10); Na2O is particularly scattered in the more silicic samples, whereas Al2O3 tends to scatter in andesite rocks. The scatter in Na2O is probably
due to sodium mobility during devitrification that systematically leaches Na2O and
increases K2O content of some ignimbrites
(Lipman, 1965; Conrad, 1984). Some of the
vitrophyres show incipient devitrification,
which could account for the variation in
Na2O content. The consistency in the CaO/
SiO2 and K2O/SiO2 trends in the same rocks
suggests, however, that devitrification did
not modify their chemical composition significantly. The andesite and basalt analyzed
show no recrystallization or posteruption alteration effects.
Relative to the arc-related andesite and
basalt, synrift rocks are characterized by a

Geological Society of America Bulletin, April 1995

417

Downloaded from gsabulletin.gsapubs.org on February 20, 2012

Figure 10. Variation of major oxides versus silica in the Puertecitos Volcanic Province. Symbols as in Figure 8. Data from Tables 4 and
5 and Martn-Barajas and Stock (1993).

418

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

Figure 11. Spider diagrams of selected mafic rock samples. Normalization values are
chondritic abundances, except those for Rb, K, and P (normalizing values from Thompson
et al., 1984). (a) Pre11 Ma rocks. (b) Late Miocene to Pliocene rocks. Shaded pattern is
arc-related basalt and basaltic andesite of Figure 12a. Data from Table 4 in the Data
Repository.

low K2O content (typically ,1.0% K2O).


Some synrift andesites and basaltic andesites show high TiO2; however, this is not
systematic, and some samples show the
same TiO2 as do the arc rocks. The MgO
contents of the youngest synrift andesite
rocks from Volcan Prieto and west of Cerro
Canelo are high when compared with the
orogenic andesite and basalt. In the former,
Mg# (Mg# 5 atomic Mg/[Mg 1 Fe]) varies
from 55 in dacite and rhyolite rocks to 70 75
in andesites with a silica content of 60% (Tables 4 and 5 in the Data Repository). In synrift andesites of Group 2 (Tba), Mg# is

lower and ranges from 46 to 56. The Al2O3


content in three andesite samples (Tba) is
anomalously high (.17 wt%).
Trace-Element Geochemistry
Arc-Related Rocks. The samples analyzed
for trace elements are two alkalic basalts
and two basaltic andesites. Three of these
samples (VC-86-99, VC-87-10, VC-87-54)
are from southern Valle Chico and were
dated by K-Ar (Stock, 1989). Sample P38-1
is a basaltic andesite from the Arroyo Los
Heme (Fig. 3). Orogenic basalt and basaltic

andesite contain large amounts of incompatible elements and pronounced negative


anomalies in Nb-Ta and Ti (Fig. 11a). The
rare-earth-element (REE) patterns are
strongly fractionated between light REE
and heavy REE (Lan/Lun ratios of 9.520)
(Fig. 12a). This fractionation is more pronounced in the alkalic basalt of Valle Chico
than in the subalkalic basaltic andesite. Alkalic basalt from Valle Chico has characteristics similar to synrift alkalic rocks from
southern Baja California. Both have some
trace-element characteristics of arc magmas
(Sawlan and Smith, 1984; Sawlan, 1991).
Synrift Rocks. Trace-element analysis of
synrift rocks includes two basaltic andesites
(Tba) from Group 2. From Group 3, we also
analyzed two andesite flows from Volcan
Prieto (VP-1, VP-2) and one andesite from
west of El Canelo (MT-91-67).
The younger synrift andesite is the leastevolved lava. Most of the incompatible elements are less concentrated in synrift andesites than in the arc-related basalts
(Fig. 11b). This is especially true of the alkaline-earth and the light rare-earth elements (LREE). Synrift andesite also shows
the same negative Nb-Ta anomaly as do the
arc-related rocks, but the negative Ti anomaly is not as well defined. In synrift andesites,
K/Rb ratios vary between 248 and 590 (Table 4 in the Data Repository). The K/Rb
ratios in Group 2 andesites (Tba, samples
2-V-2 and 30-IV-4) are extremely low (248
282), whereas andesitic rocks of Volcan Prieto and west of Cerro Canelo have similar
low K2O contents but have K/Rb ratios from
490 to 590. The Ni content of two samples
from Volcan Prieto (VP-1 and VP-2) is high
for andesitic rocks (120 ppm); these rocks
also have high MgO and TiO2. This andesite
volcano erupted the least-differentiated lavas among the Pliocene and late Miocene
synrift volcanic rocks.
Most trace elements exhibit similar behavior in Group 2 and Group 3 andesite.
Variations among synrift andesite within
each group occur in the Ni, Sr, and TiO2
contents. Nevertheless, a homogeneous and
distinctive characteristic of synrift andesite
is its REE pattern. Both Group 2 and Group
3 andesites have the same REE patterns
characterized by less fractionation of LREE
than arc rocks, with Lan/Lun values between
3.8 and 7.2 (Figs. 12a and 12b). The traceelement composition in andesites is also
similar in both synrift andesite groups and is
closely clustered in the ratio of some traceelement versus Zr diagrams (Fig. 13). Despite the scatter in these diagrams, a trend is

