Quaternary adakite—Nb-enriched basalt association in the western

Transcription

Quaternary adakite—Nb-enriched basalt association in the western
Contrib Mineral Petrol
DOI 10.1007/s00410-007-0274-9
ORIGINAL PAPER
Quaternary adakite—Nb-enriched basalt association
in the western Trans-Mexican Volcanic Belt: is there any
slab melt evidence?
Chiara M. Petrone Æ Luca Ferrari
Received: 5 January 2007 / Accepted: 18 December 2007
Ó Springer-Verlag 2008
Abstract A spatial and temporal association between
adakitic rocks and Nb-enriched basalts (NEB) is recognised
for the first time in the western sector of the Trans-Mexican
Volcanic Belt in the San Pedro–Cerro Grande Volcanic
Complex (SCVC). The SCVC is composed of subalkalic
intermediate to felsic rocks, spanning in composition from
high-silica andesites to rhyolites, and by the young transitional hawaiite and mugearite lavas of Amado Nervo shield
volcano. Intermediate to felsic rocks of the SCVC show
many geochemical characteristics of typical adakites, such
as high Sr/Y ratios (up to 180) and low Y (\18 ppm) and Yb
contents. Mafic Amado Nervo rocks have high TiO2 (1.5–
2.3 wt%), Nb (14–27 ppm), Nb/La (0.5–0.9) and high
absolute abundances of HFSE similar to those shown by
NEB. However, the Sr and Nd isotopic signature of SCVC
rocks is different from that shown by typical adakites and
NEB. Although the adakites–NEB association has been
traditionally considered as a strong evidence of slab-melting,
we suggest that other processes can lead to its generation.
Here, we show that parental magmas of adakitic rocks of the
SCVC derive their adakitic characteristic from high-pressure
crystal fractionation processes of garnet, amphibole and
pyroxene of a normal arc basalt. On the other hand, Amado
Nervo Na-alkaline parental magmas have been generated by
Communicated by T.L. Grove.
C. M. Petrone (&)
CNR-Istituto Geoscience e Georisorse, Sezione di Firenze,
Via G. La Pira 4, 50121 Firenze, Italy
e-mail: [email protected]
L. Ferrari
Centro de Geociencias,
Universidad Nacional Autónoma de México,
Campus Juriquilla, Queretaro 76230, Mexico
sediment melting plus MORB-fluid flux melting of a heterogeneous mantle wedge, consisting of a mixture of depleted
and an enriched mantle sources (90DM + 10EM). We cannot exclude a contribution to the subduction component of
slab melts, because the component signature is dominated by
sediment melt, but we argue that caution is needed in interpreting the adakites–NEB association in a genetic sense.
Keywords Adakite–Nb-enriched basalt association Slab melts and fluids Sediment melts Western Mexico
Introduction
Adakites are characterised by high silica contents
([56 wt%), high Sr/Y and La/Yb ratios ([40 and [20,
respectively) and low Y and Yb contents (\15–18 ppm and
\1–1.5 ppm, respectively; e.g., Drummound and Defant
1990; Defant and Drummond 1990; Defant et al. 1992;
Peacock et al. 1994; Defant and Kepezhinskas 2001). In
recent years, there has been a considerable debate on the
significance of adakitic lavas found in several volcanic arcs
worldwide. Many authors have suggested that adakites
represent partial melting of subducted plates (e.g., Drummound and Defant 1990; Defant and Drummond 1990;
Defant et al. 1992; Peacock et al. 1994; Yogodzinski et al.
1994), whereas others proposed that they have negligible
connection to slab-melts as they derive their geochemical
signature from mixing between mafic and differentiated
lavas through crystal fractionation and/or crustal contamination processes (Castillo et al. 1999; Garrison and
Davidson 2003, Grove et al. 2005). Still others studies
demonstrate that they can be produced by crustal thickening
and/or eclogitization and delamination of the lowermost
crust (Kay and Kay 1993; Petford and Atherton 1996;
123
Contrib Mineral Petrol
Chung et al. 2003). The notable interest on adakites as slab
melt is probably related to the fact that they can be used as a
proxy to infer particular geodynamic settings. Indeed, the
high thermal regime required to obtain the melting of the
slab can be found in specific subduction settings, such as
subduction of very young (\25 Ma) oceanic crust (Peacock
et al. 1994), prolonged slab residence in the shallow mantle
as result of flat subduction (Gutscher et al. 2000), initiation
of subduction causing melting of the leading edges of newly
subducted slabs as a consequence of large thermal contrast
between the new slab and the mantle (Sajona et al. 1993),
subduction of an active ridge (Lagabrielle et al. 2000;
Aguillón-Robles et al. 2001).
Basalts are rarely found associated with adakites; when
this occurs they commonly have high Nb/La and are
defined as high-Nb and/or Nb-enriched basalts (NEB). The
association of adakites with NEB found in northern Kamchatka (Kepezhinskas et al. 1996), Panama and Costa Rica
(Defant et al. 1992), Cascades (Defant and Drummond
1993), and Baja California (Aguillón-Robles et al. 2001)
has been also considered to be an evidence for slab-melts.
As proposed by Kepezhinskas et al. (1996), slab melting
can generate both adakites and NEB in the same subduction zone, originating the so-called adakites–NEB
connection of Sajona et al. (1996). According to these
authors, adakitic liquids are formed by slab-melting at
depths of 70–100 km. Some of these are emplaced on the
Fig. 1 Schematic map of
western Mexico showing the
present plate boundaries and the
main volcanic and tectonic
features of the western TransMexican Volcanic Belt (heavy
grey). The study area, depicted
in Fig. 2, is also shown. SJ San
Juan volcano, PV Puerto
Vallarta, GDL Guadalajara, TZR
Tepic-Zacoalco rift, CR Colima
rift, ChR Chapala rift
123
surface with no interaction with the mantle wedge, giving
rise to adakites. In some cases, these adakitic liquids
derived by slab melting react with mantle peridotite. Later
melting of this metasomatised mantle produces NEB in the
mantle wedge. However, although the adakites–NEB
association seems to represent a solid evidence of slab-melt
phenomenon, others studies do not completely agree with
this view (Castillo et al. 2002; Mcpherson et al. 2006).
In this paper, we report for the first time the presence of
an adakite–NEB association in western Mexico, and discuss its possible origin. We show that this association does
not necessarily represent slab-melt evidence.
Geological setting
Western Mexico is characterised by the subduction of the
young Rivera micro-plate (*9 Ma, DeMets and Traylen
2000) beneath the North America plate (Fig. 1) with a
steep angle of *50° after 40 km of depth (Pardo and
Suarez 1995; Bandy et al. 1999). The Trans-Mexican
Volcanic Belt (TMVB) here starts at *180 km from the
trench and is characterized by intra-arc extension that
began in the late Miocene, at the time of the initial rifting
of the southern Gulf of California (Ferrari 1995), and
continued in Plio-Quaternary times possibly due to plate
boundary forces (Rosas-Elguera et al. 1996; Ferrari and
Contrib Mineral Petrol
Rosas-Elguera 2000). This part of the TMVB consists
mainly of abundant calc-alkaline rocks and minor alkaline
volcanics (e.g. Nelson and Carmichael 1984; Righter 2000;
Petrone et al. 2003). Quaternary adakites were previously
recognised by Luhr (2000) at the San Juan volcano and
other nearby monogenetic centres, all located in the westernmost part of the arc (Fig. 1). Luhr (2000) related the
occurrence of these adakites to the presence of the young
and hot Rivera slab beneath the region. Late Miocene
adakites were also reported in the eastern TMVB by
Gomez-Tuena et al. (2003) who suggested that they are
related to a period of flat subduction. More recently,
Gomez-Tuena et al. (2007) reconciled the presence of
Quaternary high-Mg andesites in the central TMVB with
the presence of slab and sediment melts interacting with the
mantle wedge.
