You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329677217

Mantle driven cretaceous flare-ups in Cordilleran arcs

Article  in  Lithos · December 2018


DOI: 10.1016/j.lithos.2018.12.007

CITATIONS READS

5 398

5 authors, including:

Ana Martinez Scott R. Paterson


Loma Linda University University of Southern California
12 PUBLICATIONS   31 CITATIONS    291 PUBLICATIONS   6,448 CITATIONS   

SEE PROFILE SEE PROFILE

Vali Memeti Pablo G. Molina


California State University, Fullerton University of Chile
105 PUBLICATIONS   560 CITATIONS    12 PUBLICATIONS   19 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Magmatic Processes in the Cretaceous Peruvian Coastal Batholith View project

Geochemistry and Geochronology of Mantle-derived mafic magmas from the American Cordillera View project

All content following this page was uploaded by Ana Martinez on 23 December 2018.

The user has requested enhancement of the downloaded file.


Lithos 326–327 (2019) 19–27

Contents lists available at ScienceDirect

Lithos

journal homepage: www.elsevier.com/locate/lithos

Mantle driven cretaceous flare-ups in Cordilleran arcs


Ana María Martínez Ardila a,b,⁎, Scott R. Paterson a, Vali Memeti b, Miguel A. Parada c, Pablo G. Molina c
a
Department of Earth Sciences, University of Southern California, 3651 Trousdale Parkway, Los Angeles, CA 90089, United States
b
Department of Geological Sciences, California State University, Fullerton, 800 N State College Blvd, CA 92831, United States
c
Departamento de Geología, Centro de Excelencia en Geotermia de los Andes (CEGA), Universidad de Chile, Plaza Ercilla 803, Santiago, Chile

a r t i c l e i n f o a b s t r a c t

Article history: Continental arcs display episodic magmatism characterized by flare-ups and lulls. Models published to explain
Received 9 July 2018 these patterns invoke (1) upper plate crustal processes driven by internal feedback; (2) episodic mantle melting
Accepted 7 December 2018 processes, or (3) external lower plate tectonic events.
Available online 15 December 2018
This study addresses the role of mantle magmas during flare-ups in Mesozoic Cretaceous continental arcs using
geochronological and geochemical data for three Cretaceous arc segments: the western Peninsular Ranges Bath-
Keywords:
Mantle
olith (wPRB), the Peruvian Coastal Batholith (PCB), and the Chilean Coastal Batholith (CCB).
Crust In all three arc segments, bedrock zircon age patterns defining a flare-up from ~125 to 90 Ma characterized by
Flare-up gabbro to granite units with Sri b 0.705, ƐNd from 0 to +7.5, 208Pb/204Pb from 38.2 to 38.7, and 206Pb/204Pb
Arcs from 18.3 to 18.7. These values project well towards a depleted mantle source. Areal measurements show that
Isotopes gabbro forms ~18% (wPRB), ~24% (PCB), and ~10% (CCB) of exposed plutonic material. AFC modeling indicates
Assimilation that these magmas have experienced fractional crystallization with only minor crustal assimilation (b20–30%),
implying that the great majority of these magmas are mantle-derived. Thus the cause of these flare-ups must
be episodic mantle processes: crustal melting was not required for triggering the flare-up, and only played a sec-
ondary role in modifying melt compositions. It remains unclear if the episodic mantle processes reflect internal
feedback(s) or external tectonically driven processes.
© 2018 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction 2015; Cao et al., 2016), and/or external forcing by tectonic events out-
side the arc (Hughes and Mahood, 2008; Zellmer, 2008; de Silva et al.,
Studies of continental magmatic arcs in the North and South 2015). Large, linked geochemical and geochronological data sets are
American Cordilleras have recognized patterns of episodic magmatism. needed to expand our knowledge about the non-steady-state magmatic
Episodicity is characterized by periods of high magma addition rates activity of continental arcs, to test the validity of the proposed models,
(MARs) called flare-ups that are separated by periods of low MARs, and to better evaluate the role that mantle-crust interactions play.
called lulls (Ducea, 2001; DeCelles et al., 2009; de Silva et al., 2015; Mantle melting drives magmatism at mid-oceanic ridges, hot spots
Paterson and Ducea, 2015; Kirsch et al., 2016). Episodic patterns are and ocean island arcs, but its role in driving magmatism in continental
documented either by (1) the relative abundance of igneous bedrock margin arcs has remained contentious. In this paper, we first briefly in-
and their ages (Condie et al., 2011; Ducea et al., 2015; Paterson and troduce proposed models that attempt to explain continental arc flare-
Ducea, 2015; Kirsch et al., 2016), (2) the relative abundances of different ups and then explore in more detail the role that mantle magma input
detrital zircon ages in sediments derived from the arc, or (3) volume es- plays. We examine the geochemical and geochronological evolution of
timates of plutonic and volcanic rocks produced during a flare-up versus mafic versus felsic plutonic rocks formed during the Late Cretaceous,
a lull (Paterson and Ducea, 2015; Kirsch et al., 2016). the volumetrically largest Mesozoic Cordilleran flare-up in the western
Why magmatism in arcs is episodic remains an unresolved and Peninsular Ranges Batholith (wPRB) in California, the Peruvian Coastal
exciting question. Episodicity may reflect a combination of internal Batholith (PCB) and the Central Chilean Coastal Batholith (CCCB).
feedback processes in the upper plate, mantle, or crust within the arc Chemical data suggest that mantle-derived magmas drive flare-ups in
(Ducea, 2001; DeCelles et al., 2009; Chin et al., 2015; DeCelles et al., these arc segments with moderate to little amounts of crustal materials
involved, implying that mantle processes drove these flare-ups. A com-
⁎ Corresponding author. parison to other subduction-related arc segments suggests that mantle
E-mail: amartinez-ardila@fullerton.edu (A.M. Martínez Ardila). processes may commonly be the dominant driver of flare-ups and that

https://doi.org/10.1016/j.lithos.2018.12.007
0024-4937/© 2018 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
20 A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27