Geological Society of America Bulletin, April 1995

419

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

Figure 12. Chondrite-normalized REE patterns for selected mafic rock samples. (a)
Pre11 Ma rocks. (b) Late Miocene to Pliocene rocks. Shaded pattern is arc-related basalt
and basaltic andesite of Figure 13a. Data from Table 4 in the Data Repository. Normalization values from Boynton (1984).
visible in diagrams of Nb/Zr, La/Zr, and
Y/Zr. The dispersion is larger in the Rb/Zr,
Ba/Zr, and Sr/Zr diagrams, and the trend is
best defined when only the andesitic to rhyolitic lavas are considered. The pyroclastic
rocks analyzed (mostly basal vitrophyres)
show the most dispersed values.
DISCUSSION
The 16 Ma andesitic rocks from Arroyo
La Cantera represent proximal facies of the
early to middle Miocene volcanic arc, which

420

is continuously exposed southward along the


eastern margin of the peninsula (Hausback,
1984). The ages obtained in our study and
those reported for arc-rocks from Valle
Chico (Stock, 1989) are consistent with previous interpretations of the earlier waning
and termination of the subduction-related
volcanism in northern Baja California at
1714 Ma, earlier than in southern Baja California (14 12 Ma, Hausback, 1984; Sawlan,
1991; Gastil et al., 1979). At the latitude of
Puertecitos, proximal facies of the volcanic
arc crop out for 30 40 km from the gulf

coast westward to the southern part of the


Sierra San Pedro Martir (Stock, 1989;
Dorsey and Burns, 1994). Farther west, only
medial to distal facies crop out. The width of
the proximal facies in the Puertecitos Volcanic Province is comparable to the width
reported in southern Baja California by
Hausback (1984). Although outcrops of the
orogenic volcanic belt are discontinuous in
northern Baja California, the Puertecitos
Volcanic Province region was a major locus
of arc-related volcanism in early to middle
Miocene time.
The ages of the post11 Ma rocks (the
older probable age of inception of extension) cluster around two periods of volcanism: one at the end of late Miocene time and
a second one at the end of early Pliocene
time. The locus of synrift explosive volcanism shifted dramatically eastward between
these two periods; we speculate that the
shifting may be related to the development
of the shoulders of the new rift. The younger
ignimbrite-forming eruptions are associated
with a second period of extension that continued along the coast after deposition of
the youngest ignimbrite.
The alignment of rhyolite domes follows
the north-northeast to north-northwest orientation of the main fault pattern here
(Martn-Barajas and Stock, 1993) and west
of the study area, along the escarpment
(Stock and Comenetz, 1992; Stock et al.,
1993). This suggests that normal faults acted
as conduits through which lava flows and
domes were extruded. Additionally, both
the rhyolite domes and the outcrops of
andesitic rocks from Los Heme acted as topographic controls on the deposition of the
pyroclastic units. There is evidence for synextensional tectonism during deposition of
both the Pliocene and late Miocene ignimbrite sequences, from progressive upward
shallowing of eutaxitic foliation and bedding
of the pyroclastic and epiclastic units. In
both cases the pyroclastic units are well
stratified and have reliable attitudes. In Arroyo Los Heme, the base of the Tuff of Los
Heme is tilted 238E, whereas higher units
dip ,108E. In Arroyo La Cantera, the unit
at the base of the Tuff of El Canelo dips
.408ENE, whereas the top of the sequence
dips ,108ENE. This suggests syndepositional growth faulting.
Mineralogy and Geochemistry
The mineralogy of the synrift pyroclastic
flows, and the rhyolite-andesite lavas, is consistent with high temperatures in the magma

