Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Turning up the Heat:

High-Flux Magmatism
in the Central Andes
Shanaka L. de Silva1 and Suzanne M. Kay2

1811-5209/18/0014-0245$2.50  DOI: 10.2138/gselements.14.4.245

T
he Neogene history of the Central Andes records one of Earth’s most preted to record the major changes
productive periods of high-flux silicic magmatism. Subduction of an in subduction rate and geometry
that triggered a massive Andean
aseismic ridge, the Juan Fernández Ridge (JFR), led to changes in mantle crustal magmatic episode and
melt productivity that initiated a transcrustal magmatic system culminating the intrusion of massive silicic
in massive caldera- and ignimbrite-forming eruptions. This volcanism is time batholiths.
transgressive, tracking the southward passage of the JFR beneath the Central
HISTORICAL
Andes. The volcanic field is underlain by a composite, arc-long mid- and DEVELOPMENT
upper-crustal granodiorite batholith that represents extensive processing of Although the concept and processes
the continental crust by mantle-derived magmas. This batholith stabilized the of ignimbrite flare-ups were not
upper crust and contributed to the extreme elevations despite a net crustal introduced into the Central Andes
until the late 1980s (e.g. de Silva
loss beneath the Puna region. 1989), a rich history of investiga-
Keywords : Neogene, subduction, Cordilleran, ignimbrite flare-up, aseismic ridge, tions had already recognized the
delamination, Altiplano-Puna Plateau extensive ignimbrite stratigraphy.

INTRODUCTION Brazil
Peru
The Andes mountains of South America are the type
locality for continental arc volcanism, and they host the 15°S
eponymous andesite, the most ubiquitous magma compo-
sition erupted and the archetypal magma of normal or La Paz Bolivia
steady-state arcs. Changes in plate tectonic conditions can 20 Ma 15 Ma
lead to episodic higher flux magmatism (DeCelles et al.
25 Ma
2015), commonly known as flare-ups, the surface signature Sucre
of which is an extensive volcanic caldera and ignimbrite 20°S 30 Ma
plateau or a large silicic igneous province dominated by 10 Ma
dacite to rhyolite compositions (Best et al. 2016). 35 Ma
APVC Paraguay
In the last several decades, the Neogene (from roughly 10 - 1 Ma
23 Ma to 1 Ma) magmatic record of the Central Andes Chile
(Fig.  1) has become a key to understanding convergent 5 Ma
margin magmatic flare-ups (e.g. de Silva 1989, de Silva et 25°S
al. 2006; Kay and Coira 2009; Kay et al. 2010; DeCelles Cerro Galán
et al. 2015). The duration of the Central Andean Neogene 6.3 - 2.0 Ma
Ignimbrite Province approximately coincides with the
last 25 million years of southward subduction of the Juan 0 Ma
Fernández Ridge on the Nazca Plate as volcanism switched Argentina
from steady state (dominantly andesite–dacite composite 30°S
cones) to flare-up mode (dominantly large-scale ignimbrites 0-2.0 Ma
and caldera complexes). The southward sweep of the ridge 2-6.5
6.5-8
(Fig. 1) resulted in a time-transgressive southward switch
8-12
from steady state to flare-up (Kay and Coira 2009; Freymuth 12-14
Santiago
et al. 2015). This Neogene volcanic record is, thus, inter- 14-18
18-24
35°S km
no age data
0 200 400

1 College of Earth, Ocean, and Atmospheric Sciences


75°W 70°W 65°W 60°W
Oregon State University The Neogene Ignimbrite Province of the Central
104 CEOAS Admin Bldg Figure 1
Andes (region shown in grey/blue). The time-trans-
Corvallis, OR 97331-5503, USA gressive position of the Juan Fernández Ridge (shown as grey lines)
E-mail: desilvas@geo.oregonstate.edu from 35 Ma to today is based on Yáñez et al. (2001) and Kay and
2 Department of Earth and Atmospheric Sciences Coira (2009). Age-specific fields of ignimbrites (colored areas) are
Cornell University based primarily on the maps in Kay et al. (2010), Kern et al. (2016),
3140 Snee Hall, Ithaca, NY 14853, USA and Freymuth et al. (2015). APVC = Altiplano–Puna Volcanic
E-mail: smk16@cornell.edu Complex.