crustal cyclic (?) processes largely play a second-order effect in the de- oceanic arcs strongly supports the idea that flare-ups can be driven by
gree to which the mantle signature is modified. mantle processes without crustal involvement (Ishizuka et al., 2011;
Reagan et al., 2013; Jicha and Jagoutz, 2015).
2. Review of flare-up models
3. Geological setting
Several models have been published in the last decade to explain
cyclic and episodic patterns of magmatism and deformation in arcs. The Cretaceous wPRB, PCB, and CCCB, arc segments of the North and
Models invoke either internal feedback driven by intra-arc cyclic pro- South American Cordillera provide examples of continental margin arc
cesses fairly independent of plate motions, or external controls caused flare-ups. In the following sections, the tectonic and magmatic history
by events outside the arc and driven by plate tectonic processes of the selected Cordilleran arc sections is briefly summarized. Isotopes
(Haschke et al., 2006; DeCelles et al., 2009; Chapman et al., 2013; and age compilations of the three arc segments presented in this
Ducea et al., 2015; Paterson and Ducea, 2015; Kirsch et al., 2016). paper (Figs. 1 and 2) are focused on the Late Cretaceous with particular
A popular crustal “internal feedback” model originally described by emphasis on documenting the mantle signature and reservoirs.
Ducea (2001), Ducea and Barton (2007) and DeCelles et al. (2009,
2015) involves periodic episodes of high flux arc magmatism fueled 3.1. Western Peninsular Ranges Batholith (35–33° N)
by underthrusting foreland lithosphere. In this model, underthrusting
drives increased melting and arc magmatism and, along with associated The northern part of the Cretaceous Peninsular Ranges Batholith be-
tectonism, produces thick crust, mountains, and dense melt residues tween about 33° and 35°N is divided into western and eastern zones
that during crustal thickening change to eclogite (Ducea, 2001; with distinct ages and chemistry (Baird et al., 1979; Kistler et al., 2014;
Sobolev et al., 2006). When the eclogite residues reach a critical mass, Kimbrough et al., 2015). The numerous western PRB plutons (and some
they founder into the mantle causing isostatic adjustment of the topo- volcanic rocks) emplaced in an oceanic arc setting have ages of ca.
graphic surface and rapid regional elevation gain in response to removal 134–108 Ma. The many central PRB plutons have ages of ca. 111–93 Ma
of the dense root (Garzione et al., 2006; Pelletier et al., 2010). Increased and record the west to east progression of subduction transitioning
topography drives renewed rapid foreland underthrusting that re- from an oceanic to a continental arc setting. The few large eastern PRB
charges the magmatic system, and the cycle is rejuvenated. plutons were emplaced in a continental arc setting at ca. 98–91 Ma
External control models involve events outside the crustal arc col- (Hildebrand and Whalen, 2014b; Morton et al., 2014). The western
umn, such as plate reconfigurations and changes in mantle flow and zone (wPRB) is characterized by gabbro (~14%) and a wide range of
magma production (e.g., Hughes and Mahood, 2008; Zellmer, 2008; more siliceous compositions with primitive island-arc geochemical af-
de Silva et al., 2015). Specifically, if arc systems are controlled by param- finities (Silver and Chappell, 1988; Todd et al., 2003; Kimbrough et al.,
eters of the down-going plate, such as convergence rate, age, and sub- 2015). In the eastern zone, gabbros are rarely exposed (~1%) and are
duction angle, then flare-ups and lulls would likely be widely dominated by a tonalite-trondhjemite-granodiorite (TTG) suite that
distributed along the arc and occur as events that coincide with periods composes ~50% of the exposed rocks (Kimbrough and Grove, 2006).
of global plate reorganization (Kirsch et al., 2016).
Mantle melt input has also been cited as playing an important role in 3.2. Peruvian Coastal Batholith (6–18° S)
driving flare-ups in which the mass transfer and heat input of mantle-
derived magmas promotes fusion of crustal rocks (Gaetani and Grove, Between 6° S and 18° S, the Cretaceous marks the intrusion of the
1998; Ulmer, 2001). Evidence supporting this idea comes from studies Peruvian Coastal batholith (PCB), a linear belt of calc-alkaline granitoids
of hybrid intermediate and felsic plutonic rocks in the Famatinian arc emplaced from 130 to 60 Ma (Pitcher et al., 1985; Mukasa, 1986;
(Otamendi et al., 2012) and mafic and felsic magmatism with isotopi- Hildebrand and Whalen, 2014a). The PCB consists of plutons with bi-
cally primitive values and their respective contemporaneity (Condie, modal magma compositions, exhibits a west to east change in chemical
1998; Pietranik et al., 2008; Sun et al., 2010; Condie et al., 2011; and isotopic signatures and increase in the thickness of the continental
Kimbrough et al., 2015; Schwartz et al., 2017). crust (Atherton and Petford, 1996), potentially due to subduction
Other potential processes, such as episodic volatile fluxing into the shallowing (Coira et al., 1982). Five segments are recognized, but only
mantle wedge and episodic melting and ascent scenarios within the two mapped in any detail: the 400 km long Lima segment and the
mantle wedge may drive episodic mantle melting (Sekine and Wyllie, 1000 km long Arequipa segment (Pitcher et al., 1985). The southern
1982; Scott and Stevenson, 1984; Nicolas, 1986; Billen and Arredondo, Arequipa segment is influenced by the rise of magmas through thicker
2018). Research conducted at subduction zones by Peacock (2001) and Precambrian and Paleozoic basement with evidence of higher crustal
Obara (2002) suggested that episodic volatile flux causes intermediate- contamination. The northern Arequipa segment is characterized by a
depth earthquakes, and instances of episodic volatile fluxing into the larger proportion of gabbro at the present erosion level (~30%) and by
mantle can be indicated by slab seismicity (Manning, 2004). Episodic a diversity of more siliceous rocks including quartz diorite, tonalite,
melt extraction events suggested by Nicolas (1986) trace successive granite, and monzogranite all with primitive geochemical affinities.
stages of melt extraction from mantle diapirs and explain how diapiric The gabbro plutons occur over the entire width of the batholith, but
rise and melting can generate the discontinuous and episodic processes most of them along the western flank (Regan, 1985).
of melt extraction. The periodicity of such events depends on the spread-
ing rate. Scott (1984) studied the process of melt migration in the Earth's 3.3. Chilean Coastal Batholith (18°– 35°S)
mantle by applying a numerical experiment and suggested that episodic
melt ascent processes would lead to episodic eruptions. Most recent In the Chilean Andes, the Cretaceous batholith is located in the west-
studies of mantle flow and subduction zone dynamics conclude that ep- ern belt and divided into two tectono-magmatic domains: central
isodic motion of the slab and subduction rates are a consequence of epi- Chilean Coastal Batholith (18°–35°S) and Patagonian Batholith (42°–
sodic mantle flow which depends on the viscosity of the asthenosphere. 55°S). The central Chilean Coastal Batholith (CCCB) is characterized by
Changes in mantle viscosity might thus result in the variation of subduc- Mesozoic plutonic belts with eastward decreasing ages and isotopic en-
tion rates that could contribute to the episodic patterns observed in arc richments that culminate with primitive Late Cretaceous plutons dom-
magmatism (Jadamec and Billen, 2010; Billen and Arredondo, 2018). inated by mafic compositions (~60%) hosted in Early Cretaceous
Also, studies of the timing and magnitudes of magmatism in oceanic volcanic-sedimentary successions (Parada et al., 1988). Emplacement
arcs show that magmatism is mantle-derived, sometimes episodic, and started in the Early Mesozoic at the end of an extensional tectonic re-
that flare-ups are driven by mantle processes. The evidence from gime (Parada et al., 1988, 1999; Hervé et al., 2007; Parada et al., 2007;
A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27 21

Fig. 1. Schematic representation of the Cretaceous arc system along the western margin of North and South America and the locations of samples and distribution of U—Pb ages of intrusive
rocks. Inset maps of the selected arc sections and distribution of U—Pb ages and isotope data. Abbreviations are as follows: WPRB = Western Peninsular Ranges Batholith, PCB = Peruvian
Coastal Batholith, CCCB = Central Chilean Coastal Batholith. See the supplementary material for data sources.
22 A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27

Fig. 2. Age histograms and magmatic flare-ups for different domains of Cretaceous arcs in North and South America. Mafic magmatism is indicated only for the selected flare-ups. Total
number of samples with U—Pb zircon ages (n). See the supplementary material for data sources.

