You are on page 1of 18

Journal of Volcanology and Geothermal Research 288 (2014) 28–45

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


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

Review

Calderas and magma reservoirs


Katharine V. Cashman a,⁎, Guido Giordano b
a
University of Bristol, UK
b
Universitá Roma Tre, Italy

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

Article history: Large caldera-forming eruptions have long been a focus of both petrological and volcanological studies; petrolo-
Received 4 June 2014 gists have used the eruptive products to probe conditions of magma storage (and thus processes that drive
Accepted 15 September 2014 magma evolution), while volcanologists have used them to study the conditions under which large volumes of
Available online 6 October 2014
magma are transported to, and emplaced on, the Earth's surface. Traditionally, both groups have worked on
the assumption that eruptible magma is stored within a single long-lived melt body. Over the past decade, how-
Keywords:
Explosive eruptions
ever, advances in analytical techniques have provided new views of magma storage regions, many of which pro-
Calderas vide evidence of multiple melt lenses feeding a single eruption, and/or rapid pre-eruptive assembly of large
Magma storage volumes of melt. These new petrological views of magmatic systems have not yet been fully integrated into vol-
Syn-eruptive melt extraction canological perspectives of caldera-forming eruptions. Here we explore the implications of complex magma res-
ervoir configurations for eruption dynamics and caldera formation. We first examine mafic systems, where
stacked-sill models have long been invoked but which rarely produce explosive eruptions. An exception is the
2010 eruption of Eyjafjallajökull volcano, Iceland, where seismic and petrologic data show that multiple sills at
different depths fed a multi-phase (explosive and effusive) eruption. Extension of this concept to larger mafic
caldera-forming systems suggests a mechanism to explain many of their unusual features, including their
protracted explosivity, spatially variable compositions and pronounced intra-eruptive pauses. We then review
studies of more common intermediate and silicic caldera-forming systems to examine inferred conditions of
magma storage, time scales of melt accumulation, eruption triggers, eruption dynamics and caldera collapse.
By compiling data from large and small, and crystal-rich and crystal-poor, events, we compare eruptions that
are well explained by simple evacuation of a zoned magma chamber (termed the Standard Model by Gualda
and Ghiorso, 2013) to eruptions that are better explained by tapping multiple, rather than single, melt lenses
stored within a largely crystalline mush (which we term complex magma reservoirs). We then discuss the impli-
cations of magma storage within complex, rather than simple, reservoirs for identifying magmatic systems with
the potential to produce large eruptions, and for monitoring eruption progress under conditions where succes-
sive melt lenses may be tapped. We conclude that emerging views of complex magma reservoir configurations
provide exciting opportunities for re-examining volcanological concepts of caldera-forming systems.
© 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction — why review calderas? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29


2. Caldera-forming eruptions: an overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3. Mafic magmatic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1. Stacked sill models of mafic magmatic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2. Caldera formation from complex mafic magma reservoirs — an example from Colli Albani . . . . . . . . . . . . . . . . . . . . . . . . 32
4. Storage and eruption from large silicic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.1. Pre-eruptive magma storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.2. Time scales of magma accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.3. Eruption triggers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.4. Eruption dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.5. Caldera collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5. Implications for recognizing eruption potential of large magmatic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

⁎ Corresponding author at: University of Bristol, School of Earth Sciences, Bristol BS8 1RJ, UK.
E-mail address: glkvc@bristol.ac.uk (K.V. Cashman).

http://dx.doi.org/10.1016/j.jvolgeores.2014.09.007
0377-0273/© 2014 Elsevier B.V. All rights reserved.
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 29

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
41
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

“However, there is a pleasure in recognizing old things from a new point crystals or crystal plumes within a large batch of liquid that cools from
of view.” Richard Feynman, 1948 the margins inward; (2) eruption initiation by injection of a vertical
and pressurized dyke, located either in an axisymmetric position or at
the chamber margin; (3) magma withdrawal starting from the top of
the melt lens and propagating downward, as evidenced by deposits
1. Introduction — why review calderas? that are reversely zoned in composition and/or pre-eruptive tempera-
ture and pressure; and (4) caldera formation by collapse of an under-
The importance of studying caldera-forming eruptions cannot be pressured magma chamber after some fraction of magma has been
under-estimated. Caldera-forming eruptions include some of the largest withdrawn, with the timing of collapse determined by the strength
volcanic events ever to affect the Earth. Many of these have produced and thickness of the overlying country rock relative to width of the
large volumes (N100 km3) of highly evolved crystal-poor melt. As a re- magma chamber.
sult, an enduring paradigm of both igneous petrology and volcanology Over the past few decades, however, detailed volcanological, petro-
has been one of melt accumulation, evolution and eruption from a single logical and geophysical studies of individual (intermediate–silicic) mag-
melt-dominated magma chamber. This conceptual metaphor has pro- matic systems have shown that (1) magma storage regions are
vided a framework for petrological models of magma evolution and dif- composed primarily of crystalline mush (crystals plus interstitial liquid;
ferentiation, and for volcanological models of eruption initiation, Fig. 2; e.g., Hildreth, 2004; Hildreth and Wilson, 2007; Lipman, 2007;
magma withdrawal and caldera collapse (Fig. 1). Key features include: Bachmann and Bergantz, 2008; Reid, 2008; Bachmann, 2010; Deering
(1) development of stably zoned magma chambers by crystal fraction- et al., 2011; Walker et al., 2013; Simon et al., 2014), (2) large melt vol-
ation, where crystal-liquid separation is driven by settling of individual umes may be assembled rapidly (Charlier et al., 2005; Wilson and
Charlier, 2009; Druitt et al., 2012; Allan et al., 2013; Gebauer et al.,
2014; Simon et al., 2014; Wotzlaw et al., 2014), (3) caldera-forming

Fig. 2. End member models of caldera-producing magmatic systems. (A) Crystal-poor


Fig. 1. The Standard Model of caldera formation. (A) Stably stratified magma chamber (CR), and often zoned, eruptions are fed from a single melt body contained within a
forms over thousands of years by crystal settling and upward migration of volatiles. much larger and more crystalline system comprising a crystal mush (~50% crystals) and
(B) Eruption starts with Plinian activity through a single vent, driven primarily by volatile surrounding rigid sponge (N65% crystals); modified from Hildreth (2004). (B) Crystal-
exsolution. (C) Evacuation of magma causes under-pressurization and destabilization of rich (CR) eruptions occur by rejuvenation of a near-rigid crystal mush (by input of melt
the reservoir. (D) Caldera forms by roof collapse. and/or gas); modified from Bachmann and Bergantz (2006); Huber et al. (2011).
30 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

eruptions may be triggered either internally (‘chamber triggered’) or are most commonly associated with stratovolcanoes, where storage re-
externally (‘roof triggered’) depending on the tectonic setting, accumu- gions tend to be vertically elongated.
lated magma volume and roof aspect ratio (Lindsay et al., 2001; Jellinek The very largest eruptions almost always involve MI magma that is
and DePaolo, 2003; de Silva et al., 2006; Gudmundsson, 2008, 2012; (relatively) homogeneous in both composition and phenocryst content
Gottsmann et al., 2009; Gregg et al., 2012), and (4) large eruptions (Fig. 4A). Resulting ignimbrites are typically composed of ash and bro-
may tap multiple and distinct melt sources and crystal populations (in- ken crystals, with only scarce pumice clasts (b10%; e.g., Carter et al.,
cluding phenocrysts, antecrysts and xenocrysts, e.g., Maughan et al., 1986; Bachmann et al., 2002; Gottsmann et al., 2009; Wright et al.,
2002; Ellis et al., 2010; Wright et al., 2011; Allan et al., 2012; Cooper 2011). CPR eruptions, in contrast, initiate with pumice-bearing deposits
et al., 2012; Ellis and Wolff, 2012; Vinkler et al., 2012; Gualda and of crystal-poor (and often high-SiO2) rhyolite magma that may vary in
Ghiorso, 2013; Fig. 3). At the same time, geophysical studies of active both crystallinity and bulk composition throughout the course of an
volcanic systems have failed to locate large volumes of crystal-poor eruption (e.g., Hildreth, 1981; Deering et al., 2011; Pamukcu et al.,
melt (e.g. Iyer et al., 1990; Masturyono et al., 2001; Schilling and 2013; Fig. 4B). The high-SiO2 rhyolite melt often lacks a counterpart in
Partzsch, 2001; Sherburn et al., 2003; Zandt et al., 2003; Lees, 2007; correlative plutonic sequences (although there are exceptions,
Chu et al., 2010). These observations are difficult to reconcile with a e.g., Walker et al., 2007), but overlaps the composition of matrix melts
classical magma chamber model sensu stricto, that is, a single, very in MI magmas (e.g., Lindsay et al., 2001; Lipman, 2007; Bachmann
large, long-lived melt-dominated magma chamber (termed the Stan- et al., 2007; Fig. 4). For this reason, both magma types are inferred to
dard Model by Gualda and Ghiorso, 2013). have a similar origin, with the difference being that MI eruptions evac-
Here we review explosive caldera-forming eruptions and their prod- uate the entire (rejuvenated) magma storage region, while CPR erup-
ucts, paying particular attention to new conceptual models of complex tions are dominated by the (segregated) melt phase (e.g., Bachmann
magma storage regions. By “complex” we refer to magma reservoirs et al., 2007). Accumulation of rhyolitic residual melt prior to CPR erup-
comprising multiple melt lenses within a partially to completely crystal- tions is inferred to occur by crystal settling, compaction and/or filter
line framework; such reservoir configurations have been advanced by pressing of the crystal mush (e.g., Sisson and Bacon, 1999; Bachmann
recent petrological studies of mafic systems, but are consistent with and Bergantz, 2004; Bea, 2010; Dufek and Bachmann, 2010).
emerging views of some silicic systems (e.g., Fig. 3). We take this view MI and CPR eruptions differ in eruption style and timing of caldera
one step further by exploring the implications of complex magma reser- formation. MI eruptions commonly lack an early Plinian (single vent,
voir geometries for syn-eruptive melt extraction, eruption dynamics high plume) phase and initiate, instead, with eruption of pyroclastic
and caldera collapse. We end by examining the implications of different density currents from faults along the caldera margin. Where the timing
magma storage models for hazard assessment, including geophysical of caldera collapse can be determined, it is coincident with the start of
‘prospecting’ for magma reservoirs capable of producing very large the eruption (Sparks et al., 1985; Lindsay et al., 2001; de Silva et al.,
eruptions. 2006; Gregg et al., 2012; Willcock et al., 2013). As a result, associated
distal ash deposits derive mostly (or exclusively) from the co-
ignimbrite plume (e.g., Chesner et al., 1991). CPR eruptions, in contrast,
2. Caldera-forming eruptions: an overview typically start with a Plinian (high plume) phase, as recorded in wide-
spread and voluminous fall deposits. With time, the vent widens
Excellent reviews of magmatic systems that produce large silicic (often by propagating ring faults) and pyroclastic flows comprise an in-
caldera-forming eruptions are provided by Hildreth (1981, 2004) and creasing proportion of the erupted volume. Caldera collapse occurs only
Lipman (2007); here we summarize some pertinent points from these after withdrawal of a critical volume of magma that can be related to the
and related studies. Large (≥100 km3 DRE) silicic eruptions can be clas- depth and geometry of the reservoir (e.g., Roche and Druitt, 2001; Geyer
sified by dominant magma type as either crystal-rich (≥ 35%) dacite et al., 2006; Geshi et al., 2014).
(often termed monotonous intermediates, MI) or crystal-poor (≤ 15%) Smaller caldera-forming eruptions (≤~100 km3) encompass a wide
rhyolite (CPR). Magma within these systems is typically stored in sill- range of magma compositions and crystallinities (e.g., Hildreth, 1981)
like bodies, that is, they have horizontal dimensions that greatly exceed and are typically associated with stratovolcanoes. When classified by
the vertical dimension. Smaller (b 100 km3) caldera-forming eruptions matrix glass (rather than bulk) compositions, these eruptions can be
assigned to one of three groups: rhyolite (SiO2 N 70%), intermediate
(phonolite/trachyte; 55% b SiO2 b 70%), or mafic (ultrapotassic;
SiO2 b 55%; Fig. 4B). These melt compositions are often buffered at
pseudo-invariant points (Fowler et al., 2007; Boari et al., 2009; Gualda