E lements , V ol . 14, pp. 245–250 245 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user
Since the mid-1980s, comprehensive studies of
A B
the calderas and ignimbrites have built on the
early remote sensing studies, with major efforts
in the Cerro Galán caldera region of Argentina
(26°S to 27°S) and the Altiplano–Puna transition
region (22°S to 24°S). Cerro Galán was the first
Andean caldera to be comprehensively investi-
gated and it inspired the recognition of multiple
C ignimbrites and calderas formed throughout the
Central Andes that together signaled the Neogene
Central Andean ignimbrite flare-up (F ig.  1).
The spatial and temporal connectivity of this
coherent regional flare-up was recognized by de
Silva (1989), who proposed the name Altiplano–
Puna Volcanic Complex. Since its recognition,
the Altiplano–Puna Volcanic Complex has been
the subject of extensive geophysical, geochrono-
logical, and petrochemical studies that are now
revealing the true structure and nature of the
Andean Neogene ignimbrite flare-up (e.g. Ward
et al. 2017; Pritchard et al. 2018).

Three views, at different scales, of one of the largest GEODYNAMICS OF THE IGNIMBRITE
Figure 2 FLARE-UP
resurgent calderas on Earth: the Cerro Guacha (CG)
caldera complex. (A) The eruption of three ignimbrites from CG:
the 1.8 Ma Puripica Chico ignimbrite (green dashed line); the
The prodigious volume of silicic volcanism that character-
3.6 Ma Tara ignimbrite (dashed yellow line); the 5.6 Ma Guacha izes an ignimbrite flare-up requires an elevated heat supply
ignimbrite (dashed orange line). The total volume of erupted from the mantle to generate a transcrustal magma system
magma is almost 2,000 km3, leaving a complex nested structure that can create some of the largest silicic eruptions on
and associated lava domes. (B) Perspective view (tilted to the
north) of the Cerro Guacha (CG) caldera complex, showing the
Earth. Kay and Coira (2009) proposed that the southward
composite central resurgent uplifts. Image from Google E arth. passage of the Juan Fernández Ridge caused the southward-
(C) Massive “yardangs” (wind-eroded aerodynamic ridges) in trending ignimbrite flare-up in the Puna due to mantle
the Tara outflow ignimbrite. Photo : Shanaka de Silva melting associated with re-steepening of the subducting
Nazca Plate. As shown in Figure 1, this concept was then
extended northwards into Peru by Freymuth et al. (2015).
As early as 1948, incandescent tuff flows in southern Peru, In conjunction, an Early Miocene acceleration in the rate of
parts of which are now known as the Arequipa Sillar, had westward drift of South America over the Nazca Plate may
been described. The ignimbrites of northern Chile had have contributed to the contractional tectonic environ-
been ascribed to the Liparitica Formation after the tuffs of ment that produced over 300 km of Central Andean short-
Lipari (Italy) that were the standard ignimbrites of the day. ening (Oncken et al. 2006) and contributed to eclogite
Early work on the volcanic stratigraphy in the San Pedro facies metamorphism at the base of a thickening ~50–70
and San Bartolo areas of northern Chile were followed km crust. Sporadic delamination of dense lower crust,
by the geomorphological, tectonic, and volcanic studies along with the subcontinental lithospheric mantle above
by the University of London groups in the 1960s. These the southwardly steepening slab, created a pathway for
were instrumental in establishing the Upper Miocene to mantle magmas to interact with the base of the crust and
Recent age of the ignimbrites as the interplay of tectonism focus the elevated thermal input needed to fuel the ignim-
and volcanism began to elegantly unravel. Chemical and brite flare-up (e.g. Kay and Coria 2009) (Fig. 3). After the
isotopic studies by German groups in the 1970s let to the hot middle crust was established, piecemeal delamination
recognition that the ignimbrites represented significant and upper-crustal deformation likely occurred episodi-
crustal processing. Hamilton (1969) proposed that the cally to produce the observed peaks in volcanic output
Central Andes was tectonically analogous to the Cretaceous in the Altiplano–Puna Volcanic Complex (e.g. de Silva et
in North America and argued that “modern analogues for al. 2015). Evidence for a delaminating lithospheric block
the great Late Cretaceous batholiths appear to be present beneath Cerro Galán caldera in the southern Puna has
only in the volcanic fields of the continental margin … been revealed by recent seismic studies (e.g. Heit et al.
particularly in the Central Andes.” He explicitly connected 2014; Liang et al. 2014).
the ignimbrites with plutonic rocks in saying “the great
elongate field of ignimbrites, beneath which crystallized TRANSCRUSTAL HIGH-FLUX MAGMATISM
granite batholiths”.
The Neogene Central Andean ignimbrite flare-up is the
surface record of a transcrustal magmatic system that itself
THE REVOLUTIONARY SYNOPTIC VIEW reflects the reaction of the crust to an elevated input of
FROM SPACE mantle heat and fluid. Geophysical and petrological foren-
The advent of satellite imagery revolutionized volcanic sics indicate that the processing of mantle input to the
studies in the Central Andes. Photographs taken by Skylab crust takes place in three different domains: the lower,
and the invaluable, ongoing, set of images taken by the middle, and upper crust (e.g. de Silva 1989; Davidson et
Landsat suite of satellites provided a synoptic perspec- al. 1991; Coira et al. 1993; de Silva et al. 2006; Kay et al.
tive that has revealed the sources of regionally extensive 1994, 2010) (Fig. 4).
Andean ignimbrites to be large calderas (Fig. 2). This, in
turn, expanded a knowledge of Andean volcanic stratig- Lower-Crust
raphy and chronology. The rheological and density trap at the Moho leads to
ponding of mantle-derived, primitive magmas. As gravi-
tational instabilities develop in this region, mantle heat