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

Figure 13. Selected trace elements versus SiO2 for synrift dacite to rhyolite lavas and ignimbrites in the northern Puertecitos
Volcanic Province. Symbols as in
Figure 11. Data from Tables 4
and 5.

chamber and/or mixing of a hot mafic


magma with a more felsic material. Most
rocks in these sequences lack phenocrysts of
hydrous minerals (biotite and hornblende),
suggesting that hydrous minerals, if present,
broke down under the high temperatures
and (or) decompressional melting under
adiabatic conditions during rifting. The association of augite, hypersthene, and fayalite in the rhyolites and dacites is indicative
of a more mafic magma. However, the clinopyroxene and orthopyroxene are euhedral and homogeneous and appear to have

been in equilibrium with the melt. Only olivine phenocrysts in rhyolite to dacite ignimbrites are often altered to iron oxides.
Assimilation is suggested by resorption
textures in plagioclase phenocrysts and normal compositional zonation with cores of
labradorite-bytownite and rims of oligoclase-andesine in synrift rocks. Additionally,
modification of the mineralogical characteristics of some pyroclastic flows is inferred to
have occurred by mechanical fragmentation
of the granitic wall rocks. The Tuff of Valle
Curbina appears to be contaminated with

biotite xenocrysts from the granitic basement, because two 40Ar/39Ar ages on biotite
concentrates yielded Cretaceous ages. Similar problems have been noted with biotite
in younger tuffs interbedded in the marine
section in the Sierra San Fermn (Lewis, 1994).
The geochemistry of the synrift rocks suggests that primary melts, probably from the
mantle, were extensively hybridized by
crustal components. Partial melting from a
mantle source is suggested by the low K2O
and by relatively high, but variable Ni, TiO2,
and Mg# in synrift andesite. The Mg# and

Geological Society of America Bulletin, April 1995

421

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.

the Ni content are particularly high in


Volcan Prieto lavas and constitute the most
primitive synrift rocks in the Puertecitos Volcanic Province. Evidence of crustal contamination is provided by the ratios of some incompatible elements that do not follow
typical differentiation trends from andesite
to rhyolite due to crystal fractionation from
a homogeneous source (Fig. 13). Crystal
fractionation alone is unlikely to produce
differentiation paths for rhyolite magmas.
We would expect, for instance, clinopyroxene fractionation to produce a decrease in
both the Al2O3 and CaO trends in the
Harker diagrams (Fig. 10). Additionally,
REE patterns show no Eu anomaly indicative of plagioclase fractionation. Except for
the lowermost units in both the Tuff of El
Canelo and the Tuff of Los Heme, synrift
rhyolite-dacite lava and tuff are chiefly crystal poor (,10%), so that it is difficult to appeal to extensive crystal fractionation as an
explanation for differentiation to produce
felsic magma bodies. Magma mixing is a
plausible explanation to produce rhyolite
and dacite rocks. Petrographic evidences for
magma mixing are few (e.g., the two glass
compositions in the Tuff of Valle Curbina,
resorbed hornblende and plagioclase phenocrysts, olivine microphenocrysts in rhyolite and dacite rocks). Additionally, variations in some trace element ratios also
suggest magma mixing processes. For instance, some 6 Ma rhyolites have the same
Ce, La, and Sr values as do coeval andesites,
and Zr in synrift rocks varies from 50 ppm to
.400 ppm independently of the degree of
differentiation.
Differences in trace element compositions in andesitic rocks probably indicate
differences in the degree of partial melting.
Andesite from Volcan Prieto and west of
Cerro Canelo shows similar low K2O contents as Tba andesite but has twice the K/Rb
ratios. On the other hand, variation in the
content of some highly incompatible elements in synrift andesite, such as Ni and
TiO2, is also consistent with different degrees of partial melting. Similar compositions of most trace elements suggest, however, a similar source.
The main differences between synrift andesite and arc-related andesite and basalt
are the lower K2O, the smaller fractionation
of LREE, the smaller negative Nb-Ta anomaly, and the lower contents of Zr and alkaline-earth elements (Ba, Rb, Sr) in synrift
rocks. These suggest that primary melts for
synrift rocks are less evolved than for arcrocks. The negative Nb-Ta anomaly in arc-