Fig. 4. Bulk rock and matrix glass compositions of CP and CR ignimbrites as a function of
Fig. 3. Schematic view of the magmatic system that fed the very large (~1200 km3) rhyo- DRE erupted volume. (A) CR data; yellow squares show matrix glass, orange bars show
litic Kidnappers eruption, Mangakino volcano, Taupo Volcanic Zone, New Zealand. The bulk compositional range. (B) CP data; blue circles show matrix glass composition of ear-
eruption tapped three separate melt bodies distributed laterally along the rift. liest erupted magma, blue bars show bulk compositional range. Data sources are listed in
Modified from Cooper et al. (2012). Table S1.
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 31

et al., 2012). By analogy with the larger systems, it is commonly as- counterparts, with an early fallout phase from a central single vent
sumed that evolved crystal-poor (CP) magma is segregated into a single followed by climactic, syn-collapse highly mobile ignimbrites with asso-
large body prior to eruption, and that late-erupted crystal-rich (CR) ciated proximal collapse breccias (e.g. Giordano and Dobran, 1994;
magma is mobilized by recharge melts (e.g., Bacon and Druitt, 1988; Watkins et al., 2002). As is typical for other stratovolcano eruptions,
Pallister et al., 1992; Allen, 2001; Bachmann, 2010). When crystal-rich none of these caldera complexes has undergone post-collapse resur-
magmas are erupted early (for example, Pinatubo 1991, Philippines; gence; instead, post-collapse subsidence, intracaldera volcanism and
Quilotoa 800 ybp, Ecuador), erupted magmas appear similar to larger sedimentation are common (Giordano et al., 2006; Acocella et al., 2012).
MI eruptions in their (general) homogeneity, bulk composition and
high crystallinity (e.g., Polacci et al., 2001; Rosi et al., 2004). Caldera col- 3. Mafic magmatic systems
lapse in these systems is attributed to magma withdrawal and under-
pressurization, and may happen at some point during the eruptive se- To explore relations between conditions of magma eruption and
quence (e.g., Druitt and Sparks, 1984). storage, we first briefly review mafic and mafic ultrapotassic magmatic
Of the caldera-forming ignimbrite family, the smallest, and in many systems, where both recent observations of eruptions (Tarasewicz et al.,
ways the oddest, group is that of ultrapotassic (SiO2 b 55%) ignimbrites 2012; Carbotte et al., 2013) and detailed petrological (e.g., Marsh, 2013;
found primarily in the Quaternary Roman Magmatic Province (QRMP; Neave et al., 2013) and volcanological (e.g., Brown et al., 2014; Vinkler
Italy). The QRMP comprises four major caldera complexes that have et al., 2012) studies all point to syn-eruptive, and sometimes explosive,
produced recurrent eruptions of tephritic to tephri-phonolitic ignim- tapping of multiple melt lenses stored within complex magma plumb-
brites with DRE (dense rock equivalent) volumes of 1–50 km3 (Fig. 5; ing systems. Geochemical data suggest, moreover, that these melt
Giordano et al., 2006; Boari et al., 2009; Masotta et al., 2010; Freda lenses may be vertically distributed throughout the crust. In fact, isoto-
et al., 2011; Vinkler et al., 2012; Acocella et al., 2012). The recurrence pic evidence for crustal assimilation of flood basalt magmas (e.g., Wolff
times for caldera-forming eruptions at each caldera complex are 40 to et al., 2008; Vye-Brown et al., 2013) requires that very large volumes of
50 ka, and caldera areas range from 30 to 300 km2. These eruptions mafic melt accumulate within the crust prior to eruption.
are curious because ultrapotassic magmas have low viscosities
(≤104.5 Pa s, even accounting for up to 30% syn-eruptive crystallization; 3.1. Stacked sill models of mafic magmatic systems
Vinkler et al., 2012; Campagnola, 2014) and are therefore susceptible
to gas escape during magma ascent. The eruption sequence of Mafic magmatic systems are commonly envisioned as sequences of
ultrapotassic ignimbrites is, however, comparable to that of their silicic melt lenses or stacked sills within largely to completely crystalline

Fig. 5. The Quaternary Roman Magmatic Province. (A) Main caldera-producing centers. (B) The Colli Albani volcano with the extent of the 355 ka Villa Senni (VSN) caldera-forming erup-
tion unit: black lines = isopachs of the basal scoria fall (shown in yellow); orange = Tufo Lionato pre-collapse ignimbrite; pink = Pozzolanelle climactic ignimbrite. (C) Chemical com-
position of VSN units: orange field = Tufo Lionato (from Freda et al., 1997); pink squares = Pozzolanelle ignimbrite (from Conticelli et al., 2010); purple circles = spatter in caldera
collapse breccia (from Conticelli et al., 2010).
32 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

zones (e.g., Marsh, 1996; Annen et al., 2006; Gudmundsson, 2012). Sills 2012; Neave et al., 2013; van der Zwan et al., 2013). In these conceptual
form when repeated magma injections are sufficiently spaced in time to models, each thermally zoned mafic melt lens evolves by the progres-
allow complete cooling between injection events (e.g., Annen et al., sion of a solidification front that is multiply saturated, such that melt
2006; Burchardt, 2008); melt lenses dominate when new inputs are composition is buffered along a eutectic/cotectic. At the same time, crys-
added to systems that are still partially molten, or within sills that are tallization generates highly differentiated melt compositions that may
sufficiently thick to allow internal redistribution of melt (e.g., Marsh, (1) segregate by compaction at intermediate crystallinities, (2) become
2002, 2013). Variants of stacked sill models have been invoked to ex- trapped within small pores in regions of high crystallinity (N 70%;
plain the petrologic diversity of mafic magma erupted during individual, e.g., Dufek and Bachmann, 2010), or (3) segregate into lenses, pockets,
often long-lived, eruptive episodes (e.g., Aarnes et al., 2008; Kelley and and bulbous masses (Marsh, 2002; Masotta et al., 2012).
Barton, 2008; Erlund et al., 2010; Dahren et al., 2012; Passmore et al., The volcanological consequences of stacked sill models have not
been thoroughly explored. Direct evidence for syn-eruptive tapping of
previously intruded sills is provided by precursory and syn-eruptive
geophysical data for the remarkable pattern of downward-
propagating seismicity that accompanied the 2010 Fimmvörðduháls-
Eyjafjallajökull eruption in Iceland (Tarasewicz et al., 2012; Fig. 6). Evi-
dence for at least two sill intrusions in the 1990s (Sigmundsson et al.,
2010) provides support for this interpretation. An important conse-
quence was an eruption that involved several distinct explosive epi-
sodes separated by times of lava effusion; another consequence was
a wide range of erupted compositions and groundmass textures
(e.g., Cioni et al., 2014). This well documented example shows that
melt does not have to be assembled pre-eruptively into a single large
body to contribute to a single eruptive episode, and supports previous
interpretations of individual mafic–intermediate eruptions fed from
complex reservoirs (e.g., Yoshimoto et al., 2004; Roman et al., 2006;
Erlund et al., 2010; de Angelis et al., 2013; Neave et al., 2013). These
data also show that changes in eruptive activity (such as pauses and
transitions between explosions and lava effusion) may reflect not only
conditions of magma transport to the surface (e.g., Melnik et al., 2005)
but also conditions of magma storage, including the over-pressure
maintained within the magma reservoir. For example, pressure within
individual melt lenses may be modulated dynamically by the interplay
between gravitational instability of the solidification front (Marsh,
2002; Humphreys and Holness, 2010), local gas build-up caused by
volatile-saturated crystallization (Tait et al., 1989), changes in melt vol-
ume caused by the balance between crystallization and interstitial melt
segregation (e.g. Sisson and Bacon, 1999; Bachmann and Bergantz,
2004), compressibility of magma with an exsolved volatile phase
(Johnson, 1987; Voight et al., 2010), and/or intrusion of ‘recharge’
melt and/or gas from deeper melt lenses.