Detailed geochemical and isotopic study in the western
TMVB led us to recognise an adakite–NEB-type association at the San Pedro–Cerro Grande Volcanic Complex
(SCVC), located to the east of San Juan volcano, in the San
Pedro–Ceboruco graben (Figs. 1, 2). Volcanism in the
graben consists of two stratovolcanoes (Ceboruco and
Tepetiltic) and several monogenetic centres aligned on its
north-eastern border, and by several cinder cones and
domes on its southwestern part (Ferrari et al. 2003)
(Fig. 2). A detailed account of the volcanic and magmatic
history of the SCVC was recently presented by Petrone
et al. (2006) and is briefly summarised here. The SCVC is
Fig. 2 Simplified geologic map
of the San Pedro–Ceboruco
graben based on Ferrari et al.
(2003) and Petrone et al. (2006)
123
Contrib Mineral Petrol
located in the central part of the Ceboruco graben and is
comprised several domes and minor pyroclastic rocks
ranging in composition from high-silica andesites to
rhyolites (Fig. 2). Volcanic activity can be subdivided into
two episodes separated by a caldera-forming event that
produced pyroclastic deposits. The San Pedro caldera is
partially covered by the young mafic lavas of Amado
Nervo shield volcano, which were erupted from a vent
located along the rim. Amado Nervo shield volcano consists of transitional (mildly Na-alkaline) hawaiite and
mugearite, and intermediate to felsic subalkalic rocks
(Fig. 3a). No genetic relationships are recognized between
subalkalic and transitional Na-alkalic rocks, which are
thought to represent different batches of magma from different mantle sources. Intermediate to felsic rocks can be
divided in amphibole-bearing and amphibole-free groups
(Fig. 3). These two groups of rocks are coeval but spatially
separated.
a
Petrography and geochemistry
The SCVC volcanic rocks are bimodal in composition,
with the transitional Na-alkalic (OIB-type, Petrone et al.
2003) Amado Nervo hawaiite to mugearite on one end, and
intermediate to felsic subalkalic rocks on the other end
(Fig. 3a). Amado Nervo lavas show seriate subporphyritic
textures with olivine ([5 vol%) being the only phenocryst
in a groundmass of plagioclase, olivine, opaques, and
clinopyroxene. Among intermediate to felsic rocks plagioclase and orthopyroxene are ubiquitous among
phenocrysts and microphenocrysts in both amphibolebearing and amphibole-free rocks, clinopyroxene is rare
and is almost exclusively found in the groundmass. The
two groups are also chemically and isotopically distinct.
The amphibole-bearing rocks are characterized by lower
Ce and other incompatible trace element contents compared to amphibole-free rocks at similar silica contents
(Fig. 3b). Amphibole-bearing rocks also show lower
87
Sr/86Sr (0.70382–0.70401) than amphibole-free rocks
(0.70411–0.70424) (Table 11 in Petrone et al. 2006)
(Fig. 4).
Amado Nervo rocks are enriched in the high field
strength elements (HFSE), and compatible elements, and
Rock/Prim Mantle
b
Fig. 3 a Total alkali versus silica classification diagram (Le Bas et al.
1992) for amphibole-free, amphibole-bearing and Amado Nervo
rocks of the San Pedro–Cerro Grande Volcanic Complex (data from
Petrone et al. 2006). The Irvine and Baragar (1971) line is also shown
(heavy dash line), b normalized incompatible element diagram for
amphibole-free (AF light grey pattern), amphibole-bearing (AB dark
grey pattern) and Amado Nervo (AN dashed line pattern) rocks
(Petrone et al. 2006). Normalization values for primordial mantle are
from Sun and McDonough (1995)
123
Fig. 4 eNd(N) versus 87Sr/86Sr(N) for amphibole-free, amphibolebearing and Amado Nervo rocks (data from Petrone et al. 2006).
MORB, OIB (White and Duncan 1995); island arc basalt (IAB)
(Wilson 1989) fields are also shown. Light grey field Western TransMexican Volcanic belt Na-alkaline rocks (Petrone et al. 2003), heavy
grey field calcalkaline rocks from western Trans-Mexican Volcanic
Belt (Verma and Nelson 1989a, b; Luhr 1997; Petrone et al. 2003):
Adakites field from Yogodzinski et al. 1995; Kepezhinskas et al.
1997; Aguillón-Robles et al. 2001; NEB field from Sajona et al. 1996;
Kepezhinskas et al. 1997; Aguillón-Robles et al. 2001. Mantle
reservoirs: EMI enriched mantle 1 and EMII enriched mantle 2 (White
1985). Other symbols as in Fig. 3. Modified from Petrone et al. 2003
Contrib Mineral Petrol
depleted in Ba and Rb with respect to intermediate-felsic
rocks (Fig. 3b). Hygromagmatophile element patterns
(Fig. 3b) are characterized by low LILE/HFSE and LREE/
HFSE ratios, intermediate between those of the sub-alkaline felsic rocks and oceanic-island basalt suites (i.e, OIBtype or ‘‘transitional series’’ of Petrone et al. 2003). Amado
Nervo rocks have lower 87Sr/86Sr (0.70351–0.70388) and
higher eNd values (4.3–5.7) with respect to all intermediate
to felsic subalkaline rocks (Table 11 in Petrone et al. 2006)
(Fig. 4).
Most of the intermediate to felsic rocks of the SCVC,
particularly the amphibole-bearing ones, show many
characteristics of typical adakites, such as high Sr/Y ratios
(up to 180) and low Y (\18 ppm) and Yb contents
(Fig. 5a). On the other hand, they show a low La/Yb ratio
(9–15; Table 10 in Petrone et al. 2006) that does not match
that of adakite. However, given the overall geochemical
similarity with adakites, the SCVC amphibole-bearing and
-free rocks can be defined as adakitic rocks sensu Castillo
(2006), i.e. rocks having chemical characteristics similar to
those of typical adakite regardless to their relation to slab
melt.
Mafic Amado Nervo rocks have high TiO2 (1.5–
2.3 wt%) and Nb (14–27 ppm) contents similar to those
shown by NEB from Vizcaino Peninsula of Baja California
(Aguillón-Robles et al. 2001) but lower than the NEB
found by Sajona et al. (1996) in Western Mindanao, which
commonly have Nb contents [20 ppm. According to
Kepezhinskas et al. (1997) and Defant and Kepezhinskas
(2001), NEB are characterised by high Nb/La ([0.5), high
absolute abundance of HFSE (Fig. 5b) and have an alkaline
or transitional alkaline character. The latter is shown by
Amado Nervo lavas. However, despite their Nb and other
HFSE enrichments as well as high Nb/La values (0.5–0.9),
Amado Nervo lavas display negative Nb anomalies with
Nb/Nb* values ranging from 0.5 to 0.8 (Nb* is calculated
as the expected value in the case of no enrichment or
depletion of Nb relative to adjacent elements, K and La, on
diagrams normalised to the primitive mantle value of Sun
and McDonough 1989). A similar observation was reported
by Kepezhinskas et al. (1996) for the Kamchatka NEBs,
which shows Nb/Nb* \ 1 and a typical arc depletions in
HFSE (Fig. 5b).
Amphibole-bearing and -free rocks have Sr and Nd
isotope ratios that fall within the field of the sub-alkaline
Western Trans-Mexican Volcanic Belt (WTMVB) rocks,
which have high and variable 87Sr/86Sr ratios and low eNd.
Amado Nervo rocks have higher 143Nd/144Nd but lower
87
Sr/86Sr values. The Sr and Nd isotopic signature of
SCVC rocks is different from that shown by typical
adakites and NEB (Kepezhinskas et al. 1997; Castillo et al.
