Thouret2016 PDF
Thouret2016 PDF
Abstract: With an area exceeding 25 000 km2 and volumes c. 5000 km3, south Peru hosts the Andes’ second largest Neogene
ignimbrite field. We document the extent, stratigraphy and chronology of 12 ignimbrite sheets in the Río Ocoña–Cotahuasi–Marán
and Colca deep canyons. Based on 74 40Ar/39Ar and U/Pb age determinations, ignimbrite-forming episodes span 25 myr. Prior to
9 Ma, eight large-volume ignimbrites were produced every 2.4 myr. After 9 Ma, average lulls between small- to moderate-volume
ignimbrites decreased to 0.85 myr. The refined volcanic stratigraphy reveals three main features. (1) Larger volume ignimbrites
were emplaced by punctuated flare-ups between 25 and 9 Ma during uplift of the Western Cordillera. (2) Numerous smaller
ignimbrites were emplaced after 9 Ma as the ignimbrite production rate decreased threefold. This decrease may be due to the
declining crustal melting rate, decreasing plate convergence rate after 9 Ma, or more magma stagnation in the shallow crust, which
promoted the growth of composite cones. (3) Growth of two volcanic arcs has added twice as much volume (c. 53 km3 Ma−1) to
the Río Ocoña–Cotahuasi–Marán volcanic field than the ignimbrites after 2.27 Ma. Estimated linear arc magma output has,
however, decreased twofold (0.15 – 0.08 km3 km−1 Ma−1) from the Early Quaternary to the Pleistocene–Holocene.
Supplementary materials: Supplementary text, tables 1–3 and figures 1–4 are available at [Link]
3147100.
Received 24 June 2015; revised 18 February 2016; accepted 18 February 2016
The abundance of silicic magmas is one of the most prominent Rio Colca region including 52 ignimbrites and 22 intercalated
characteristics of the Neogene magmatic arcs in the Central Andes. Late Miocene to Quaternary lava flows, both of which are
The Andean Central Volcanic Zone produced two of the world’s complemented by 34 new 40Ar/39Ar ages and 10 U/Pb ages on
largest Neogene ignimbrite provinces. The largest, the Altiplano zircons in addition to 30 reported by Thouret et al. (2007). Despite
Puna Volcanic Complex, straddles the border between Chile, uncertainties in the volume estimates, erupted volume and Volcanic
Bolivia and Argentina (de Silva 1989; Salisbury et al. 2011). South Explosivity Index (VEI) are used to subdivide the pyroclastic
Peru hosts the second largest Andean Neogene ignimbrite field, deposits into three categories: extremely large-volume ignimbrites
with widespread ignimbrite sheets covering at least 25 000 km2 (50 – 880 km3, VEI 8 and 9); very large-volume ignimbrites
among c. 54 000 km2 of Tertiary–Quaternary volcanic rocks in the (10–50 km3, VEI 7 and 8); large-volume ignimbrites and PDC
region 15–18°N, 70 – 74°W (Figs 1 and 2; Mamani et al. 2010; deposits (1–10 km3, VEI 6 and 7).
Acosta et al. 2011). Despite being widespread, well preserved in an First, we reconstruct the history of volcanism, in particular that of
arid environment and of economic interest (Noble et al. 1974, 1984; ignimbrites, and, second, we relate the volcanic history to the
Skinner 2009), these ignimbrites have been the focus of detailed evolution of the Western Cordillera. Reconstructing the ignimbrite
geological studies only recently as one of multiple proxies for eruptions and volcanism for the past 25 myr in the Peruvian Central
estimating the valley incision and uplift in Peru and Bolivia Andes requires the determination of the following: (1) the spatial
(Schildgen et al. 2007, 2009, 2010; Thouret et al. 2007; Barnes & extent and characteristics of 12 ignimbrites, half of them being
Ehlers 2009; Gunnell et al. 2010; De La Rupelle 2013). either unknown or poorly described; (2) the stratigraphic and
The ignimbrites of the Río Ocoña–Cotahuasi–Marán (OCM) chronological correlations of the OCM ignimbrite sheets with
region near the northernmost tip of the Peruvian volcanic range (15° ignimbrites identified farther south to harmonize our volcanic
20’S, 73°30’W) are better exposed than the Altiplano Puna stratigraphy with that of Peru; (3) the geographical sources
Volcanic Complex ignimbrites owing to the exceptional 3.5 km (calderas) in the Western Cordillera and its divide with the
deep incision of the north–south-trending OCM canyon system, Altiplano. We seek to understand to what extent ignimbrites can
which runs perpendicular to the Andean arc. The extent (200 km be used to unravel the volcanic history of the Central Andes.
north–south and 110 km east–west) and geometry of the deep OCM
and Colca canyon catchments (Figs 1, 2, 3 and 4) allow correlations
of ignimbrite sheets that either mantle the pre-25 Ma high plateaux Geomorphological and geological setting of the region
or are confined in deep valleys cut in the pre-Tertiary bedrock.
Geomorphological setting and regional geology
Twelve distinct rhyolitic ignimbrites and pyroclastic density current
(PDC) deposits from the OCM canyons are used as marker horizons This overview of the region summarizes a more complete
and allow us to trace the history of volcanism of the west Central description by Benavides-Cáceres (1999) and Gunnell et al.
Andes for the past 25 myr (Figs 1–4, Tables 1 and 2). We develop a (2010). The OCM area includes parts of the Coastal Cordillera,
chronostratigraphic sequence for the OCM and neighbouring the Western Cordillera and the western edge of the Altiplano at the
© 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: [Link]
Publishing disclaimer: [Link]/pub_ethics
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
northern end of the modern Central Andean steep subduction sandstone and siltstone (Huaman 1985; Roperch et al. 2006; Decou
segment (Figs 1, 2 and 4). et al. 2011, 2013; Sempere et al. 2014). The 0.7 km thick upper
In the Peruvian OCM canyon territory, the 1.5–2.0 km high sequence of the Moquegua Formation forms an upward coarsening
Coastal Cordillera consists of a Precambrian to Palaeozoic unit of nonmarine conglomerate, sandstone and siltstone with an
metamorphic basement, the Arequipa Massif (Ramos 2008), increasing frequency of tuff intercalations (Fig. 3). Clasts in the
covered by a sequence of Cenozoic forearc sediments. Instead of lower Moquegua A sequence include coastal batholith rocks
an abrupt topographic step typical of many mountain fronts, the (Huaman 1985; Roperch et al. 2006), demonstrating that the
Western Cordillera of Peru forms a large monocline (Figs 2 and 4) plutons were already emplaced by Oligocene time. A stratigraphic
that extends from Nazca to the Chilean border. Here the orogenic discontinuity exists between the Moquegua B unit (45–30 Ma),
plateau connects to the Pacific coastline via a 60 km long which is almost devoid of volcanic material, and the Moquegua C
topographic ramp, descending from c. 4.4 km above sea level (30–15/13 Ma) and D (14/10–4.5 Ma) units. The top of the
(a.s.l.) to sea level (Fig. 4). The slope of such a large, continuous Moquegua C unit is diachronous along the Western Cordillera,
monocline is primarily an effect of crustal deformation of the pre- which led Sempere et al. (2014) to distinguish two subunits C1 and
Chuquibamba land surface in response to crustal mechanisms that C2. The units C and D contain abundant volcaniclastic and
have been discussed by Gunnell et al. (2010). Where rivers have pyroclastic material (Roperch et al. 2006; Decou et al. 2011)
incised the ramp, the canyon drainage tends to expand laterally, consistent with an increase in large silicic ignimbrite-forming
developing a WNW–ESE-trending anticlinal valley and exposing eruptive episodes.
the strike-parallel batholith (Figs 1, 2 and 4). Widespread ignimbrites overlie high-relief erosional surfaces
The oldest arc rocks belong to the Chocolate (195–190 Ma) and sloping toward the Pacific Ocean that are polygenetic peneplains
Toquepala (78–50 Ma) Formations. In the study area, where the levelling the pre-Tertiary basement. Other smaller ignimbrites fill
Chocolate Formation is absent, granodiorite and tonalite units, valleys cut in the high, uplifted plateaux whereas distal, thinner
collectively known as the Andean batholiths, have been ascribed to layers are intercalated with conglomerates and alluvial strata of
middle Cretaceous and late Cretaceous–Palaeogene magmatism basins. The bulk volume of ignimbrites and PDC deposits coincides
(e.g. Pecho Gutierrez 1983a,b; Beckinsale et al. 1985; Demouy with the 30 myr period of major uplift and crustal thickening of the
et al. 2012). These crystalline rocks crop out where the studied Central Andes. Mamani et al. (2010) described rocks and
canyons and their tributaries have cut deepest into the crust (Fig. 4). formations developed over the past 30 myr; namely, the Tacaza
The Cenozoic inner-forearc sediments that separate the Western (30–24 Ma), Huaylillas (24–10 Ma), Lower Barroso (10–3 Ma) and
Cordillera from the Pacific coastline are known as the Moquegua Upper Barroso (3–1 Ma) units, and the frontal arc <1 Ma. Uplift
Formation. Deposited in extensional basins west of the Cordillera, began in the Western and Coastal Cordilleras c. 50 myr ago, perhaps
this formation comprises >1 km of Eocene to Miocene continental as early as 70 Ma (McQuarrie et al. 2005; Oncken et al. 2006), and
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 2. NASA Shuttle photograph showing the Andean structures that parallel the Pacific Trench from the forearc (south) to the Altiplano (north), the two
deep canyon systems (OCM and Colca), and the frontal volcanic range. This map also indicates the principal ríos (rivers) and cities or towns mentioned in
the text. Continuous and dashed lines delineate the Pleistocene–Holocene volcanic range and the Early Quaternary range, respectively. HC: Huarcaya
caldera; LTCH: La Torta-Huancarama caldera.
continued later and more slowly in the Eastern Cordillera Ignimbrite definition, mapping and correlation
(40–10 Ma) while undergoing a marked interruption in the
Western Cordillera (Decou et al. 2013). Uplift resumed in both We use the term ‘ignimbrite’ for voluminous (>1 km3) deposits of
Cordilleras and in the Altiplano c. 30–20 Ma, in phase with periods explosive volcanism, which have been generated by the emplace-
of higher convergence rates between the Nazca and South American ment of PDCs (Wilson & Hildreth 2003). A compound ignimbrite is
plates (Sébrier et al. 1988), but accelerated significantly after 10 Ma one that has multiple flow units, which are distinguished from each
(Schildgen et al. 2007, 2009; Thouret et al. 2007; Garzione et al. other by temporal boundaries (Wilson & Hildreth 2003). A simple
2008; Schildgen 2009) in a context of decreasing convergence cooling unit is a body that has been emplaced and undergone
velocity between the Farallón–Nazca plate and South America continuous cooling while its earliest material was still hot, whereas a
(Iaffaldano & Bunge 2008; Martinod et al. 2010). composite sheet is a deposit that has significant time-breaks
between cooling units (Smith 1960; Branney & Kokelaar 2002). We
define each of the 12 ignimbrites as either a voluminous sheet
Chronology methods and ignimbrite correlation (single or compound cooling unit including multiple flow units) or a
succession of non-welded PDC deposits, which are multi-layered
Ar/Ar and U/Pb analytical methods
and small in volume (<5 km3).
The 40Ar/39Ar geochronology was determined using ignimbrite We distinguish and correlate cooling units and, where identifi-
pumice and lava flows and U–Pb geochronology was determined by able, flow units emplaced between c. 24.43 and c. 0.91 Ma on the
laser ablation inductively coupled plasma mass spectrometry on basis of stratigraphic unconformities, presence of vitrophyres,
zircons. The 40Ar/39Ar analyses from 54 ignimbrite ( pumice) changes in lithofacies, and 40Ar/39Ar and U/Pb age determinations
samples and 22 lava flows, and results of 280 U/Pb analyses of (Tables 1 and 2). We show how these ignimbrites fit in the
zircons from 10 ignimbrite units are summarized in Table 2. 30 myr OCM stratigraphic scheme compared with the Neogene
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Fig. 3. Composite stratigraphy of the 12 ignimbrite sheets and PDC deposits of the OCM region, lava flows and their setting in volcaniclastic and
continental siliciclastic deposits. Correlations are shown with other formations; in particular, with ignimbrites of Orcopampa east of the OCM region
(Swanson et al. 2004), Arequipa (Thouret et al. 2001; Paquereau-Lebti et al. 2006, 2008) and in southernmost Peru. A detailed stratigraphy of the
Moquegua Formation has been given by Sempere et al. (2014).
ignimbrite stratigraphy in adjacent Orcopampa and southernmost pyroclastic unit were calculated using the spatial analyst tools of
Peru (Fig. 3). the software. Sampling focused on both unweathered pumice and
A geological map and cross-sections show the extent of the whole-rock from ignimbrite units. Petrographical, mineralogical
ignimbrites, lava flows and bedrock in the OCM region (Figs 4–6). and geochemical analyses are described in Figure 4c. Petrographic
Digital geological maps have been created using the ArcGIS examinations including the volume distribution of volcanic versus
platform and checked by field surveys. Surface areas of each lithic components, vesicles and groundmass together with
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 4. Two geological maps, (a) north and (b) south of the OCM region, draped on the NASA Shuttle Radar Topography Mission (SRTM) digital
elevation model (DEM), highlighting the 12 ignimbrites and PDC deposits together with the Late Miocene to Pleistocene lava flows. The legend for the
geological maps is shown in (c). Non-volcanic bedrock has been simplified (Benavides-Cáceres 1999; Ramos 2008; Gunnell et al. 2010). Sites of principal
logs and dated samples are indicated. (d) OCM ignimbrites in a total alkali v. silica diagram after Le Bas et al. (1986).