3.2. Caldera formation from complex mafic magma reservoirs — an


example from Colli Albani

The Eyjafjallajökull eruption produced a small volume of magma,


was only moderately explosive, and did not produce a caldera. Could a
magma system comprising multiple melt lenses produce a large,
caldera-forming eruption? To address this question we examine the
mafic ultrapotassic volcano Colli Albani (QRMP; Fig. 5), which has pro-
duced several large (30–50 km3) caldera-forming eruptions (Giordano
and the CARG team, 2010). Best characterized is the 355 ka Villa Senni
eruption (VSN; Vinkler et al., 2012; Figs. 5, 7). The architecture of the
VSN deposit is similar to those of many silicic caldera-forming erup-
tions, with a basal fall deposit of 0.3 km3 (all volumes in dense rock
equivalent, or DRE) followed by the N 10 km3 tephri-phonolitic,
crystal-poor Tufo Lionato ignimbrite (VSN1). VSN1 is overlain by the
N20 km3 Pozzolanelle ignimbrite (VSN2), which is tephri-phonolitic to
tephritic in composition and (mostly) crystal-rich (Watkins et al.,
2002). Evidence of caldera collapse can be found in the sudden appear-
ance between VSN1 and VSN2 of an intercalated co-ignimbrite breccia
with up to 30% deep-seated lithics (from ~1 to 6 km depth, including in-
trusives and cumulates). The breccia also contains an isotopically dis-
tinct K-foidite juvenile magma that does not appear elsewhere in the
Fig. 6. Time-progressive unloading of a stacked sill sequence beneath Eyjafjallajökull vol-
cano, Iceland, 2010; illustrates response to a downward propagating decompression front
VSN stratigraphy (Conticelli et al., 2010). The deposit characteristics,
of a partially molten crustal magmatic system. as well as the ubiquitous vapor-phase lithification of the lower VSN1 ig-
Redrafted from Tarasewicz et al. (2012). nimbrite, suggest that a (short?) hiatus preceded caldera collapse.
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 33

Fig. 7. Stratigraphic and textural characteristics of VSN (compiled from Vinkler et al., 2012); color scheme used is the same as in Fig. 5. Juvenile types (by number) refer to Fig. 8. Reference
lines for vesicularity and vesicle number density (VND) are from Rust and Cashman (2011). Arrow on right hand side labeled f represents the volumetric ratio of pre-collapse and post-
collapse deposits.

One unusual feature of the eruptive deposit relates to the pyroclast the ejection of both cognate cumulates and loose cumulate crystals
vesicularities, which are generally very low (b~40% bulk, and b ~60%, (up to 35% of mm- to cm-sized leucite, clinopyroxene and biotite;
melt-referenced; Fig. 7), and contrast with typical crystal-poor silicic Fig. 8c, d) during and after caldera collapse. Incorporation of cumu-
pumice vesicularities of N 70% (e.g., Klug et al., 2002; Gurioli et al., late crystals into the erupted melt explains the observed change in
2005; Adams et al., 2006; Houghton et al., 2010). At the same time, ves- bulk composition from tephri-phonolite to phono-tephrite, and
icle number densities of ≤ 108 cm3 reach those of other mafic plinian suggests that progressive disruption of the crystal framework
eruptions (Vinkler et al., 2012). These vesicle characteristics suggest ex- accompanied caldera collapse. Additionally, eruption of K-foidite in
tensive syn-eruptive outgassing, although concomitant with high rates association with syn-caldera breccias shows that at least one isolat-
of magma decompression. And yet tens of cubic kilometers of low vis- ed melt lens was tapped at the time of caldera formation.
cosity magma were erupted explosively to form two separate, large- The data reviewed above – low pyroclast vesicularity, time gaps,
volume ignimbrites. This apparent conundrum strongly suggests that multiple explosive episodes, variable melt and bulk compositions –
volatile exsolution is not necessarily the primary driving force for erup- are difficult to reconcile with magma extraction from a single, pressur-
tive activity. ized, well mixed, melt-dominated magma chamber. Instead, the
At the same time, geochemical examination of the eruptive prod- eruptive sequence has many elements that suggest involvement of a
ucts shows that extensive (≤ 60%) crystallization of clinopyroxene complex magma reservoir composed of both laterally and vertically dis-
and leucite are required to produce Colli Albani magma (Freda tributed melt lenses within a variably crystalline mush (Fig. 9). Most im-
et al., 1997, 2011; Peccerillo, 2005; Boari et al., 2009). This implies portantly, both the explosive nature of the eruptive activity and the
that the bulk of the magma storage system was highly crystalline, poorly vesicular scoria appear to require sustained overpressures to
even though both early-erupted (pre-collapse) and very evolved drive the eruption. We suggest that a complex magma reservoir can ex-
(syn-collapse foidite) magma are crystal-poor (Fig. 8a, b). Further plain these observations if overpressure is accommodated within indi-
support for a largely crystalline magma reservoir can be found in vidual melt lenses (e.g., Bagdassarov and Dorfman, 1998). In this
34 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

Fig. 8. Photomicrographs of component types in VSN deposit (labeled by number to correspond with labels in Figs. 6 and 8). (1) Crystal-poor scoria with round vesicles.
(2) Microphenocryst-rich scoria with irregular vesicles. (3) Phenocryst-rich scoria with microcrystalline matrix and highly irregular vesicles. (4) Cognate xenolith containing crystals of
leucite, clinopyroxene and biotite. All images are 3.3 mm across.

scenario, elastic relaxation of the crystalline framework could maintain magma reservoir model, time gaps reflect modulation of magma with-
a sufficiently high driving pressure to sustain fast withdrawal of even drawal by the strength of the crystal framework (Fig. 10), and the
volatile-poor magma. As a consequence, sustained melt extraction time required for connection of melt lenses to the main conduit
would not require volatiles to provide the only driving force for erup- (e.g., Fig. 6).
tion. Instead, as for oil or water extraction from an over-pressured res- To summarize, magma storage within a complex melt/mush reser-
ervoir, melt out-flow could be driven initially by release of stored voir helps to explain many unusual characteristics of the Villa Senni
deformational energy within the (bubble-bearing?) magma, the (visco- eruption, including the stability afforded by distributing, rather than
elastic) crystalline parts of the reservoir, and the country rock. Vesicula- concentrating melt, the potential for isolated lenses to develop both
tion triggered by magma ascent and decompression would then high overpressures and highly evolved compositions, and the opportu-
enhance magma ascent, particularly at low pressures. In this complex nity for sequential tapping of melt lenses to sustain explosive activity

Fig. 9. Conceptual model for the VSN magma reservoir, which is composed of CP melt lenses within a CR magma matrix. Melt lenses are variably interconnected prior to eruption; during
eruption, isolated melt pockets may be tapped either as intervening septa are ruptured or when intercepted by propagation of the caldera-bounding fault. During eruption, magma flows
initially from within CP lenses, where viscosity is lower, and progressively involves larger portions of the reservoir eventually including the crystal rich framework. The enlarged box to the
left depicts the possible geometry of an upper and a lower solidification front with embedded melt accumulation. Numbers 1 to 4 are the same as the zoned juvenile types in Fig. 7, and
show the inferred original positions of magma parcels erupted sequentially in the Villa Senni ignimbrite succession, from crystal poor type (1) to progressively crystal richer types (2, 3, 4).
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 35