422

related andesite and basalt rocks is generally


associated with low values of TiO2 (Arculus,
1987) and with the presence of a refractory
residual phase (rutile, ilmenite, or perovskite) (Saunders et al., 1980). In synrift
rocks, this negative anomaly is least pronounced and may also indicate contribution
of crustal material. Elsewhere, calc-alkaline
volcanism not related to an active subduction setting is reported as a dominant type
of volcanism in the southwestern United
States. There, crustal contamination previously modified in a paleosubduction setting
appears to play a major role in the geochemistry of late Cenozoic magmas (Glazner,
1990).
Variations in the ratios of incompatible
elements within the synrift lavas reflect a
complex history of magma mixing and/or
different degrees of partial melting. On the
other hand, the enrichment of incompatible
elements in the orogenic andesites and the
presence of alkalic basalts with typical orogenic signatures (negative Nb-Ta and Ti
anomaly) may be attributed to the tectonic
history of the continental margin, which was
a convergent plate boundary during most of
Mesozoic and Cenozoic time (e.g., Atwater,
1970, 1989).
Comparison with Other Synrift
Calc-Alkaline Volcanic Centers
in the Gulf Region
The major recognized center of Plio-Quaternary calc-alkaline volcanism in the gulf

depression is the Tres VrgenesSanta Ana


caldera area, located in the south-central
Gulf coast (Sawlan and Smith, 1984; Sawlan,
1991). The Tres Vrgenes volcanic center
has a similar geochemical and mineralogic
composition to orogenic andesites (Sawlan,
1981, 1991). Another Pliocene calc-alkaline
volcanic sequence crops out in Isla San Esteban (Desonie, 1992; Fig. 2a). There, a
suite of rhyolite to andesite domes, lavas,
and pyroclastic rocks has similar trace element compositions to the Tres Vrgenes volcanoes. The Isla San Esteban sequence is
coeval with the Group 3 sequence from Puertecitos. Nevertheless, the former has a distinct mineralogic and geochemical composition. Andesite lavas from Isla San Esteban
are more fractionated and have higher K2O,
lower LREE/HREE ratios, and lower Ni
and Cr, but similar high-field-strength elements as compared with synrift andesites
from Puertecitos (Fig. 14). In Isla San Esteban, the dominant mafic phase is either
hornblende or pyroxene, and andesite and
dacite rocks have up to 40% phenocrysts.
The source for these volcanic rocks remains
unknown, and it has been proposed that
postsubduction calc-alkaline volcanism may
be related to an asthenospheric window created by subduction of the spreading ridge
(Desonie, 1992).
Although speculative, we propose another explanation. Partial melting of a transitional tholeiitic basalt may have produced
the synrift andesite rocks from Puertecitos.
Holocene tholeiitic volcanism occurs in the

Figure 14. N-MORB-normalized spider diagrams of synrift andesite, compared with


spider diagrams of synrift calc-alkaline andesite and dacite from Isla San Esteban (cross),
and Cerro Prieto tholeiitic basalt (Herzig, 1990), andesite, and dacite (shaded pattern).
N-MORB normalizing constants are from Pearce (1983).