representative mineral assemblage have been performed on 38 Tacaza Group (Newell 1949), the ‘volcánico’ Huilacollo Formation
samples to distinguish lithofacies including mineral suites of each (Wilson & Garcia 1962), the ‘volcánico’ Alpabamba Formation
ignimbrite. Lithofacies at the outcrop and sample scales help to (Guevara & Dávila 1983), the Huaylillas Formation (Wilson &
distinguish the OCM Neogene and Quaternary rhyolitic ignimbrites Garcia 1962), and the ‘Volcánico’ Sencca, thought by Mendívíl
(Fig. 5). Most ignimbrites are rhyolitic in composition but some (1965) to overlie both lower and upper members of the ‘Barroso
samples of Upper Sencca and Lomas lie in the trachyte and dacite Group’ (Table 2). The latter encompasses mostly lava flows from
fields in the total alkali v. silica diagram after Le Bas et al. (1986) two composite volcanic arcs of Quaternary age (Figs 3 and 4). All
(Fig. 4c). quoted formations include ignimbrites indicated as tuffs in
In Peru, the abundance of poorly dated volcanic formations with INGEMMET geological maps. We named some ignimbrites
local terminology renders the correlation between pyroclastic according to the Peruvian nomenclature but we distinguished
deposits difficult. Peruvian geologists have distinguished the additional ones, either unknown or poorly defined, based on
40
following stratigraphic nomenclature for <30 Ma volcanic deposits Ar/39Ar and U/Pb dates. The first method was also applied to
in geological maps across south Peru: the volcano-sedimentary 23 lava flows that bracket the ignimbrites (Table 2). However,
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Fig. 4. Continued
correlating ignimbrites over SW Peru is challenging owing to C2’ boundary (Decou et al. 2011). The Nazca tuffs are distal beds of
overlaps in geographical locations or diachronous positions, large-volume ignimbrites sourced from the NW. These distal beds,
lithofacies and age determinations. Margins of error associated each corresponding to a single flow unit deposited near sea level in
with the published K–Ar analyses (0.3–1.3 myr) are significantly the western Moquegua basin, crop out in three places in the south
larger than that of our 40Ar/39Ar and U/Pb age determinations OCM region on the cliffs of the Cuno Cuno homocline (Fig. 5a), the
(0.15 myr on average). We therefore retain chronological intervals Pampa Gramadal plateau south of Caravelí (Fig. 5b) and closer to
instead of the formation nomenclature (Fig. 3). this town the Pampa Bonbón plateau (Figs 4, 5d and 6, section A–
A′). The actual Nazca extent is poorly known as exposures are made
visible only by valley incision through the Neogene sediment fill of
Extent, stratigraphy and chronology of ignimbrites
the Caravelí depression. The tuff beds thin out towards the south and
The OCM chronostratigraphy and correlations suggest that 12 SE but thicken towards the WNW and NW (Figs 4 and 6, sections
ignimbrite sheets and PDC deposits have erupted every 1.9 myr on A–A′ and E–E′). Nazca ignimbrites may reflect a change in source
average over the past c. 25 myr: Nazca 1, Nazca 2, Alpabamba, from the distant ‘Tacaza arc’ to the closer ‘Huaylillas arc’ in the
Majes, Chuquibamba, Huarcaya, Caravelí, Arma, Lower Sencca, Western Cordillera (Decou et al. 2011). Abundant xenoliths from
Upper Sencca, Las Lomas and Capilla (Tables 1 and 2 and Figs 3–5). the batholiths located on the western flank of the Western Cordillera
Additional ignimbrites between c. 30 and 2.7 Ma identified in the reflect incipient erosion into far-flung mountains of Cretaceous–
adjacent Orcopampa region east of the OCM region (Swanson et al. Palaeogene age upstream in the OCM region or farther towards
2004; Fig. 3) support the fact that pyroclastic activity became more WNW (Figs 1, 2 and 4).
sustained while the Cordillera uplift was taking place. No significant We argue that two Nazca ignimbrites exist (Table 2) in the OCM
(>0.6 myr) break occurred after 5.1 Ma; instead, quasi-continuous region, Puquio area, and southernmost Peru. Figures 1, 4 and 6,
volcanism produced three generations of composite volcanoes with section A–A′, show that the Nazca Group encompasses one older,
four intercalated ignimbrite sheets and PDC deposits, and plateau-forming, Nazca 1 ignimbrite sheet near Torata in SW Peru
Pleistocene monogenetic fields. Figure 4 displays the extent of and younger tuff beds of Nazca 2 intercalated in the Moquegua C1
each ignimbrite or PDC unit. Figure 6 shows the position of the conglomerates in the south OCM region. The Nazca ignimbrite at
principal ignimbrites in cross-sections. Table 2 gives the chrono- Pampa Gramadal south of Caravelí (22.61 ± 0.42 Ma), and in the
logical data acquired in this study. Cuno Cuno cliff was emplaced between c. 22 and 23 Ma (Fig. 5a;
Thouret et al. 2007; Schildgen 2009). Imprecise K–Ar age
determinations in the same region push the Nazca ignimbrite ages
The ‘Nazca Group’: two ignimbrites
back to c. 24.5 Ma at Cuno Cuno (Cruzado & Rojas 2005) and
The Nazca ignimbrites in the OCM region (Tables 1 and 2) are 24.5–25.5 Ma at Gramadal (Noble et al. 1985) (Table 2). The
≤10 m thick, non-welded, whitish, ash- and lithic-rich, massive younger Nazca 2 ignimbrite is contemporaneous with K–Ar dated
beds of coarse pumice lapilli tuff intercalated with the Moquegua ignimbrites between 22.1 and 18.7 Ma between Puquio and Nazca
conglomerates (Roperch et al. 2006, 2011) not far above the ‘C1– (Noble et al. 1979); these ages coincide with our 40Ar/39Ar ages
Table 1. Principal characteristics of the twelve ignimbrite sheets and PDC deposits in the OCM region
Ignimbrite sheet and Preserved Average Minimum Aspect Stratigraphic position and Depositional Lithofacies and Welding intensity5 Grain size; sorting and Lithological
PDC deposits; age surface thickness volume ratio3 location with respect to environment; structure4 and weathering texture;6 colour components
range1 in the OCM area (km2) (m) range2 probable source deposit structure (ranked by
region and bedforms abundance)
Capilla unit
0.91 ± 0.05 Ma 250 15–20 2.75–5.0; 4 × 10–3; Base; proximal or inflow In caldera?; confined; Base: vitrophyre; Welded base VI; Glass; fine texture; poor Glass; pumice;
Ignimbrite sheet and Preserved Average Minimum Aspect Stratigraphic position and Depositional Lithofacies and Welding intensity5 Grain size; sorting and Lithological
PDC deposits; age surface thickness volume ratio3 location with respect to environment; structure4 and weathering texture;6 colour components
range1 in the OCM area (km2) (m) range2 probable source deposit structure (ranked by
J.-[Link] et al.
Nazca Group (distal)
Nazca 2: 22.61– 80 (exposed) 6–10 0.50–0.80; 7× Distal layer intercalated Intercalated layer or Massive bed, lithic- Non-welded I; Lapilli; coarse texture, Large, fibrous
22.16 Ma; Nazca 1: SVI 10−4; with Moquegua C27 bed, unconfined, rich lapilli tuff slightly indurated poor sorting; pinkish pumice; crystals;
24.43–23.92 Ma LAR conglomerates non-welded (but fractured) lithic fragments
ignimbrite
Names or parts of names in bold italics are those we have introduced; names in bold regular font are from the national geological maps of Peru although the nomenclature changes from one map to another.
1
Age range includes minimum and maximum age determinations (see Table 2 and Fig 7).
2 2 3
Total area (measured on map Fig. 4 with ArcGIS) is c. 7050 km . Minimum bulk volume (based on observed thickness and mapping) is c. 404–1186 km . Estimated total volume range is that of outflow sheets. Volume of intra-caldera ignimbrites has not been
3 3
accounted for, as calderas boundaries have not been checked through geophysical surveys and depth of inflow ignimbrite is unknown. Ignimbrites are subdivided into large-volume (LVI; >40 km , VEI >7), moderate-volume (MVI; 2.5–40 km , VEI 5 and 6) and
3
small-volume (SVI; <2.5 km , VEI <5) deposits. The poorly known volume of the Alpabamba ignimbrites (intercalated with volcano-sedimentary layers) may be larger than that indicated here on the basis of the >400 m high cliffs in the northern OCM region (Figs
4a, 5c and 6B–B′ and C–C′).
3 −4
Aspect ratio (AR) is the ratio of the average thickness (T ) to horizontal extent (H ), where H is taken as the diameter of a circle whose area is equal to that of the flow. High aspect ratio ignimbrites (HAR I) show AR >1 × 10 .
4
Lithofacies after Branney & Kokelaar (2002, modified). Structures include internal layering, gas-escape structures, imbrication, and features related to temperature and viscosity. Lithofacies can vary locally owing to flow dynamics and irregular palaeotopography of
the basement.
5
Welding intensity based on Quane & Russel (2005) nomenclature (I (non-welded) to VI (densely welded)) and physical properties.
6
Texture includes grain-size distribution, sorting, componentry (proportions of pumice, lithic fragments and crystals), vertical size grading, segregation of crystals and lithic fragments, welding, compaction, devitrification, recrystallization and weathering. Texture
can vary locally owing to flow dynamics and palaeotopography.
7
Roperch et al. (2006, 2011); Decou et al. (2011, 2013); Sempere et al. (2014).