in laterally extensive (rather than vertically elongated) reservoirs


(e.g., Fig. 3).
Monotonous (crystal-rich) ignimbrites (MI) may also preserve evi-
dence of melt segregation. In fact, although the very name “monoto-
nous” denotes homogeneity, careful examination of some large MI
deposits has shown that these systems may also be spatially heteroge-
neous. The best-documented example is the Lund Tuff (USA), where
studies of individual pumice clasts show that the erupted magma was
inhomogeneous in temperature, phenocryst proportions, and mineral
compositions (Maughan et al., 2002). Compositional heterogeneity has
also been documented in the Cerro Galan ignimbrite, NW Argentina,
where different sectors of the ignimbrite outflow sheet have distinct
compositional characteristics (Francis et al., 1989; Kay et al., 2011),
and compositionally distinct white and gray pumice clasts provide evi-
dence for at least two magmas involved in the eruption (Folkes et al.,
Fig. 10. Changes in magma strength as a function of melt volume fraction. Dashed lines are 2011; Wright et al., 2011). In fact, discrete evolved melt pockets are in-
fit to experimental data on the Westerly granite (upper) and Delegate aplite (lower). ferred even for systems that lack evidence for diverse melt compositions
Adapted from Rosenberg et al. (2007).
(e.g., Huber et al., 2012; Willcock et al., 2013).
Compositionally zoned ignimbrites are perhaps most representative
throughout an eruptive sequence. Deposits from the VSN eruption fur- of the Standard Model, in that they are interpreted to record top-down
ther suggest that the timing and style of caldera collapse may be con- evacuation of individual magma chambers. They also demonstrate the
trolled by processes within the reservoir (specifically, failure of parts fundamental role of mafic magma in providing either heat (to partially
of the crystal network), in addition to processes external to the reservoir melt roof rocks) or evolved melt (from cooling and crystallization). Ex-
(such as the geometry, thickness and mechanical properties of the roof amples of the former include Gran Canaria (Freundt and Schmincke,
rock and caldera faults). More broadly, we highlight emerging views of 1995) and Iceland (Askja, Sigurdsson and Sparks, 1981), where mafic
mafic magma reservoirs as vertically extensive and comprising both inputs are hot and water-poor. Examples of the latter are common in
melt-rich and melt-poor (or melt-absent) regions. This view derives water-rich environments, where (often cooler) crystal-poor silicic
not only from the geophysical, petrologic and volcanologic studies magma overlies hotter (often more crystalline) mafic magma
reviewed above, but also from new thermal models that examine condi- (e.g., Hildreth, 1981; Bacon and Druitt, 1988; Druitt and Bacon, 1989;
tions required to develop complex storage regions (e.g., Annen et al., Deering et al., 2011; Pamukcu et al., 2013). As both lower temperatures
2006; Annen, 2011; Solano et al., 2012). More important from a and lower PH2O promote crystallization (Blundy and Cashman, 2008;
volcanological perspective, however, are the implications for conditions Cashman and Blundy, 2013), the lower temperature of the dominant
leading to, and evolving during, volcanic eruptions from magma crystal-poor rhyolite requires commensurate zoning in water unless
reservoirs that contain multiple and variably connected melt lenses compositional differences are large. Zoned magmas are particularly
(e.g., Gudmundsson, 2012). characteristic of (although not unique to) vertically elongated magmat-
ic systems that feed arc stratovolcanoes, perhaps because vertically
4. Storage and eruption from large silicic systems elongated (and water-rich) reservoirs are less susceptible to convective
mixing than sill-like melt bodies (e.g., Maughan et al., 2002; Blundy and
We now address the question of the extent to which complex (melt- Cashman, 2008).
lens-dominated), as compared to simple (single melt body), magma A more unusual example of a zoned magma reservoir is provided by
reservoirs can provide insight into processes that contribute to the the Novarupta–Katmai eruption of 1912, which involved 7.5 km3 of
much more common eruptions of intermediate to silicic magmas. Our crystal-poor high-silica rhyolite and 5.5 km3 of crystal-rich continuously
goal is not to dismiss the Standard Model, but instead to evaluate the ex- zoned dacite to andesite magma. The unusual aspect of the eruption is
tent to which emerging, and sometimes conflicting, observations about not the compositional zonation but the observed caldera collapse at
very large explosive eruptions can be reconciled by broadening our Mt. Katmai, which lies 10 km from the eruptive vent of Novarupta. De-
views of magmatic systems. tailed studies by Hildreth and Fierstein (2012) demonstrate that the
rhyolite was extracted from the intermediate composition magma and
4.1. Pre-eruptive magma storage that extraction and lateral transport of crystal-poor rhyolite from the
storage region beneath Mt. Katmai was necessarily rapid, perhaps oc-
Several recent studies of crystal-poor rhyolitic ignimbrites suggest curring during the 5 days of recorded precursory activity. This raises
that multiple, rather than single, melt batches were tapped during indi- some interesting questions. First, to what extent does zoning observed
vidual caldera-forming eruptions. Evidence for multiple melt batches is in ignimbrites provide direct evidence of zoning within magma reser-
particularly common in extensional environments such as the Snake voirs? Second, how can large amounts of rhyolite melt be extracted
River Plain (US; Ellis et al., 2010; Ellis and Wolff, 2012) and the Taupo from crystal mush zones both efficiently (without accompanying crys-
Volcanic Zone (TVZ, New Zealand; Brown et al., 1998; Charlier et al., tals) and rapidly (in days)? An alternative suggestion is that the rhyolite
2003; Gravley et al., 2007; Wilson and Charlier, 2009; Bégué et al., magma intruded as an ascending rhyolite dike that intersected the res-
2014), and demonstrates the importance of crustal forcing on both ident intermediate composition magma reservoir under Mount Katmai
magma storage and eruption (e.g., Lindsay et al., 2001; Gottsmann (Eichelberger and Izbekov, 2000). Neither model provides a good expla-
et al., 2009; Cooper et al., 2012; Allan et al., 2013). Two laterally nation for the lateral displacement of the eruptive vent from the magma
displaced (and non-communicating) melt lenses may also have been storage region.
tapped during the 600 km3 Bishop Tuff eruption from the Long Valley
caldera (Gualda and Ghiorso, 2013), which lies within a trans- 4.2. Time scales of magma accumulation
tensional setting at the eastern edge of the Sierra Nevada and has long
been considered the iconic example of the Standard Model of a single The time scale of magma accumulation prior to large eruptions has
zoned magma chamber. In all cases, melt lenses are similar in bulk, been the subject of numerous recent studies that apply new diffusion
but distinct in trace element and isotopic, composition and were stored chronometers to observed crystal zoning patterns. A surprising result
36 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

of these studies is the suggestion that large volumes of silicic melt may evolved than the resident magma in the uppermost part of the system
accumulate over short (centuries or less) time scales (e.g., Charlier (Eichelberger et al., 2006; Hildreth and Wilson, 2007; Wright et al.,
et al., 2007; Gualda et al., 2012; Allan et al., 2013; Fig. 11). Other studies 2011). The geometry of the magma reservoir may also control both
describe very short time scales (decades to years or even months) of the extent of interaction of new hotter melt inputs with cooler mush
late-stage crystallization (e.g., Wark et al., 2007; Saunders et al., 2010; (e.g., Humphreys et al., 2009) and the interaction of volatiles with the
Druitt et al., 2012; Gualda et al., 2012; Matthews et al., 2012) and/or in- mush (Wright et al., 2012).
corporation of xenocrysts (e.g., Gardner et al., 2002) prior to large erup- Late-stage disturbances to magmatic systems may also be recorded
tions. Short timescales of magma accumulation are difficult to reconcile as selective crystal dissolution (e.g., feldspar and/or quartz; Bachmann
with long times (104–105 years) required for compaction-driven melt et al., 2002), phenocryst rim growth (Wark et al., 2007; Saunders
extraction (e.g., McKenzie, 1984; Bachmann and Bergantz, 2004; Dufek et al., 2010; Druitt et al., 2012; Matthews et al., 2012; Allan et al.,
and Bachmann, 2010). An alternative mechanism for segregating silicic 2013) or microlite formation (Pamukcu et al., 2012). Both dissolution
magma invokes formation of melt channels or dikes within more crystal- and phenocryst rim growth are commonly interpreted to reflect intru-
line parts of the magma reservoir; melt extraction in this scenario is sion of mafic magma into the system; in the latter crystal growth results
driven by either pore pressure response to an anisotropic stress field from cooling of the mafic input. In the absence of evidence for mafic in-
(e.g., Eichelberger et al., 2006; Allan et al., 2013) or rapid inter- puts, however, selective dissolution and new crystal growth can also be
connection of isolated melt lenses (e.g., Eichelberger and Izbekov, explained by changes in PH2O in response to decompression or addition
2000; Fig. 6). Rapid extraction and shallow accumulation of melt may of volatiles (Bachmann et al., 2002; Wark et al., 2007; Blundy and
also be driven by perturbations of local stress fields surrounding crystal Cashman, 2008; Matthews et al., 2012; Cashman and Blundy, 2013). Ev-
mush zones. Perturbations could be caused by the arrival of new idence for volatile transfer underlies the concept of gas sparging, where-
magma inputs, gas exsolution (Sisson and Bacon, 1999), or tectonic by sufficient heat to unlock crystal networks is transferred by an influx
stresses (particularly extension related to rifting; Allan et al., 2012). of volatiles to the system (Bachmann and Bergantz, 2006). The time
scales required for unlocking by heat transfer alone are long and
are similar to those required for melt extraction by compaction
4.3. Eruption triggers (e.g., Gottsmann et al., 2009). Fluxing with H2O-rich fluids could unlock
crystal networks more rapidly, however, by resorbing anhydrous
Critical to understanding caldera-forming eruptions is the consider- phases. Introduction of CO2-rich fluids, in contrast, would promote crys-
ation of processes that cause a stable melt/mush system to destabilize. tal growth, particularly of feldspar (e.g., Cashman and Blundy, 2013),
By definition, an eruption starts when the magmatic system becomes thereby strengthening crystal networks.
connected to the surface. This connection can be established by Early (precursory) phases of eruptive activity provide insight into
upward-propagating dikes driven by magmatic overpressure (an inter- conditions required to initiate and sustain an eruption. Interestingly,
nal trigger), or downward-propagating faults generated by thermo- many eruptions are preceded by ‘leaks’ from the magma reservoir. Ex-
mechanical instabilities in the roof rocks (an external trigger; amples include the eruption of the 200 km3, largely degassed, Pagosa
e.g., Gudmundsson, 2008; Gregg et al., 2012). Upward-propagating Peak dacite just before the very Fish Canyon Tuff eruption (FCT;
dikes are commonly invoked when there is evidence for intrusion of e.g., Bachmann et al., 2000), the explosive-to-effusive Cleetwood erup-
hotter ‘recharge’ melt (and/or volatile phase). Evidence of magma influx tion that preceded the c. 50 km3 caldera-forming (zoned) eruption of
may be preserved in the form of magmatic inclusions, banded pumice or Crater Lake, USA by weeks to months (Bacon, 1983; Kamata et al.,
multiple pumice populations (e.g., Hildreth, 1981; Pallister et al., 1992; 1993), and the small (~ 0.3 km3) explosive eruption that preceded,
Polacci et al., 2001; Rosi et al., 2004). Although most easily recognized probably by months, the 530 km3 (crystal-poor) Oruanui eruption in
when mafic magma is intruded into a silicic system, the intruding New Zealand (Allan et al., 2012). In all three cases, precursory eruptions
magma may be silicic, in which case it is typically hotter and less clearly tapped the primary magmatic system, and yet did not immedi-
ately trigger the climactic eruption. In both the FCT and Oruanui exam-
ples, precursory activity has been linked to tectonism in the form of
block faulting (FCT) or rifting (Oruanui), with the latter inducing lateral
melt migration from an isolated melt lens. The dynamics of the
Cleetwood eruption have not been explained, although Crater Lake
also lies within an extensional region (Bacon et al., 1999) and may
have been subject to tectonic stressing.
What, then, causes transitions from precursory activity to climactic
eruptions? Interestingly, the crystal-rich FCT preserves evidence of
pre-eruptive crystal breakage interpreted to record rapid decompression
of the magma storage region (and explosive expansion of phenocryst-
hosted melt inclusions) during either the Pagosa Peak eruption or
early ignimbrite eruptions from the southern part of the FCT caldera
(Lipman et al., 1997). Pre-eruptive crystal breakage may have been nec-
essary to fully mobilize magma from the crystal-rich reservoir. Pre-
eruptive crystal breakage may also occur in response to migration of re-
charge melt through overlying crystal mush (e.g., Pallister et al., 1992),
as illustrated by the association of broken crystals with a geochemically
distinct and partially degassed magma in the crystal-rich Cerro Galan ig-
Fig. 11. Time constraints on melt accumulation prior to the Oruanui eruption, Taupo nimbrite (Wright et al., 2011). Heating accompanying melt migration
Volcanic Zone, New Zealand. Orange curve represents the time required to construct the can also cause crystal rupture by volatile expansion within melt inclu-
primary magma reservoir (high-Si rhyolite), using Fe–Mg interdiffusion in orthopyroxene sions (Gualda et al., 2004; Bindeman, 2005). Finally, it has been sug-
(opx). Purple curve represents late-stage re-equilibration of opx incorporated into the gested that crystal breakage could be a response to seismic shaking
main magma body. Green inset curve represents the distribution of opx diffusion ages
within low-Si rhyolite magma, which is interpreted as growth in isolated melt pockets
(Gottsmann et al., 2009). In all cases, physical breakage of the crystal
that were intersected during the eruption. framework would help to mobilize magma preparatory to the climactic
Complied from Allan et al. (2013)). event.
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 37