Parada et al., 2005). Gabbro, diorite, granodiorite, granite, and spectrometry (LA-ICP-MS) following the procedure outlined by
monzogranite are the dominant lithologies. Very low initial 87Sr/S6Sr ra- Gehrels et al. (2008) and Johnston et al. (2009). The spectrometer is
tios and positive ƐNd for these CCCB plutons and associated volcanic coupled to an Excimer laser system operating at a wavelength of
rocks (mostly ca. 0.7035 and between +4.0 and + 5.0) confirm deriva- 193 nm. This method does not involve chemical abrasion pretreatment
tion of the magmas from the upper mantle with little to no continental on the zircons. Final ages reported and discussed throughout this paper
crustal involvement (Parada et al., 1988). Isotopic variations recorded in are concordia ages calculated using the Isoplot Excel® macro of Ludwig
Cretaceous igneous rocks are thought to be a consequence of continuous (2003). Ages given in the text and figures are quoted at a 2σ confidence
lithospheric delamination and associated asthenosphere upwelling and level. All geochronological data are included in supplementary material,
melting that started in the Early Jurassic (Parada et al., 1999; Parada et and errors are reported at ±2σ (see the supplementary data).
al., 2005).
4.2. New whole rock element and isotope geochemistry
4. Methods and results
We present whole-rock geochemical data of major and trace ele-
The role of the mantle for triggering flare-ups was examined using ments, and Sr, Nd, and Pb isotope ratios from 31 samples of the PCB
geochronological and geochemical datasets from all three arc segments: (see the supplementary data). The samples used for geochemical analy-
wPRB, PCB, and CCB. The compiled geochemical and geochronological sis had no visible inclusions. Whole rock samples were analyzed for
data (see the supplementary data) are mainly derived from published major element chemistry at the SGS laboratories in Canada using a
data and supplemented by the Central Andes geochemical and geochro- Thermo Jarrell Ash Enviro II simultaneous and sequential ICP with a de-
nology database (http://andes.gzg.geo.uni-goettingen.de). Data for the tection limit from 0.001 to 0.01% for major elements, from 0.002 to
PCB are also supplemented with new data from the lead author and 0.05 ppm for REE, and from 0.01 to 20 ppm for other trace elements.
from Ben Clausen at Loma Linda University. Two instrumentation techniques were used by SGS Laboratories to ob-
tain the chemical data: ICM90A using sodium peroxide fusion analyzed
4.1. New U—Pb zircon geochronology via ICP-MS for trace elements, and ICP95A using lithium metaborate fu-
sion analyzed via ICP atomic emission spectroscopy (AES) for major and
U—Pb single zircon ages were obtained for the PCB from nine plu- some trace elements.
tonic samples at the Arizona Laserchron Center, University of Arizona. The isotopic ratios of 87Sr/86Sr and 143Nd/144Nd and the trace
Measured spots (20 μm size) included cores and rims in all samples. element concentrations of Rb, Sr, Sm, and Nd were measured by
All U-Th/Pb isotopic measurements were performed by Element2 thermal ionization mass spectrometry in the Geochronology and
single-collector laser ablation-inductively coupled plasma mass Thermochronology Lab of the University of Arizona and were
A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27 23

performed on a VG Sector TIMS instrument using the techniques 4.4. Flare-up and Bedrock Ages
described by Ducea (1998) and Otamendi et al. (2009). The common
isotopes of lead were analyzed on separate batches of dissolved sam- Fig. 2 shows the age histograms from compiled geochronological
ples. Lead was extracted using an anion exchange procedure modified dataset for ages between 150 and 60 Ma for the wPRB, PCB, and CCCB.
after Chen and Wasserburg (1981). Bedrock zircon age patterns from the wPRB area define a flare-up
from ~125 to 90 Ma and with a peak at ~108 Ma. Zircon age patterns
4.3. Databases for cretaceous Arcs from the PCB area define a flare-up from ~110 to 87 Ma. Smaller peaks
are observed in the bedrock age histogram plot, two older maxima at
We synthesized U—Pb zircon age data from 333 plutonic and volca- 130 Ma and 115 Ma, and one younger maxima at 68 Ma. Age patterns
nic samples with an age spectrum of 150–60 Ma including analyses de- for the CCCB define a flare-up from ~105 to 90 Ma. Three older peaks
termined by thermal ionization mass spectrometry (TIMS), laser are seen at 138 Ma, 128 Ma, and 117 Ma, and two younger maxima at
ablation-inductively coupled plasma-mass spectrometer (LA-ICP-MS), 85 and 66 Ma. The timing of the Late Cretaceous flare-up thus can be de-
sensitive high-resolution ion microprobe (SHRIMP), and secondary fined from ~125 to 90 Ma collectively for the three arc segments.
ion mass spectrometry (SIMS) analyses. Areal measurements (km2) The histogram plot includes zircon U—Pb crystallization ages from
were made of outcrop exposures for mafic and felsic rocks within the mafic and felsic compositions and shows that at the beginning, during,
study areas and calculated using ArcMap 10.6 by compiling geological and at the end of the flare-up mafic and felsic magmatism is coeval for
map data and attributing polygons to all plutonic bodies. The purpose the three arc segments.
of this analysis was to determine the area percentage represented by
mafic and felsic compositions formed during the Cretaceous magmatic 4.5. Radiogenic Isotopes
flare-up in each arc segment. In this case, calculations of the area rather
than the volume of the plutons in the three arc segments is more appro- In Sri versus εNd, and Pb isotope plots, both mafic and felsic compo-
priate, as the compositional variation beneath the exposed surface of sitions of the wPRB, PCB, and CCCB lie along the mantle array trending
the intrusive units is unknown. The presented compositional features towards a depleted mantle (DM) signature, and show minor systematic
and areas calculated for mafic and felsic intrusions agree with previ- changes in Sri (Fig. 3). The mafic compositions with restricted Sri values
ously published research (Pitcher et al., 1985; Parada et al., 2002; ranging from 0.703 to 0.705, ƐNd values between 0 and + 6, 208Pb/204Pb
Kimbrough et al., 2015). We rule out tilting because there is little evi- from 38 to 39.2, and 206Pb/204Pb from 18.2 to 19. The felsic compositions
dence in the arcs examined and we estimate the respective volume of have Sri values ranging from 0.703 to 0.706, ƐNd values between −2
magma added during each pulse by using the present-day exposures. and + 7, 208Pb/204Pb from 38 to 39.6, and 206Pb/204Pb from 18 to 19.4.
However, this simplification may introduce a significant bias because These values are attributed to a depleted mantle source with no to an in-
tilting could increase or decrease apparent sizes (Petford et al., 2000; creasing amount of crust involved.
Karlstrom et al., 2017; Takasuka et al., 2018).
The selected whole rock geochemical dataset includes data from 222 4.6. Assimilation and Fractional Crystallization (AFC)
samples of the Late Cretaceous flare-up accompanied by U—Pb zircon
data. SiO2, MgO, and Mg# are used as proxies to evaluate the degree of Isotope AFC modeling after DePaolo (1981) was used to estimate the
differentiation and to provide a filter for the least differentiated potential crustal contribution to the isotopic signatures for mafic and
mantle-derived rocks. Samples with SiO2 values b45 wt% were classified felsic rocks. The first step is the selection of the input parameters that
as ultramafic, 45–55 wt% classified as mafic (gabbro), and N 55 wt% clas- involve the end members and the bulk partition coefficient (D). The
sified as felsic rocks. Using MgO and Mg# the group of mafic rocks is de- primary concern for this modeling was the characterization of the
fined by Mg# N52, MgO N 3.5 wt%, and SiO2 b 55 wt%. These categories chemistry of the end members, the parental mantle melt composition
are consistent with the observed compositional groups defined by the (C0) and the assimilant (Ca).
SiO2 proxy. Sr, Nd, and Pb isotopes were used to evaluate the role of man- In general, the most primitive sample in the dataset is selected as the
tle versus crust for triggering flare-ups, and to identify mantle reservoirs C0, while the average composition of the crustal basement units is
(e.g., DePaolo, 1981; Zindler and Hart, 1986; Chapman et al., 2017). regarded as the Ca (Keskin 2012). In this study, for all three arc