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Table 2. Summary of 42 40Ar/39Ar laser fusion analyses and 10 U/Pb age determinations for 12 Neogene–Quaternary ignimbrites and 22 40Ar/39Ar analyses of
lava flows in the OCM region, upper Río Colca valley and SW Peru
Deposit and map Sample Material Method Age (Ma) x coordinate y coordinate Elevation Rock type Description
number* number (m)
Pyroclastic deposits
Sara Sara PDC
40
1 OCO-06-09 Feldspar Ar/39Ar 0.62 ± 0.08 672 685 8 292 840 3750 Trachydacite Base PDC unit
Capilla ignimbrite
40
2 PIG-08-40 Feldspar Ar/39Ar 0.91 ± 0.05 767 199 8 361 944 4850 Rhyolite Base: vitrophyre
Lomas PDCs
40
3 PIG-08-34 Glass Ar/39Ar 1.26 ± 0.01 731 263 8 329 485 4460 Rhyolite Top unit
40
4 PIG-03-123 Feldspar Ar/39Ar 1.36 ± 0.27 705 449 8 277 851 2300 Trachydacite Lower unit
40
5 PIG-10-48 Feldspar Ar/39Ar 1.46 ± 0.01 705 141 8 278 110 2350 Trachydacite Middle unit
6 PIG-13-12 Zircon U–Pb 1.54 ± 0.04 731 050 8 279 644 3970 Rhyolite Valley fill top
flow unit
40
7 PIG-02-84 Feldspar Ar/39Ar 1.56 ± 0.32 728 085 8 313 279 4020 Trachydacite Top unit below
lava flow
Upper Sencca
ignimbrite
40
8 PIG-10-32B Feldspar Ar/39Ar 1.80 ± 0.04 717 200 8 274 150 3450 Rhyolite Second upper
unit
40
9 PIG-00-06 Feldspar Ar/39Ar 1.81 ± 0.11 662 050 8 271 200 2550 Rhyolite Top unit
10 PIG-11-43 Zircon U–Pb 1.90 ± 0.11 731 052 8 279 779 4490 Rhyolite Pumice in sole
layer
40
11 PIG-00-25 Feldspar Ar/39Ar 1.93 ± 0.04 749 792 8 253 371 3600 Rhyolite Sole layer
40
12 OCO-05-06 Feldspar Ar/39Ar 1.95 ± 0.17 719 137 8 274 358 3550 Rhyolite Upper unit
40
Biotite Ar/39Ar 1.76 ± 0.17
40
13 PIG-10-37B Feldspar Ar/39Ar 1.95 ± 0.02 708 953 8 275 893 2300 Rhyolite Base of upper
unit
40
14 OCO-04-08 Feldspar Ar/39Ar 1.96 ± 0.06 698 800 8 243 130 1675 Rhyolite Base of unit
40
15 PIG-10-20A Feldspar Ar/39Ar 1.97 ± 0.01 720 737 8 256 663 2420 Rhyolite Vitrophyre
40
16 OCO-06-02 Feldspar Ar/39Ar 2.00 ± 0.09 704 612 8 269 233 1795 Rhyolite Residual valley-
fill unit
40
17 PIG-04-02 Feldspar Ar/39Ar 2.04 ± 0.14 725 754 8 312 325 3570 Rhyolite Middle unit
40
18 OCO-04-05 Feldspar Ar/39Ar 2.05 ± 0.29 704 400 8 237 700 750 Rhyolite Base of unit
40
19 PIG-07-11 Feldspar Ar/39Ar 2.06 ± 0.06 718 631 8 332 981 4480 Rhyolite Base of unit
20 PIG-10-51 Zircon U–Pb 2.10 ± 0.12 731 052 8 279 779 4490 Rhyolite Pumice-fall
below A2
40
21 OCO-05-04 Feldspar Ar/39Ar 2.09 ± 0.06 716 172 8 251 325 1840 Rhyolite Upper unit
40
Biotite Ar/39Ar 2.02 ± 0.04
40
22 PAT-04-01 Feldspar Ar/39Ar 2.20 ± 0.15 8 210 150 8 260 826 4675 Rhyolite Base of lower
unit
Lower Sencca
ignimbrite
23 PIG-13-27 Zircon U–Pb 2.82 ± 0.04 709 048 8 343 593 4900 Rhyolite Inflow top unit,
caldera
40
24 PIG-00-24 Feldspar Ar/39Ar 3.16 ± 0.04 752 633 8 262 442 3890 Rhyolite Upper unit
40
25 PIG-00-28 Feldspar Ar/39Ar 3.76 ± 0.14 727 371 8 316 576 2800 Rhyolite Base: vitrophyre
26 PIG-13-24 Zircon U–Pb 3.84 ± 0.06 735 336 8 357 634 4850 Rhyolite High plateau top
unit
27 PIG-13-25 Zircon U–Pb 3.84 ± 0.06 735 337 8 357 633 4850 Rhyolite High plateau
upper unit
28 PIG-13-29 Zircon U–Pb 4.32 ± 0.06 768 139 8 361 213 4750 Rhyolite Inflow lower
unit, caldera
40
29 OCO-06-05 Feldspar Ar/39Ar 4.36 ± 0.16 706 099 8 268 291 875 Rhyolite Single reworked
unit
40
30 PIG-07-10 Biotite Ar/39Ar 4.65 ± 0.11 723 406 8 322 861 4350 Rhyolite Top unit
40
31 OCO-05-12 Feldspar Ar/39Ar 4.84 ± 0.07 704 831 8 276 933 1665 Rhyolite Sole layer base
unit
40
32 PIG-08-28B Feldspar Ar/39Ar 5.09 ± 0.03 749 550 8 351 979 4300 Rhyolite Base: vitrophyre
40
33 PIG-08-33 Glass Ar/39Ar 5.13 ± 0.01 748 418 8 347 547 4750 Rhyolite Lower unit
Arma PDCs
40
34 PIG-08-21 Feldspar Ar/39Ar 7.97 ± 0.10 738 644 8 289 629 4020 Rhyolite Upper unit
40
35 PIG-10-56 Feldspar Ar/39Ar 7.99 ± 0.04 731 105 8 280 158 4390 Rhyolite Middle unit
Caraveli ignimbrite
40
36 PIG-00-07 Feldspar Ar/39Ar 8.97 ± 0.06 677 489 8 312 058 2420 Rhyolite Upper unit
40
37 CARA-05-07 Biotite Ar/39Ar 8.98 ± 0.15 684 535 8 234 701 1750 Rhyolite Upper unit
40
Feldspar Ar/39Ar 9.40 ± 0.83
(continued)
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Table 2. (Continued)
Deposit and map Sample Material Method Age (Ma) x coordinate y coordinate Elevation Rock type Description
number* number (m)
40
38 PIG-00-04 Feldspar Ar/39Ar 9.02 ± 0.11 669 064 8 257 871 2550 Rhyolite Uppermost unit
40
39 PIG-10-02B Sanidine Ar/39Ar 9.10 ± 0.02 670 993 8 260 424 2190 Rhyolite Base: vitrophyre
40
40 PIG-00-03 Feldspar Ar/39Ar 9.15 ± 0.31 670 283 8 257 708 2350 Rhyolite Uppermost unit
40
41 PIG-07-05 Feldspar Ar/39Ar 9.35 ± 0.06 691 184 8 275 407 3900 Rhyolite Upper unit to the
NE
Huarcaya ignimbrite
42 PIG-13-22 Zircon U–Pb 10.78 ± 0.13 745 981 8 338 349 4040 Rhyolite Valley fill in
Alpabamba
Chuquibamba
ignimbrite
40
43 PIG-03-126 Biotite Ar/39Ar 13.19 ± 0.07 704 629 8 229 204 2275 Rhyolite Distal upper unit
40
Feldspar Ar/39Ar 13.21 ± 0.53
44 PIG-10-16 Zircon U–Pb 13.60 ± 0.14 704 635 8 229 205 2260 Rhyolite Top unit on
plateau
40
45 PIG-00-31 Feldspar Ar/39Ar 14.23 ± 0.07 730 375 8 309 352 4300 Rhyolite Medial upper
unit
40
46 PIG-00-33 Feldspar Ar/39Ar 14.25 ± 0.08 807 050 8 190 500 1000 Rhyodacite Distal unit in
Moq C2
Majes ignimbrite
40
47 PIG-00-32 Feldspar Ar/39Ar 16.25 ± 0.10 773 850 8 194 450 850 Rhyolite Bed at Moq C1–
C2
Alpabamba ignimbrite
40
48 PIG-00-11 Feldspar Ar/39Ar 18.23 ± 0.17 552 154 8 374 953 3200 Rhyolite Nazca Group
upper unit
40
49 PIG-00-41 Feldspar Ar/39Ar 18.90 ± 0.50 8 301 650 8 120 150 2900 Trachydacite Lower unit
Moquegua
50 PIG-13-21 Zircon U–Pb 20.13 ± 0.17 746 091 8 332 512 3590 Dacite Base of
ignimbrite
Nazca ignimbrite
40
51 PIG-00-10 Feldspar Ar/39Ar 22.16 ± 0.34 550 933 8 374 911 3500 Rhyolite Nazca Group
lower unit
40
52 PIG-10-11 Plagioclase Ar/39Ar 22.61 ± 0.42 681 947 8 230 672 1765 Rhyolite Unit 2 in Moq C1
OCM lava flows
Pleistocene composite
cone
40
53 SAR-07-03 Feldspar Ar/39Ar 0.02 ± 0.01 681 430 8 309 048 2540 Trachyandesite Most recent lava
flow
Nevado Sara Sara
40
54 SAR-07-05 Feldspar Ar/39Ar 0.05 ± 0.01 678 090 8 322 409 2635 Trachyandesite Upper lava flow
40
55 SAR-07-06 Groundmass Ar/39Ar 0.14 ± 0.02 678 595 8 320 300 2440 Trachyandesite Base of lava
flows
40
56 COTA-05-10 Feldspar Ar/39Ar 0.34 ± 0.06 706 959 8 299 105 2235 Andesite Upper lava flow
Monogenetic cones
40
57 COTA-05-11 Feldspar Ar/39Ar 0.45 ± 0.01 708 072 8 301 599 2345 Trachyandesite Lava flow from
cone
40
58 COTA-05-15 Feldspar Ar/39Ar 0.66 ± 0.01 715 085 8 302 926 4180 Trachyandesite Lava flow from
cone
40
59 COTA-05-06 Feldspar Ar/39Ar 0.68 ± 0.03 712 734 8 300 308 3940 Trachyandesite Lava flow from
cone
Early Quaternary lava
field
40
60 PIG-10-42 Feldspar Ar/39Ar 2.04 ± 0.06 705 209 8 277 540 1960 Andesite Base of lava
flow
40
61 OCO-06-06 Feldspar Ar/39Ar 2.24 ± 0.45 706 157 8 264 226 950 Andesite Base of lava
flow
40
62 OCO-05-11 Feldspar Ar/39Ar 2.27 ± 0.05 706 950 8 253 980 830 Andesite Base of lava
flow
Miocene composite
cones
40
63 COTA-06-05 Feldspar Ar/39Ar 5.36 ± 0.12 729 400 8 321 250 3010 Trachyandesite Lava flow in
cliff
40
64 BAR-03-01 Feldspar Ar/39Ar 5.80 ± 0.10 745 259 8 331 693 3510 Trachydacite Lava flow on
canyon wall
40
65 BAR-07-04 Groundmass Ar/39Ar 7.32 ± 0.05 690 794 8 273 299 3950 Andesite Base of lava
flows
(continued)
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Table 2. (Continued)
Deposit and map Sample Material Method Age (Ma) x coordinate y coordinate Elevation Rock type Description
number* number (m)
Rio Colca lava flows
40
66 COL-04-17 Groundmass Ar/39Ar 0.13 ± 0.02 794 950 8 261 980 1885 Trachyandesite Lower lava flow
40
67 BAR-01-62 Groundmass Ar/39Ar 0.15 ± 0.01 8 221 050 8 270 350 3445 Andesite Upper lava flow
40
68 COL-04-16 Groundmass Ar/39Ar 0.54 ± 0.06 817 939 8 273 569 2240 Andesite Base of valley
fill
40
69 COL-04-01 Groundmass Ar/39Ar 0.61 ± 0.01 809 790 8 267 148 3460 Andesite Lava flow on
lake deposits
Pleistocene composite
cones
40
70 BAR-01-61 Groundmass Ar/39Ar 0.65 ± 0.01 8 219 670 8 270 980 3445 Andesite Lower lava flow
40
71 COL-04-14 Groundmass Ar/39Ar 0.65 ± 0.02 815 841 8 267 571 3600 Basaltic Intermediate
trachyandesite lava flow
40
72 COL-04-03 Groundmass Ar/39Ar 0.67 ± 0.01 819 425 8 274 042 3770 Trachyandesite Middle lava flow
valley fill
40
73 COL-04-07 Groundmass Ar/39Ar 1.06 ± 0.04 803 300 8 264 075 2980 Trachyandesite Upper lava flow
valley fill
40
74 COL-04-10 Groundmass Ar/39Ar 1.07 ± 0.02 815 528 8 270 769 3350 Trachyandesite Upper lava flow
40 39
The Ar/ Ar and U/Pb age ranges take 2σ analytical errors into account. Map numbers 1–74 are indicated in Figures 1, 2, 4 and 6G–G’. Italicized text indicates Ar/Ar ages acquired
after Thouret et al. (2007).
between c. 22.16 and c. 18.23 Ma in the same area (Table 2; Thouret succession yielded two Ar/Ar ages of c. 18.23 and 18.9 Ma in two
et al. 2007). The Nazca 2 ignimbrite may be contemporaneous with localities outside the OCM region (Figs 1 and 4 and Table 2).
the c. 19 Ma Manto Tuff in the Orcopampa area (Swanson et al. Comparable Ar/Ar ages c. 19.47 and 19.7 Ma have been published
2004; Fig. 3). by Swanson et al. (2004) for two tuffs in the Orcopampa region east
Nazca includes the oldest ignimbrite sheets in the OCM region of the OCM area (Fig. 3). In geological maps for the OCM region,
(Fig. 3) based on chronostratigraphic links with ignimbrites of the the ‘Volcanico Alpabamba’ was attributed to the middle and upper
‘Nazca Group’ c. 40 km west of Puquio (40Ar/39Ar age 22.16 Ma; Miocene as it unconformably overlies the Tacaza Group (19–
Thouret et al. 2007) and the lower ignimbrite unit (K–Ar ages 22.1 18 Ma: Pecho Gutierrez 1983a), but rocks from the Oligocene
and 20.5 ± 0.6 Ma) described by Noble et al. (1979). Late Tacaza arc, bracketed between 30 and 24 Ma by Mamani et al.
Oligocene–Early Miocene ignimbrites yielded a wide range (2010), underlie the ‘Huaylillas Formation’. The Alpabamba unit
of ages between c. 22.7 and 25.5 Ma in southernmost Peru was assigned to the Pliocene in the Chuquibamba and Cotahuasi
(Tosdal et al. 1981; Swanson et al. 2004; Quang et al. 2005; maps (Olchauski & Dávila 1994), although it underlies c. 7.3–
Thouret et al. 2007) but Tosdal et al. (1981) termed these 5.3 Ma lava flows (Fig. 3 and Table 2) and the c. 14.2–13.2 Ma
ignimbrites ‘Huaylillas’. ‘Huaylillas’ is equivalent to the Chuquibamba ignimbrite in both
areas (Thouret et al. 2007).