4.4. Eruption dynamics horizontally propagating yield surfaces (Karlstrom et al., 2012) and
around subsiding caldera blocks (Kennedy et al., 2008). These processes
The triggering event can also determine the nature of initial eruptive would promote extensive syn-eruptive mixing, which could explain the
activity. For example, very large MI eruptions inferred to be triggered by apparent dichotomy between the broad homogeneity of MI pyroclast
roof collapse (e.g., Jellinek and DePaolo, 2003; Gregg et al., 2012; de hand samples and the extreme complexity commonly recorded within
Silva and Gregg, 2014) lack an early single vent (Plinian) phase phenocryst populations. Alternatively, mobilization of crystal-rich
(e.g., Druitt and Sparks, 1984; Sparks et al., 1985; de Silva et al., 2006; magma is common in late stage (syn- or post-collapse) eruptive activity
Cas et al., 2011; Chesner, 2012). They initiate instead with eruption of from (inferred) vertically zoned magma reservoirs (e.g., Bacon and
poorly expanded pyroclastic density currents along bounding ring faults Druitt, 1988; Druitt and Bacon, 1989; Deering et al., 2011; Hildreth
(Willcock et al., 2013) that are sustained by very high mass fluxes (Cas and Fierstein, 2012; Pamukcu et al., 2013), where eruptions are well ex-
et al., 2011; Lesti et al., 2011). In contrast, (often smaller) eruptions of plained by the Standard Model of top-down magma withdrawal
crystal-poor magma typically have protracted single vent phases prior (e.g., Bacon and Druitt, 1988; Allen, 2001; Mandeville et al., 2009). In
to caldera collapse (e.g., Crater Lake, USA (Bacon, 1983); AD 161 these examples, however, crystals are not pervasively broken.
Taupo, New Zealand (Wilson and Walker, 1985); Minoan Santorini, The scarcity of pumice clasts in MI ignimbrites makes it difficult to
Greece (Druitt and Bacon, 1989; Sparks and Wilson, 1990); 39 ka establish details of magma extraction. Deposits from crystal-poor rhyo-
Campanian Ignimbrite, Italy (Rosi et al., 1999)). Initial vents may be lite eruptions, in contrast, often produce abundant pumice that allows
located either on marginal ring faults when eruptions tap large sills individual parcels of magma to be related directly to their pre-
(e.g., Hildreth and Mahood, 1986), or the summits of stratovolcanoes. eruptive storage conditions. Most informative are detailed studies of
In these eruptions, the mass eruption rate probably increases with phenocryst-hosted melt inclusions, which preserve dissolved volatiles
time, particularly after caldera collapse, and pressure changes within that can be used to estimate entrapment pressures, and major and
the magma reservoir may be preserved in pyroclast textures trace element compositions that can be used to track magma evolution.
(e.g., Bacon, 1983; Klug et al., 2002; Gurioli et al., 2005). From a volcanological perspective, an interesting observation is that
The mode of eruption will also affect the nature of the eruptive prod- melt inclusion studies often indicate magma extraction from a large
ucts. Ignimbrite deposits from large MI eruptions contain mostly ash pressure range, even very early in the eruptive sequence (e.g., Wallace
and pervasively shattered individual crystals, with limited abundance et al., 1999; Liu et al., 2006; Mangiacapra et al., 2008; Roberge et al.,
of (often low vesicularity) pumice (e.g., Carter et al., 1986; Bachmann 2013). One explanation for this could be pre-eruptive mixing of crystals
et al., 2002; Gottsmann et al., 2009; Wright et al., 2011). Shattering of from throughout the magma storage region. Alternatively, magma
crystals provides evidence of extensive disruption of magma during ex- could be extracted from a large depth (pressure) range syn-eruptively
traction from the reservoir (e.g., Maughan et al., 2002). When combined by lateral melt migration to a vertically extensive feeder dike.
with the low vesicularity of rare pumice clasts, these textural character- The most thoroughly documented example is that of the Bishop Tuff,
istics suggest that vesiculation played a limited role in the eruption pro- where detailed stratigraphic and volcanological studies of the eruptive
cess, which was probably dominated instead by catastrophic deposits (Wilson and Hildreth, 1997) provide exceptional constraints
decompression (Gottsmann et al., 2009). This scenario is similar to on both the timing and location of magma extraction from the underly-
that outlined above for Colli Albani. At the same time, large shear strains ing reservoir (Fig. 12A). Here quartz-hosted melt inclusions record pres-
may be imposed by magma mobilization along vertically extensive and sures of b100 to N200 MPa throughout the eruption, although a slightly

Fig. 12. Melt inclusion constraints and magma storage and withdrawal of the Bishop Tuff magma, Long Valley. (A) Map of the aerial distribution of Bishop Tuff deposits; right red = Plinian
vent, dark red = ignimbrites erupted from the southeastern margin of the caldera, dark and light blue represent ignimbrites erupted from the northwest and north rim of the caldera,
respectively (modified from Wilson and Hildreth, 1997). (B) Magma storage pressures inferred from melt inclusion data as a function of eruption time. (C) Variations in incompatible
element ratio U/Ce in melt inclusions as a function of eruption time. Colors are coded to the map in (A).
Melt inclusion data from Wallace et al. (1999); Roberge et al. (2013); time constraints from Wilson and Hildreth (1997).
38 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