Fig. 3. Magma sources and compiled isotopic data from mafic and felsic rocks for three Cretaceous arc segments: wPRB, PCB, and CCCB. (a) Isotope ratios of Nd and Sr. (b) 206Pb/204Pb
versus 208Pb/204Pb. Color fields represent data for felsic rocks (wPRB = yellow, PCB = red, CCCB = blue). For comparison, the field end members are included, and these are depleted
mantle (DM), high μ (HIMU), enriched mantle with low initial 87Sr/86Sr (EM1), and enriched mantle with high initial 87Sr/86Sr (EM2). Other end member fields are from Ducea
(2001), and these are C—old lower crust; M—mantle; S—sedimentary. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)
24 A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27

segments, depleted mantle (DM) values (Stracke et al., 2003) are 2 which indicate that most of the mafic and felsic compositions
interpreted to approximate C0 representing the most primitive mantle have experienced fractional crystallization (r = 0) combined
source for the modeling. To characterize Ca, two different values were with variable degrees of assimilation (with r N 0). However these
considered to estimate a likely maximum amount of crust assimilated: likely maximum amounts of needed crustal assimilation are low:
(1) the felsic sample with the most isotopically evolved Sri value from wPRB rocks incorporated about 30% crustal materials (Fig. 5a),
our dataset and (2) the respective Sri value for the crustal basement in the PCB rocks about 25% in average (Fig. 5b), and the CCCB rocks
each area. Whereas one AFC model (Fig. 4) assumes that the most isoto- about ca. 30% in average (Fig. 5c).
pically evolved felsic sample is the crustal contaminant assimilated dur- For the gabbro compositions used in the AFC modeling, we are con-
ing fractional crystallization (model 1), another model assumed that cerned about the chemical changes as a result of partial melting of DM
the basement crustal values are the most likely contaminants and mafic crust, crustal assimilation, and fractional crystallization. For
(model 2). For model 2, the appropriate Sri (~0.708) crustal basement this reason, we calculated the final SiO2 wt% compositions assuming
value for the wPRB is from Early Jurassic to Triassic metavolcanic values for an initial melt ranging from 45 to 54 wt% and for a crustal
and metasedimentary rocks (Kistler et al., 2014), for the PCB the Sri assimilant ranging from 55 to 68 wt% of SiO2 and a maximum of crustal
(~0.710) value is from our own Precambrian and Paleozoic samples, assimilation of 25%. The data suggest that most of the resulting compo-
and the Sri (~0.708) value for the CCCB is from Paleozoic rocks sitions (~85%) can be classified as gabbros with final compositions hav-
(Parada et al., 1999; Gonzalez et al., 2017). The Sr bulk partition coeffi- ing b55 wt% of SiO2.
cient (D) used in the modeling is from Rollinson (1993) and the Because each data point on the AFC plots has an associated age, these
GERM Kd database for basalts (https://earthref.org/GERM/). data can be used to broadly estimate the relative amounts of mantle and
The modeled curves correspond to different r-values;r being the crustal magmas through time (Fig. 2). We have matched ages to mapped
ratio of mass assimilation rate to fractional crystallization rate. The units to roughly predict the area of units of different ages, compositions,
input parameters control the positions and shapes of the modeled and isotopic values. If the AFC results are accepted, then these results in-
curves, and the model is modified if any parameter is changed. In this dicate that mantle magmas dominated at the initiation of the flare-up
way, we can run the two models using different parameters for Ca and and the amounts of mantle magma continued to increase up to the
observe the results (Figs. 4 and 5). peak of the flare-up. And since we used conservative input values, we
After running models, we consider that the most representative would argue that these present maximum amounts of crustal magmas
results from AFC modeling that fit the isotopic data are from model implying that mantle magma input dominated throughout the flare-ups.

Fig. 4. AFC modeling for Sri vs. Sr plot. The initial magma composition uses a Sri and Sr values suggested for DM (Gale et al., 2003; Faure, 2009). Color fields represent the distribution of the
felsic samples, and the star symbol indicates their respective averages. The maximum amounts of crust assimilated to form the gabbros are: (a) an average of 30% for the wPRB; (b) an
average of 25% for PCB, and (c) an average of 30% for CCCB.
A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27 25