The Alpabamba ignimbrite
The Majes ignimbrite
The Alpabamba sheet is the most widespread and voluminous,
composite ignimbrite in the OCM region. It consists of cooling units Massive and indurated, pumice and ash-rich, pinkish beds only
and multiple flow units of whitish to yellowish grey, indurated to 60 km2 in area and 10–15 m thick were observed in the Río
slightly welded, massive and bedded ash-and-pumice lapilli tuffs. Majes canyon (this is the name of the Río Colca down-valley of
Preserved from erosion by subsequent sheets, the Alpabamba Aplao in the SE corner of the OCM region, Fig. 1) and in cliffs
ignimbrite forms broad and high plateaux up to 0.5 km thick NNE bordering the depression of Caravelí (Figs 4, 5d and 6, section
of Cotahuasi towards Río Huarcaya (Figs 4 and 6, sections B–B′ to E–E′). The Majes lithofacies and deposition environment are
D–D′), which thin out towards the east and south as the result of similar to those of the Nazca ignimbrite but the Majes unit is
inherited depositional thickness. Towards the WNW (Pacapausa intercalated with the Moquegua C2 conglomerate. The thickness
map area) a ‘c. 1000 m-thick sequence’ has been described by of the distal tuff bed in the west bank of Río Majes increases
Guevara & Dávila (1983) at the type locality, at San Javier de towards the WNW (near Caravelí), but does not change
Alpabamba across the canyon of Río Huanca Huanca upstream of substantially towards the ENE, including the Río Sihuas valley
Río Marán. The thickness of the Alpabamba unit near this locality 30 km east of Río Majes. This suggests a source on high
has been overestimated as the top whitish tuffs are Lower Sencca plateaux to the NW and WNW beyond the OCM region (Figs 1,
deposits (see below) and intercalated volcano-sedimentary beds 2 and 6, section E–E′).
with tephra and breccia belong to the Aniso Formation (Fig. 4a and We Ar/Ar dated the Majes ignimbrite at 16.35 ± 0.10 and 16.25 ±
legend in b). Vertical cliffs display three types of deposits totalling 0.10 Ma on feldspars from the layer intercalated in the upper third
400–500 m in thickness in the Río Huarcaya, Pampamarca, Oyolo of Moquegua conglomerates on the east bank of the Río Majes
and Marán canyons (Figs 5b and 6, sections B–B′ and C–C′). valley (Fig. 1). Schildgen (2009) reported a similar age (16.26 ±
Exposure thickness decreases southward to 100–200 m on the 0.08 Ma) but also two ages of 14.11 Ma on feldspar and 14.35 Ma
Solimana east flank (Río Arma valley). A minimum surface area of on biotite, suggesting the presence of Chuquibamba distal beds in
c. 2200 km2 and a bulk volume range of 220–880 km3 were the Río Majes canyon. Another group of ages of 16.8 Ma reported
estimated (Table 1). for the ‘Huaylillas Formation’ by Pecho Gutierrez (1983b) and ages
The base of the massive, thick, composite Alpabamba succession c. 16.26 and 16.12 ± 0.04 Ma measured by Schildgen (2009) on
near the village of Puica has a lower intercept U/Pb age of 20.13 ± distal ignimbrites in the Río Sihuas Valley also probably represent
0.17 Ma based on 33 analysed zircon crystals. The top of the Majes deposits (Fig. 1).
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Fig. 5. Selected stratigraphical sections of the principal ignimbrites showing sub-units, lithofacies, welding degree and lithological components. Sampling
sites and dated samples are indicated. Locations of sections are indicated with stars in Figure 4a and b. (a) Cuno Cuno cliff cut in Moquegua Formation
with intercalated Nazca 1 and 2 distal tuff beds. (b) Pampa Gramadal south of Caravelí and Quebrada Seca showing Nazca 2 and Caravelí ignimbrites.
(c) Alpabamba multi-layered ignimbrite sheet and Sencca ignimbrites near Pampamarca (Río Huanca Huanca) and Oyolo (Río Marán). (d) Two Caravelí
cooling units at Quebrada Chuñuño D1 and Pampa Bonbón D2 north and SSE of the town of Caravelí. (e) Complex stratigraphy at Salamanca cliff (east Río
Arma) including Alpabamba, Chuquibamba, Arma PDC deposits and Upper Sencca. (f ) Lower Sencca and Upper Sencca ignimbrites and Quaternary lava
flows at Chaucalla Lomas ridge (confluence of Río Ocoña and Chichas) covered by Lomas PDC deposits. (g) Chuquibamba ignimbrite upstream of the
town of Chuquibamaba. (h) Alpabamba and Lower Sencca ignimbrites at Chulca in the upper Río Huarcaya valley. (i) Lower(?) and Upper Sencca
ignimbrites at Antonieta (Río Chichas valley) near the mine of Arirahua. ( j) Upper Sencca ignimbrites near Caravelí ( j1), between Atico and Caravelí ( j2),
and near Yanaquihua and Ispacas (in east tributaries of the Río Ocoña valley). (k) Early Pleistocene lava flow and Las Lomas PDC deposits overlying
Alpabamba ignimbrite upstream of Cotahuasi and Huaynacotas (Río Cotahuasi valley).
The Chuquibamba ignimbrite Cuno) to the south at c. 13.19 – 13.59 Ma (Table 2, Figs 4 and 5a,
g). Thirty analysed zircon grains yielded a lower intercept U/Pb age
The well-preserved, compound Chuquibamba ignimbrite is the of 13.60 ± 0.14 Ma, which is very close to the Ar/Ar age for the
second most extensive (≥1850 km2) ignimbrite sheet in the OCM similar ignimbrite at Cuno Cuno. The Chuquibamba composite
region albeit less thick than the Alpabamba sheet (Figs 4, 5g and sheet consists of two cooling units and up to four flow units, based
6, section D–D’). The strongly welded Chuquibamba sheet on lithofacies and two age clusters 1.02 myr apart. Swanson et al.
contributed to high-relief, gently sloping but broadly bent (2004) reported a similar Ar/Ar age of 14.16 ± 0.025 Ma for the
plateaux over 80 – 90 km distance between the frontal volcano Chipmo Tuff near Orcopampa east of the OCM region. We discard
range and the Pacific coast across the clastic continental wedge the name ‘Huaylillas’ given by Ingemmet to that ignimbrite dated at
(Fig. 4; Gunnell et al. 2010, fig. 4). Chuquibamba is overlain by 13.8 ± 0.3 Ma near Chuquibamba (Pecho Gutierrez 1983a). In fact,
Late Miocene lava flows and Sencca ignimbrites forming radial the ‘Huaylillas’ Formation, a term coined by Wilson & Garcia
aprons around Nevado Coropuna. To the north of the OCM (1962) in southernmost Peru, comprises Early Miocene ignimbrites
region, three wine red-coloured units exhibit a strongly welded, in the Moquegua–Tacna region (Tosdal et al. 1981, 1984), where
crystal-rich and lithic-poor, coarse lithofacies in 80 m cliffs outcrops coincide with the 25.6 ± 0.9 – 19.0 ± 0.6 Ma Oxaya
crowning the Pampa Puca Puca high plateau 6 km NW of Oyolo, ignimbrites in adjacent Chile (Wörner et al. 2000). Other
in the 50 – 80 m landslide scar upstream of Chuquibamba and in ‘Huaylillas’ outcrops north and east of the OCM region dated at
the east side of the upper Río Arma (Figs 4 and 5e). At 60 km to 18.35 – 18.40 ± 0.5 Ma by Lefèvre (1979) precede the
the SSW, the ignimbrite thins out to a pinkish grey, massive, Chuquibamba eruptions.
incipiently welded, ash-rich lapilli tuff 30 m thick at the top of
the Cuno Cuno cliff (Fig. 5a).
The Huarcaya ignimbrite
We have dated the top of proximal outcrops to the north
(Cotahuasi) at c. 14.21 – 14.25 Ma, the base and top of medial, thick We coined the term Huarcaya for an ignimbrite found in the canyon of
exposures (Chuquibamba) and the top of thin distal outcrops (Cuno Rio Huarcaya between the villages of Puica and Chulca (Fig. 4). This
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 5. Continued
greyish, welded, massive, ash-rich lapilli tuff in the north cliff of the outside the Río Huarcaya canyon only in the lowermost unit in the
canyon >300 m above the valley bottom is inset at the base of a 400 m Huancamarca depression (Figs 4 and 6, section B–B′).
thick pile of ignimbrites attributed to the Alpabamba and Lower U/Pb analyses of 27 zircons yielded a lower intercept age of 10.78 ±
Sencca. The poorly defined Huarcaya ignimbrite was identified 0.13 Ma (Table 2), which is interpreted as the time of crystallization of
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Fig. 5. Continued
the zircons in the magma reservoir. The Huarcaya age of c. 10.78 Ma We gave the name Caraveli name (Thouret et al. 2007) to an
can be correlated with the 10–11 Ma Cerro Hospicio tuff in the ignimbrite previously included by Tosdal et al. (1984) and Sébrier
Orcopampa region (Swanson et al. 2004; Fig. 3). The position of this et al. (1988) in the Chuntacala Formation (Fig. 3). This ignimbrite
ignimbrite indicates that as early as the Late Miocene, the north OCM extensively covers high plateaux with a 4% slope from the edge of
region was cut by 200–400 m deep valleys, and that Huarcaya deposits Lake Parinacocha to the NNW to as far as 80 km SSE at Pampa
filled a valley cut in the Alpabamba ignimbrites. Gramadal. It now forms a NNW–SSE cradle-like topography across
broad (3–4 km) and shallow (200–400 m) valleys incised in the pre-
Chuquibamba S1 surface. The high plateau around the Huancarama
The Caravelí ignimbrite
depression is mantled by Caraveli and Lower Sencca ignimbrites at
The widespread Caravelí compound ignimbrite sheet covers at least least 50 km northward (Fig. 4a). The Caravelí sheet thins out south
810 km2 with a thickness of 50 m to a maximum of 120 m in the and west of the Caraveli depression but is absent on the east side of
SW and west OCM region (Figs 4 and 6 , section E–E′, and Table 1). the Río Ocoña canyon.
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 5. Continued
The compound, strongly welded ignimbrite sheet near Caravelí by reddish scoria-rich flow deposits or thin lava flows. Based on its
comprises at least two cooling units both underlain by vitrophyre, location, thickness and grain size we relate the Arma PDC deposits
and overlies Moquegua C2 conglomerates (Fig. 5d). The to neighbouring, deeply eroded Late Miocene composite cones or
lower cooling unit is represented by columnar flow units separated small calderas east of Nevado Solimana and WSW of Coropuna.
by planar discontinuities. In the 120 m high cliff exposures, a The Ar/Ar ages of the Arma PDC valley fill is c. 7.97–7.99 Ma. A
second c. 1 m thick vitrophyre is overlain by the 40–50 m thick preserved volume of 1 km3 is a minimum value for the Río Arma
upper cooling unit with a brownish, columnar, welded crystal-rich and Oyolo valleys (Fig. 5c and e). Other Arma-like, 100 m thick
lapilli tuff. South of the Caravelí depression, the Pampa PDC deposits crop out in a similar headvalley position east of
Gramadal plateau shows a grey, vacuolar, ash-rich unit covered Oyolo (Río Uchubamba) and 400 m below the Cerro Huanipaco
by scattered pumices, which yielded an age of c. 9 Ma (Table 2 and high plateau (Fig. 4). Arma PDCs can be correlated with 7–8 Ma
Fig. 5b). tuffs reported in the Orcopampa region (Swanson et al. 2004;
The chronology of the Caravelí ignimbrite sheet is one of the Fig. 3).
best constrained among the OCM ignimbrites. Seven 40Ar/39Ar The Lower Sencca Group is a composite ignimbrite sheet, which
age determinations range between c. 9.40 and 8.97 Ma (Table 2) encompasses several whitish, nonwelded flow units 50–150 m thick
on samples collected from the base to the top and 78 km apart capping the north OCM high plateaux (Fig. 4). Outcrops thin out
between Pampa Gramadal south of Caravelí and Lampa near Pausa towards the south and SW but form 100 m thick terraces hanging
to the north (Fig. 4). The narrow age cluster (Fig. 7) at c. 9.05 Ma high above the canyon bottoms (Fig. 5f) or in tributary headvalleys
suggests a relatively brief period of voluminous eruptions. Pecho (Fig. 6, section C–C′). Thick inflow deposits are observed in
Gutierrez (1983b) assigned this ignimbrite to the Huaylillas depressions such as Chulca and Huarcaya (Fig. 5f ) in the upper
Formation on the geological map of Caravelí, and Quispesivana reach of the Río Huarcaya canyon, and the outflow Lower Sencca
& Navarro (2001) to the Alpabamba Formation on the revised 2001 pumice-rich units are observed 32 km SW on the cliff looming over
map. Both attributions are at odds with field observations, Pampamarca (Figs 4 and 5c). Because the Lower Sencca units crop
lithofacies and two K–Ar ages of 10.8 ± 0.24 and 8.72 ± 0.19 Ma out in high cliffs, lithological correlation is difficult. We distinguish
indicated on the 2001 map 10 km SE of Caraveli and 15.5 km four cooling and/or flow units A–D on the basis of lithofacies
farther south. characteristics, age distribution and variation of mineral assem-
blages (Table 2, and Figs 3 and 6, sections B–B′, D–D′ and F–F′).
The Ar/Ar age distribution from 5.1 to 2.83 Ma suggests that five
The Arma PDC deposits
distinct pulses over the Pliocene formed the multi-layered Lower
The small-volume Arma PDC deposits have been observed along Sencca sheet (Fig. 3 and Table 2). The U/Pb ages on zircons fall in
the upper Río Arma valley between Nevado Solimana and the range of c. 5 to 2.8 Ma. Zircon crystals from sample PIG-13-29
Coropuna with an area c. 35 km2 (Fig. 4). A succession of 15– yield a lower intercept age of 4.32 ± 0.06 Ma, and grains from both
30 m thick, whitish PDC deposits on top of the Alpabamba PIG-13-25 and PIG-13-24 define the same lower intercept age at
ignimbrite has filled the high Río Arma valley that cut into the 3.84 ± 0.06 Ma. Zircon crystals of PIG-13-27 yield a lower intercept
Alpabamba, Chuquibamba and Sencca cliffs (e.g. Fig. 5e). These age of 2.82 ± 0.04 Ma.
loose, poorly to well-sorted deposits encompass bedded, stratified The Upper Sencca compound ignimbrite includes more scattered
and thin tephra-fall deposits, cross-stratified surge deposits, and but valley-confined outcrops, and hence is more accessible. Upper
stratified and thin-bedded flow deposits. This succession is overlain Sencca deposits have a pinkish or yellowish, recrystallized facies
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Fig. 6. Geological cross-sections located on a NASA SRTM DEM. Section A–A′, from Pampa Gramadal, Caravelí to Cuno Cuno across the Río Ocoña canyon; section B–B′, from the high plateau west of Río Pampamarca to
the east edge of the upper Río Cotahuasi valley and Nevado Firura; section C–C′, from Nevado Sara Sara to the Huancarama high plateau across Río Huanca Huanca valley; section D–D′, the Caravelí depression and adjacent
Pampa plateaux; section E–E′, the Río Cotahuasi canyon across the town of Cotahuasi; section F–F′, from Caravelí to Salamanca cliff across the confluence of the three OCM canyons; section G–G′, composite cross-sections
across the upper Río Colca reaches near Chivay–Achoma and Huambo–Condorsaya–Ayo. Numbers in section G–G′ refer to dated samples listed in Table 2.