Geological Society of America Bulletin, April 1995

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


PUERTECITOS VOLCANIC PROVINCE, BAJA CALIFORNIA, MEXICO

Cerro Prieto Geothermal Field and in the


Salton Trough (Herzig and Elders, 1988;
Herzig, 1990). There, differentiation occurred mainly by crystal fractionation and,
to a lesser extent, by contamination with
crustal rocks (Herzig, 1990). Differentiation
of this tholeiitic magma produced a strong
fractionation of incompatible elements
(Fig. 14). The Cerro Prieto basalt-to-dacite
suite has a higher high-field-strength element concentration (e.g., Zr, Y, except
Hf) and higher content of HREE than synrift andesite from Puertecitos, but Ni in
lavas from Volcan Prieto have values similar to MORB (mid-oceanic-ridge basalt,
Fig. 14).
In the Puertecitos Volcanic Province, extension during the rifting may have induced
a high geothermal gradient that caused partial melting of a lithospheric, MORB-type
mantle. Subsequently, andesitic magma
bodies where contaminated with crustal
melts to produce rhyolite and dacite rocks.
However, isotopic studies are needed to
evaluate this hypothesis inferred from our
mineralogical and geochemical data.
CONCLUSIONS
The volcanic succession in the eastern and
northern Puertecitos Volcanic Province
comprises three main sequences that record
the transition from subduction to rifting of
the Gulf of California. The lower sequence
(Group 1) consists of middle Miocene andesite of the volcanic arc, formed during the
final subduction of the Farallon-Guadalupe
Plate beneath northern Baja California. The
older synrift sequence (Group 2) consists of
a series of rhyolite domes, a sequence of
.300 m of pyroclastic flows (the Tuff of El
Canelo), and a terminal phase of andesitic
lavas volumetrically small compared to the
rhyolite. This sequence crops out in the
northern Puertecitos Volcanic Province,
where it is covered discordantly by Pliocene
ignimbrites. Isotopic ages limit this sequence to between 6.4 and 5.8 Ma in age.
In early Pliocene time (3.24 2.7 Ma), further synrift volcanism (Group 3) occurred
east of the locus of the first synrift volcanism
(Group 2) and produced a series of pyroclastic flows up to 200 m thick (Tuff of Los
Heme and Tuff of Mesa El Tabano). Their
source appears to have been located in the
southeastern part of the Puertecitos Volcanic Province. This Pliocene phase of explosive
volcanism culminated with the formation of
a monogenetic andesitic volcano and various local mafic lava flows (Volcan Prieto

and flows west of Cerro Canelo). The synrift


nature of the upper two groups is indicated
by the local alignment of rhyolite domes
along north-south to north-northwest
south-southeaststriking faults, the upward
shallowing of dips in the volcaniclastic units
(indicative of growth faulting), and the faultrelated topographic control on deposition of
the pyroclastic and lava flows.
The mineralogic composition of the riftrelated rocks suggests high-temperature
magmas. The major and the trace element
signatures of most synrift and arc rocks are
calc-alkaline. However, synrift andesites are
low-K and have a lower content of incompatible elements than the arc-related andesite and basalt. The lower content in incompatible elements in synrift andesite suggests
mantle-derived melts hybridized by crustal
contamination.
ACKNOWLEDGMENTS
Thanks to J. Kimbrough and M. Walawender for use of the XRF analytical facilities at San Diego State University. Special
thanks to H. Delgado for fruitful discussions
and for a preliminary revision of the draft
manuscript. Thanks to Ch. Herzig and
D. Kimbrough for multiple suggestions that
also improved the manuscript, to V. Frias
and G. Rendon for drafting, and to S. Augustin and G. Axen for improvement of the
English usage. Special thanks to M. Sawlan
and P. Weigand for their constructive reviews. A. Martn-Barajas was supported by
CONACYT, Mexico, Grant 1224-T9203.
J. Stock was partially supported by National
Science Foundation Grants EAR-89-04022
and EAR-92-18381. Additional support for
J. Stock and B. Hausback came from the
Petroleum Research Fund of the American
Chemical Society via Grant No. 21291-G.
REFERENCES CITED
Arculus, R. J., 1987, The significance of source versus process in
the tectonic controls of magma genesis, in Weaver, S. D.,
and Johnson, R. W., eds., Tectonic controls on magma
chemistry: Journal of Volcanology and Geothermal Research, v. 32, p. 112.
Atwater, T. M., 1970, Implications of plate tectonics for the Cenozoic evolution of western North America: Geological Society of America Bulletin, v. 81, p. 35133536.
Atwater, T. M., 1989, Plate tectonic history of northeast Pacific
and western North America, in Winterer, E. L., Hussong,
D. M., and Decker, R. W., eds., The eastern Pacific Ocean
and Hawaii: Boulder, Colorado, Geological Society of
America, The Geology of North America, v. N, p. 2178.
Bard, J. P., 1986, Microtextures of igneous rocks: Dordrecht, Netherlands, D. Reidel Publishing Co., 264 p.
Boynton, V. W., 1984, Cosmochemistry of rare earth elements:
Meteorite studies, in Henderson, P., ed., Rare earth geochemistry: Amsterdam, Netherlands, Elsevier, p. 63114.
Bryant, B., 1986, Geology of the Sierra Santa Rosa Basin, Baja
California, Mexico [Masters thesis]: San Diego, California,
San Diego State University, 75 p.
Cas, R. A., and Wright, J. V., 1987, Volcanic successions: London,
Allen & Unwin, 528 p.
Conrad, W., 1984, The mineralogy and petrology of composition-