5. Discussion “TESTING MANTLE DRIVEN FLARE-UPS” consideration. Under this scenario the internal feedback model pre-
sented by DeCelles et al. (2009) may play an important role in magmatic
A Late Cretaceous flare-up event was characterized using zircon bed- contamination, but not for triggering the flare-ups.
rock ages for the wPRB from 125 to 90 Ma, in the PCB from 110 to 87 Ma, Our data strongly suggest that during the Late Cretaceous flare-up
and for the CCCB from 105 to 90 Ma. In the three arc segments, most the mantle-derived magmas are dominant over the crustal-derived
mafic and felsic rocks are contemporaneous, and the isotopic composi- magmas and therefore the episodic patterns of arc magmatism origi-
tions of the mafic rocks do not change much through time (Fig. 3). These nated from episodic mantle processes (e.g., melting, ascent, and volatile
observations are supported by previous studies conducted in other con- fluxing) as described above. Important evidence for episodic melting in
tinental magmatic arcs (Sun et al., 2010; Kimbrough et al., 2015; Kirsch the mantle is presented in studies of subducting slab and mantle flow
et al., 2016; Schwartz et al., 2017). dynamics suggesting that episodic folding of the slab is a consequence
Almost all samples from the three arc segments have Sri b ~ 0.705 of increasing viscosity in the asthenosphere. Billen and Arredondo
and higher values of Sri are rare, with relatively few data points N (2018) presented a 2D dynamic model of subduction to model the
0.706, and εNd values mostly from −2 to +7 (Fig. 3a). In the Sri versus time-evolution of slab deformation and thermal structure. They found
εNd plot, both mafic and felsic compositions lie along the mantle array that rapid sinking of the slab and folding causes a reduction in astheno-
trending towards a depleted mantle (DM) signature and show minor sphere viscosity, which allows the overriding plate to move in the oppo-
systematic changes in Sri. The Pb isotopes plot shows all samples in site direction of the asthenosphere. Increase in viscosity leads to
the DM reservoir field and mafic and felsic compositions have the fol- episodic folding of the slab and causes episodic motion of the trench
lowing ranges: 206Pb/204Pb ratios from 18 to 19.5 and 208Pb/204Pb ra- and plates in which velocities increase and then decrease during each
tions from 38 to 39.5 (Fig. 3b). The Sr, Nd, and Pb isotope values episodic folding event.
correspond to mantle-derived magmas source defining a trend from Convergence rate has been proposed to influence the extent of melt-
the DM towards the older crust emphasizing the idea of a mantle source. ing beneath arcs (Hughes and Mahood 2008; Zellmer 2008; Hebert et al.
Since the associated felsic granitoids and volcanic rocks have overlap- 2009; England and Katz 2010; Turner and Langmuir 2015, 2015). How-
ping or slightly more crustal isotopic signatures (Figs. 3, 4) and are tem- ever, some studies have rejected such a relationship, providing evidence
porally associated with gabbros maintaining primitive isotopic that flare-up events in some parts of the Cordillera are seemingly out of
signatures (see also Kimbrough et al., 2015), we conclude that these sync with peaks in convergence rates (e.g., Ducea 2001; DeCelles et al.,
felsic units are the result of similar mantle magmas that incorporated 2009, 2015; Cao et al., 2016).
some crustal rocks. Our interpretation of existing isotopic data is Higher convergence rates have been shown to: (1) lead to more vig-
complemented with AFC modeling which indicates that both mafic orous hydration of the mantle wedge causing increased melting (e.g.,
and felsic compositions consist of ~85% to 70% mantle-derived magmas Cagnioncle et al., 2007; Plank et al., 2009), and/or (2) increase the flux
and no more than ~25 to 30% recycled crust in all three arc segments. of hot mantle into the wedge corner, raising the temperature and caus-
If 70% to 100% of these Late Cretaceous arc rocks originated from a ing increased melt formation beneath the arc (England and Wilkins
DM reservoir and contain only 0% to 30% crust, it implies that crust is 2004; England and Katz 2010; Turner and Langmuir 2015a, 2015b).
not required for triggering nor playing a dominant role in an arc flare- Convergence rates during the Cretaceous for the Farallon and Nasca
up. These conclusions, based on the Late Cretaceous flare-up events doc- plates are between 100 and 200 mm/y and 60–100 mm/y respectively.
umented here, are remarkably different from the internal feedback These relatively high convergence rates may be related to (1) the suc-
model for the evolution of the Mesozoic North American and South cessive emplacement of a sequence of large igneous provinces during
American Cordilleras presented by Ducea (2001), Ducea and Barton 140 and 120 Ma and (2) a major plate reorganization event occurred
(2007) and DeCelles et al. (2009, 2015). In their model, upper plate ma- at ca. 100 Ma (Müller et al., 2016; Shephard et al., 2013; Matthews et
terials play a significant role in triggering and driving flare-ups and the al., 2012). For our three arc segments, we consider a possible link be-
Sr, Nd, and Pb isotope values correspond to crustal-derived magma tween arc-external events and magmatic episodicity because the corre-
sources. Although our isotope data in this study are limited to three lation between plate convergence rate and episodicity observed for the
arc segments, our conclusion is in agreement with research done in selected time scale about ca. 90–120 Ma.
other magmatic arcs like the Median Batholith in New Zealand In summary, MOR's and island arc magmatism are clear examples
(Decker et al., 2017; Schwartz et al., 2017). In the Median Batholith, zir- where mantle melts are the source of sometimes episodic crustal
con chronology and isotope data strongly support an externally trig- magma additions. Synchronous ages of fairly primitive gabbros and felsic
gered, mantle-generated process such as ridge subduction or a slab- granitoids and isotopic ƐNd, Sri, and Pb data from Cretaceous Cordilleran
breakoff event, leading to the surge of mafic and intermediate continental arc segments discussed herein support the idea that episodic
magmatism from 128 to 114 Ma with only limited contributions from mantle processes, leading to the episodic development of mantle melts,
evolved lithospheric sources. Their data suggest that crustal signatures are playing the major role for triggering and driving continental arc
might reflect intracrustal partial melting due to elevated geothermal flare-ups. Upper plate cyclic crustal processes may play a secondary
gradients resulting from increasing mantle melt influx to the base of role in modifying these mantle magmas. The potential for episodic man-
the crust and/or assimilation of mantle-derived and hybrid magmas tle processes is well supported by published models of mantle flow near
during ascent through the crustal column (Schwartz et al., 2017). subduction zones. An implication of this observation is the need to fur-
The wPRB and the PCB arc segments exhibit the typical spatial isoto- ther examine the potential links between oceanic plate motions and ep-
pic trend for continental arcs, with magmatism closest to the trench isodic mantle melting above subduction zones, which in turn lead to
having an isotopic composition comparable to the depleted mantle episodicity in continental arc magmatism. A more extensive linked geo-
and felsic magmatism increasingly evolved landward (Chapman et al., chemical and geochronological data compilation, complemented with
2017). More efficient assimilation as result of hotter lower crust or a plate reconstructions and numerical simulations of mantle processes
more prolonged crustal assimilation during magma ascent, lead to are needed to test the mantle-trigger model for other arc segments.
more evolved isotopic compositions towards the continent (Farmer
and DePaolo, 1983; Hildreth and Moorbath, 1988). In the CCCB, the iso- Acknowledgments
topic compositions are less evolved landward and explained by the pro-
gressive crustal delamination and tectonic extension during that period The authors acknowledge the financial support of this study from
(Parada et al., 1999; Parada et al., 2005). However, the more evolved Loma Linda University and the Geoscience Research Institute (Projects
magma signature is linked to crustal thickness, and thus the result GRI2014-BC01 and GRI2015-AM10). We want to thank Orlando Poma,
of crustal assimilation from evolved lithospheric sources needs Lance Pompe, Italo Payacan, and Fabian Figueroa for their valuable help
26 A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27