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 7. Temporal distribution of 40Ar/39Ar age determinations for the OCM ignimbrites compared with ignimbrites in SW Peru. Histograms show the
minimum (continuous lines) and maximum (dashed lines) volume estimates for each ignimbrite sheet (see Table 1). Width of bars expresses the age
determination with 2σ error. Three generations of lava flows dated in the OCM region are indicated by brown boxes. The duration of eruptive lulls between
the ignimbrite sheets is indicated below the time axis. Distinct eruptive pulses are suggested for the Lower Sencca ignimbrite. *Peruvian nomenclature
(INGEMMET & Mamani et al. 2010). **Proposed chronostratigraphic nomenclature. GR, Gramadal (this study); T, Tosdal et al. 1981; NG, Nazca Group,
Puquio (Noble et al. 1979); S, Schildgen et al. 2009; M, Majes (this study and Schildgen et al. 2009); CH, Chuquibamba; RCI, Rio Chili (Paquereau-Lebti
et al. 2006, 2008); CC, Cuno Cuno (Fornari, unpublished data); La Joya ignimbrite (Paquereau-Lebti et al. 2006, 2008); LP, Lauca–Perez ignimbrite
(Wörner et al. 2000); PP, Patapampa ignimbrite (R. Colca, Thouret et al. 2007); A, Arequipa; AAI, Arequipa Airport ignimbrite; YT, Yura Tuff
(Paquereau-Lebti et al. 2006, 2008).
containing large pumice fragments form eroded terraces between are indistinguishable from the Ar/Ar ages of similar Upper Sencca
200 and 800 m in elevation above the bottom of deep canyons (Río units.
Ocoña) and 400–600 m thick ridges (Chaucalla) and hanging The ‘Volcánico Sencca’ in Peruvian maps encompasses
terraces that encroach on headvalleys (Río Arma, Figs 4 and 6, ignimbrites dated between 3 and 2 Ma such as the eponymous
section F–F′). The Upper Sencca ignimbrite is often found as Sencca near the Chilean border, the ‘white sillar’ in the basin of
narrow but thick, dark grey or pinkish deposits (>100 m) in the Arequipa, and Ayacucho (Jenks & Goldich 1956; Guevara 1969;
upper reaches of deep gorges carved in granodiorite batholiths Vargas 1979; Kaneoka & Guevara 1984; Barker 1996). The Sencca
(Quebrada Tranca near Yanaquihua), and near the bottom of ignimbrite has been divided into two composite ignimbrite sheets,
shallow valleys or depressions (Caravelí, Fig. 6, section D–D′). the Lower Sencca group and Upper Sencca unit (Thouret et al.
Upper Sencca ignimbrites filled shallow valleys incised in the 2007), in contrast to the stratigraphic nomenclature after Mendivíl
Late Miocene–Early Pliocene high plateau lavas and valleys cut in (1965) for three reasons: (1) the Upper Sencca ignimbrite
the Chuquibamba ignimbrite. It also forms 10–30 m thick aprons systematically yields younger ages between c. 1.76 and 2.09 Ma
around recent composite volcanoes (e.g. Nevado Coropuna). whereas the collectively termed Lower Sencca group yield older
The Upper Sencca deposits cover c. 2.24–2.27 Ma lava flows ages scattered between c. 5.13 and 2.83 Ma (Fig. 3 and Table 2); (2)
on both sides of the Ocoña valley and on the cliff of Chaucalla the Lower Sencca composite group is more widespread (c.
ridge 3 km south of the confluence of three canyons. This 950 km2; c. 47.5–76 km3), less confined and contains more units
demonstrates that the Pliocene Lower Sencca ignimbrites and than the smaller Upper Sencca unit (c. 640 km2; 13 – 32 km3;
Quaternary Upper Sencca ignimbrites must be distinguished. The which usually consists of a single cooling unit), which thicken in
0.29 Ma age range in the Upper Sencca suggests distinct pulses as gorges (east slope of Ocoña canyon, Fig. 4b) that have channelled
reflected by two or three flow units with distinct massive, welded thicker cooling units; (3) Late Miocene ‘Lower Barroso’ lava flows
and eutaxitic lapilli tuff lithofacies in the 100–150 m thick valley- between c. 7.32 and 5.36 Ma in the OCM region underlie the Lower
fill. The lower intercept ages of 2.10 ± 0.12 Ma and 1.90 ± 0.11 Ma Sencca ignimbrite (e.g. east cliff of the Chulca depression, Fig. 5f ),
from zircons of samples PIG-10-51 and PIG-11-43, respectively, whereas part of the early Quaternary (<2 Ma) and all Pleistocene
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
lava flows ≤1.3 Ma overlie the Upper Sencca ignimbrite (e.g. Ar/Ar ages obtained from this group. The Chuquibamba age of
Chuquibamba plateau, Thouret et al. 2007). 13.59 Ma lies within the range of five Ar/Ar Chuquibamba dates
from this unit. The Alpabamba base at 20.10 Ma is coherent with the
stratigraphy and overlying, younger units of the composite ignimbrite
The Lomas PDC deposits
sheet despite considerable distances between outcrops. A U/Pb date
The Las Lomas valley fill consists of a succession of non- of 10.78 Ma for the Huarcaya ignimbrite is the only age
welded, crystal-poor, pumice-rich PDC deposits that can be determination from this unit, but its age is in agreement with
traced along the edge of the Cotahuasi canyon up to 10 km north similar ignimbrites dated c. 10.80 Ma in the Orcopampa region
of the Huaynacotas presumed caldera (Figs 4 and 5a). We named further east. A U/Pb age of 2.83 ± 0.03 Ma for the Lower Sencca
this ignimbrite after the Las Lomas–Chaucalla ridge near the Group suggests that ignimbrite production ceased ≥0.7 myr before
confluence of Río Ocoña and Chichas (Fig. 6, section F–F′). the older unit (2.09 ± 0.06 Ma) of the upper Sencca ignimbrite. This
Position, lithofacies and source help to distinguish the Lomas exceeds the 0.6 myr interval defined as a quiescent period based on
ignimbrite from the Upper Sencca. Preserved PDC deposits crop the entire dataset (Fig. 7). Together with stratigraphy and depositional
out on top of high interfluves between the canyons of Río characteristics, this time interval has allowed us to separate
Cotahuasi–Marán and Río Cotahuasi–Arma or in broad and ignimbrites from the Lower Sencca group and the Upper Sencca unit.
shallow valleys cut in Quaternary edifices. (2) 40Ar/39Ar dates on biotite and feldspar from the same
This small-volume ignimbrite extends over c. 185 km2 and ignimbrite unit are indistinguishable from one another when the
comprises multiple pumice-flow deposits Ar/Ar dated between c. analytical uncertainties are considered. For example, feldspars from
1.56 and 1.26 Ma. The U/Pb age obtained from PIG-13-12 is very one of the Upper Sencca ignimbrite units yielded an age of 1.95 ±
close to the base of the Lomas succession with a lower intercept age of 0.17 Ma whereas biotites yielded 1.76 ± 0.17 Ma (Thouret et al.
1.54 ± 0.04 Ma. The highest outcrops (e.g. east of Cotahuasi on the 2007; Table 2). This youngest age is similar to the minimum age,
lower plateau edge) were protected by Pleistocene lava flows of 1.80 ± 0.04 Ma, of 14 Upper Sencca dated units.
<1.3 Ma, but they are now cut by landslides on the steep canyon edges (3) The narrow 40Ar/39Ar and U/Pb age range of ignimbrites
(Fig. 6, section D–D′). The Lomas deposits filled the valleys and indicates that at least four sheets were erupted during one time interval
preceded the last re-incision of the canyons during the Pleistocene. shorter than the measured, minimum repose time of 0.6 myr between
pulses (Fig. 7): Lomas PDCs 0.3 myr, Upper Sencca unit 0.29 myr,
Arma PDCs 0.02 myr and Caraveli sheet 0.38 myr (with analytical
The Capilla ignimbrite
uncertainty of 0.10–0.22 myr). There are three exceptions: (a) the long,
The Capilla ignimbrite is one of the young, small-volume, single 1.06 myr time interval of the Chuquibamba sheet includes two cooling
cooling units that crop out in glacially shaped valleys above the units, one proximal c. 13.2 Ma and and another distal c. 14.2 Ma,
Lower Sencca unit and below dome clusters and lava flows of although analytical uncertainty (0.18 myr) and one U/Pb 13.59 Ma age
Pleistocene age to the north of the OCM region (Figs 4a and 5b). Its suggest that the time interval may have been shorter; (b) the longer,
extent on the Cordillera–Altiplano divide is poorly known but 1.87 myr time interval is reflected by the thickest (200–500 m) OCM
similar outcrops are observed near Capilla and Culipampa in the succession of the Alpabamba multi-layered, composite sheet, which
eastern area of the geological map of Chulca (8 km north of Laguna lacks a detailed stratigraphy owing to unaccessible cliffs; (c) the
Ecma, Fig. 4a). The base is a 10 m thick vitrophyre (one only Ar/Ar longest, 2.30 myr time interval characterizes the eight scattered, dated
dated at 0.91 Ma), which grades into a pinnacle-shaped, weathered, units of the Lower Sencca Group. Table 2 suggests the existence of five
grey unit totalling 40 m. This vitrophyre is overlain by a grey, distinct pulses (based on the average analytical uncertainty), or four (as
welded, columnar jointed, massive lapilli tuff about 10 m thick. the stratigraphic A–D units) if we take the average 0.6 myr repose time
from the entire chronological record into account (Fig. 7).
Chronological constraints and limitations
Volcanological implications
Two geochronological datasets (40Ar/39Ar and U/Pb) have been used
to constrain the chronology and stratigraphy of the OCM Neogene and Three implications stem from the chronostratigraphy of the OCM
Quaternary ignimbrites. There are no discrepancies between Ar–Ar or ignimbrites (Table 2 and Figs 3 and 7): (1) the ignimbrite-forming
U/Pb dates for the same ignimbrite unit, and there are no age eruption rate changed after 9 Ma; (2) the growth of volcanic arcs and
discrepancies between chronology and stratigraphy. All 74 dated ignimbrites was intertwined after 9 Ma; (3) potential sources
ignimbrites and lavas are in stratigraphic succession within analytical (calderas) have been identified for the majority of ignimbrite
uncertainty (Table 2 and Fig. 4). Dates were obtained from multiple sheets. These implications show how ignimbrites contribute to
units within the majority of the ignimbrite sheets except for the Capilla, unravelling the volcanic history of the west Central Andes over the
Huarcaya and Majes ignimbrites. Published 40Ar/39Ar chronology on a past 25 myr.
number of ignimbrites and lava flows in the OCM region (Schildgen
et al. 2007, 2009; Schildgen 2009) are within uncertainty (within ‘Flare-ups’ before 9 Ma contrast with sustained explosive
0.1 myr) of the dataset presented here. Limitations in correlating
activity after 5 Ma
stratigraphy with chronology stem from a very limited number of
ignimbrite units, in particular where deposits drape steep valley slopes. New Ar/Ar and U/Pb age determinations and correlations (Table 2
These points are addressed separately below. and Fig. 7) are the basis for the statistical analysis of ignimbrite
(1) Both Ar/Ar and U/Pb dates from the same ignimbrite units but recurrence. Volume–time relationships and average eruptive rates
from different sites show no major discrepancy. For example, the allow us to address some of the chronostratigraphic disagreements
1.53 ± 0.02 Ma U/Pb age of the Las Lomas unit in the Río Arma that exist in the literature.
valley lies between the oldest Lomas PDC deposits Ar/Ar dated at Lulls in eruptive activity, defined as the absence of sizeable
1.46 ± 0.01 and 1.56 ± 0.32 Ma. The Upper Sencca unit U/Pb dated (<1 km3) pyroclastic deposits, reveal that the repose times between
1.87 ± 0.08 and 2.08 ± 0.09 Ma (Salamanca cliff, Fig. 5e) lies in the very large-volume (>50 km3) and large-volume (5–50 km3 VEI 6)
40
Ar/39Ar age range of this unit (2.09–1.80 Ma). Distal outcrops of ignimbrites are significantly longer than those between their small-
the Lower Sencca ignimbrite yielded four scattered 2.83–4.34 Ma U/ volume (<5 km3) counterparts. Recorded ignimbrite pulses after
Pb ages, but all are close to or within the 3.16–5.13 Ma range of seven 9 Ma, interspersed with the growth of composite volcanoes after
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Fig. 8. Cumulative eruptive volumes of OCM ignimbrites v. time. Maximum and minimum cumulative curves are based on liberal and conservative
estimates of thicknesses and original areas of the mapped units (Table 1 and Fig. 4). Bulk volumes are underestimated because of erosion, burial of older
units and the unknown extent of large sheets offshore. Timing of three volcanic ranges growth is shown together with processes that may explain the
apparent decline in ignimbrite production after 9 Ma. MF: monogenetic field.