E lements 246 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user
AA steady-state subduction regime Middle- to Upper-Crust
frontal arc volcanism Subandean Small basaltic andesite to andesite volcanic centers with
CVZ thrusting
phase assemblages indicating equilibration at depths of
0 AFC 30 km to 10 km throughout the flare-up region provide
MASH
km evidence of progressive mid-crustal differentiation of
100 parental magmas derived from the lower crustal MASH
zone (e.g. Kay et al. 1994, 2010; Risse et al. 2013; Burns
200 Stage 4 et al. 2015). These magmas pond in the mid-crust and
fractionate into the more silicic melts that ascend to higher-
0 200 400 600 800 level pre-eruptive magma chambers from which they
eventually erupt. Extensive regional-scale, low-velocity
near & backarc regions identified in seismic images at this level (Fig. 4A)
BB calderas and ignimbrites are thought to represent magmatic mush containing up to
22% melt (e.g. Altiplano–Puna Magma Body, or the Cerro
steepening Subandean Galán Magma Body) (Ward et al. 2017). Transfer of melts
slab thrusting from the mush to the upper crust appears to be episodic,
0 Upper Crustal AFC
controlled either by internal instabilities, deep recharge,
AFC
km MASH
and/or upper crustal tectonics (e.g. Riller et al. 2001; Kay
100 and Coira 2009; de Silva et al. 2015).
Stage 3 delam
200 inatio
n
Upper Crust
Phase petrology of the ignimbrites indicate that magmas
200 400 600 800
0 were stored pre-eruptively over a depth range from about
3 km to 8 km (e.g. Folkes et al. 2011; Grocke et al. 2017).
broad arc Puna & Cordillera Further differentiation at these uppermost crustal levels is
initial slab small stocks/domes Oriental thrusting
CC steepening
indicated by equilibration pressure estimates from amphi-
bole, experimental evidence of late crystallization of quartz
0 AFC and sanidine, and zircon with oxygen isotopic and trace-
MASH
km element compositions characteristic of upper crustal differ-
100 entiation. Age and thermochemical modeling of zircon
characteristics suggest that eruptions tapped magmatic
200 Stage 2 zones that contained melt for periods of 100 ky to over
200 400 600
1 My before eruption (see Kern et al. 2016) and were respon-
0 800
sible for local upper-crust ratios of plutonic to volcanic
rocks up to 70:1 (Tierney et al. 2016). Such long melt-
ridge subduction Puna & Cordillera
DD magmatic gap Oriental thrusting
present lifetimes are the consequence of elevated tempera-
tures and storage-enhancing rheological conditions in the
0 crust upper crust (e.g. de Silva and Gregg 2014). Building such
km lithosphere large long-lived magma reservoirs has a profound effect
100 Ridge on wall-rock rheology and promotes storage over eruption,
leading to reservoir growth rather than eruption under
Stage 1 conditions of constant recharge flux and/or crystalliza-
200
tion. The eventual eruptions are triggered by the failure
200 400 600 800
0 Km of the weakened and extended roof as faults propagate
Schematic representations of the geodynamics of the from the surface into the magma system. Locally, exten-
Figure 3 sional regional tectonics, due to plateau uplift in response
Central Andean arc during the Neogene as a conse-
quence of the passage of the aseismic Juan Fernández Ridge. to crustal shortening and lower crustal flow (Kay and Coira
Abbreviations are as follows: AFC (assimilation, fractional crystalliza- 2009), may have played a critical role in triggering the
tion); MASH (mixing, assimilation, storage, homogenization); CVZ
(Central Volcanic Zone). Blue is subducting slab; yellow is crust; eruptions (e.g. Riller et al. 2001).
black is subducted ocean ridge; oblique lined is subcontinental
mantle lithosphere. Panels are not specific to any particular latitude DOES HIGH-FLUX MAGMATISM CREATE
and depict an integration of the processes that occur throughout
the arc region. Present-day situation at top (A). (A) Present day –
OR DESTROY CRUST?
steady state arc; steep slab. (B) Slab steepening, delamination, The volumes of ignimbrites, their plutonic underpin-
ignimbrite flare-up. (C) Post–ridge passage, slab steepening. nings, and the ignimbrites’ “crustal” isotopic composi-
(D) Initial condition - ridge subduction, flat slab. Modified from
K ay  and Coira (2009). tions evoke speculations about the crustal evolution of
the Central Andes and the relative role of flare-up modes
in the balance of continental crust created and recycled
back into the mantle. This has been a recurrent issue in
input is transferred upwards by advection. Injection and the Central Andes because crustal structure and magmatic
ponding of primitive magmas at the base of the crust architecture conspire to filter out geochemical signatures
generate and mix with crustal melts, leaving behind dense of primitive mantle, thereby leaving a critical ambiguity as
cumulates that, in turn, can be removed by delamination to the balance between new mantle addition and recycled
(foundering). Thermal and mechanical feedbacks result in crust in the erupted magmas. According to mass balance,
a lower-crustal zone of mixing, assimilation, storage, and the ignimbrites are roughly 50:50 volume mixtures of
homogenization (MASH) where baseline isotopic composi- mantle-derived basalts and regionally variable crust (e.g.
tions are set (Hildreth and Moorbath 1988; Davidson et Kay et al. 2010; Freymuth et al. 2015). Yet, uncertainty
al. 1991). and debate persists as to the isotopic characteristics of the
mantle source. While a typical depleted island arc mantle
source has generally been inferred under the Altiplano–
Puna Plateau (e.g. Davidson et al. 1991; de Silva et al.