during fieldwork activities. We also would like to thank the editor Xian- Gehrels, G.E., Valencia, V.A., Ruiz, J., 2008. Enhanced precision, accuracy, efficiency, and
spatial resolution of U-Pb ages by laser ablation–multicollector–inductively coupled
Hua Li and reviewers Emily J. Chin, Tetsuo Kawakami, and anonymous plasma–mass spectrometry. Geochemistry, Geophysics, Geosystems 9, 3.
reviewer for their useful comments that helped improve the quality of González, J., Oliveros, V., Creixell, C., Velásquez, R., Vásquez, P., Lucassen, F., 2017. The Tri-
the manuscript. assic magmatism and its relation with the Pre-Andean tectonic evolution: Geochem-
ical and petrographic constrains from the High Andes of north central Chile (29° 30’-
30° S). Journal of South American Earth Sciences.
Appendix A. Supplementary data Haschke, M., Günther, A., Melnick, D., Echtler, H., Reutter, K.-J., Scheuber, E., Oncken, O.,
2006. Central and Southern Andean Tectonic Evolution Inferred from Arc
Magmatism. Springer, The Andes, pp. 337–353.
Supplementary data to this article can be found online at https://doi. Hebert, L.B., Antoshechkina, P., Asimow, P., Gurnis, M., 2009. Emergence of a low-viscosity
org/10.1016/j.lithos.2018.12.007. channel in subduction zones through the coupling of mantle flow and thermodynam-
ics. Earth and Planetary Science Letters 278 (3–4), 243–256.
Hervé, F., Pankhurst, R.J., Fanning, C.M., Calderón, M., Yaxley, G.M., 2007. The South Pata-
References gonian batholith: 150 my of granite magmatism on a plate margin. Lithos 97 (3–4),
373–394.
Atherton, and Petford, 1996. Plutonism and the growth of Andean Crust at 9’S from 100 to Hildebrand, R.S., Whalen, J.B., 2014a. Arc and Slab-failure Magmatism in Cordilleran Bath-
3 Ma. J. S. Am. Earth Sci. 9, 1–9. oliths I - the cretaceous Coastal Batholith of Peru and its Role in South American Oro-
Baird, A., Baird, K., Welday, E., 1979. Batholithic Rocks of the Northern Peninsular and genesis and Hemispheric Subduction Flip. p. 28.
Transverse Ranges, Southern California: Chemical Composition and Variation: Meso- Hildebrand, R.S., Whalen, J.B., 2014b. Arc and slab-failure magmatism in Cordilleran bath-
zoic Crystalline Rocks: Peninsular Range Batholith and Pegmatites, Point Sal oliths II—The cretaceous Peninsular Ranges batholith of Southern and Baja California.
Ophiolite. Department of Geological Sciences, San Diego State University, San Paul Hoffman Volume: : Geoscience Canada 41, 339–458.
Diego, pp. 111–132. Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andes of
Billen, M.I., Arredondo, K.M., 2018. Plate-Asthenosphere Decoupling Caused by Non- Central Chile. Contributions to Mineralogy and Petrology 98, 455–489.
linear Viscosity during Slab Folding in the Transition Zone. Hughes, G.R., Mahood, G.A., 2008. Tectonic controls on the nature of large silicic calderas
Cagnioncle, A.-M., Parmentier, E.M., Elkins-Tanton, L.T., 2007. Effect of solid flow above a in volcanic arcs. Geology 36 (8), 627–630.
subducting slab on water distribution and melting at convergent plate boundaries. Ishizuka, O., Tani, K., Reagan, M.K., Kanayama, K., Umino, S., Harigane, Y., Sakamoto, I.,
Journal of Geophysical Research: Solid Earth 112, B9. Miyajima, Y., Yuasa, M., Dunkley, D.J., 2011. The timescales of subduction initiation
Cao, W., Paterson, S., Saleeby, J., Zalunardo, S., 2016. Bulk arc strain, crustal thickening, and subsequent evolution of an oceanic island arc. Earth and Planetary Science Let-
magma emplacement, and mass balances in the Mesozoic Sierra Nevada arc. Journal ters 306 (3), 229–240.
of Structural Geology 84, 14–30. Jadamec, M.A., Billen, M.I., 2010. Reconciling surface plate motions with rapid three-
Chapman, A.D., Saleeby, J.B., Eiler, J., 2013. Slab flattening trigger for isotopic disturbance dimensional mantle flow around a slab edge. Nature 465 (7296), 338–341.
and magmatic flare-up in the southernmost Sierra Nevada batholith. California: Geol- Jicha, B., Jagoutz, O., 2015. Magma production rates for intraoceanic arcs. Elements 11,
ogy 41 (9), 1007–1010. 105–111.
Chapman, J., Ducea, M., Kapp, P., Gehrels, G., Decelles, P., 2017. Spatial and Temporal Ra- Johnston, S., Gehrels, G., Valencia, V., Ruiz, J., 2009. Small-volumeU–Pb zircon geochronol-
diogenic Isotopic Trends of Magmatism in Cordilleran Orogens: Gondwana Research. ogy by laser ablation-multicollector-ICP-MS. Chemical Geology 259 (3–4), 218–229.
Chen, J.H., Wasserburg, G.J., 1981. Isotopic determination of uranium in picomole and Karlstrom, L., Paterson, S.R., Jellinek, A.M., 2017. A reverse energy cascade for crustal
subpicomole quantities. Analytical Chemistry 53 (13), 2060–2067. magma transport: Nature Geosci, v. advance online publication.
Chin, E.J., Lee, C.-T.A., Blichert-Toft, J., 2015. Growth of upper plate lithosphere controls Keskin, M., 2012. AFC-Modeler: a Microsoft Excel Workbook Program for Modelling As-
tempo of arc magmatism. Constraints from Al-diffusion kinetics and coupled Lu-Hf similation combined with Fractional Crystallization (AFC) Process in Magmatic Sys-
and Sm-Nd chronology: Geochemical Perspective Letters 1, 20–32. tems by Using Equations of DePaolo. Turkish Journal of Earth Sciences 21, 1–18.
Coira, B., Davidson, J.P., Mpodozis, C., Ramos, V., 1982. Tectonic and magmatic evolution of Kimbrough, D.L., Grove, M., Morton, D.M., 2015. Timing and significance of gabbro em-
the Andes of northern Argentina and Chile. Earth Science Reviews 18, 303–332. placement within two distinct plutonic domains of the Peninsular Ranges batholith,
Condie, K.C., 1998. Episodic continental growth and supercontinents: a mantle avalanche southern and Baja California. Geological Society of America Bulletin 127 (1–2), 19–37.
connection? Earth and Planetary Science Letters 163 (1), 97–108. Kirsch, M., Paterson Scott, R., Wobbe, F., Martinez-Ardila, A.M., Clausen Benjamin, L.,