1999; Aguillón-Robles et al. 2001) being higher in eNd ([6)
and having lower (although variable) 87Sr/86Sr (Fig. 4).
a
b
Fig. 5 a Sr/Y versus Y plot for the amphibole-free and amphibolebearing rocks (Petrone et al. 2006). Fields represent rocks filtered for
SiO2 [ 56% and are based on data from adakite and normal andesitedacite-rhyolite (Defant and Drummond 1990; Drummond et al. 1996);
Caminguin rock (Castillo et al. 1999); San Juan volcano (Luhr 2000);
Southern Volcanic Chain, Tepetiltic and Ceboruco volcanoes (Petrone 1998; Petrone et al. 2001); Colima volcano (Luhr and
Carmichael 1980; Allan 1986; Robin et al. 1991; Luhr 1992); Valle
de Bravo-Zitacuaro area (Gomez-Tuena et al. 2007); b La/Nb versus
Nb/Nb* for Amado Nervo rocks (Petrone et al. 2006); Na-alkaline
(down filled triangle) data from Petrone et al. (2003); NEB field is
based on data from Sajona et al. (1996), Kepezhinskas et al. (1997)
and Aguillón-Robles et al. (2001)
Discussion
Several authors proposed that adakite magmas represent
silicic melts produced by partial melting of subducted
basaltic rocks which have been metamorphosed to eclogite
or amphibolite at the depth of arc magma generation
(Drummound and Defant 1990; Defant and Drummond
1990; Defant et al. 1992; Peacock et al. 1994; Castillo
2006). Partial melting of such subducted basalts will give
melts characterised by a strong depletion in Y and Yb and
high Sr/Y and La/Yb if garnet or amphibole is left as
123
Contrib Mineral Petrol
residue. Other workers, however, put forward different way
to produce adakites magmas. Yogodzinski et al. (1994)
proposed an indirect link to the slab melt suggesting that
the adakites in Adak Island and the Piip-type magnesian
andesites were derived from a mantle wedge metasomatised by melts of subducted basalts. On the other hand,
Castillo et al. (1999) showed that the origin of adakitic
lavas from Caminguin Island (Philippines) was unrelated to
slab melting but it would rather come from a mixing of
mafic and differentiated magmas plus crystal fractionation
processes. Finally, Grove et al. (2005) argued that high-Mg
andesite and dacite from northern California were generated through fractional crystallization of H2O-rich mantle
melts.
The association of NEB with adakites in Kamchatka and
Philippines (Kepezhinskas et al. 1996; Sajona et al. 1996),
also called NEB-adakite connection (Sajona et al. 1996),
has been interpreted as an evidence for their petrogenetic
link. In this scenario, NEB melts are interpreted as due to
the melting, at depths of 70–100 km, of mantle wedge
peridotites previously metasomatized by adakite melts
(Kepezhinskas et al. 1996; Sajona et al. 1996). Given the
similarity in incompatible element contents and ratios, as
well as isotope ratios between NEB and OIB rocks, several
authors invoked the presence of an OIB-type component in
the sub-arc mantle source (e.g., Leeman et al. 1990;
Righter and Rosas-Elguera 2001; Castillo et al. 2002;
Petrone et al. 2003). In the rest of the paper, we discuss the
petrogenensis of the SCVC rocks, which seems an example
of the NEB-adakite connection.
The SCVC amphibole-free dacites are characterised by
Sr/Y values that mostly fall in the adakite field, along with
San Juan (Luhr 2000), Valle de Bravo-Zitacuaro (GomezTuena et al. 2007) and Caminguin rocks (Castillo et al.
1999), in the ‘‘classic’’ adakite Sr/Y versus Y plot (Fig. 5
a). A few of them have slightly higher Rb/Sr ratios compared to San Juan rocks and Valle de Bravo-Zitacuaro
(Luhr 2000; Gomez-Tuena et al. 2007) (Fig. 6a), which is
at odds with a slab-melt process, as it should produce low
Rb/Sr values (Luhr 2000). However, the Rb/Sr ratio is also
sensitive to crustal assimilation. Thus, the higher Rb/Sr
values shown by amphibole-free rocks might be due to
crustal assimilation processes (i.e., AFC process outlined in
Petrone et al. 2006), which might have overprinted the
slab-melt signature.
Amphibole-bearing rocks are characterised by Rb/Sr
values higher than those of the San Juan and Valle de
Bravo-Zitacuaro rocks and closer to the values shown by
the calc-alkaline rocks of Ceboruco and Tepetiltic volcanoes (Fig. 6a). Their Sr/Y (Fig. 5a) ratios are higher than
those of San Juan and Valle de Bravo-Zitacuaro rocks, and
are also similar to adakites (e.g. Defant and Drummond
1990). However, their Sr/Y ratios correlate negatively with
123
a
b
Fig. 6 Sr/Y versus Rb/Sr and SiO2 wt% (water-free basis) for the
amphibole-free (open circle) and amphibole-bearing rocks (filled
circle) (Petrone et al. 2006). The following field are also shown:
TMVB (Trans-Mexican Volcanic Belt) normal andesite-rhyolite
(Nelson and Hegre 1990; Wallace and Carmichael 1994; Mahood
1981; Petrone 1998); San Juan (Luhr 2000); SVA (Southern Volcanic
Chain) (Petrone 1998; Petrone et al. 2001); VBZ (Valle de BravoZitacuaro) (Gomez-Tuena et al. 2007), adakites (Kepezhinskas et al.
1997; Aguillón-Robles et al. 2001); Caminguin rock (Castillo et al.
1999). Rocks are filtered for SiO2 [ 56 wt%
SiO2 (Fig. 6b) and this is more akin to a normal arc
andesite-dacite-rhyolite than to an adakite.
The presence of amphibole is also noteworthy. Most
adakites are characterised by the presence of amphibole
and ubiquitous plagioclase. Amphibole is commonly found
in rocks located along the volcanic front of the western
Trans-Mexican Volcanic Belt, such as horneblende-bearing
andesite to rhyodacite of San Juan, Colima, Los Volcanes
and Mascota (Luhr and Carmichael 1980; Wallace and
Carmichael 1992), but it is rare or absent in the composite
volcanoes located in the rear part of the arc like Ceboruco
and the main vent of Tequila (Luhr 2000), with the
exception of some dacitic rocks from Tepetiltic volcano
(DeRemer 1986; Petrone 1998). Amphibole-bearing
dacites are found associated with the monogenetic
Contrib Mineral Petrol
magmatism at the southern edge of the San Pedro–Ceboruco graben (the southern volcanic chain of Petrone et al.
2001 and Ferrari et al. 2003) in close spatial association
with SCVC rocks (Fig. 2). These southern volcanic chain
dacites clearly represent the more evolved end-member of
a normal arc basalt–andesite–dacite series although they
have high Sr/Y ratios (Fig. 5a) similar to San Juan adakites
and SCVC amphibole-bearing and amphibole-free rocks.
Consequently, the link between the high Sr/Y, low Y and
Yb signature of amphibole bearing rocks and their slabmelt origin is not straightforward and perhaps dubious.
For a given silica content, the amphibole-bearing rocks
of San Juan and Colima are generally characterised by
lower pre-eruptive temperature, higher pre-eruptive water
contents, lower incompatible elements contents and higher
values of subduction-zone diagnostic incompatible element
ratios (e.g., Ba/La, K/La), with respect to the amphibolefree rocks of Ceboruco (Luhr 1992, 2000). Such a correlation between temperature, H2O and incompatible
elements contents has been attributed by Luhr (1992) to a
larger addition of subduction fluids to the mantle beneath
Colima volcano. The pre-eruptive temperatures of the
amphibole-bearing rocks in the SCVC are between 800 and
980°C on average (Petrone et al. 2006), which fall in the
range of variation of San Juan rocks, at the same silica
content. Amphibole-bearing and amphibole-free rocks
have incompatible element contents (e.g., Th, Ti, P, La)
similar to those shown by hornblende-bearing andesitedacite of San Juan and Colima, but amphibole-bearing
rocks have higher K/La and Ba/La in respect with amphibole-free rocks which might be suggestive of a greater
subduction-fluid addition in the former group. The depletion in Tb with increasing SiO2 observed for the latter
group is consistent with strong fractionation of amphibole
in a way similar to what observed for the volcanic-front
suites of San Juan and Colima (Luhr 2000).