E lements 247 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user
A B
Resurgent Composite
Minor Caldera Volcano
-5 Center
0

20

40

60

80
Subduction flux-modified mantle wedge
km
Island arc tholeiite Enriched mantle
87
Sr/86Sr = 0.704; Sr = 200 ppm OR 87
Sr/86Sr = 0.7055; Sr = 500 ppm
e.g. Davidson et al. 1991 e.g. Kay et al. 2010

Interpretations of geophysical and petrochemical seismic tomography from the APVC in Figure 4A. These data suggest
Figure 4
forensics of the Central Andean subcaldera crust. three staging levels in the crust for the magma that eventually
(A) Two seismic tomographic vertical slices through the lithosphere erupts to produce the ignimbrites. The seismic data most clearly
beneath the Altiplano–Puna Volcanic Complex (APVC) and Cerro image the upper and middle crustal melt storage zone, labeled as
Galán regions. Seismic images suggest the mid-crustal melt zone is the two MASH (mixing, assimilation, storage, homogenization)
deeper in the Cerro Galán region [CGMB = Cerro Galán Magma zones. The magmas become progressively more differentiated as
Body] than in the Altiplano-Puna region [APMB = Altiplano–Puna magmatism propagates upward through the crust, with the final
Magma Body]. The APVC and Galán regions are indicated by the erupted ignimbrites representing an approximately 50:50 mantle-
dashed black lines. Black triangles are volcanoes. Green lines show to-crust melt mix, although the nature of the mantle source region
the borders between Chile, Bolivia, and Argentina. From Ward et al. is debated. Figure 4B is based on the APVC region (APMB) in Figure
(2018). (B) Central Andean crustal magmatism pictorially summa- 4A. Data sources and other references can be found online at elements-
rized. Petrological and geochemical evidence layered over the magazine.org