Condie, K.C., Bickford, M.E., Aster, R.C., Belousova, E., Scholl, D.W., 2011. Episodic zircon Alasino Pablo, H., 2016. Temporal histories of Cordilleran continental arcs: testing
ages. Hf isotopic composition, and the preservation rate of continental crust: Geolog- models for magmatic episodicity. American Mineralogist 101, 2133.
ical Society of America Bulletin 123 (5–6), 951–957. Kistler, R.W., Wooden, J.L., Premo, W.R., Morton, D.M., Miller, F., 2014. Pb-Sr-Nd-O isotopic
de Silva, S.L., Riggs, N.R., Barth, A.P., 2015. Quickening the Pulse: Fractal Tempos in Conti- characterization of Mesozoic rocks throughout the northern end of the Peninsular
nental Arc Magmatism. Elements 11 (2), 113–118. Ranges batholith: Isotopic evidence for the magmatic evolution of oceanic arc–
Decelles, P.G., Ducea, M.N., Kapp, P., Zandt, G., 2009. Cyclicity in Cordilleran orogenic sys- continental margin accretion during the Late Cretaceous of southern California: Pen-
tems. Nature Geosci 2 (4), 251–257. insular Ranges Batholith, Baja California and Southern California. Geological Society of
Decelles, P.G., Zandt, G., Beck, S.L., Currie, C.A., Ducea, M.N., Kapp, P., Gehrels, G.E., Carrapa, America Memoirs 211, 263–316.
B., Quade, J., Schoenbohm, L.M., 2015. Cyclical orogenic processes in the Cenozoic Manning, C.E., 2004. The chemistry of subduction-zone fluids. Earth and Planetary Science
central Andes. Geological Society of America Memoirs 212. Letters 223 (1), 1–16.
Decker, M., Schwartz, J., Stowell, H., Klepeis, K., Tulloch, A., Kitajima, K., Valley, J., Matthews, K.J., Seton, M., Müller, R.D., 2012. A global-scale plate reorganization event at
Kylander-Clark, A., 2017. Slab-triggered arc flare-up in the cretaceous median Batho- 105–100 Ma. Earth and Planetary Science Letters 355, 283–298.
lith and the growth of lower arc crust, Fiordland. New Zealand: Journal of Petrology Morton, D.M., Miller, F.K., Kistler, R.W., Premo, W.R., Lee, C.-T.A., Langenheim, V.E.,
58 (6), 1145–1171. Wooden, J.L., Snee, L.W., Clausen, B.L., Cossette, P., 2014. Framework and petrogenesis
Depaolo, D.J., 1981. A neodymium and strontium isotopic study of the Mesozoic calc- of the northern Peninsular Ranges batholith, southern California. Geological Society
alkaline granitic batholiths of the Sierra Nevada and Peninsular Ranges. California: of America Memoirs 211, 61–143.
Journal of Geophysical Research: Solid Earth 86 (B11), 10470–10488. Mukasa, S.B., 1986. Zircon U-Pb ages of super-units in the Coastal batholith, Peru: Impli-
Ducea, M., 1998. A petrologic investigation of deep-crustal and upper-mantle xenoliths cations for magmatic and tectonic processes. Geological Society of America Bulletin
from the Sierra Nevada, California; constraints on lithospheric composition beneath 97 (2), 241.
continental arcs and the origin of cordilleran batholiths. PhD Disseration. California Müller, R.D., Seton, M., Zahirovic, S., Williams, S.E., Matthews, K.J., Wright, N.M., Shephard,
Institute of Technology, p. 364. G.E., Maloney, K.T., Barnett-Moore, N., Hosseinpour, M., 2016. Ocean basin evolution
Ducea, M.N., 2001. The California arc Thick granitic batholiths, eclogitic residues, and global-scale plate reorganization events since Pangea breakup. Annual Review of
lithospheric-scale thrusting, and magmatic flare-ups. GSA Today 11, 4–10. Earth and Planetary Sciences 44, 107–138.
Ducea, M.N., Barton, M.D., 2007. Igniting flare-up events in Cordilleran arcs. Geology 35 Nicolas, A., 1986. A melt extraction model based on structural studies in mantle perido-
(11), 1047–1050. tites. Journal of Petrology 27 (4), 999–1022.
Ducea, M.N., Paterson, S.R., Decelles, P.G., 2015. High-volume Magmatic Events in Subduc- Obara, K., 2002. Nonvolcanic deep tremor associated with subduction in southwest.
tion Systems. Elements 11 (2), 99–104. Japan: Science 296 (5573), 1679–1681.
England, P.C., Katz, R.F., 2010. Melting above the anhydrous solidus controls the location Otamendi, J.E., Ducea, M.N., Tibaldi, A.M., Bergantz, G.W., de la Rosa, J.D., Vujovich, G.I.,
of volcanic arcs. Nature 467 (7316), 700. 2009. Generation of Tonalitic and Dioritic Magmas by coupled Partial Melting of Gab-
England, P., Wilkins, C., 2004. A simple analytical approximation to the temperature broic and Metasedimentary Rocks within the Deep Crust of the Famatinian Magmatic
structure in subduction zones. Geophysical Journal International 159 (3), Arc. Argentina: Journal of Petrology 50 (5), 841–873.
1138–1154. Otamendi, J.E., Ducea, M.N., Bergantz, G.W., 2012. Geological, Petrological and Geochem-
Farmer, G.L., Depaolo, D.J., 1983. Origin of Mesozoic and Tertiary granite in the western ical evidence for Progressive Construction of an Arc Crustal Section, Sierra de Valle
United States and implications for Pre-Mesozoic crustal structure: 1 Nd and Sr isoto- Fértil, Famatinian Arc. Argentina: Journal of Petrology 53 (4), 761–800.
pic studies in the geocline of the Northern Great Basin. Journal of Geophysical Re- Parada, M.A., Rivano, S., Sepulveda, P., Herve, M., Herve, F., Puig, A., Munizaga, F., Brook,
search: Solid Earth 88 (B4), 3379–3401. M., Pankhurst, R., Snelling, N., 1988. Mesozoic and cenozoic plutonic development
Gaetani, G.A., Grove, T.L., 1998. The influence of water on melting of mantle peridotite. in the Andes of central Chile (30°30′–32°30′S). Journal of South American Earth Sci-
Contributions to Mineralogy and Petrology 131 (4), 323–346. ences 1 (3), 249–260.
Garzione, C.N., Molnar, P., Libarkin, J.C., MacFadden, B.J., 2006. Rapid late Miocene rise of Parada, M.A., Nyström, J.O., Levi, B., 1999. Multiple sources for the Coastal Batholith of
the Bolivian Altiplano: Evidence for removal of mantle lithosphere. Earth and Plane- central Chile (31–34°S): geochemical and Sr–Nd isotopic evidence and tectonic im-
tary Science Letters 241 (3–4), 543–556. plications. Lithos 46 (3), 505–521.
A.M. Martínez Ardila et al. / Lithos 326–327 (2019) 19–27 27