7.3 Ma, suggest a steady eruptive activity. Alpabamba, Growth of three volcanic ranges after 9 Ma intertwined
Chuquibamba and Caravelí were emplaced by very large (VEI 7) with ignimbrites
or ‘super-eruptions’ (Self & Blake 2008) separated by long lulls of
c. 2–3.8 myr. The 3.7–3.8 myr lulls exist between the voluminous The stratigraphy and extent of lava flows depicted in Figures 3 and
Alpabamba and Chuquibamba ignimbrites, the Chuquibamba and 4a, b and age determinations (Table 2) demonstrate that eruptive
Caravelí eruptions, and the onset of Lower Sencca activity. activity has been sustained after 9 Ma.
The large-magnitude–low-frequency pattern changed after 9 Ma. The oldest OCM volcanic range, defined between 8 Ma Arma
The maximum eruption rate decreased from 66.9 km3 myr−1 prior to PDC deposits and 7–5.36 Ma lava flows, differs from the 10 to 3 Ma
9 Ma to 26.3 km3 myr−1 after 9 Ma (Fig. 8 and Table 1). Explosive interval of the ‘Lower Barroso Arc’ fixed by Mamani et al. (2010).
volcanism shifted to more sustained activity at c. 5.1 Ma producing The oldest OCM range of volcanoes is inferred from a thick
moderate- to small-volume but frequent ignimbrites separated by lulls succession of lava flows overlapping the Chuquibamba and
that did not exceed 0.6 myr (Fig. 7). Decrease in quiescence duration Caravelí ignimbrites. Eroded piles of thin, low-angle andesitic
between pyroclastic episodes has been observed after 7 Ma in the lava flows (e.g. between Río Marán and Río Cotahausi, Figs 1 and
Orcopampa area east of the OCM region and after 5 Ma in the 4a) and glacially eroded and hydrothermally altered edifice cores
Arequipa region (Paquereau-Lebti et al. 2006, 2008). (e.g. south of Nevado Sara Sara) reflect the growth of shield-like
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
stratovolcanoes north and south of the present-day frontal range. south of the Arcata rim. Monogenetic fields such as the
The large Late Miocene edifices were deeply cut by valley incision Orcopampa–Andahua–Huambo field 30 km east of the OCM
and eroded by glaciers and landslides. Both Arma PDC deposits and region (Delacour et al. 2007; Galas 2014) are associated with
7.3–5.3 Ma lava flows (Río Marán and Río Cotahuasi, Fig. 4a) deep-seated faults or calderas (Mamani et al. 2008).
allow us to better constrain the growth of Late Miocene volcanoes in
the OCM region. As composite volcanoes are not long-lived
Revised time constraints on volcanic history in south Peru
(<1 myr) systems (Hildreth & Lanphere 1999), the deeply eroded,
compound edifices claimed to have ages >9–14 Ma NW of the Our revised chronology of OCM Neogene and Quaternary
OCM region (Lefèvre 1979; Mamani et al. 2010) belong to the ignimbrites and lava flows together with correlations across SW
Middle Miocene ‘Huaylillas’ period. Peru replaces the formation nomenclature used in INGEMMET
Lava-flow chronology suggests that a 3 myr interval separated the maps and in the synthesis by Mamani et al. (2010) (Table 2 and
Late Miocene range from the Quaternary volcano range. During this Fig. 7). This leads to a complex volcanic history that intertwines
long interval, the Lower Sencca sheet (5.1–2.8 Ma) shortly ignimbrites, composite cone ranges and monogenetic fields. Figures
followed the Miocene range whereas the Upper Sencca and 3 and 7 propose a chronostratigraphic nomenclature of seven main
Lomas ignimbrites (2.09–1.26 Ma) coexisted with Quaternary volcanic stages linked with the evolution of the Western Cordillera
edifices. Instead of the ‘Upper Barroso arc’, two categories of in the south Peruvian Central Andes.
volcanic range can be distinguished (Fig. 2), as follows. (1) The (1) The Late Oligocene–Early Miocene (30–20 Ma) includes the
OCM Quaternary volcanoes are better preserved than their Miocene Nazca group of ignimbrites. We bracket the ‘Tacaza Group’
counterparts, although they display amphitheatres formed by between 30 and c. 20 Ma instead of 30–24 Ma (Mamani et al.
landslides or flank failures (e.g. Nevado Solimana) and eroded by 2010). The upper limit of Oligocene volcaniclastic rocks attributed
glaciers. These edifices point to an activity (2.27–1.30 Ma) shorter to this group has been loosely dated between c. 24 or 21.7 and
than the ‘Upper Barroso Group’ approximated between 3 and 1 Ma 15.85 Ma (Bellon & Lefèvre 1976; Lefèvre 1979). The c. 30 Ma
(Mamani et al. (2010). (2) The Pleistocene and present-day volcanic Jallua tuff (Swanson et al. 2004) east of the OCM region near
range straddles the west flank of the Western Cordillera (Figs 2 and Orcopampa (Fig. 2), the oldest known ignimbrite outside the
4a). From available Ar/Ar age determinations and geological maps, tuff beds intercalated in the Moquegua Formation, is used here as
the Pleistocene range did not start before 1.3 ± 0.1 Ma and the base of the Tacaza Group. The uppermost limit coincides with
encompasses several coeval composite cones and dome clusters; the base of the Alpabamba ignimbrite dated at c. 20.10 Ma
in chronological order, these are Nevados Coropuna, Firura and (Table 2). Because both Nazca tuffs reflect interspersed explosive
Sara Sara (Figs 2, 4 and 6). In SW Peru, the Pleistocene range activity increasing before (Nazca 1, c. 24–25 Ma) and after (Nazca
includes (a) volcanic complexes including extinct, eroded strato- 2, c. 22–23 Ma) the top of Tacaza Group, we extend its boundary to
volcanoes and dome coulée clusters (Nocarane in Chachani, Hualca 20 Ma.
Hualca), (b) dormant but youthful eruptive centres (Nevado (2) The Early–Middle Miocene stage (20–13 Ma) is bracketed by
Chachani, Coropuna, Firura), none having an age >1 Ma, and (c) the Alpabamaba composite sheet at the base and the Chuquibamba
active composite volcanoes all with ages <0.6 Ma. Nevado Sara compound sheet at the top. We bracket the ‘Huaylillas’ Formation
Sara, the northernmost Pleistocene composite volcano in the Central between c. 20 and c. 13 Ma (instead of 24–10 Ma; Mamani et al.
Andes, is considered as active (Morche & Nuñez Juárez 1998) but 2010) on the basis of the two most extensive and voluminous
no Holocene deposits have been dated so far. The most recent lava ignimbrites in south Peru (Alpabamba c. 20.2–18.23 Ma and
flows on its east flank have been dated between c. 146 and 20 ka Chuquibamba c. 14.3–13.2 Ma), and because the Chuquibamba
(Table 2). ignimbrite pulse preceded the Late Miocene uplift and deep river
incision (Schildgen et al. 2007; Thouret et al. 2007). The Majes tuff
(16.25 Ma) and MoqC1–C2 volcaniclastic material heralded the
Quasi-steady eruptive activity after 2.27 Ma
increase in eruptive activity recorded in the Huaylillas Formation
We argue that the eruptive activity was quasi-steady after 2.27 Ma. (Roperch et al. 2006; Decou et al. 2011) at a time when the forearc
Lava flows of that age in Rio Ocoña, probably sourced at Nevado was not experiencing warping and uplift. The forearc, now 2.2 km
Solimana, were immediately followed by 2.09–1.8 Ma Upper Sencca a.s.l., was at sea level at 25 Ma as shown by marine sediment of that
ignimbrites. In South Peru, ubiquitous Upper Sencca ignimbrites age at Pampa Gramadal and Cuno Cuno (Figs 5a, b and 6; Cruzado
encompass a collection of flow units between c. 2.20 Ma (Río Colca, & Rojas 2005). The lithofacies of Nazca and Majes tuff layers in
Fig. 6, section G–G′) and c. 1.62 Ma (Arequipa Airport ignimbrite: conglomerates suggests emplacement in shallow water bodies. All
Paquereau-Lebti et al. 2008). The lowermost lava flows of the eroded younger ignimbrites such as Chuquibamba show no evidence for
Quaternary volcanoes overlie the top Lomas deposits (c. 1.36 Ma) emplacement into water.
indicating almost no lull until the growth of the next generation of (3) The Middle–Late Miocene (11–5 Ma) pyroclastic stage
Pleistocene volcanoes (≤1.3–0.6 Ma; Thouret et al. 2001, 2005; Gerbe includes the 10.78 Ma Huarcaya ignimbrite, the c. 9 Ma Caraveli
& Thouret 2004). In the Arequipa region, PDC deposits of c. 1.41 Ma ignimbrite and the c. 8 Ma PDC deposits, and the oldest
overlie the upper unit of the Arequipa Airport Ignimbrite. The quasi- recognizable range of shield volcanoes (7.3–5.3 Ma). This coin-
contemporary Capillune Formation has been geochemically correlated cides with the most recent pulse of surface uplift of the Western
with the Yura tuffs c. 1.02 Ma (Paquereau-Lebti et al. 2008). In the Cordillera (Schildgen et al. 2007, 2010; Thouret et al. 2007). The
north OCM region, Capilla is one of the youngest inflow ignimbrites, Caraveli episode was the most recent ignimbrite flare-up. After
dated at 0.91 Ma in the Huarcaya caldera. These Early to Middle 7.3 Ma, magma output became steadier albeit smaller and
Pleistocene ignimbrites clearly post-date the Upper Sencca ignimbrite contributed to the intertwined growth of volcanic ranges of
and pre-date large volcanic complexes (e.g. Chachani; Aguilar et al. composite cones with moderate volumes of ignimbrite sheets.
2015) and active composite volcanoes (<0.8–0.6 Ma; base of Misti, (4) The Pliocene stage is characterized by the emplacement of the
Sabancaya and Ubinas). composite, multi-layered sheet of the Lower Sencca Group (5.3–
After 0.7 Ma, abundant monogenetic cones (e.g. the Auquihuato 2.83 Ma) similar to the composite sheet of Alpabamba. The latest
cone with lava flows of 11 km length north of Oyolo) and lava fields units of the Group were confined in deep valleys in contrast to the
coexisted with composite cones (Fig. 4a). Unglaciated mafic lava earliest units, which often crown the Alpabamba or Chuquibamba
flows erupted near calderas during the Holocene, such as one 6 km successions.
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
(5) The Early Quaternary volcanic range grew after 2.27 Ma and is still sizeable but each pulse produced a smaller amount of
their deposits intertwined with the Upper Sencca unit at c. 1.97 Ma ignimbrites and/or PDCs. Short-lived lulls between moderate to
on average. Nevado Solimana, one of these edifices, is a probable small volumes of ignimbrites after 5 Ma suggest the growth of
source of Upper Sencca units and Lomas PDC deposits. The relatively large and shallow magma reservoirs in the upper crust.
Nevado Coropuna dome cluster has grown on one of the sources of This coincided with the eruption of large volumes of andesite lava
Upper Sencca units. flows after 7.3 Ma, which became steady after 2.27 Ma and
(6) The Pleistocene volcanic range <1.3 Ma was contemporan- contributed to the growth of three generations of andesite–dacite
eous with, and followed the 1.56–1.26 Ma ‘Lomas’ PDC deposits. composite volcanoes (7.3–5.3, 2.3–1.5 and <1.3 Ma) together with
The latter is slightly younger than the Arequipa Airport ignimbrite (at <0.7 Ma) more mafic monogenetic fields.
(c. 1.65 Ma) and contemporaneous with subsequent PDC deposits The change at 9 Ma near the onset of the major uplift period
near Arequipa. The most recent ignimbrites of sizeable volume coincided with a prevailing extensional tectonic regime (Sempere
(Capilla in the north OCM region, Yura tuffs near Arequipa) erupted et al. 2008) allowing the ascent of more mafic magma. Low-
between 1.1 and 0.9 Ma. magnitude eruption activity may be linked to decreasing plate
(7) Finally, the Middle to Late Pleistocene monogenetic fields of convergence rate and obliquity since the Late Miocene (Somoza
andesite and basaltic andesite lava flows and cones grew together 1998; Norabuena et al. 1999; Martinod et al. 2010; Quinteros &
with the most recent, large composite cones at <0.6 Ma, including Sobolev 2013), whereas increased eruption frequency of small,
the active Sara Sara composite cone and Coropuna dome cluster. mafic magma batches may reflect magma recharge into large mid-
crustal magma reservoirs (Caricchi et al. 2014; Malfait et al. 2014)
or andesite magma stagnation in the upper crust before final
Discussion
eruption (Mamani et al. 2010).