ally zoned ash flow tuffs and related silicic volcanic rocks,
from the McDermitt caldera complex, Nevada-Oregon:
Journal of Geophysical Research, v. 89, no. B10, p. 86398664.
Deino, A., and Potts, R., 1990, Single-crystal 40Ar/39Ar dating of
the Olorgesailie Formation, Southern Kenya Rift: Journal
of Geophysical Research, v. 95, no. B6, p. 8453 8470.
Delgado-Argote, L., Lo
pez Martnez, M., and Perrillat, M. C.,
1992, Edad del vulcanismo y de la fauna en sedimentos asociados de Baha de Los Angeles, Golfo de California: Geos,
v. 12, no. 5, p. 81.
Demant, A., 1981, Plio-Quaternary volcanism of the Santa Rosalia
area, Baja California, Mexico, in Orlieb, L., and Roldan-Q,
J., eds., Geology of northwestern Mexico and southern Arizona: Geological Society of America, Cordilleran Section,
Annual Meeting, Hermosillo, Sonora, Field guides and papers, p. 295305.
Desonie, D. L., 1992, Geologic and geochemical reconnaissance of
Isla San Esteban: Post-subduction orogenic volcanism in the
Gulf of California: Journal of Volcanology and Geothermal
Research, v. 52, p. 123140.
Dokka, R. K., and Merriam, R. H., 1982, Late Cenozoic extension
of northeastern Baja California, Mexico: Geological Society
of America Bulletin, v. 93, p. 371378.
Dorsey, R. J., and Burns, B., 1994, Regional stratigraphy, sedimentology and tectonic significance of Oligocene-Miocene
sedimentary and volcanic rocks, northern Baja California,
Mexico: Sedimentary Geology, v. 88, p. 231251.
Fisher, R. V., 1961, Proposed classification of volcaniclastic sediments and rocks: Geological Society of America Bulletin,
v. 72, p. 14091414.
Gastil, G., and Krummenacher, D., 1977, Reconnaissance geology
of coastal Sonora between Puerto Lobos and Bahia Kino:
Geological Society of America Bulletin, v. 88, p. 189198.
Gastil, R. G., Phillips, R. P., and Allison, E. C., 1975, Reconnaissance geology of the State of Baja California: Geological
Society of America Memoir 140, 170 p.
Gastil, R. G., Krummenacher, D., and Minch, J., 1979, The record
of Cenozoic volcanism around the Gulf of California: Geological Society of America Bulletin, Part I, v. 90,
p. 839 857.
Glazner, A. F., 1990, Recycling of continental crust in Miocene
volcanic rocks from the Mohave block, southern California,
in Anderson, J. L., ed., The nature and origin of Cordilleran
magmatism: Geological Society of America Memoir 174,
p. 147168.
Hausback, B. P., 1984, Cenozoic volcanism and tectonic evolution
of Baja California Sur, Mexico, in Frizzell, V. A., Jr., ed.,
Geology of the Baja California Peninsula: Society of Economic Paleontologists and Mineralogists (SEPM), Pacific
Section, v. 39, p. 219236.
Herzig, T. C., 1990, Geochemistry of igneous rocks from the Cerro
Prieto geothermal field, northern Baja California, Mexico:
Journal of Volcanology and Geothermal Research, v. 42,
p. 261271.
Herzig, T. C., and Elders, W. A., 1988, Nature and significance of
igneous rocks cored in the State 2-14 research borehole:
Salton Sea Scientific Drilling Project: Journal of Geophysical Research, v. 93, p. 13069 13080.
Irvine, T. N., and Baragar, W. R. A., 1971, A guide to chemical
classification of the common volcanic rocks: Canadian Journal of Earth Science, v. 8, p. 523548.
Le Bas, M. J., Le Maitre, R. W., Streckeisen, A., and Zanettin, B.,
1986, A chemical classification of volcanic rocks based on
the total alkali-silica diagram: Journal of Petrology, v. 77,
p. 2437.
Lewis, J. C., 1994, Constraints on extension in the Gulf Extensional Province from Sierra San Fermin, northeastern Baja
California, Mexico [Ph.D. thesis]: Cambridge, Massachusetts, Harvard University, 359 p.
Lipman, P. W., 1965, Chemical comparison of glassy and crystalline volcanic rocks: U.S. Geological Survey Bulletin 1201-D,
p. D1D24.
Martn-Barajas, A., and Stock, J. M., 1993, Estratigrafa y petrologa de la secuencia volcanica de Puertecitos, noreste de
Baja California. Transicio
n de un arco volcanico a rift, in
Delgado-Argote, L. A., and Martn-Barajas, A., eds., Contribuciones a la Tecto
nica del Occidente de Mexico:
Monografas de la Unio
n Geofsica Mexicana, no. 1,
p. 66 89.
Martn-Barajas, A., Tellez-Duarte, M., and Rendo
n-Marquez, G.,
1993, Estratigrafa y ambientes de depo
sito de la secuencia
marina de Puertecitos, NE de Baja California. Implicacines
sobre la evolucio
n de la margen Occidental de la depresio
n
del Golfo, in Delgado-Argote, L. A., and Martn-Barajas,
A., eds., Contribuciones a la Tecto
nica del Occidente de
Mexico: Monografas de la Union Geofsica Mexicana,
no. 1, p. 90 114.
McDonald, G. A., and Katsura, T., 1964, Chemical composition of
Hawaiian lavas: Journal of Petrology, v. 5, p. 82133.
Mora-Alvarez, G., 1993, Relaciones estratigraficas y geocronolo
gicas entre las unidades volcanicas de la Sierra Santa Ursula
en Sonora, y el magmatismo de la regio
n del Golfo de California, in Delgado-Argote, L. A, and Martn-Barajas, A.,
eds., Contribuciones de Tecto
nica del Occidente de Mexico:
Monografas de la Unio
n Geofsica Mexicana, no. 1,
p. 123147.
Neuhaus, J., 1989, Volcanic and non-marine stratigraphy of southwest Isla Tiburo
n, Gulf of California, Mexico [Masters thesis]: San Diego, California, San Diego State University,
170 p.
Pearce, J. A., 1983, The role of sub-continental lithosphere in