Parada, M.A., Larrondo, P., Guiresse, C., Roperch, P., 2002. Magmatic Gradients in the Cre- Shephard, G.E., Müller, R.D., Seton, M., 2013. The tectonic evolution of the Arctic since
taceous Caleu Pluton (Central Chile): Injections of Pulses from a Stratified Magma Pangea breakup: Integrating constraints from surface geology and geophysics with
Reservoir. Gondwana Research 5 (2), 307–324. mantle structure. Earth-Science Reviews 124, 148–183.
Parada, M.A., Féraud, G., Fuentes, F., Aguirre, L., Morata, D., Larrondo, P., 2005. Ages and Silver, L.T., Chappell, B.W., 1988. The Peninsular Ranges Batholith: an insight into the
cooling history of the Early Cretaceous Caleu pluton: testimony of a switch from a evolution of the Cordilleran batholiths of southwestern North America. Earth and En-
rifted to a compressional continental margin in central Chile. Journal of the Geological vironmental Science Transactions of the Royal Society of Edinburgh 79 (2–3),
Society 162 (2), 273–287. 105–121.
Parada, M.A., López-Escobar, L., Oliveros, V., Fuentes, F., Morata, D., Calderón, M., Aguirre, L., Sobolev, S.V., Babeyko, A.Y., Koulakov, I., Oncken, O., 2006. Mechanism of the Andean
Féraud, G., Espinoza, F., Moreno, H., 2007. Andean magmatism (The geology of Chile). Orogeny: Insight from Numerical Modeling. Springer, The Andes, pp. 513–535.
Paterson, S.R., Ducea, M.N., 2015. Arc magmatic tempos: gathering the evidence. Elements Stracke, A., Bizimis, M., Salters, V.J., 2003. Recycling oceanic crust: Quantitative con-
11 (2), 91–98. straints: Geochemistry, Geophysics, Geosystems. 4 (3).
Peacock, S.M., 2001. Are the lower planes of double seismic zones caused by serpentine Sun, J.-F., Yang, J.-H., Wu, F.-Y., Li, X.-H., Yang, Y.-H., Xie, L.-W., Wilde, S.A., 2010. Magma
dehydration in subducting oceanic mantle? Geology 29 (4), 299–302. mixing controlling the origin of the early cretaceous Fangshan granitic pluton,
Pelletier, J.D., Decelles, P.G., Zandt, G., 2010. Relationships among climate, erosion, topog- North China Craton: in situ U–Pb age and Sr-, Nd-, Hf- and O-isotope evidence. Lithos
raphy, and delamination in the Andes: A numerical modeling investigation. Geology 120 (3–4), 421–438.
38 (3), 259–262. Takatsuka, K., Kawakami, T., Skrzypek, E., Sakata, S., Obayashi, H., Hirata, T., 2018. Spatio-
Petford, N., Cruden, A.R., McCaffrey, K.J.W., Vigneresse, J.L., 2000. Granite magma forma- temporal evolution of magmatic pulses and regional metamorphism during a creta-
tion, transport, and emplacement in the Earth's crust. Nature 408, 669–673. ceous flare-up event: Constraints from the Ryoke belt (Mikawa area, Central
Pietranik, A., Hawkesworth, C., Storey, C., Kemp, A., Sircombe, K., Whitehouse, M., Bleeker, Japan). Lithos 308-309, 428–445.
W., 2008. Episodic, mafic crust formation from 4.5 to 2.8 Ga: New evidence from de- Todd, V.R., Shaw, S.E., Hammarstrom, J.M., 2003. Cretaceous plutons of the Peninsular
trital zircons, Slave craton, Canada. Geology 36 (11), 875–878. Ranges batholith, San Diego and westernmost Imperial counties, California: Intrusion
Pitcher, W.S., Atherton, M.P., Cobbing, E.J., Beckinsale, R.D., 1985. Magmatism at a Plate across a Late Jurassic continental margin. SPECIAL PAPERS-GEOLOGICAL SOCIETY OF
Edge: The Peruvian Andes. Wiley. AMERICA 185–236.
Plank, T., Cooper, L.B., Manning, C.E., 2009. Emerging geothermometers for estimating Turner, S.J., 2015. What processes control the chemical compositions of arc front strato-
slab surface temperatures. Nature Geoscience 2 (9), 611. volcanoes? Geochemistry, Geophysics, Geosystems 16 (6), 1865–1893.
Reagan, M.K., McClelland, W.C., Girard, G., Goff, K.R., Peate, D.W., Ohara, Y., Stern, R.J., 2013. Turner, S.J., Langmuir, C.H., 2015. The global chemical systematics of arc front stratovol-
The geology of the southern Mariana fore-arc crust: Implications for the scale of Eo- canoes: Evaluating the role of crustal processes. Earth and Planetary Science Letters
cene volcanism in the western Pacific. Earth and Planetary Science Letters 380, 41–51. 422, 182–193.
Regan, P.F., 1985. The Early Basic Intrusions, Blackie Glasgow. Magmatism at a Plate Edge, Ulmer, P., 2001. Partial melting in the mantle wedge—the role of H2O in the genesis of
The Peruvian Andes. mantle-derived ‘arc-related'magmas. Physics of the Earth and Planetary Interiors
Rollinson, H., 1993. Using Geochemical Data. Longman, London 352 pp. 127 (1–4), 215–232.
Schwartz, J.J., Klepeis, K.A., Sadorski, J.F., Stowell, H.H., Tulloch, A.J., Coble, M.A., 2017. The Zellmer, G.F., 2008. Some first-order observations on magma transfer from mantle wedge
Tempo of Continental Arc Construction in the Mesozoic Median Batholith. Litho- to upper crust at volcanic arcs: Geological Society, London. Special Publications 304
sphere, Fiordland, New Zealand. (1), 15–31.
Scott, D.R., Stevenson, D.J., 1984. Magma solitons. Geophysical Research Letters 11 (11), Zindler, A., Hart, S., 1986. Chemical geodynamics. Annual Review of Earth and Planetary
1161–1164. Sciences 14 (1), 493–571.
Sekine, T., Wyllie, P.J., 1982. Phase relationships in the system KAlSiO 4-Mg 2 SiO 4-SiO 2-
H 2 O as a model for hybridization between hydrous siliceous melts and peridotite.
Contributions to Mineralogy and Petrology 79 (4), 368–374.

View publication stats

You might also like