Twelve ignimbrites cover about 7050 km2 with a minimum bulk
volume of about 400–1200 km3 in the OCM region, which
Calderas as potential sources for several ignimbrite sheets
represents almost 30% of the widespread Neogene ignimbrites of
SW Peru (Tosdal et al. 1981; Mamani et al. 2010). Despite a decline Two categories of geographical sources for ignimbrites include
in OCM ignimbrite volume after 9 Ma, a large silicic volcanic probable calderas identified in the field and possible calderas
province grew as multiple ignimbrites were erupted until 0.9 Ma, suggested by remote sensing using satellite imagery, on which
intertwined with the growth of composite volcano arcs. systematic anisotropy of magnetic susceptibility measurements will
shed more light (Bréard 2011; De La Rupelle 2013). Several factors
Volume of erupted magmas and fluctuations of eruption hinder the identification of OCM calderas: substantial post-caldera
volcanism, combined with glacial erosion and faults, blurs caldera
rates
boundaries in the north OCM region; also, the oldest ignimbrites are
Figure 8 shows the minimum and maximum cumulative volumes of either distal tuff layers with no visible link with the Western
the OCM ignimbrites. On the basis of available deposits onland and Cordillera or thick sheets buried by huge volumes of subsequent
geographic information system measurements, the area covered by volcanic deposits. However, the exposure distribution of recent
ignimbrites is 7050 km2 (Table 1 and Fig. 4) and the minimum ignimbrites points to depressions assumed to be eruptive centres,
volume ranges between c. 404 and 1186 km3, the latter probably based on geological evidence such as collapse rims and floors,
being closer to the initial erupted volume. A significant volume of outflow deposits and/or radially sloping ignimbritic aprons (e.g.
ignimbrite may have been transported offshore and co-ignimbrite Nevado Coropuna).
ash has not been observed in the study area, which makes our The first category of ‘young’ calderas includes three probable
volume estimation a minimum figure. Erosion has also removed at sources for the Lomas and Upper Sencca ignimbrites. The
least 25–30% of the Pliocene and Quaternary ignimbrites during Huynacotas 6 × 4.5 km piecemeal collapse caldera, 6 km NNW of
uplift and river incision after 13 Ma. the Huynacotas village, is the likely source of the Lomas ignimbrite:
The cumulative volume of OCM ignimbrites is smaller by a factor outflow deposits can be traced from the rim 22 km to the SSE on the
of 3–5 than that of the ignimbrite fields of Yellowstone and Taupo high-plateau edge upstream of Cotahuasi and 58 km farther on the
(Salisbury et al. 2011, fig. 10). From the table in Figure 8, however, Lomas–Chaucalla ridge. The large Arcata caldera (10.5 km in size)
the OCM average eruption rate of c. 17–50 km3 myr−1 over 25 myr includes a resurgent dome 9 km NNW of the Arcata Mine. The dip
belies high production during flare-ups: c. 127–171 km3 myr−1 of the thick pile of Sencca ignimbrites forming the Lomas–
(Caraveli) and c. 117–467 km3 myr−1 (Alpabamba). This rate slowed Chaucalla ridge points to a source located beneath the Quaternary
markedly to c.12–76 km3 myr−1 after 9 Ma. Magma production also Nevado Solimana (Fig. 6, section F–F′).
contributed to the growth of central shield volcanoes between 7.3 and The second category of large and ‘older’ calderas stems from
5.3 Ma and two ranges of composite volcanoes after 2.27 Ma. As a satellite images and geological interpretations. The widespread,
result, the arc magma eruption rate was higher than the ignimbrite shallow-valley filling Caraveli ignimbrite and outcrops near Pausa
production after 2.27 Ma (53 km3 v. 29.5–46 km3), but the estimated north of Río Marán point to a source between the Laguna
linear rate decreased from 0.15 to 0.08 km3 km−1 ka−1 between the Parinacocha and Nevado Sara Sara (De La Rupelle 2013) or to
Early Quaternary and late Pleistocene volcanic ranges. the NE, the Huancarama presumed caldera, at a time when the Late
Miocene volcano range did not exist (Figs 4 and 6, section B–B′).
Huarcaya, the largest OCM caldera, which is 27 km in diameter and
Change in crustal melting rate after 9 Ma?
is on the divide between the Western Cordillera and the Altiplano, is
Contrasting magnitude–frequency patterns before and after 9 Ma poorly visible from the ground as the Lower Sencca inflow
raise the issue of what drove the change in crustal melting rate since ignimbrites, Quaternary cones and dome clusters, the Capilla
the Late Miocene. Super-eruptions before 9 Ma suggest increase in ignimbrite and glaciers have blurred the shape and boundaries of the
magma genesis at the subduction zone and in the deep crust owing depression. Extensive aprons of Sencca ignimbrites with inter-
to asthenospheric flux towards the mantle wedge induced by rapid calated Quaternary lavas dip gently over distances of 15–25 km
convergence rate at the trench and/or increasing magma ascent rates around Nevado Coropuna, the highest and most voluminous dome
initiating large crustal melting in the middle crust. After the most cluster in south Peru. This may also be one source of the
recent Caraveli flare-up at 9 Ma, the erupted volume of ignimbrites Chuquibamba ignimbrites (Bréard 2011).
Downloaded from [Link] at University of Birmingham on May 6, 2016
J.-[Link] et al.
Conclusions Beckinsale, R.D., Sanchez-Fernandez, A.W., Brook, M., Cobbing, E.J., Taylor,
W.P. & Moore, N.D. 1985. Rb–Sr whole-rock isochron and K–Ar age
(1) A set of 74 40Ar/39Ar and U/Pb age determinations together with determinations for the Coastal Batholith of Peru. In: Pitcher, W.S., Atherton,
M.P., Cobbing, E.J. & Beckinsale, R.D. (eds) Magmatism at a Plate Edge: The
lithofacies analysis and mapping allowed us to refine the volcanic Peruvian Andes. Blackie, Glasgow; Halsted Press, New York, 177–202.
stratigraphy and history of the Central Andes in Peru. In the OCM Bellon, H. & Lefèvre, C. 1976. Données géochronométriques sur le volcanisme
region, 12 ignimbrite sheets and PDC deposits have been emplaced andin dans le sud du Pérou: implications volcanotectoniques. Comptes Rendus
every 1.9 myr on average over the past c. 25 myr: Nazca 1, Nazca 2, de l’Académie des Sciences, Série D, 283, 1–4.
Benavides-Cáceres, V. 1999. The Andean cycle. In: Skinner, B.J. (ed.) Geology
Alpabamba, Majes, Chuquibamba, Huarcaya, Caravelí, Arma, and Ore Deposits of the Central Andes. Society of Economic Geologists,
Lower Sencca, Upper Sencca, Las Lomas and Capilla. Additional Special Publications, 7, 61–107.
ignimbrites between c. 30 and 2.7 Ma identified in the adjacent Branney, M.J. & Kokelaar, P. 2002. Pyroclastic Density Currents and
Sedimentation of Ignimbrites. Geological Society, London, Memoirs, 27.
Orcopampa region east of OCM support the fact that pyroclastic Bréard, E. 2011. Localisation des sources d’ignimbrites de la région des canyons
activity became more sustained while the Cordillera uplift was Ocona–Cotahuasi–Maran (Sud du Pérou) par l’étude de l’Anisotropie de
taking place. No significant (>0.6 myr) break occurred after 5 Ma. Susceptibilité Magnétique (ASM). Masters thesis, Université Blaise Pascal,
Instead, quasi-continuous volcanism produced three generations of Clermont-Ferrand.
Caricchi, L., Annen, C., Blundy, J., Simpson, G. & Pinel, V. 2014. Frequency and
composite volcanoes with four intercalated ignimbrite sheets and magnitude of volcanic eruptions controlled by magma injection and
PDC deposits, and Pleistocene monogenetic fields. buoyancy. Nature Geosciences, 7, 126–130.
(2) Minimum bulk volumes of c. 400–1200 km3 for ignimbrites Cruzado, G.H. & Rojas, M.C. 2005. Estratigrafía de la Fm. Moquegua superior
en el area del Cerro Cuno Cuno y Pampa del Gramadal. Tesis de grado,
emplaced over 25 myr in the OCM region represent a minimum Universidad Mayor de San Marcos, Lima.
average volume output of c. 17–50 km3 myr−1. Repeated pulses or Decou, A., von Eynatten, H., Mamani, M., Sempere, T. & Wörner, G. 2011.
‘flare-ups’ of ignimbrites occurred on average every 2–3.8 myr Cenozoic forearc basin sediments in Southern Peru (15–18°S): Stratigraphic
between 24.5 and 9.0 Ma. In contrast, quasi-continuous volcanism and heavy mineral constraints for Eocene to Miocene evolution of the Central
Andes. Sedimentary Geology, 237, 55–72.
after 5 Ma produced four smaller ignimbrite sheets and PDC Decou, A., Von Eynatten, H., Dunkl, I., Frei, D. & Wörner, G. 2013. Late Eocene
deposits. As a result, the ignimbrite production rate decreased to Early Miocene Andean uplift inferred from detrital zircon fission track and
markedly after 9 Ma (from c. 127–171 to c. 12–76 km3 myr−1). U–Pb dating of Cenozoic forearc sediments (15–18°S). Journal of South
American Earth Sciences, 45, 6–23.
Output became less punctuated after 5 Ma when repose time Delacour, A., Gerbe, M.-C., Thouret, J.-C., Wörner, G. & Paquereau-Lebti, P.
decreased to 0.85 myr. The two Early Quaternary and Pleistocene 2007. Magma evolution of Quaternary minor volcanic centres in southern
volcano ranges have added more bulk volume (c. 120 km3, i.e. c. Peru, Central Andes. Bulletin of Volcanology, 69, 581–608.
53 km3 myr−1) than the 29.5–46 km3 of ignimbrites after 2.27 Ma. De La Rupelle, A. 2013. Le volcanisme ignimbritique des canyons Ocona–
Cotahuasi (sud du Pérou): chronostratigraphie, sources et liens avec la
Estimated linear magma output has, however, apparently decreased surrection andine. PhD thesis, Université Blaise Pascal, Clermont-Ferrand.
twofold, 0.15–0.08 km3 km−1 myr−1, from the Early Quaternary to Demouy, S., Paquette, J.-L., et al. 2012. Spatial and temporal evolution of Liassic
the Pleistocene volcano range. to Paleocene arc activity in southern Peru unraveled by zircon U–Pb and Hf in-
situ data on plutonic rocks. Lithos, 155, 183–200.
(3) We attribute the change in ignimbrite production between 9 De Silva, S.L. 1989. Altiplano–Puna volcanic complex of the Central Andes.
and 5 Ma to declining crustal melting. This change coincided with Geology, 17, 1102–1106.
the arrival of the Nazca Ridge at the trench at 5.9 Ma and flat Galas, A. 2014. Petrology and new data on the geochemistry of the Andahua
volcanic group (Central Andes, southern Peru). Journal of South American
subduction under the overriding South America plate thereafter Earth Sciences, 56, 301–315.
(Hampel 2002), and with the steady decline of the convergence rate Garzione, C.N., Hoke, G.D., et al. 2008. Rise of the Andes. Science, 320,
after 10 Ma (Martinod et al. 2010). Although the interplay of slower 1304–1307.
plate convergence rate with magma recharge or stagnation in the Gerbe, M.-C. & Thouret, J.-C. 2004. Role of magma mixing in the petrogenesis
of lavas erupted through the 1990–98 explosive activity of Nevado Sabancaya
crust must be considered, the mechanisms that drove changes in in south Peru. Bulletin of Volcanology, 66, 541–561.
crustal melting after 9 Ma remain to be addressed. However, the Guevara, C. 1969. Geología del cuadrángulo de Characato. Boletin, 23. Servicio
ignimbrite record together with lava flows can be used as proxies to de Geología y Minería, Lima.
Guevara, C. & Dávila, D. 1983. Estratigrafía y tectónica terciaria del area de
reconstruct the surface uplift and incision history of the west Central Cora Cora y Pacapausa. Boletín de la Sociedad Geológica del Perú, 71,
Andes over the past 25 myr. 281–289.
Gunnell, Y., Thouret, J.C., Brichau, S., Carter, A. & Gallagher, K. 2010. Low-
temperature thermochronology in the Peruvian Central Andes: implications
Acknowledgements and Funding for long-term continental denudation, timing of plateau uplift, canyon incision
and lithosphere dynamics. Journal of the Geological Society, London, 167,
This work has been supported by grants from INSU ‘CT Aléas’ programmes,
803–815, [Link]
PICS CNRS and IRD collaborative projects with IGP and INGEMMET, ANR
Hampel, A. 2002. The migration history of the Nazca Ridge along the Peruvian
RiskNat Laharisk and the French Embassy in Peru. E. Cubukcu thanks Clermont
active margin: a re-evaluation. Earth and Planetary Science Letters, 203,
University for a post-doctorate period in 2011. Laboratory analyses have been
665–679.
performed by L. Akin, D. Auclair, C. Bosq, M. Benbakkar, J.-L. Devidal,
Hildreth, W. & Lanphere, M.A. 1999. Potassium–argon geochronology of a
C. Fonquernie, J. Marin (CRPG), P. Boivin, M.C. Gerbe and A. De La Rupelle.
basalt–andesite–dacite arc system: The Mount Adams volcanic field, Cascade
L. Thouret has provided field assistance and artwork and E. Romero has been a
Range of southern Washington. Geological Society of America Bulletin, 106,
helpful field assistant. Discussion with T. Sempere and review comments from
1426–1439.