deeper magma level may have been tapped late in the eruption incoherent when R N ~1.4. Conditions of f = 0, that is, where collapse
(Fig. 12B). Phase equilibrium constraints, in contrast, allow both early is synchronous with the start of an eruption, require coupling between
and late Bishop magma to span the entire (100–250 MPa) pressure shallow, laterally extensive magma chambers and the overlying roof
range (Gualda and Ghiorso, 2013). Here continuous magma extraction rocks (Gregg et al., 2012). Model predictions have been tested experi-
from a large pressure range can be explained by lateral magma supply mentally with analogue “magmas” that are withdrawn steadily from
to vertically extensive feeder dikes located on the caldera margin “chambers”. Fluids used in analogue experiments include air (μ =
(e.g., Gardner et al., 1991). Horizontally directed melt flow is also sug- 10−5 Pa s), water (μ = 10− 3 Pa s), and silicone (μ = 104 Pa s), and
gested by lateral propagation of ring faults during caldera collapse thus span a range of viscosities. All experiments, however, investigate
(Wilson and Hildreth, 1997). Corresponding trace element analyses steady withdrawal of fluid from a single (simple) reservoir; caldera for-
show that late (post-collapse) eruptive activity tapped magma that mation by fluid extraction from complex reservoirs has not yet been
was both more and less evolved than prior to collapse (Fig. 12C), explored.
which suggests late stage involvement of both less evolved melt lenses Variations in f can be evaluated as a function of both eruption mag-
(as also suggested from zircon data; Chamberlain et al., 2014) and more nitude (DRE volume) and eruption type (crystal-rich [CR] or crystal-
evolved matrix melt, the latter perhaps extracted during caldera poor [CP]; Fig. 14A; Table S1 in Supplementary material). These data
collapse. show that f is small (or 0) in large CR eruptions (e.g., Cerro Galan [CG],
La Pacana [LP] and Fish Canyon Tuff [FCT]), and variable (but non-
4.5. Caldera collapse zero) in moderate to large CP eruptions (e.g., Long Valley [LV], Taupo
[TP] and Villa Senni [VSN]). Collapse may occur very late (large f) in stra-
Comprehensive reviews of caldera collapse are provided in Lipman tovolcano eruptive sequences (e.g., Vesuvius [VS], Tambora [TMB] and
(1997); Cole et al. (2005); Acocella (2007); Marti et al. (2008). These re- Aso [AS]), even for CR magma (e.g., Pinatubo [PN]) or substantial
views focus largely on structural controls on caldera collapse styles, a erupted volumes (e.g., AS). This compilation shows that collapse
topic that is beyond the scope of this review. Instead we explore the re- can occur over a wide range of f for similar total erupted volumes
lation between melt storage and caldera formation, as indicated by var- (i.e., after very different volumes of magma extraction from the reser-
iations in the timing of caldera formation within an eruptive sequence. voir), and at similar f for erupted volumes that vary over two orders of
Caldera formation after evacuation of substantial magma volumes has magnitude. Some of this variation can be explained by differences in
been interpreted as a consequence of under-pressurization of the magma chamber geometry (R), but the extreme variability must reflect
magma storage region (Druitt and Sparks, 1984; Scandone, 1990; other controlling factors, including the distribution of melt within a
Martí, 1991; Branney, 1995; Lipman, 1997; Cole et al., 2005). Caldera crystalline reservoir.
collapse coincident with the onset of eruptive activity, in contrast, To relate f directly to conditions of magma withdrawal requires that
suggests that pressurization and pre-eruptive doming caused by shal- we know the thickness (pressure) of the melt lens evacuated during col-
low magma accumulation may trigger collapse of large calderas lapse. If magma is assumed to be withdrawn from a single, sill-like res-
(e.g., Gudmundsson, 2008; Gregg et al., 2012; de Silva and Gregg, ervoir, and if the surface expression of the caldera can be taken as the
2014). Also important is the tectonic stress field, which can control “footprint” of that reservoir, then the total thickness of magma extract-
the location of caldera-bounding faults (e.g., Holohan et al., 2008a). ed during an eruption can be estimated using the erupted (DRE) volume
Models of caldera collapse (e.g., Roche and Druitt, 2001; Geyer et al., and caldera area, a value commonly referred to as the collapse height.
2006; Stix and Kobayashi, 2007; Geshi et al., 2014) typically measure Early studies identified a linear relation between DRE volume and col-
the timing of collapse by f, the fraction of the total (DRE) magma volume lapse area (Smith, 1979; Spera and Crisp, 1981) that suggested a con-
that is erupted prior to the onset of collapse. The value of f can be related stant thickness of magma evacuation characterized many collapse
to the roof rock strength and magma chamber aspect ratio (or roof as- events. This thickness could be related to the tensile strength of the
pect ratio R = thickness/width; Roche and Druitt, 2001; Fig. 13). Col- crust (Walker, 1984; Scandone, 1990). Our re-examination of calculated
lapse is assumed to occur when the roof can no longer support the collapse heights for all sufficiently well characterized eruptions (120 in
(under-pressured) magma chamber. Theoretical analysis suggests that total) in the Collapse Caldera Data Base (Geyer and Marti, 2008;
collapse begins earlier (smaller f) for magma chambers that are shallow Table S1; Fig. 14B) shows that many eruptions do have collapse heights
and wide compared to those that are deep and narrow; these models that cluster at a single value ~ 1000 m, although the data are highly
also predict that collapse will be piston-like when R b ~ 1.4, and variable. In general, collapse heights ≥ 2000 m are relatively rare and
occur primarily in large CR eruptions, where total evacuation of the
magma reservoir is expected. Calculated collapse heights for many
large (N100 km3) CP eruptions, in contrast, are surprisingly small
(b 1000 m), possibly because the measured caldera areas exceed the
magma reservoir footprint. Where f (and thus the volume of magma
withdrawn prior to caldera collapse) is known (Fig. 14A), the thickness
of magma withdrawn from the reservoir prior to collapse can also be in-
ferred. Eruptions with large f have correspondingly large pre-eruptive
magma extraction depths, with a maximum value of ~ 1000 m (TMB
and Campanian [CMP]). Not surprisingly, eruptions with small f have
small pre-eruptive magma extraction depths (b0.2 km for Aira [AR],
KOS, VSN and Santorini Minoan [SAN]).
The compilation shown in Fig. 14B also provides insight into erup-
tion triggers. The dashed line on the diagram defines a caldera area
A = 100 km2, which has been identified as a thermomechanical bound-
ary that separates eruptions triggered by (external) roof collapse or (in-
ternal) chamber collapse (Gregg et al., 2012; de Silva and Gregg, 2014).
Fig. 13. Volume ratio f of pre-collapse to total eruption volume as a function of R, the ratio CR eruptions tend to lie above this boundary (in the roof-collapse re-
of roof thickness to roof area. High f means that caldera collapse was late in the eruptive
sequence, under these conditions it is likely that only part, rather than all, of the magma
gion), a placement that is consistent with the observed synchroneity be-
reservoir was evacuated. Specific eruptions are labeled. tween eruption initiation and caldera collapse. Data from many CP
Modified from Roche and Druitt (2001). eruptions also lie above this line, however, even for eruptions with
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 39

exceeds by a factor of 3 the collapse height of 1500 m calculated for


the entire erupted volume (Table S1), and by a factor of ~2 the collapse
height of 2–3 km indicated by analysis of drill core samples taken from
within the caldera (Hildreth and Mahood, 1986). At face value, this dis-
crepancy implies that on average, melt was extracted from less than 1/
2–1/3 of the magma reservoir. Possible explanations for this mismatch
include under-estimation of collapse heights, prior incorporation of
deep crystals (and their melt inclusions) into the eruptible melt lens,
or withdrawal of magma from stacked lenses that comprise only part
of the total magma reservoir (Fig. 15). We prefer the latter explanation,
as data presented in Hildreth (1979) argues against mixing by large-
scale convective overturn of the magmatic system prior to eruption.
If collapse height records the combined thickness of individual melt
lenses and compaction, but not evacuation, of the crystal framework,
and if the structural support provided by the reservoir plays a role in
controlling collapse height, then we would expect the largest discrepan-
cies between caldera collapse heights and magma drainage heights in
CP eruptions that tap complex reservoir geometries, where the strength
of the crystal framework would allow decoupling between melt drain-
age and collapse. By contrast, collapse height and drainage height
should coincide in eruptions that involve complete evacuation of the
reservoir (f = 0). This prediction is supported by data that indicate
that subsidence associated with large volume CR eruptions is typically
3–4 km, consistent with complete evacuation of these crystal-rich
magma reservoirs (Lipman, 1997). In CP eruptions, in contrast, partial
involvement of the crystal framework is suggested by a transition
during and after caldera collapse from initially crystal-poor magma
to crystal-rich magma. The latter often contains both cognate
glomerocrysts and antecrysts with melt inclusions that are more
evolved in composition than the matrix glass (that is, from a cooler, or
more evolved, part of the magma reservoir; Beddoe-Stephens et al.,
1983; Wolff et al., 1999; Charlier et al, 2007; Saunders et al., 2010;
Roberge et al., 2013).
In summary, there is growing evidence that caldera-forming erup-
Fig. 14. Relations between erupted volume (DRE), fraction of magma erupted prior to col- tions are not all fed by single magma bodies, and that accumulations
lapse (f) and inferred collapse height (calculated as erupted volume/caldera area). (A) f vs. of eruptible melt do not necessarily require assembly over long time pe-
volume; (B) volume vs. inferred collapse height. Symbols are the same in both, with blue riods. Instead, some systems may store melt within multiple sills (with-
circles CP magma and yellow squares CR magma. Lines in (A) show pre-eruptive DRE vol-
umes (labeled, in km3). Dashed line in (B) shows the contour for a caldera area of 100 km2,
in a rigid framework) or lenses (within a crystal mush) that may either
which Gregg et al. (2012) suggest as the bounding limit between chamber-trigged and amalgamate into a single melt body shortly before eruption, or may be
roof-triggered (yellow shading) eruptions. Labeled eruptions are as follows: AR = Aira, tapped syn-eruptively, particularly in extensional environments. In
AS = Aso, CEB = Ceboruco, CG = Cerro Galan, CL = Crater Lake, CMP = Campanian these complex magma reservoirs, pulsed interconnection of isolated
(Campi Flegrei), KT = Katmai, KOS = Kos, KRA = Krakatau, LG = La Garita (Fish Canyon
melt lenses promoted by syn-eruptive depressurization could both pro-
Tuff), LV = Long Valley (Bishop Tuff), NYT = Neapolitan Yellow Tuff (Campi Flegrei),
PN = Pinatubo, SAN = Santorini (Minoan), TMB = Tambora, TP = Taupo (181 AD), long explosive activity and explain commonly observed hiatuses in
VICO = Vico, VS = Vesuvius (79 AD), VSN = Villa Senni (Colli Albani). All data from eruptive sequences (e.g., Aramaki, 1984; Allen, 2004; Palladino and
Table S1. Simei, 2005; Bear et al., 2009; Vinkler et al., 2012), particularly if adjust-
ments within the reservoir are required to mobilize more crystalline
(and viscous) magma to newly formed caldera-bounding fractures.
The internal geometry and properties of a complex magma reservoir
very large f values (e.g., Campanian [CMP] and Aso [AS]) that denote late may also bear an unexplored influence on the development of ring
stage collapse (which is not consistent with a roof trigger). These dis- faults (e.g., Kennedy et al., 2004; Holohan et al., 2008b; Burchardt and
crepancies indicate that caldera area is not the sole control on eruption Walter, 2010) and extent of collapse in different sectors of a caldera.
triggering, which will also be affected by magma storage depth, reser- In fact, it seems likely that the spatial distribution of melt within a com-
voir configuration, magma input and tectonic triggering (e.g. Lindsay plex reservoir will contribute to the identified spectrum of piston,
et al., 2001; Allan et al., 2012). Support for the latter includes the relative downsag, trapdoor and piecemeal collapse styles (e.g., Cole et al.,
placement of small to moderate (b100 km3) stratovolcano eruptions 2005). Taken together, we suggest that the full range of eruptive condi-
(e.g., Tambora [TM], Krakatau [KR] Crater Lake [CL] and Santorini [SN]; tions (precursors, triggers and eruption dynamics) created by tapping
A b 100 km2) compared with eruptions of similar sizes in extensional complex storage regions has yet to be explored, and represents exciting
settings (e.g., Taupo [TP]; A N 100 km2). opportunities for future research.
Ideally, the entire vertical extent of magma extraction (the drainage
height) should be recorded by the volatile contents of phenocryst- 5. Implications for recognizing eruption potential of large
hosted melt inclusions (Wallace et al., 1999) and/or the stability of phe- magmatic systems
nocryst assemblages (e.g., Hammer et al., 2002; Gualda and Ghiorso,
2013). There are very few caldera-forming eruptions, however, for Effective volcano monitoring requires identification of systems
which the drainage height is well constrained. An exception is the Bish- capable of producing very large eruptions. For this reason, several
op Tuff, where both melt inclusion and phase equilibria data suggest potentially active volcanic regions have been the recent targets of
magma withdrawal over 130 MPa (~ 5000 m; Fig. 12B). This value geophysical surveys to search for large melt bodies. With only a
40 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