In summary both the amphibole-bearing and amphibolefree rocks from SCVC have some characteristics of adakites. However, they share their high Sr/Y and low Y
signature with the southern volcanic chain amphibolebearing dacites, which belong to a normal basalt–dacite arc
series. Thus, no firm conclusion can be drawn about their
genesis, since our data seem to fit equally well in an
‘‘adakite model’’ (i.e., slab-melt) and a ‘‘typical arc magma
model’’ (i.e., fractional crystallization products of melt
from a subduction-fluid metasomatised mantle wedge). In
the first model, amphibole-bearing and amphibole-free
rocks are truly silicic slab-melts with limited or no interaction with the mantle-wedge, but are subsequently
modified by interaction with the crust. In this case, the least
differentiated (i.e., high-silica andesite and dacite) compositions should have the strongest adakite signatures
(highest Sr/Y) and should be the least modified since they
represent the closest derivatives from the adakite parental
magma. The less differentiated SCVC rocks apparently fit
the model since they have the highest Sr/Y ratios (especially the amphibole-bearing rocks) and, thus they have the
strongest adakite signatures (Fig. 6b). Following the classical model of adakite genesis (e.g., Defant and Drummond
1990), their parental magma originated from partial melting of an oceanic crust in amphibolite or eclogite facies.
Low degree partial melts of basalt with an eclogite residual
mineralogy are rich in SiO2, alkali and Sr, but poor in Y as
shown by SCVC amphibole-bearing rocks (grey dashed
line Fig. 7). However, as shown in Fig. 8a, this model fails
to match both the amphibole-bearing and amphibole-free
high-silica andesite and dacites trace element contents and
Fig. 7 Sr/Y versus Y (ppm) for the amphibole-free and amphibolebearing rocks (Petrone et al. 2006). Field of adakite and normal
andesite-dacite-rhyolite from Defant and Drummond (1990); Southern Volcanic Chain from Petrone (1998) and Petrone et al. (2001).
The thick black solid line with white stars illustrate fractional
crystallization of high pressure mineral assemblage from a Southern
Volcanic Chain basaltic melt (PM white star parental magma)
initially containing 691 ppm Sr and 26 ppm Y (sample SPC135,
Petrone 1998). The high-pressure assemblage is clinopyroxene,
garnet, amphibole and apatite in the proportions 65:27:8:0.5, based
on the equilibrium assemblage calculated with pMELTS Software
Package (Ghiorso et al. 2002) at 1.2 GPa, Tliq 1,180°C, fO2 at NNO
condition and 6H2O wt%. Partition coefficients from Luhr and
Carmichael (1980), Watson and Ryerson (1986), Bacon and Druitt
(1988), Ewart and Griffin (1994). White stars indicate the composition of the remaining melt fraction at each step. Grey dashed line with
multiplication symbol is partial melting of EPR basalt (grey filled dot)
with 103 ppm Sr and 44 ppm Y. These contents were calculated as
mean value of Sr and Y contents of EPR basalts from DSDP site 485a
leg 65 (Saunders 1982). Multiplication marks indicate the extension
of partial melting, which were calculated assuming a residual
mineralogy of garnet, clinopyroxene and rutile in the proportions
42:27:0.5. Partition coefficients from Rapp and Watson (1995)
123
Contrib Mineral Petrol
a
b
Fig. 8 Normalized incompatible element diagram showing the
results of slab melt (a) and HP fractionation model (b) compared
with normalized incompatible element diagram for less silica-rich
compositions of amphibole-free (AF light grey pattern) and amphibole-bearing (AB dark grey pattern) groups (Petrone et al. 2006).
Normalization values for primordial mantle are from Sun and
McDonough (1995). a Adakite model (i.e. slab melt): solid lines
are variation patterns at different extension of partial melting of EPR
basalts from DSDP site 485a leg 65 (Saunders 1982) assuming a
residual mineralogy of garnet, clinopyroxene and rutile in the
proportions 42:27:0.5, and partition coefficients from Rapp and
Watson (1995); b HP pressure fractionation model: solid lines with
stars are variation patterns at different fraction of residual melt
(F = 0.9: thick black line with white star; F = 0.6: thick dashed grey
line with grey stars) of a HP fractionation assemblage as resulted
from calculation with pMELTS Software Package (Ghiorso et al.
2002) at 1.2 GPa, Tliq 1,180°C, fO2 at NNO condition and 6 H2O
wt%. Fractionation assemblages are as follows: F = 0.9–0.8 cpx;
F = 0.7 cpx0.8:amph0.1:grt0.1; F = 0.6 cpx0.6:grt0.3:amph0.1:apatite0.5. Partition coefficients from Luhr and Carmichael (1980),
Watson and Ryerson (1986), Bacon and Druitt (1988), Ewart and
Griffin (1994)
patterns, strongly suggesting that SCVC high-silica
andesite rocks cannot be generated through eclogite melting. This is also confirmed by the isotopic data, which
suggest that SCVC high-silica andesite rocks were not
generated by melting of the East Pacific Rise (EPR) or any
MORB (Fig. 4) as their Sr and Nd isotopic data plot outside the EPR or MORB fields.
123
In the second model, amphibole-bearing and amphibolefree rocks are not slab melts but owe their adakitic signatures to crystal fractionation. Fractionation of amphibole
and/or garnet (or pyroxene) depletes the melt in Y relative
to Sr giving high Sr/Y, whereas fractionation of plagioclase
depletes the melt in Sr relative to Y, giving low Sr/Y
(Castillo et al. 1999). As also shown by Garrison and
Davidson (2003), high Sr/Y rocks do not necessarily
require an adakite parent, but they might be explained by
high-pressure (HP) fractionation of garnet, amphibole and
pyroxene. We modelled this HP fractionation starting from
a typical arc basalt of the Southern Volcanic Chain as a
starting composition. Major elements were modelled using
pMELTS Software Package (Ghiorso et al. 2002) and the
results are shown in Fig. 9. Chemical composition of less
evolved amphibole-bearing and amphibole-free dacites are
matched by 20–40% crystallization of a high-pressure
(12 kbar) assemblage constituted by 0.6cpx + 0.3grt +
0.1amph ± apatite (trace) under water-saturated condition.
Clinopyroxene is the first crystallizing phase, joined by
amphibole (from 20 to 30% of crystallization) and subsequently by garnet (from 26%) and few amount of apatite.
Residual liquid composition and amount of crystallizing
phases were evaluated in step of 10% of crystallization,
and the results were used to model trace elements using the
Rayleigh algorithm accordingly evaluating the total partition coefficients at each step. The results are shown in
Fig. 7 and 8b, where it is possible to see a very good match
between the HP fractionation model and trace element
variation patterns of both amphibole-bearing and amphibole-free high-silica andesite and dacites. Starting from a
typical arc basalt composition (PM in Fig. 7), fractionation
of clinopyroxene, garnet and amphibole will move liquid
compositions along a path of increasing Sr/Y and lowering
Y contents (black solid line in Fig. 7). The more garnet and
amphibole is fractionated, the more the Sr/Y ratio increases
(from point 0.8 onward in Fig. 7). This normal arc basalt
fractionation path is able to produce high Sr/Y values in
both groups (i.e., the parental magmas of SCVC rocks)
after 20–40% crystallization of the high-pressure assemblage. Subsequently, the liquid may follow a fractionation
path controlled more by plagioclase than by amphibole and
garnet, which will lower the Sr/Y ratios according to what
observed in the SCVC rocks. The initial fractionation of
garnet and amphibole indicates a deep environment, in
which plagioclase does not play a key role. Given the
higher Sr/Y of the parental magma of the amphibolebearing group of SCVC, compared to that of the amphibole-free group, it is possible to suggest that the former
underwent an extended or deeper fractionation.
The two different scenarios have important implication
for the genetic relationship between amphibole-bearing and
amphibole-free rocks. The adakite model prevents a
Contrib Mineral Petrol
Fig. 9 MgO and CaO versus TiO2 wt% diagrams for amphibole-free
(open circle) and amphibole-bearing (filled circle) (Petrone et al.