R. Lease, one anonymous reviewer and the Editor, Dr. Crowley, greatly improved the
Huaman, D. 1985. Evolution tectonique cénozoïque et néotectonique du Piémont
paper. This is CLERVOLC Contribution 193.
pacifique dans la région d’Arequipa (Andes du Sud du Pérou). PhD Thèse,
Scientific editing by Quentin Crowley Université Paris XI Orsay.
Iaffaldano, G. & Bunge, H.P. 2008. Strong plate coupling along the Nazca–South
America convergent margin. Geology, 36, 443–446.
Jenks, W.F. & Goldich, S.X. 1956. Rhyolitic tuff flows in southern Peru. Journal
References of Geology, 64, 156–172.
Acosta, H., Alván, O., Oviedo, M., Mamani, M. & Rodriguez, J. 2011. Mapa Kaneoka, I. & Guevara, C. 1984. K–Ar age determinations of late Tertiary and
geológico de la costa sur del Perú, Proyecto GR-1, Geología de la costa Sur y Quaternary Andean volcanic rocks, southern Peru. Geochemical Journal, 18,
de la vertiente Oeste de la Cordillera Occidental. INGEMMET, Lima. 233–239.
Aguilar, R., Thouret, J.-C., Suaña, E., Samaniego, P., Jicha, B. & Rivera, M. Le Bas, M.J., Le Maitre, R.W., Streckeisen, A. & Zanettin, B. 1986. A chemical
2015. Evolution of a long-lived volcanic complex: The Chachani case study classification of volcanic rocks based on the total alkali–silica diagram.
(Peru). In: Foro Internacional sobre gestion del riesgo geológico, Arequipa, Journal of Petrology, 27, 745–750.
Libro de resumenes, INGEMMET, OVI, Gobierno Regional. 14–16 October Lefèvre, C. 1979. Un exemple de volcanisme de marge active dans les Andes du
2015, Moquegua, Peru, 138–141. Pérou (Sud) du Miocène à l’Actuel. PhD thesis, Université Sciences et
Barker, S.D. 1996. El sillar ocioso. Bulletin of Volcanology, 58, 317–318. Techniques du Languedoc, Montpellier.
Barnes, J.B. & Ehlers, T.A. 2009. End member models for Andean Plateau uplift. Malfait, W.J., Seifert, R., et al. 2014. Supervolcano eruptions driven by melt
Earth-Science Reviews, 97, 105–132. buoyancy in large silicic magma chambers. Nature Geoscience, 7, 84–85.
Downloaded from [Link] at University of Birmingham on May 6, 2016
Ignimbrites and volcanic history of Central Andes
Mamani, M., Wörner, G. & Thouret, J.-C. 2008. Tracing a major crustal domain Roperch, P., Carlotto, V., Ruffet, G. & Fornari, M. 2011. Tectonic rotations and
boundary based on the geochemistry of minor volcanic centres in southern transcurrent deformation south of the Abancaya deflection in the Andes of
Peru. In: 7th International Symposium on Andean Geodynamics, Extended south Peru. Tectonics, 30, TC2010.
Abstracts, Institut de Recherche pour le Développement Editions, Nice, Salisbury, M.J., Jicha, B.R., de Silva, S.L., Singer, B.S., Jimenez, N.C. & Ort, M.
France, 298–301. H. 2011. 40Ar/39Ar chronostratigraphy of Altiplano–Puna volcanic complex
Mamani, M., Wörner, G. & Sempere, T. 2010. Geochemical variations in igneous ignimbrites reveals the development of major magmatic province. Geological
rocks of the Central Andean orocline (13°S to 18°S): Tracing crustal Society of America Bulletin, 123, 821–840.
thickening and magma generation through time and space. Geological Society Schildgen, T.F. 2009. Late Cenozoic structural and tectonic development of the
of America Bulletin, 122, 162–182. western margin of the central Andean plateau in southwest Peru. Tectonics, 28,
Martinod, J., Husson, L., Roperch, P., Guillaume, B. & Espurt, N. 2010. TC 4007.
Horizontal subduction zones, convergence velocity and the building of the Schildgen, T.F., Hodges, K.V., Whipple, K.X., Reiners, P.W. & Pringle, M.S.
Andes. Earth and Planetary Science Letters, 299, 299–309. 2007. Uplift of the western margin of the Andean plateau revealed from
McQuarrie, N., Horton, B.K., Zandt, G., Beck, S. & DeCelles, P.G. 2005. canyon incision history, southern Peru. Geology, 35, 523–526.
Lithospheric evolution of the Andean fold–thrust belt, Bolivia, and the origin Schildgen, T.F., Ehlers, T.A., Whipp, D.M., Jr, van Soest, M.C., Whipple, K.X.
of the central Andean plateau. Tectonophysics, 399, 15–37. & Hodges, K.V. 2009. Quantifying canyon incision and Andean Plateau
Mendivíl, E.S. 1965. Geología de los cuadrángulos de Maure y Antajave. surface uplift, southwest Peru: A thermochronometer and numerical
Commisión Carta Geológica Nacional, Boletín, 10. modeling approach. Journal of Geophysical Research: Earth Surface, 114,
Morche, W. & Nuñez Juárez, S. 1998. Estudio del riesgo geológico del volcán FO4014.
Sara Sara. INGEMMET, Lima. Schildgen, T.F., Balco, G. & Shuster, D.L. 2010. Canyon incision and knickpoint
Newell, N.D. 1949. Geology of the Lake Titicaca Region, Peru and Bolivia. propagation recorded by apatite 4He/3He thermochronometry. Earth and
Geological Society of America, Memoirs, 36. Planetary Science Letters, 293, 377–387.
Noble, D.C., McKee, E.H., Farrar, E. & Petersen, U. 1974. Episodic Cenozoic Sébrier, M., Lavenu, A., Fornari, M. & Soulas, J. 1988. Tectonics and uplift in the
volcanism and tectonism in the Andes of Peru. Earth and Planetary Science Central Andes (Peru, Bolivia and northern Chile) from Eocene to present.
Letters, 21, 213–220. Géodynamique (ORSTOM), 3, 85–106.
Noble, D.C., Farrar, E. & Cobbing, E.J. 1979. The Nazca group of south central Self, S. & Blake, S. 2008. Consequences of explosive supereruptions. Elements,
Peru: age, source, and regional volcanic and tectonic significance. Earth and 4, 41–46.
Planetary Science Letters, 45, 80–86. Sempere, T., Folguera, A. & Gerbault, M. 2008. New insights into the Andean
Noble, D.C., McKee, E.H., Eyzaguirre, V.R. & Marocco, R. 1984. Age and evolution: An introduction to contributions from the 6th IASAG Symposium
regional tectonic and metallogenetic implications of igneous activity and (Barcelona, 2005). Tectonophysics, 459, 1–13.
mineralization in the Andahuaylas–Yauri Belt of southern Peru. Economic Sempere, T., Noury, M., García, F. & Bernet, M. 2014. Elementos para una
Geology, 79, 172–176. actualización de la estratigrafia del Grupo Moquegua, sur del Peru. In XVII
Noble, D.C., Sébrier, M., Mégard, F. & McKee, E.H. 1985. Demonstration of two Congreso Peruano de Geología, Lima, Extended Abstracts, digital file
pulses of Paleogene deformation in the Andes of Peru. Earth and Planetary “Sempere, T”, Sociedad Geológica del Perú.
Science Letters, 73, 345–349. Skinner, B.J. 2009. Geology and Ore Deposits of the Central Andes. Society of
Norabuena, E., Dixon, D.H., Stein, S. & Harrison, C.G.A. 1999. Decelerating Economic Geologists, Special Publications, 7.
Nazca–South America and Nazca–Pacific Plate motions. Geophysical Smith, R.L. 1960. Ash flows. Geological Society of America Bulletin, 71,
Research Letters, 26, 3405–3408. 795–842.
Olchauski, E. & Dávila, D. 1994. Geología de los cuadrangulos de Somoza, R. 1998. Updated Nazca (Farallon)–South America relative
Chuquibamba y Cotahuasi, hojas 32q & 31q, 1:100,000. Carta Geológica motions during the last 40 My: implications for mountain building in the
Nacional, Boletin 50, Serie A. INGEMMET, Lima. central Andean region. Journal of South America Earth Sciences, 11,
Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P. & Schemmann, K. 2006. 211–215.
Deformation of the Central Andean upper plate system—facts, fiction, and Swanson, K.E., Noble, D.C., Connors, K.A., Mayta, T.O., McKee, E.H.,
constraints for plateau models. In: Oncken, O., Chong, G. & Franz, G. (eds) Sanchez, P.A. & Heizler, M.T. 2004. Mapa geológico del cuadrángulo de
The Andes: Active Subduction Orogeny. Springer, Berlin, 3–27. Orcopampa (Sur del Perú). Carta Geológica Nacional, Boletín 137, Serie A.
Paquereau-Lebti, P., Thouret, J.C., Wörner, G. & Fornari, M. 2006. Neogene and INGEMMET, Lima.
Quaternary ignimbrites in the area of Arequipa, Southern Peru: Stratigraphical Thouret, J.-C., Suni, J., Finizola, A., Fornari, M., Legeley-Padovani, A. &
and petrological correlations. Journal of Volcanology and Geothermal Frechen, M. 2001. Geology of El Misti volcano near the city of Arequipa,
Research, 154, 251–275. Peru. Geological Society of America Bulletin, 113, 1593–1610.
Paquereau-Lebti, P., Fornari, M., Roperch, P., Thouret, J.C. & Macedo, O. 2008. Thouret, J.-C., Rivera, M., Wörner, G., Gerbe, M.-C., Finizola, A., Fornari, M. &
Paleomagnetism, magnetic fabric, and 40Ar/39Ar dating of Pliocene and Gonzales, K. 2005. Ubinas: the evolution of the historically most active
Quaternary ignimbrites in the Arequipa area, southern Peru. Bulletin of volcano in southern Peru. Bulletin of Volcanology, 67, 557–589.
Volcanology, 70, 977–997. Thouret, J.C., Wörner, G., Gunnell, Y., Singer, B., Zhang, X. & Souriot, T. 2007.
Pecho Gutierrez, V. 1983a. Geología de los cuadrángulos de Chalhuanca, Geochronologic and stratigraphic constraints on canyon incision and Miocene
Antabamba y Santo Tomas, Hojas 29p, 29q y 29r, 1:100,000. Carta Geológica uplift of the Central Andes in Peru. Earth and Planetary Science Letters, 263,
Nacional, Boletin 37, Serie A. INGEMMET, Lima. 151–166.
Pecho Gutierrez, V. 1983b. Geología de los cuadrángulos de Pausa y Caravelí, Tosdal, R.M., Farrar, E. & Clark, A.H. 1981. K–Ar geochronology of the Late
Hojas 31p y 32p, 1:100,000. Carta Geológica Nacional, Boletin 39, Serie A. Cenozoic volcanic rocks of the Cordillera Occidental, southernmost Peru.
INGEMMET, Lima. Journal of Volcanology and Geothermal Research, 10, 157–173.
Quane, S.L. & Russel, J.K. 2005. Ranking welding intensity in pyroclastic Tosdal, R.M., Clark, A.H. & Farrar, E. 1984. Cenozoic polyphase landscape and
deposits. Bulletin of Volcanology, 67, 129–143. tectonic evolution of the Cordillera Occidental, southernmost Peru.
Quang, C.X., Clark, A.H., Lee, J.K.W. & Hawkes, N. 2005. Response of Geological Society of America Bulletin, 95, 1318–1332.
supergene processes to episodic Cenozoic uplift, pediment erosion, and Vargas, L.V. 1979. Geología del cuadrangulo de Arequipa, color geologic map
ignimbrite eruption in the porphyry copper province of southern Peru. 1:100,000. Boletín 24. Servicio de Geología y Minería, Lima.
Economic Geology, 100, 87–114. Wilson, C.J.N. & Hildreth, W. 2003. Assembling an ignimbrite: Mechanical and
Quinteros, J. & Sobolev, S.V. 2013. Why has the Nazca plate slowed since the thermal building blocks in the Bishop Tuff, California. Journal of Geology,
Neogene? Geology, 41, 31–34. 111, 653–670.
Quispesivana, L. & Navarro, P.C. 2001. Geologic map of Caraveli no. 32p: Wilson, J. & Garcia, W. 1962. Geología de los cuadrángulos de Pachía y Palca.
1:100,000. INGEMMET, Lima. Boletin 4, Serie A. Comisión Carta Geológica Nacional, Lima.
Ramos, V.A. 2008. The basement of the Central Andes: The Arequipa and related Wörner, G., Hammerschmidt, K., Henjes-Kunst, F., Lezaun, J. & Wilke, H. 2000.
terranes. Annual Review of Earth and Planetary Sciences, 36, 289–324. Geochronology (40Ar/39Ar, K–Ar and He-exposure ages) of Cenozoic
Roperch, P., Sempere, T., et al. 2006. Counterclockwise rotation of late Eocene– magmatic rocks from Northern Chile (18–22°S): implications for magmatism
Oligocene fore-arc deposits in southern Peru and its significance for oroclinal and tectonic evolution of the central Andes. Revista Geológica del Chile, 27,
bending in the central Andes. Tectonics, 25, TC3010. 205–240.