2006), an enriched mantle is favored by others and is oceanic Aleutian arc and greater than that for the Sierra
required under the southern Puna where Mg- and Cr-rich Nevada (USA) flare-up when only new crustal addition is
late Neogene mafic lavas have enriched 87Sr/ 86Sr isotopic considered (see Jicha and Jagoutz 2015) (Table 1).
ratios (>0.705) (e.g. Kay et al. 1994; Risse et al. 2013). As In addressing the question of crustal volume change during
this enriched signature is strongest in late Neogene mafic magmatic flare-ups, new crustal input from the mantle
magmas, the recent enrichment has been interpreted to must be balanced against crustal loss due to removal.
reflect the introduction of crustal material into the mantle And this removal can be caused by (1) delamination (also
wedge through a combination of delamination and forearc referred to as foundering or dripping) of mafic cumulates
subduction erosion (see Kay et al. 2010; Risse et al. 2013; and lower crust whose density exceeds that of the under-
Goss et al. 2013). lying mantle; (2) forearc subduction erosion. The bottom
Whereas the prevailing view has been that flare-ups repre- of Table 1 addresses the possible crustal loss under the Puna
sent a time of crustal growth through magmatic addition, from 21°S to 27°S over the last 11 My. A rough volume
the important—maybe even dominant—role for delamina- estimate of high-velocity mantle regions, interpreted as
tion and subduction erosion, and its impact on the mass delaminated blocks on seismic tomographic images under
balance of crustal addition and loss, needs to be considered. the region (e.g. the Galán region in Liang et al. 2014),
As the compositions and erupted volumes of the ignimbrites permits the loss of more than one million km3 of basal
can be observed and calculated to acceptable precision, and crust. Adding in crustal loss by forearc subduction erosion
the crust-to-mantle ratio in the erupted magmas can be near 27°S (Goss et al. 2013) produces an additional loss
assumed to be near 50:50, the critical unknowns relevant to of ~992 km3. Averaging this crustal loss over the region
mass balance are the plutonic versus volcanic ratio and the from 21°S to 27°S leads to removal rates of −155 km3/km/
percent of crust recycled back into the mantle. Petrological My for delamination and of −15 km3/km/My for forearc
and volcanological approaches have traditionally adopted subduction erosion.
overall ratios of plutonic-to-volcanic ratios of 3:1 to 10:1, A comparison of the volume of the crust that could have
respectively. However, recent ambient noise crustal-scale been lost with the volume of existing crust between 21°S
tomographic seismic images under the Altiplano–Puna and 27°S shows an approximate 5.5% crustal loss over the
Plateau suggest average plutonic-to-volcanic ratios as high region during the last 11 My (see Table 1). Therefore, from
as 35:1 based on interpreted volumes of magma mush zones a simplified viewpoint, the Puna has been a site of net
containing an ~22% melt fraction (Ward et al. 2017). crustal loss rather than crustal gain during the flare-up over
Table 1 shows estimates of new crustal volumes added to the the last 11 My. Furthermore, recycling of basal and forearc
Puna crust from the mantle during the magmatic flare-up crust removed by delamination and forearc subduction
that happened over the last 11 My, based on representative erosion into the mantle wedge provides a mechanism for
plutonic-to-volcanic ratios of 10:1 and 35:1 and having a enriching the mantle magma source beneath the plateau
50% new mantle contribution. These new crustal volumes during the Neogene. Overall, the Central Andean plateau
correspond to addition rates of 14–45 km3/km/My for the is an area where crustal processing is producing an increas-
Puna from 21°S to 27°S (Galán region) and 33–107 km3/km/ ingly more evolved (granodioritic) crust as it evolves from a
My for the Altiplano–Puna Volcanic Complex. The latter continental margin to a region of stable continental crust.
rate is about half that suggested for the initial stages of the

E lements 248 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user
Table 1 ESTIMATES OF MAGMA ADDITION AND CRUSTAL GAIN AND LOSS UNDER THE ANDEAN PLATEAU (21°S–27°S). Estimates
focus on the best-studied and constrained portion of the Central Andean plateau from 21°S to 27°S latitude, which includes
the Cerro Galán and Altiplano–Puna Volcanic Complex (APVC) regions (see Fig. 1 for locations). Erupted volumes are converted to crustal
magmatic volumes based on a plutonic-to-volcanic ratio of 10:1; the extreme ratio of 35:1 by Ward et al. (2017), among others, is based
on joint inversions of Rayleigh-wave dispersion from ambient seismic noise with P-wave receiver functions. All estimates assume 50:50
crust-to-mantle ratio that is constrained by trace element and O and Sr isotopic ratios (e.g. Kay et al. 2010). Estimates of crustal loss by
delamination under the Andean Plateau are guided by seismic P-wave and attenuation tomographic images. The forearc subduction
erosion rate at 26°S to 27°S is from Goss et al. (2013). Note that these are integrated rates and that episodic rates can be considerably
higher (de Silva et al. 2015). Data sources and other referencecs can be found at elementsmagazine.org.