2006) showing the results of the HP pressure fractionation model
(normal arc basalt model). The thick black solid line with white stars
illustrates fractional crystallization of high-pressure mineral assemblage from the Southern Volcanic Chain basaltic melt (PM white star
parental magma) (sample SPC135, Petrone 1998). The high-pressure
assemblage is clinopyroxene, garnet, amphibole and apatite in the
proportions 65:27:8:0.5, based on the equilibrium assemblage calculated with pMELTS Software Package (Ghiorso et al. 2002) at 1.2
GPa, Tliq 1,180°C, fO2 at NNO condition and 6H2O wt%
comagmatic origin for the two groups of SCVC rocks,
whereas the typical arc magma model predicts a common
source for both groups of rocks. Note that the less differentiated rocks of both groups show very similar
compositions in major and trace elements, which strongly
points to a comagmatic origin. There are also differences
between the two groups (e.g. different Ce contents, different 87Sr/86Sr) but these differences can be explained by a
different crustal evolution and crustal contamination processes acting on a common parental magma (Petrone et al.
2006).
Therefore, based on the available petrographic, geochemical, and isotopic data, and on the comagmatic nature
of amphibole-bearing and amphibole-free SCVC rocks, we
conclude that their adakitic signature can be explained by
the typical arc magma model, in which these rocks differentiate from a common parental magma originating
from the metasomatised mantle wedge. Note that we cannot fully exclude melting of the slab, since it might be part
of the complex subduction component fluxing the mantle
wedge in this part of the western TMVB, which is characterised by variable hydrous fluid and sediment melt
(Petrone et al. 2003). As suggested by Gomez-Tuena et al.
(2003), subducted slab melt may accompany sediment melt
and they may interact in a complex way with the mantle
wedge, giving rocks with adakitic characteristics. In this
case the parental magma is derived from the mantle wedge
fluxed by sediment plus slab melt. However, the signature
of the slab melt is masked by that of sediment melt, and its
discrimination is not easy especially when trace element
and isotopic ratios of rocks are subsequently modified by
crustal contamination, as is the case of SCVC high silica
andesite-rhyolite rocks.
To further discriminate between the two models and
asses the possible presence of a slab melt it is useful to look
at the Amado Nervo mafic lavas which accompanied the
SCVC rocks. These Amado Nervo rocks resemble NEB,
which are considered to be derived from the melting,
at depths of 70–100 km, of mantle wedge peridotites
previously metasomatised by slab melts (Kepezhinskas
et al. 1996, Sajona et al. 1996). At the same time, NEB are
similar to OIB like rocks as shown by Luhr (1997) and
Ferrari et al. (2001). In particular, low LILE/HFSE and
LREE/HFSE basalts in Western Mexico range from
transitional (OIB-type, i.e. Amado Nervo) to truly OIB
(Na-alkaline basalts) (Petrone et al. 2003). The presence of
alkali basalts have been linked to partial melting of an
enriched mantle component (EM) advected from behind
the arc (Luhr 1997) or laterally introduced via corner flow
(Ferrari et al. 2001; Petrone et al. 2003) or a slab window
formation after detachment of the lower part of the slab
(Ferrari 2004). In addition, Cervantes and Wallace (2003)
have shown that in central Mexico magmas with low LILE
and LREE relative to HFSE have relative low H2O and
they have been formed by decompression melting of
unmodified mantle. Thus, the presence of OIB-type rocks
seems to represent a strong evidence against the presence
of slab melts or even sediment melts since these would
have elevated water contents. Neverthless, Petrone et al.
(2003) pointed out that the genesis of Amado Nervo rocks
can be reconciled with partial melting of a mixture of
depleted and enriched mantle source (95–90%DM +
5–10% EM) fluxed by fluid/melt of subducted sediment.
Indeed, Sr and Pb isotope ratios strongly suggest the
presence of a mixed DM-EM mantle component in the
genesis of Amado Nervo OIB-type basalts (see, Petrone
et al. 2003, Fig. 11), but geochemical and isotope
123
Contrib Mineral Petrol
characteristics also indicate the presence of a subduction
component. Their high Ba/Nb ratios (24–52) suggest the
involvement of fluid and, at the same time, the high
207
Pb/204Pb (Fig. 10) and 208Pb/204Pb ratios indicate a
sediment contribution to the genesis of these rocks,
whereas Sr and Nd isotope data are intermediate between
those of Na-alkaline and calc-alkaline rocks of the TMVB
and different from those of typical adakites and Nb-enriched basalts (Fig. 4). Thus, while isotope data do not
support an origin of Amado Nervo rocks from melting of a
mantle wedge metasomatised by slab melt, and slab melt
cannot be advocated as part of the subduction component
fluxing the mantle wedge; the sediment melt alone does not
explain the geochemical characteristics of Amado Nervo
rocks.
Insights into the nature of the subduction component can
be obtained by combining Nd/Pb and Nb/Y values with
isotope data (Fig. 11). Nd, Nb and Y are considered to be
fluid immobile, whereas Pb is mobile in aqueous fluids
(Brenan et al. 1995). A strong depletion in Y is considered
to be mainly associated with MORB (i.e., subducted basalt
crust) melt with the presence of residual garnet or amphibole (Castillo 2006). Nd and Nb budgets of magmas are
mainly controlled by mantle wedge plus subducted sediment (Elliot 1997; Class et al. 2000). Pb contribution is
mainly controlled by fluids (Class et al. 2000). Thus, high
Fig. 10 207Pb/204Pb versus 206Pb/204Pb plot for the Amado Nervo
rocks (upper open triangle) (Petrone et al. 2003) of the San Pedro–
Cerro Grande Volcanic Complex. White field with down filled black
triangle Trans-Mexican Volcanic Belt Na-alkaline rocks (Petrone
et al. 2003); TMVB calc-alkaline: Trans-Mexican Volcanic Belt calcalkaline rocks (Luhr et al. 1989; Luhr 1997, Petrone et al. 2003).
NHRL north hemisphere reference line (Hart 1984); Pacific oceanic
sediments from Church and Tatsumoto (1975) and Plank and
Langmuir (1998); Pacific alkaline seamounts from Graham et al.
(1988); EPR tholeiites from Smith (1999); NEB from Kepezhinskas
et al. (1997); Aguillón-Robles et al. (2001); adakite from Yogodzinski
et al. (1995); Kepezhinskas et al. (1997); Aguillón-Robles et al.
(2001); Caminguin rock from Castillo et al. (1999). Modified from
Petrone et al. (2003)
123
Pb/Nd ratios reflect fluid contributions, whereas high Nb/Y
ratios should reflect melt contributions of MORB or sediment. At the same time, as shown by Cervantes and
Wallace (2003), EM mantle source has high Nb/Y, as well
as high Nd and Nb. Neverthless, the low 143Nd/144Nd,
206
Pb/204Pb and Nd/Pb (Fig. 11) shown by Amado Nervo
transitional basalts indicate that they cannot be reconciled
solely by an EM source. In particular, there is no way to
explain their Nb/Y and 206Pb/204Pb lower than the EM
source without the involvement of a DM-EM mixed source
fluxed by a subduction component. The 206Pb/204Pb versus
Nd/Pb plot (Fig. 11a) clearly shows that the subduction
component is constituted by mixing of sediment melt and
fluids derived from dehydratation of subducted MORB (i.e,
MORB fluid), which is characterised by low Nd/Pb ratios
and 206Pb/204Pb composition similar to EPR tholeiites. The
presence of MORB fluid is also confirmed by combining
Pb and Nd isotope data (Fig. 11b). A subduction-related
MORB fluid added to sediment melt produces low
206
Pb/204Pb isotope composition without significantly
changing the Nd isotope composition of the sediment melt
(black dashed line in Fig. 11b). Moreover, because Nd is a
fluid-immobile element whereas Pb is highly mobile in
fluid, adding a slab melt to the sediment melt causes a
faster decrease in Nd than in Pb isotope composition (black
dotted line in Fig. 11b), which replicates that of the sediment + mantle wedge (heavy black line in Fig. 11b) but
Fig. 11 a 206Pb/204Pb versus Nd/Pb; b 206Pb/204Pb versus c
143
Nd/144Nd; c 206Pb/204Pb versus Nb/Y plots for the Amado Nervo
rocks (upper open triangle) (Petrone et al. 2003) of the San Pedro–
Cerro Grande Volcanic Complex. The following field are also shown:
TMVB (Trans-Mexican Volcanic Belt) sub-alkaline (Luhr and Carmichael 1980, 1985, 1990; Wallace and Carmichael 1994; Luhr 2000;
Petrone et al. 2003; Verma and Hasenaka 2004; Maldonaldo-Sanchez
and Schaaf 2005; Blatter et al. 2007; Gomez-Tuena et al. 2007);
TMVB Na-alkaline rocks (Luhr and Carmichael 1985; Gomez-Tuena
et al. 2003; Petrone et al. 2003; Verma and Hasenaka 2004; Blatter
et al. 2007); EPR tholeiites (Saunders 1982; Smith 1999); NEB
(Kepezhinskas et al. 1997; Aguillón-Robles et al. 2001); Whole
sediment composition (black circle) (McLennan et al. 1990). Mantle
component from Petrone et al. (2003) and as follows: DM (depleted
mantle) (White et al. 1987; Borg et al. 1997); and EM (enriched
mantle, down filled triangle) (Borg et al. 1997, Petrone et al. 2006).