Geological Society of America Bulletin, April 1995

423

Downloaded from gsabulletin.gsapubs.org on February 20, 2012


MARTIN-BARAJAS ET AL.
magma genesis at destructive plate margins, in Hawkesworth, C. J., and Norry, M. J., eds., Continental basalt and
mantle xenoliths: Cheshire, England, Shiva Publishing,
p. 230249.
Renne, P. R., Ernesto, M., Pacca, I. G., Coe, R. S., Glen, J. M.,
Prevot, M., and Perrin, M., 1992, Age of Parana flood volcanism, rifting of Gondwanaland, and the Jurassic-Cretaceous boundary: Science, v. 258, p. 975979.
Saunders, A. D., 1983, Geochemistry of basalts recovered from
Gulf of California during Leg 65 of the Deep Sea Drilling
Project, in Robinson, P., Lewis, B. T., and others, Initial
reports of the Deep Sea Drilling Project, Volume 65: Washington, D.C., U.S. Government Printing Office, p. 591 621.
Saunders, A. D., Tarney, J., and Weaver, S. D., 1980, Transverse
geochemical variations across the Antarctic Peninsula: Implications for the genesis of the calc-alkaline magmas: Earth
and Planetary Science Letters, v. 46, p. 344360.
Saunders, A. D., Fornari, D. J., and Morrison, M. A., 1982, The
composition and emplacement of basaltic magmas produced during the development of continental margin basins:
The Gulf of California, Mexico: Journal of the Geological
Society of London, no. 139, p. 335346.
Saunders, A. D., Rogers, G., Marriner, G. F., Terrell, D. J., and
Verma, S. P., 1987, Geochemistry of Cenozoic volcanic
rocks, Baja California, Mexico: Implications for the petrogenesis of post-subduction magmas, in Weaver, S. D., and
Johnson, R. W., eds., Tectonic controls on magma chemistry: Journal of Volcanology and Geothermal Research,
v. 32, p. 223245.
Sawlan, M. G., 1981, Late Cenozoic volcanism in the Tres Virgenes area, in Orlieb, L., and Roldan-Q, J., eds., Geology of
northwestern Mexico and southern Arizona: Geological Society of America, Cordilleran Section, Annual Meeting,
Hermosillo, Sonora, Field guides and papers, p. 309319.