Fig. 15. Schematic diagram showing magma evacuation from a complex reservoir. (A) Prior to caldera collapse, magma is extracted from melt-dominated lenses throughout the reservoir
(the drainage depth), and magma migration is lateral, as well as vertical. (B) Collapse initiates when sufficient melt has been withdrawn that the strength of the crystal framework is
reduced and the framework itself partially disrupted. Here the collapse height is less than the drainage depth (as seen in many CP eruptions), and the erupted magma often contains
antecrysts from the framework; transition from a single vent to a ring vent phase is often accompanied by a hiatus in eruptive activity. (C) When the crystal mush is completely evacuated
along with any constituent melt lenses, the collapse height equals the drainage height (the case for many MI eruptions).

few exceptions (e.g., Zollo et al., 2008), interpretations of resulting (~ 200 km3) of magma was erupted under conditions of f N 0.9 but
geophysical images assess melt contents at ≤~30% (e.g., Schilling R b 1 (estimated from reported caldera area of 200 km3 and H2O con-
and Partzsch, 2001; Zandt et al., 2003; Chu et al., 2010; Luttrell tents ~ 5–7 wt.% (Kaneko et al., 2007), which places the magma at
et al., 2013), a number that appears safely at odds with the ≥ 50% depths no greater than 8–10 km, R ~ 0.5–0.67). Also suggestive is ev-
melt considered necessary for melt to be eruptible (e.g., Bachmann idence (from both melt inclusion and phase equilibria) for magma
and Bergantz, 2008). New views of magmatic systems, however, extraction over much larger pressure (depth) ranges than the col-
show that large volumes of melt may accumulate rapidly, and that lapse height calculated from erupted volume and caldera area. With-
multiple magma lenses may be tapped during a single eruptive epi- in this framework, we suggest that magma reservoirs such as those
sode. Syn-eruption melt extraction from a largely crystalline reser- imaged beneath Yellowstone may actually be capable of producing
voir seems likely for eruptions such as Aso, where a large volume a large eruption (e.g., Wotzlaw et al., 2014).
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 41

Another challenge relates to monitoring complex magmatic sys- (3) crystals carried by the transporting melt have been stored at a
tems. An increasingly important and effective volcano monitoring tech- range of pressures and temperatures, and (4) eruptions of crystal-rich
nique is measurement of surface deformation, particularly using magma are probably driven by roof collapse and fragmented by sudden
satellite-based Interferometric Synthetic Aperture Radar (InSAR; decompression (with a limited role for volatile exsolution and expan-
e.g., Sparks et al., 2012; Pyle et al., 2013). Surface deformation over sion). These observations require new models to explain and anticipate
large, shallow and sill-like magma bodies (that is, those susceptible to triggering and eruption of magma from complex storage reservoirs.
roof triggers) should be substantial; this makes them particularly good We address this question by first considering eruptions from mafic
targets for monitoring by InSAR. One complication, however, is that systems, which are commonly modeled as stacked sills. Here both direct
large magmatic reservoirs often have well-developed active hydrother- observations and petrologic studies of recently active systems show that
mal systems that may show extensive deformation related to shallow single eruptions may tap multiple melt lenses. One consequence is
changes in pore-pressure and water levels (e.g., Chiodini et al., 2003; pulsatory (and often protracted) eruptive activity and alternation be-
Husen et al., 2004; Chang et al., 2007). Pore-pressure-generated defor- tween explosion and lava effusion; another is eruption of a range of
mation signals may either mask or mimic magmatic activity. Surface de- magma compositions. We extend these modern examples to larger,
formation can also provide evidence of deep intrusive activity, such as caldera-forming eruptions using an example from the ultrapotassic
that currently ongoing at Uturuncu volcano, Bolivia (e.g., Sparks et al., Quaternary Roman Magmatic Province (QRMP), where we suggest
2008), and thus provides a potential tool for tracking long-term migra- that apparently contradictory observations of protracted explosivity
tion of magma inputs at different crustal levels. and low pyroclast vesicularity can be reconciled if the eruption tapped
Magmatic activity that precedes internally triggered eruptions (that a complex reservoir containing multiple melt lenses.
is, triggers involving intrusion of gas or hot melt from below, or over- We then turn to the more common, and larger, eruptions of silicic
pressurization caused by crystallization and associated gas exsolution) crystal-poor (CP) and crystal-rich (CR) magma, and review conditions
may be more difficult to recognize. One potential precursor is early of magma storage, time scales of melt accumulation, eruption triggers,
magma leakage from large reservoirs (e.g., Bacon, 1983; Duffield, eruption dynamics and conditions of caldera collapse. We show where
1990; Duffield et al., 1990; Hildreth, 2004; Fabbro et al., 2013). An inter- the Standard Model (single magma chamber) appears consistent with
esting observation is that precursory leaks from large magmatic systems the nature and stratigraphy of the eruptive products, as well as exam-
are often sourced from shallow levels and may produce either unusual ples where a complex (melt-lens-dominated) magma reservoir may
low energy fountains (e.g., Duffield, 1990; Bachmann et al., 2000) or better explain both petrological and volcanological observations. Most
lava flows (e.g., Bacon, 1983), despite tapping volatile-rich components important is the growing evidence that in many systems that erupt (ini-
of the magmatic system (e.g., Duffield and Dalrymple, 1990; Bacon et al., tially) crystal-poor silicic magma, particularly those in extensional envi-
1992; Mandeville et al., 2009). A good illustration of this phenomenon, ronments, melt may be stored within multiple isolated and/or partially
and a cautionary tale for event-tree-based hazard analysis, is provided connected lenses. This type of complex storage geometry could provide
by the eruptive sequence at Crater Lake, OR (Bacon, 1983). Here a com- a mechanism for rapid assembly of large melt bodies, as well as a local
posite eruption (Llao Rock; 1.7 km3 DRE pumice fall and 0.5 km3 DRE source of triggering (recharge) magma. In fact, Eichelberger et al.
lava flow) tapped the main magma reservoir about 200 years before (2006) extend this concept to suggest that a combination of porous
caldera formation. Another composite eruption (Cleetwood; 1.5 km3 media and triggered dike flow could allow sufficiently rapid syn-
DRE pumice fall and 0.6 km3 DRE lava flow) preceded the main (caldera- eruptive melt extraction to feed a silicic Plinian eruption, as we have in-
forming) phase of the eruption by only weeks to months (Kamata ferred for the QRMP example. From a hazard perspective, eruptions fed
et al., 1993). In both cases, the magma apparently came from the from complex magma reservoirs may show abrupt changes in eruptive
climactic reservoir (Bacon et al., 1992; Mandeville et al., 2009) and yet processes (such as pauses and transitions between explosive and effu-
each eruption transitioned from explosive to effusive. Why, then, was sive activity) that pose challenges for volcano monitoring and forecast-
the Cleetwood eruption followed so promptly by a very large ing efforts.
(~ 50 km3 DRE) explosive eruption from the same magmatic system? From a heuristic perspective, magma storage in complex, rather than
One possible explanation is that the precursory eruptions tapped rela- simple, magma reservoirs resolves several existing paradoxes about
tively shallow and partially to fully isolated melt lenses within a larger conditions of both pre-eruptive magma storage and syn-eruptive
reservoir. Magma withdrawal from the Cleetwood melt lens could magma withdrawal and caldera collapse. This perspective also repre-
then have triggered the climactic event by either downward or lateral sents a logical extension of recent attempts to reconcile petrological
propagation of a decompression wave capable of connecting the isolat- views of incremental assembly of plutons with volcanological require-
ed lens to the larger reservoir. This scenario illustrates the importance of ments of instantaneous availability of very large melt volumes
developing methods to monitor processes internal to magma reservoirs (Hildreth, 2004; Bachmann et al., 2007; Lipman, 2007; Walker et al.,
(e.g., induced seismicity; Catalli et al., 2013) during, as well as prior to 2007). Melt accumulation within, and syn-eruptive extraction from, in-
eruptions, and to distinguish between precursors and the main event terconnected melt lenses also alleviates problems related to maintain-
(e.g., Allan et al., 2012). ing large and stable volumes of crystal-poor melt in the upper crust
(Freda et al., 2011; Gualda et al., 2012; Vinkler et al., 2012) and places
Summary within a single coherent framework apparently disparate observations
related to magma extraction and caldera collapse. Finally, such a
It has long been known that caldera-forming eruptions may evacu- model is consistent with geophysical observations that even active
ate large volumes of either crystal-rich or crystal-poor magma magma storage reservoirs often contain ≤ 30% melt (Zandt et al.,
(e.g., Hildreth, 1981). The past ten years have seen a growing number 2003; Chu et al., 2010); the latter observation has important
of studies that relate the chemical and physical conditions in magma implications for recognizing the eruption potential of large magma
storage regions to the conditions under which different parts of the sys- reservoirs.
tem may be erupted (e.g., Jellinek and dePaolo, 2003; Bachmann and
Bergantz, 2004; Gottsmann et al., 2009; Bachmann, 2010; Allan et al.,
2012; Cooper et al., 2012; Druitt et al., 2012; Ellis and Wolff, 2012; Acknowledgments
Gregg et al., 2012; Hildreth and Fierstein, 2012; Huber et al., 2012;
Karlstrom et al., 2012; Gualda and Ghiorso, 2013). Key observations This work was supported by the AXA Research Fund through a
arising from these studies include: (1) many large eruptions tap multi- Research Professorship to KC and by Regione Lazio (818000-2009-R-
ple melt sources, (2) large melt bodies are probably transient features, M-R.N.C.T_001) for GG. We are grateful for the very thoughtful reviews
42 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