Also shown is the composition of mantle wedge giving origin to the
Amado Nervo magmas as proposed by Petrone et al. (2003) and
constituted by mixing of 90% DM and 10% of EM component (small
grey field marked 90DM + 10EM). Small black field MORB fluid
(MF) composition form Class et al. (2000); small dark grey field
subducted MORB melt (slab melt) composition as proposed in this
study. Lines show the result of proposed mixing model: thin black line
mixing line between DM and EM mantle components; black heavy
dashed line mixing line between MORB fluid (MF) and sediment
melt; black heavy line mixing line between 90DM + 10EM mantle
source and sediment melt; black dot line mixing line between
sediment melt and slab melt; grey dashed line mixing line between
the mixed mantle source (90DM + 10EM) and the subduction
component as resulted from this study: mixing between sediment
melt (SM) and MORB fluid
Contrib Mineral Petrol
requires a subduction component with low Nd/Pb and low
143
Nd/144Nd. The latter is represented by the MORB fluid
plus sediment melt as confirmed by the 206Pb/204Pb versus
Nb/Y plot (Fig. 11c). Sediment alone cannot account for
the high Nb/Y of Amado Nervo rocks, which requires a
component with high Nb/Y. This can be represented either
by slab melt with residual amphibole or garnet or by
sediment melt with residual amphibole. Amado Nervo
rocks are characterised by low Nb/Ta ratios (12–14), close
to those shown by EPR tholeiites and they have relatively
low (Tb/Yb)N ratios (1.2–1.5), which excludes a significant
amount of residual garnet. In summary, OIB-like composition of Amado Nervo rocks can be matched by a model
involving sediment melt (with residual amphibole) plus
MORB fluid fluxing the Amado Nervo mixed mantle
source (Petrone et al. 2003) (grey-dashed line in Fig. 11).
Thus, a slab melt is not required and results confirm the
conclusions of Petrone et al. (2003), regarding the nature of
the subduction component.
a
Conclusions
b
c
A spatial and temporal association between adakitic rocks
and OIB-type rocks has been recognised for the first time in
the western sector of the TMVB in the San Pedro–Cerro
Grande Volcanic Complex. Contrary to a widely held
belief, the presence of this association does not necessarily
imply the occurrence of slab-melting. In fact, slab-melts, if
any, play a limited role in the genesis of both amphibolefree and amphibole-bearing adakitic rocks of the SCVC
and Amado Nervo OIB-type basalts. Parental magmas of
amphibole-free and amphibole-bearing adakitic rocks of
the SCVC derive their adakitic characteristic from highpressure crystal fractionation of clinopyroxene, garnet and
amphibole from a normal arc basalt composition. Similarly, Amado Nervo Na-alkaline parental magma have been
generated by sediment melt plus MORB-fluid flux melting
of a heterogeneous mantle wedge consisting of a mixture of
depleted and enriched mantle sources (90DM + 10EM).
Thus, although the adakite–OIB-type association is considered by others as a strong evidence of slab-melting, this
study suggests that other processes can lead to its generation. Therefore caution is needed in interpreting the genetic
significance of the adakites–OIB-type association.
Acknowledgments The authors would like to thank Pat Castillo,
Jim Luhr and Lorella Francalanci for thoughtful comments on early
drafts of the manuscript. A special thought is devoted to Jim Luhr by
C.M.P. Our fruitful discussions on the Mexican arc at Carnegie and
Smithsonian are still alive in my mind. Andrea Orlando is thanked for
his precious advices in several occasions. We thank Paul Wallace and
an anonymous reviewer for their criticisms that led to significant
improvements in the manuscript. Editorial handling by Professor
Timothy L. Grove is also highly appreciated. Our work was supported
by a CNR (Italy)–CONACyt (Mexico) bilateral grant.
References
Aguillón-Robles A, Calmus T, Benoit M, Bellon MH, Maury RC,
Cotten J, Bourgois J, Michaud F (2001) Late Miocene adakites
and Nb-enriched basalts from Vizcaino Peninsula, Mexico:
indicators of East Pacific Rise subduction below southern
California? Geology 19(6):531–534
123
Contrib Mineral Petrol
Allan JF (1986) Geology of the Colima and Zacoalco grabens, SW
Mexico: Late Cenozoic rifting in the Mexican Volcanic Belt.
Geol Soc Am Bull 97:473–485
Bacon CR, Druitt TH (1988) Compositional evolution of the Zoned
Calcalkaline Magma Chamber of Mount-Mazama, Crater Lake,
Oregon. Contrib Mineral Petrol 98(2):224–256
Bandy W, Kostoglodov V, Hurtado-Diaz A, Mena M (1999) Structure
of the southern jalisco subduction zone, Mexico, as inferred from
gravity and seismicity. Geof Int 38:127–136
Blatter DL, Lang Farmer G, Carmichael ISE (2007) A north–south
transect across the Central Mexican Volcanic Belt at *100°W:
spatial distribution, petrological, geochemical, and isotopic
characteristics of Quaternary volcanism. J Petrol 48(5):901–950
Borg LE, Clynne MA, Bullen TD (1997) The variable role of slabderived fluids in the generation of a suite of primitive
calcalkaline lavas from the southernmost Cascades, California.
Can Mineral 35:425–452
Brenan JM, Shaw HF, Ryerson FJ, Phinney DL (1995) Experimental
determination of trace-element partitioning between pargasite
and a synthetic hydrous andesitic melt. Earth Planet Sci Lett
135:1–11
Castillo PR (2006) An overview of adakite petrogenesis. Chin Sci
Bull 51(1):1–12
Castillo PR, Janney PE, Solidum RU (1999) Petrology and geochemistry of Caminguin Island, southern Philippines: insights to the
source of adakites and other lavas in a complex arc setting.
Contrib Mineral Petrol 134:33–51
Castillo PR, Solidum RU, Punongbayan RS (2002) Origin of high
field strength element enrichment in the Sulu Arc, Southern
Philippines, revisited. Geology 30:707–710
Cervantes P, Wallace PJ (2003) Role of H2O in subduction-zone
magmatism: new insights from melt inclusions in high-Mg
basalts from central Mexico. Geology 31:235–238
Chung SL, Liu D, Ji J, Chu MF, Lee HY, Wen DJ, Lo CH, Lee TY,
Qian Q, Zhang Q (2003) Adakites from continental collision
zones: melting of thickened lower crust beneath southern Tibet.
Geology 31:1021–1024
Church SE, Tatsumoto M (1975) Lead isotope relations in oceanic
ridge basalts from the Juan de Fuca-Gorda ridge area, N.E.
Pacific Ocean. Contrib Mineral Petrol 53(4):253–279
Class C, Miller DM, Goldstein SL, Langmuir CH (2000) Distinguishing melt and fluid subduction components in Umnak
Volcanics, Aleutian Arc. Geochem Geophys Geosyst 1. doi:
1999GC000010
Defant MJ, Drummond MS (1990) Derivation of some modern arc
magmas by melting of young subducted lithosphere. Nature
347:662–665
Defant MJ, Drummond MS (1993) Mount St. Helens: potential
example of the partial melting of the subducted lithosphere in a
volcanic arc. Geology 21:547–550
Defant MJ, Kepezhinskas P (2001) Evidence suggests slab melting in
arc magmas. Eos Trans AGU 82(6):65–68
Defant MJ, Jackson TE, Drummond MS, de Boer JZ, Bellon H,
Feigenson MD, Maury RC, Stewart RH (1992) The geochemistry of young volcanism throughout western Panama and
southern Costa Rica, an overview. Geol Soc Lond J 149:569–579
DeMets C, Traylen S (2000) Motion of the Rivera plate since 10 Ma
relative to the Pacific and North American plates and the mantle.