New Magma Addition under Central Andean Plateau from 21°S to 27°S in last 11 My
Plateau from 21°S to 27°S All Galán region APVC region
Length of arc in km 600 100 230
Time of eruption in million years 11 6 11
Erupted volume in km3 16,490 1,490 15,000
Added crustal volumes and new crustal addition rates
with a new mantle contribution of 50%
At a plutonic to volcanic ratio of 10 to 1
Volume of new crust from the mantle in km3 90,695 8,195 82,500
3
New crustal growth rate (km /km of arc/My) 14 14 33
At a plutonic-to-volcanic ratio of 35 to 1
Volume of new crust from the mantle in km3 593,640 53,640 540,000
3
New crustal growth rate (km /km of arc/My) 45 45 107
Comparative new crustal growth rates in km3/km/My in other arcs
(see Jicha and Jagoutz 2015)
Initial growth-rate of oceanic Aleutian arc ~210 assumes all new mantle contribution
Sierra Nevada flare-up growth rate ~90 ~45 if new mantle contribution is 50%

Crustal Loss under Central Andean Plateau from 21°S to 27°S in last 11 My
Estimate of crustal gain and loss rate Modern Puna crustal volume
All in km3/km of arc per My at 22°S to 27°S per km of arc east–west
New crust created by magmatic addition at plutonic-to-volcanic Width kms east to
45 500
ratio of 35 to 1 west (65°W to 71°W)
Average crustal
Delamination of 20 km of basal crust (1,025,000 km3 lost) –155 50
thickness (km)
Forearc subduction erosion at 28°S to 26°S rate of 192 km3/km/ Crustal volume in
–15 25,000
My for 5 My km3 per km today
3
km of crust
Time Period of Loss % of crust removed
Lost Remaining
last 1 million years –125 24,875 –0.5%
last 8 million years –997 24,003 –4.0%
last 11 million years –1.370 23,630 –5.5%

IMPLICATIONS AND FUTURE DIRECTIONS in continental arc evolution in which significant crustal
The Neogene ignimbrite flare up of the Central Andes differentiation occurs and in which andesitic (granodio-
records a change from steady state to high-flux magmatism. ritic) crust is produced at the same time as dense mafic
The drive for this transition appears primarily to be subduc- crust is lost to the underlying mantle. This process further
tion of an aseismic ridge and the attendant changes in the enhances the differentiation of the continental crust to
crust–mantle interface. The details of the time lag between more silicic compositions. Identifying and quantifying the
the changes in the subduction zone and the magmatic overall crustal recycling processes remains a fertile field for
response remain to be understood. future investigations.

A transcrustal magmatic system that has lower-, middle-, The thermomechanical consequences of progressive crustal
and upper-crustal stages is the signature of high-flux weakening during the peaks of a flare-up are critical to
magmatism and supports the original view of Hamilton developing a long-lived system from which andesitic and
(1969) that a crustal-scale batholith underpins the Central dacitic magmas can periodically rise and erupt as ignim-
Andes, the modern modification being that this batho- brites. Linking the development and timing of shallow
lith is probably time transgressive. Parts of this batholith pre-eruptive magma reservoirs to deeper stages of differen-
have been imaged geophysically, but much remains to be tiation is a continuing challenge. Given that surface uplift
imaged and resolved locally and regionally, particularly in may be partly a consequence of upper-crustal batholith
the deeper parts of the crust. building (Perkins et al. 2016), understanding links between
magmatism, uplift, stress regimes on the plateau, crustal
The thermomechanical consequence of a transcrustal thickening and shortening, lower crustal flow, delami-
magmatic system is an upward transgressive wave of nation, and eruptions is likely to be an exciting area of
differentiation that leads to a more granodioritic crustal research.
composition. A flare-up stage is an evolutionary phase