Sawlan, M. G., 1991, Magmatic evolution of the Gulf of California


rift, in Dauphin, J. P., and Simoneit, B. R., eds., The Gulf
and Peninsular Province of the Californias: American Association of Petroleum Geologists Memoir 47, p. 301369.
Sawlan, M. G., and Smith, J. G., 1984, Petrologic characteristics,
age and tectonic setting of Neogene volcanic rocks in northern Baja California Sur, Mexico, in Frizzell, V. A., Jr., ed.,
Geology of the Baja California Peninsula: Society of Economic Paleontologists and Mineralogists, Pacific Section,
v. 39, p. 237251.
Sommer, M. A., and Garca, J., 1970, Potassium-argon dates for
Pliocene rhyolite sequences [sic] east of Puertecitos, Baja
California: Geological Society of America Abstracts with
Programs, v. 2, no. 2, p. 146.
Stock, J. M., 1989, Sequence and geochronology of Miocene rocks
adjacent to the Main Gulf Escarpment, southern Valle
Chico, Baja California: Geofsica Internacional, v. 28-5,
p. 851 856.
Stock, J. M., 1993, Geologic map of southern Valle Chico, Baja
California, Mexico: Geological Society of America Map and
Chart Series MCH076.
Stock, J. M., and Comenetz, J., 1992, Use of Landsat-TM data to
distinguish among volcanic rock types, Puertecitos Volcanic
Province, Baja California, Mexico: Eos (Transactions,
American Geophysical Union), v. 73, no. 43, p. 613.
Stock, J. M., and Hodges, K. V., 1990, Miocene to Recent structural development of an extensional accommodation zone,
NE Baja California, Mexico: Journal of Structural Geology,
v. 12, p. 315328.
Stock, J. M., and Molnar, P., 1988, Uncertainties and implications
of the Late Cretaceous and Tertiary position of North
America relative to the Farallon, Kula and Pacific plates:
Tectonics, v. 7, p. 13391384.
Stock, J. M., Martn-Barajas, A., Suarez-Vidal, F., and Miller, M.,

1991, Miocene to Holocene extensional tectonics and volcanic stratigraphy of NE Baja California, Mexico, in Walawender, M., and Hanan, B., eds., Geological excursions in
southern California and Mexico (Geological Society of
America Meeting guidebook, Field trip 2): San Diego, California, Department of Geological Sciences, San Diego
State University, p. 44 67.
Stock, J. M., Miller, M., Lewis, J. C., and Martn, A. B., 1993,
Estudios tecto
nicos y vulcanolo
gicos de la Provincia Volcanica de Puertecitos, NE de Baja California, utilizando datos
de Landsat (Mapeador Tematico): Geos Boletn Unio
n
Geofsica Mexicana, v. 13, no. 5, p. 29.
Thompson, R. N., Morrison, M. A., Hendry, G. L., and Parry, S. J.,
1984, An assessment of the relative roles of crust and mantle
in magma genesis: An elemental approach: Philosophical
Transactions of the Royal Society of London, ser. A, v. 310,
p. 549 590.
York, D., 1969, Least squares fitting of a straight line with correlated errors: Earth and Planetary Science Letters, v. 5,
p. 320 324.

MANUSCRIPT RECEIVED BY THE SOCIETY DECEMBER 23, 1993


REVISED MANUSCRIPT RECEIVED JULY 30, 1994
MANUSCRIPT ACCEPTED SEPTEMBER 14, 1994

Printed in U.S.A.

424

Geological Society of America Bulletin, April 1995

You might also like