Tectonophysics 318:119–159
Deremer LA (1986) The geologic and chemical evolution of Volcan
Tepetiltic, Nayarit, Mexico. MS thesis, Univ. of Tulane, Tulane,
pp 158
Drummound MS, Defant MJ (1990) A model for trondhjemite-tonalitedacite genesis and crustal growth via slab melting: Archean to
modern comparisons. J Geophys Res 95:21503–21521
123
Drummond MS, Defant MJ, Kepezhinskas PK (1996) Petrogenesis of
slab-derived trondhjemite-tonalite-dacite/adakite magmas. In:
Brown M et al (eds) The third Hutton symposium on the origin
of granites and related rocks. Geol. Soc. Am. Sp. Paper, vol 315,
pp 205–215
Elliot T (1997) Fractionation of U and Th during mantle melting: a
reprise. Chem Geol 139:165–183
Ewart A, Griffin WL (1994) Application of Proton-Microprobe Data
to Trace-Element Partitioning in Volcanic-Rocks. Chem Geol
117(1–4):251–284. doi:10.1016/0009-2541(94)90131-7
Ferrari L (1995) Miocene shearing along the northern bundary of the
Jalisco block and the opening of the Southern Gulf of California.
Geology 23:751–754
Ferrari L (2004) Slab detachment control on mafic volcanic pulse and
mantle heterogeneity in central Mexico. Geology 32:77–80
Ferrari L, Rosas-Elguera J (2000) Late Miocene to Quaternary
extension at the northern boundary of the Jalisco block, western
Mexico: the Tepic-Zacoalco rift revised. In: Delgado-Granados
H, Aguirre-Diaz G, Stock JM (eds) Cenozoic Tectonics and
Volcanism of Mexico. Geol. Soc. Am. Sp. Paper, vol 334(03),
pp 41–64
Ferrari L, Petrone CM, Francalanci L (2001) Generation of oceanicisland basalt-type volcanism in the western Trans-Mexican
volcanic belt by slab rollback, asthenosphere infiltration, and
variable flux-melting. Geology 29:507–510
Ferrari L, Petrone CM, Francalanci L, Tagami T, Eguchi M,
Conticelli S, Manetti P, Venegas-Salgado SV (2003) Geology
of the San Pedro–Ceboruco graben, western Trans Mexican
Volcanic Belt. Rev Mex Cienc Geolog 20(3):165–181
Garrison JM, Davidson JP (2003) Dubious case for slab melting in the
Northern volcanic zone of the Andes. Geology 31(6):565–568
Ghiorso, Mark S, Hirschmann, Marc M, Reiners, Peter W, Kress,
Victor C III (2002) The pMELTS: a revision of MELTS aimed at
improving calculation of phase relations and major element
partitioning involved in partial melting of the mantle at pressures
up to 3 GPa. Geochem Geophys Geosyst 3(5). doi:10.1029/
2001GC000217
Gomez-Tuena A, La Gatta AB, Langmuir CH, Goldstein SL, OrtegaGutiérrez F, Carrasco-Núnez G (2003) Temporal control of
subduction magmatism in the Eastern Trans-Mexican Volcanic
Belt: mantle sources, slab contributions and crustal contamination.
Geochem Geophys Geosyst 4(8). doi:10.1029/2003GC000524
Gomez-Tuena A, Langmuir CH, Goldstein SL, Straub SM, OrtegaGutierrez F (2007) Geochemical evidence for slab-melting in the
Trans-Mexican Volcanic Belt. J Petrol 48(3):537–562
Graham DW, Zindler A, Kurz MK, Jenkins WJ, Batiza R, Staudigel H
(1988) He, Pb, Sr and Nd isotope constraints on magma genesis
and mantle heterogeneity beneath young Pacific seamounts.
Contrib Mineral Petrol 99:446–463
Grove TL, Baker MB, Price RC, Parman SW, Elkins-Tanton LT,
Chatterjee N, Müntener O (2005) Magnesian andesite and dacite
lavas from Mt. Shasta, northern California: products of fractional
crystallization of H2O-rich mantle melts. Contrib Mineral Petrol
148:542–565
Gutscher MA, Spakman W, Bijwaard H, Engdahl ER (2000)
Geodynamics of flat subduction: seismicity and tomographic
constraints from the Andean margin. Tectonics 19(5):814–833
Hart SR (1984) A large scale isotope anomaly in the Southern
Hemisphere mantle. Nature 309:753–757
Kay RW, Kay SM (1993) Delamination and delamination magmatism. Tectonophysics 219:177–189
Kepezhinskas P, Defant MJ, Drummond MS (1996) Progressive
enrichement of island arc mantle by melt-periodotite interaction
inferred from Kamchatcka xenoliths. Geoch Cosmoch Acta
60(7):1217–1229
Contrib Mineral Petrol
Kepezhinskas P, McDermott F, Defant MJ, Hochstaedter A, Drummond MS, Hawkesworth CJ, Koloskov A, Maury RC, Bellon H
(1997) Trace element and Sr–Nd–Pb isotopic constraints on a
three-component model of Kamchatka arc petrogenesis. Geoch
Cosmoch Acta 61(3):577–600
Irvine TN, Baragar WRA (1971) A guide to the chemical classification of the common rocks. Can J Earth Sci 8:523–548
Lagabrielle Y, Guivel C, Maury RC, Bourgois J, Fourcade S, Martin
H (2000) Magmatic-tectonic effects of high-thermal regime at
the site of active ridge subduction: the Chile Triple Junction
model. Tectonophysics 326(3, 4):255–268
Le Bas MJ, Le Maitre RW, Woolley AR (1992) The construction of
the Total Alkali-Silica chemical classification of the volcanic
rocks. Mineral Petrol 46:1–22
Leeman WP, Smith DR, Hildreth W, Palacz Z, Rogers N (1990)
Compositional diversity of late Cenozoic basalts in a transect
across the southern Washington Cascades: implications for
subduction zone magmatism. J Geophys Res 95(B12):19561–
19582
Luhr JF (1992) Slab-derived fluids and partial melting in subduction
zones: insights from two contrasting Mexican volcanoes (Colima
and Ceboruco). J Volcanol Geotherm Res 54:1–18
Luhr JF (1997) Extensional tectonics and the diverse primitive
volcanic rocks in the western Mexican Volcanic Belt. Can
Mineral 35:473–500
Luhr JF (2000) The geology and petrology of Volcan San Juan
Nayarit, Mexico and the compositionally zoned Tepic Pumice.
J Volcanol Geotherm Res 95:109–156
Luhr JF, Carmichael ISE (1980) The Colima Volcanic Complex,
Mexico: I. Post-caldera andesites from Volcán Colima. Contrib
Mineral Petrol 71:343–372
Luhr JF, Carmichael ISE (1985) Jorullo Volcano, Michocoan, Mexico
(1759–1774): the earliest stages of fractionation in calc-alkaline
magmas. Contrib Mineral Petrol 90:142–161
Luhr JF, Carmichael ISE (1990) Petrological monitoring of cyclical
eruptive activity at Volcan Colima, Mexico. J Volcanol
Geotherm Res 23:235–260
Luhr J, Allan J, Carmichael ISE, Nelson SA, Hasenaka T (1989)
Primitive calc-alkaline and alkaline rock type from the western
Mexican Volcanic Belt. J Geophys Res 94:4515–4530
Mahood G (1981) A summary of the geology and petrology of the
Sierra La Primavera, Jalisco, Mexico. J Geophys Res 86:10137–
10152
Maldonaldo-Sanchez G, Schaaf P (2005) Geochemical and isotope
data from the Acatlán Volcanic Field, western Trans-Mexican
Volcanic Belt: origin and evolution. Lithos 82:455–470
McLennan SM, Taylor SR, McCulloch MT, Maynard JB (1990)
Geochemical and Nd–Sr isotopic composition of deep sea
turbidites: crustal evolution and plate tectonic associations.