E lements 249 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user
Although the prevailing model of crustal growth through Andean and subduction-related magmatism. Funding from
magmatic addition holds sway, the possibility that the National Aeronautical and Space Administration and
magmatic flare-ups in areas of thickened crust can be times the National Science Foundation have been critical for this
of net crustal loss, principally through the removal of basal work. Support and encouragement by Guest Editor Gerhard
crust, requires further investigation. Understanding the Wörner and reviews by Robert Trumbull, an anonymous
mechanisms and resolving the quantity of crustal loss reviewer, Gerhard Wörner, and Principal Editor Jon Blundy
throughout the entire Andean region remains an area of have clarified and improved the form and content of this
future investigation, one that will require resolving the paper significantly. Suzanne Kay acknowledges the many
nature of the mantle source. insightful contributions of Beatriz Coira (Argentina),
Constantino Mpodozis (Chile) and other South American
ACKNOWLEDGMENTS colleagues that have lead to our current understanding of
Shan de Silva acknowledges the vital roles of Peter Francis the magmatic and tectonic processes of the central Andean
and Jon Davidson (both deceased) in his understanding of plateau.

REFERENCES Freymuth H, Brandmeier M, Wörner G crustal and mantle sources and evolu-
(2015) The origin and crust/mantle tion of central Andean Puna plateau
Best MG, Christiansen EH, de Silva S, mass balance of Central Andean ignim- ignimbrites. Journal of Volcanology and
Lipman PW (2016) Slab-rollback ignim- brite magmatism constrained by oxygen Geothermal Research 198: 81-111
brite flareups in the southern Great and strontium isotopes and erupted
Basin and other Cenozoic American Kern JM and 5 coauthors (2016)
volumes. Contributions to Mineralogy Geochronological imaging of an episod-
arcs: a distinct style of arc volcanism. and Petrology 169, doi: 10.1007/
Geosphere 12: 1097-1135 ically constructed subvolcanic batho-
s00410-015-1152-5 lith: U-Pb in zircon chronochemistry of
Burns DH, de Silva SL, Tepley F III, Goss AR, Kay SM, Mpodozis C (2013) the Altiplano-Puna Volcanic Complex
Schmitt AK, Loewen MW (2015) Andean adakites-like highMg andesites of the Central Andes. Geosphere 12:
Recording the transition from flare-up on the northern margin of the Chilean- 1054-1077
to steady-state arc magmatism at the Pampean flat-slab (27-28.5°S) associated
Purico–Chascon volcanic complex, Liang X and 9 coauthors (2014)
with frontal arc migration and forearc Delamination of southern Puna
northern Chile. Earth and Planetary subduction erosion. Journal of Petrology
Science Letters 422: 75-86 lithosphere revealed by body wave
11: 2193-2234 attenuation tomography. Journal of
Coira B, Kay SM, Viramonte J (1993) Grocke SB, de Silva SL, Iriarte R, Lindsay Geophysical Research: Solid Earth 119:
Upper Cenozoic magmatic evolu- JM, Cottrell E (2017) Catastrophic 549-566
tion of the Argentine Puna—a model caldera-forming (ccf) monotonous
for changing subduction geometry. Oncken O and 5 coauthors (2006)
silicic magma reservoirs: geochemical Deformation of the Central Andean
International Geology Review 35: and petrological constraints on
677-720 upper plate system—facts, fiction, and
heterogeneity, magma dynamics, and constraints for plateau models. In:
Davidson JP, Harmon RS, Wörner G eruption dynamics of the 3.49 Ma Tara Oncken O and 7 coeditors (eds) The
(1991) The source of Central Andean Supereruption, Guacha II Caldera, Andes—Active Subduction Orogeny.
magmas; some considerations. In: SW Bolivia. Journal of Petrology 58: Springer-Verlag, Berlin, pp3-27
Harmon RS, Rapela CW (eds) Andean 227-260
Magmatism and its Tectonic Setting. Perkins JP and 5 coauthors (2016) Surface
Hamilton W (1969) The volcanic Central uplift in the Central Andes driven by
Geological Society of America Special Andes—a modern model for the
Paper 265, pp 233-244 growth of the Altiplano Puna magma
Cretaceous batholiths and tectonics of body. Nature Communications 7, doi:
DeCelles PG and 9 coauthors (2015) western North America. In: McBirney 10.1038/ncomms13185
Cyclical orogenic processes in the AR (ed) Proceedings of the Andesite
Cenozoic central Andes. In: DeCelles Conference, Oregon, July 1968, Pritchard ME and 30 coauthors (2018)
PG, Ducea MN, Carrapa B, Kapp PA International Upper Mantle Project, Synthesis: PLUTONS: investigating the
(eds) Geodynamics of a Cordilleran Science Report Number 16. Oregon relationship between pluton growth
Orogenic System: The Central Andes Department of Mineral Industries and volcanism in the Central Andes.
of Argentina and Northern Chile. Bulletin 65, pp 175-184 Geosphere 14: 954-982
Geological Society of America Memoir Heit B and 9 coauthors (2014) Structure Riller U, Petrinovic I, Ramelow J, Strecker
212, pp 459-490 of the crust and the lithosphere beneath M, Oncken O (2001) Late Cenozoic
de Silva SL (1989) Altiplano-Puna the southern Puna plateau from tectonism, collapse caldera and plateau
volcanic complex of the Central Andes. teleseismic receiver functions. Earth formation in the central Andes. Earth
Geology 17: 1102-1106 and Planetary Science Letters 385: 1-11 and Planetary Science Letters 188:
299-311
de Silva SL, Gregg PM (2014) Hildreth W, Moorbath S (1988) Crustal
Thermomechanical feedbacks in contributions to arc magmatism in the Risse A, Trumbull RB, Kay SM, Coira B,
magmatic systems: implications for Andes of central Chile. Contributions to Romer RL (2013) Multi-stage evolu-
growth, longevity, and evolution of Mineralogy and Petrology 98: 455-489 tion of Late Neogene mantle-derived
large caldera-forming magma reservoirs magmas from the Central Andes
Jicha B, Jagoutz O (2015) Magma produc- back-arc in the southern Puna plateau
and their supereruptions. Journal of tion rates in intraoceanic arcs. Elements
Volcanology and Geothermal Research of Argentina. Journal of Petrology 54:
11: 105-111 1963-1995
282: 77-91
Kay SM, Coira BL (2009) Shallowing and Tierney CR, Schmitt AK, Lovera OM, de
de Silva SL and 5 coauthors (2006) Large steepening subduction zones, conti-
ignimbrite eruptions and volcano- Silva SL (2016) Voluminous plutonism
nental lithospheric loss, magmatism, during volcanic quiescence revealed by
tectonic depressions in the Central and crustal flow under the central
Andes: a thermomechanical perspec- thermochemical modeling of zircon.
Andean Altiplano-Puna Plateau. In: Kay Geology 44: 683-686
tive. In: Troise C, de Natale G, Kilburn SM, Ramos VA, Dickinson WR (eds)
CRJ (eds) Mechanisms of Activity and Backbone of the Americas: Shallow Ward KM, Delph JR, Zandt G, Beck SL,
Unrest at Large Calderas. Geological Subduction, Plateau Uplift, and Ridge Ducea MN (2017) Magmatic evolu-
Society, London, Special Publications and Terrane Collision. Geological tion of a Cordilleran flare-up and its
269: 47-63 Society of America Memoir 204, pp role in the creation of silicic crust.
de Silva SL, Riggs NR, Barth AP (2015) 229-259 Scientific Reports 7, doi: 10.1038/
Quickening the pulse: fractal tempos in s41598-017-09015-5c
Kay SM, Coira B, Viramonte J (1994)
continental arc magmatism. Elements Young mafic back arc volcanic rocks as Yáñez GA, Ranero CR, von Huene R,
11: 113-118 indicators of continental lithospheric Díaz J (2001) Magnetic anomaly inter-
Folkes CB, de Silva SL, Wright HM, Cas delamination beneath the Argentine pretation across the southern central
RAF (2011) Geochemical homogeneity Puna plateau, central Andes. Journal of Andes (32°-34°S): the role of the Juan
of a long-lived, large silicic system: Geophysical Research: Solid Earth 99: Fernández Ridge in the late Tertiary
evidence from the Cerro Galán caldera, 24323-24339 evolution of the margin. Journal of
NW Argentina. Bulletin of Volcanology Geophysical Research: Solid Earth 106
Kay SM, Coira BL, Caffe PJ, Chen C-H (B4): 6325-6345
73: 1455-1486 (2010) Regional chemical diversity,

E lements 250 A ugus t 2018


Downloaded from https://pubs.geoscienceworld.org/msa/elements/article-pdf/14/4/245/4486155/gselements-14-4-245.pdf
by Yonsei University user

You might also like