Geochim Cosmochim Acta 34:2015–2050
Mcpherson CG, Dreher ST, Thirlwall MT (2006) Adakites without
slab melting: high pressure differentiation of island arc
magma, Mindanao, the Philippines. Earth Planet Sci Lett
BI:581–593
Nelson SA, Carmichael ISE (1984) Pleistocene to recent alkalic
volcanism in the region of Sanganguey volcano, Nayarit,
Mexico. Contrib Mineral Petrol 85:321–335
Nelson SA, Hegre J (1990) Volcan Las Navajas, a Pliocene–
Pleistocene trachyte–peralkaline rhyolite volcano in the northwestern Mexican Volcanic Belt. Bull Volcanol 52:186–204
Pardo M, Suarez G (1995) Shape of the subducted Rivera and Cocos
plates in the southern Mexico: seismic and tectonic implications.
J Geophys Res 100(B7):12357–12373
Peacock SM, Rushmer T, Thompson AB (1994) Partial melting of
subduction oceanic crust. Earth Planet Sci Lett 121:227–244
Petford N, Atherton MP (1996) Na-rich partial melts from newly
underplated basaltic crust: the Cordillera Blanca Batholith, Peru.
J Petrol 37:1491–521
Petrone CM (1998) Studio magmatologico dei sistemi vulcanici nel
graben San Pedro–Ceboruco (Nayarit, Messico): coesistenza di
magmi a diversa affinità petrologica (in Italian). PhD thesis,
University of Florence, Florence, Italy, p 280
Petrone CM, Tagami T, Francalanci L, Matsumura A, Sudo M (2001)
Volcanic systems in the San Pedro–Ceboruco graben (Nayarit,
Mexico) in the light of new K–Ar geochronological data.
Geochem J 35:77–88
Petrone CM, Francalanci L, Carlson RW, Ferrari L, Conticelli S
(2003) Unusual coexistence of subduction-related and intraplatetype magmatism: Sr, Nd and Pb isotope and trace elements data
from the magmatism of the San Pedro–Ceboruco graben
(Nayarit, Mexico). Chem Geol 193:1–24
Petrone CM, Francalanci L, Ferrari L, Schaaf P, Conticelli S (2006)
The San Pedro–Cerro Grande Volcanic Complex (Nayarit,
Mexico): inferences on volcanology and magma evolution. In:
Siebe C, Aguirre-Dı̀az G, Macı̀as JL (eds) Neogene-Quaternary
continental margin volcanism: a perspective form Mexico. Geol
Soc Am Sp Paper, vol 402(03), pp 65–98
Plank T, Langmuir CH (1998) The chemical compositon of subducting sediment and its consequences for the crust and mantle.
Chem Geol 145:325–394
Rapp RP, Watson EB (1995) Dehydration melting of metabasalt at 8–
32 kbar: implications for continental growth and crust-mantle
recycling. J Petrol 36:891–931
Righter K (2000) A comparison of basaltic volcanism in the Cascades
and western Mexico: compositional diversity in continental arcs.
Tectonophysics 318:99–117
Righter K, Rosas-Elguera J (2001) Alkaline Lavas in the Volcanic
Front of the Western Mexican Volcanic Belt: Geology and
Petrology of the Ayutla and Tapalpa Volcanic Fields. J Petrol
42:2333–2361
Robin R, Camus G, Gougaurd A (1991) Eruptive and magmatic
cycles at Fuego de Colima volcano. J Volcanol Geotherm Res
45(3, 4):209–225
Rosas-Elguera J, Ferrari L, Garduño-Monroy VH, Urrutia-Fucugauchi J (1996) Continental boundaries of the Jalisco block and their
influence in the Pliocene-Quaternary kinematics of western
Mexico. Geology 24(10):921–924
Sajona FG, Maury RC, Bellon H, Cotton J, Defant MJ, Pubellier M
(1993) Initiation of subduction and the generation of slab melts
in western and eastern Mindanao, Philippines. Geology
21:1007–1110
Sajona FG, Maury RC, Bellon H, Cotten J, Defant M (1996) High
field strength element enrichment of Pliocene-Pelistocene island
arc basalts, Zomboanga Peninsula, Western Mindanao Philippines. J Petrol 37:693–726
Saunders AD (1982) Geochemistry of basalts recovered from the Gulf
of California during LEG65 of the deep sea drilling project. In:
Lewis BTR, Robinson P et al (eds) Init. Repts. DSDP 65i (U.S.
Govt. Printing Office), Washington, pp 591–621
Smith AD (1999) The Nd–Sr–Pb isotopic record in abyssal tholeiites
from the Gulf of California region, Western Mexico: no evidence
for a gulf mouth plume. Int Geol Rev 41:921–931
Sun SS, McDonough WF (1995) The composition of the Earth. Chem
Geol 120(3–4):223–253
Sun SS, McDonough WF (1989) Chemical and isotopic systematics
of oceanic basalts: implications or mantle composition and
processes. In: Saunders AD, Norry MJ (eds) Magmatism in the
ocean basins, Geol Soc Sp Pubbl vol 42, pp 313–346
Verma SP, Nelson SA (1989a) Isotopic and trace element constraints
on the origin and evolution of alkaline and calc-alkaline magmas
123
Contrib Mineral Petrol
in the northwestern Mexican Volcanic Belt. J Geoph Res
94(B4):4531–4544
Verma SP, Nelson SA (1989b) Correction to ‘‘Isotopic and trace
element constraints on the origin and evolution of alkaline and
calc-alkaline magmas in the northwestern Mexican Volcanic
Belt’’. J Geophys Res 96(B6):7679–7681
Verma SP, Hasenaka T (2004) Sr, Nd, and Pb isotopic and trace
element geochemical constraints for a veined-mantle source of
magmas in the Michoacán–Guanajuato Volcanic Field, westcentral Mexican Volcanic Belt. Geoch J 38:43–65
Wallace PJ, Carmichael ISE (1992) Alkaline and calc-alkaline lavas
near Los Volcanes, Jalisco, Mexico: geochemically diversity
and its significance in volcanic arcs. Contrib Mineral Petrol
111:423–439
Wallace PJ, Carmichael ISE (1994) Petrology of Volcán Tequila,
Jalisco, Mexico: disequilibrium phenocryst assemblages and
evolution of the subvolcanic magma system. Contrib Mineral
Petrol 117:345–361
Watson EB, Ryerson FJ (1986) Partitioning of Zirconium between Clinopyroxene and Magmatic Liquids of Intermediate
123
Composition. Geoch Cosmoch Acta 50(11):2523–2526
doi:10.1016/0016-7037(86)90035-9
White WM (1985) Sources of oceanic basalts: radiogenic isotope
evidence. Geology 13:115–118
White WM, Duncan RA (1995) Geochemistry and geochronology of
the Society island: new evidences for deep mantle recycling. In:
Basu A, Hart SR (eds) Isotope Studies of Crust–Mantle Evolution.
AGU, Washington, DC, Geophys Monogr, vol 95, pp 1–23
White WM, Hofmann AW, Puchelt H (1987) Isotope geochemistry of
Pacific Mid-Ocean Ridge basalt. J Geophys Res 92(B6):4881–
4893
Wilson M (1989) Igneous Petrogenesis. Chapman & Hall, London
Yogodzinski GM, Kay RW, Volynets ON, Koloskov AV, Seliverstov
NI, Matvenkov VV (1994) Magnesian andesites and the
subduction component in a strongly calc–alkaline series at Piip
volcano, far western Aleutians. J Petrol 35:163–204
Yogodzinski GM, Kay RW, Volynets ON, Koloskov AV, Kay SM
(1995) Magnesian andesite in the western Aleutian Komandorsky region: implications for slab melting and processes in the
mantle wedge. Geol Soc Am Bull 107:505–519