by G. Gualda and S. Burchardt and for the encouragement of the editor Bindeman, I.N., 2005. Fragmentation phenomena in populations of magmatic crystals.
Am. Mineral. 90, 1801–1815.
(L. Wilson) to write this review. Blundy, J., Cashman, K., 2008. Petrologic reconstruction of magmatic system variables and
processes. Rev. Mineral. Geochem. 69, 179–239.
Appendix A. Supplementary data Boari, E., Avanzinelli, R., Melluso, L., Giordano, G., Mattei, M., De Benedetti, A.A., Morra, V.,
Conticelli, S., 2009. Isotope geochemistry (Sr–Nd–Pb) and petrogenesis of leucite-
bearing rocks from “Colli Albani” volcano, Roman Magmatic Province, Central Italy:
Supplementary data to this article can be found online at http://dx. inferences on volcanic evolution. Bull. Volcanol. 71 (9), 977–1005.
doi.org/10.1016/j.jvolgeores.2014.09.007. Branney, M.J., 1995. Downsag and extension at calderas: new perspectives on collapse geom-
etries from ice-melt, mining, and volcanic subsidence. Bull. Volcanol. 57 (5), 303–318.
Brown, S.J.A., Burt, R.M., Cole, J.W., Krippner, S.J.P., Price, R.C., Cartwright, I., 1998. Plutonic
References lithics in ignimbrites of Taupo Volcanic Zone, New Zealand; sources and conditions of
crystallisation. Chem. Geol. 148 (1–2), 21–41.
Aarnes, I., Podladchikov, Y.Y., Neumann, E.-R., 2008. Post-emplacement melt flow induced Brown, R.J., Blake, S., Thordarson, T., Self, S., 2014. Pyroclastic edifices record vigorous lava
by thermal stresses: implications for differentiation in sills. Earth Planet. Sci. Lett. 276 fountains during the emplacement of a flood basalt flow field, Roza Member, Colum-
(1–2), 152–166. bia River Basalt Province, USA. Geological Society of America Bulletin 126, 875–891.
Acocella, V., 2007. Understanding caldera structure and development: an overview of an- Burchardt, S., 2008. New insights into the mechanics of sill emplacement provided by
alogue models compared to natural calderas. Earth Sci. Rev. 85 (3–4), 125–160. field observations of the Njardvik Sill, Northeast Iceland. J. Volcanol. Geotherm. Res.
Acocella, V., Palladino, D.M., Cioni, R., Russo, P., Simei, S., 2012. Caldera structure, amount 173, 280–288.
of collapse, and erupted volumes: the case of Bolsena caldera, Italy. Geol. Soc. Am. Burchardt, S., Walter, T.R., 2010. Propagation, linkage, and interaction of caldera
Bull. 124 (9–10), 1562–1576. ringfaults: comparison between analogue experiments and caldera collapse at
Adams, N.K., Houghton, B.F., Hildreth, W., 2006. Abrupt transitions during sustained ex- Miyakejima, Japan, in 2000. Bull. Volcanol. 72 (3), 297–308.
plosive eruptions: examples from the 1912 eruption of Novarupta: Alaska. Bull. Campagnola, S., 2014. Large scale Plinian eruptions of the Colli Albani and the Campi
Volcanol. 69 (2), 189–206. Flegrei volcanoes: insights from textural and rheological studies. Università Roma
Allan, A.S.R., Wilson, C.J.N., Millet, M.-A., Wysoczanski, R.J., 2012. The invisible hand: tec- Tre, (Unpublished PhD thesis).
tonic triggering and modulation of a rhyolitic supereruption. Geology 40 (6), Carbotte, S.M., Marjanović, M., Carton, H., Mutter, J.C., Canales, J.P., Nedimović, M.R., et al.,
563–566. 2013. Fine-scale segmentation of the crustal magma reservoir beneath the East Pacif-
Allan, A.R., Morgan, D., Wilson, C.N., Millet, M.-A., 2013. From mush to eruption in centu- ic Rise. Nat. Geosci. 6, 866–870.
ries: assembly of the super-sized: Oruanui magma body. Contrib. Mineral. Petrol. 166 Carter, N.L., Officer, C.B., Chesner, C.A., Rose, W.I., 1986. Dynamic deformation of volcanic
(1), 143–164. ejecta from the Toba caldera: possible relevance to Cretaceous/Tertiary boundary
Allen, S.R., 2001. Reconstruction of a major caldera-forming eruption from pyroclastic de- phenomena. Geology 14 (5), 380–383.
posit characteristics: Kos Plateau Tuff, eastern Aegean Sea. J. Volcanol. Geotherm. Res. Cas, R.A., Wright, H.M., Folkes, C.B., Lesti, C., Porreca, M., Giordano, G., Viramonte, J.G.,
105 (1–2), 141–162. 2011. The flow dynamics of an extremely large volume pyroclastic flow, the 2.08-
Allen, S.R., 2004. Complex spatter- and pumice-rich pyroclastic deposits from an andesitic Ma Cerro Galán Ignimbrite, NW Argentina, and comparison with other flow types.
caldera-forming eruption: the Siwi pyroclastic sequence, Tanna, Vanuatu. Bull. Bull. Volcanol. 73 (10), 1583–1609.
Volcanol. 67, 27–41. Cashman, K., Blundy, J., 2013. Petrological cannibalism: the chemical and textural conse-
Annen, C., 2011. Implications of incremental emplacement of magma bodies for magma quences of incremental magma body growth. Contrib. Mineral. Petrol. 166 (3),
differentiation, thermal aureole dimensions and plutonism–volcanism relationships. 703–729.
Tectonophysics 500, 3–10. Catalli, F., Meier, M., Wiemer, S., 2013. The role of Coulomb stress changes for injection-
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicic magmas induced seismicity: the Basel enhanced geothermal system. Geophys. Res. Lett. 40, 72–77.
in deep crustal hot zones. J. Petrol. 47, 505–539. Chamberlain, K.J., Wilson, C.J.N., Wooden, J.L., Charlier, B.L.A., Ireland, T.R., 2014. New per-
Aramaki, S., 1984. Formation of the Aira Caldera, Southern Kyushu, ∼22,000 years ago. J. spectives on the Bishop Tuff from zircon textures, ages and trace elements. J. Petrol.
Geophys. Res. 89, 8485–8501. 55 (2), 395–426.
Bachmann, O., 2010. The petrologic evolution and pre-eruptive conditions of the rhyolitic Chang, W.L., Smith, R.B., Wicks, C., Farrell, J., Puskas, C.M., 2007. Accelerated uplift
Kos Plateau Tuff (Aegean arc). Cent. Eur. J. Geosci. 2 (3), 270–305. and magmatic intrusion of the Yellowstone caldera, 2004 to 2006. Science 318,
Bachmann, O., Bergantz, G.W., 2004. On the origin of crystal-poor rhyolites: extracted 952–956.
from batholithic crystal mushes. J. Petrol. 45 (8), 1565–1582. Charlier, B.L.A., Peate, D.W., Wilson, C.J.N., Lowenstern, J.B., Storey, M., Brown, S.J.A., 2003.
Bachmann, O., Bergantz, G.W., 2006. Gas percolation in upper-crustal magma bodies as a Crystallisation ages in coeval silicic magma bodies: 238U–230Th disequilibrium evi-
mechanism for upward heat advection and rejuvenation of silicic crystal mushes. J. dence from the Rotoiti and Earthquake Flat eruption deposits, Taupo Volcanic Zone,
Volcanol. Geotherm. Res. 149, 85–102. New Zealand. Earth Planet. Sci. Lett. 206 (3–4), 441–457.
Bachmann, O., Bergantz, G.W., 2008. Rhyolites and their source mushes across tectonic Charlier, B.L.A., Wilson, C.J.N., Lowenstern, J.B., Blake, S., Van Calsteren, P.W., Davidson, J.P.,
settings. J. Petrol. 49, 2277–2285. 2005. Magma generation at a large, hyperactive silicic volcano (Taupo, New Zealand)
Bachmann, O., Dungan, M.A., Lipman, P.W., 2000. Voluminous lava-like precursor to a revealed by U–Th and U–Pb systematics in zircons. J. Petrol. 46 (1), 3–32.
major ash-flow tuff: low-column pyroclastic eruption of the Pagosa Peak Dacite, Charlier, B.L.A., Bachmann, O., Davidson, J.P., Dungan, M.A., Morgan, D.J., 2007. The upper
San Juan volcanic field, Colorado. J. Volcanol. Geotherm. Res. 98 (1–4), 153–171. crustal evolution of a large silicic magma body: evidence from crystal-scale Rb–Sr iso-
Bachmann, O., Dungan, M.A., Lipman, P.W., 2002. The Fish canyon magma body, San Juan topic heterogeneities in the Fish Canyon magmatic system, Colorado. J. Petrol. 48
Volcanic Field, Colorado: rejuvenation and eruption of an upper-crustal batholith. J. (10), 1875–1894.
Petrol. 43 (8), 1469–1503. Chesner, C.A., 2012. The Toba caldera complex. Quat. Int. 258, 5–18.
Bachmann, O., Miller, C.F., de Silva, S.L., 2007. The volcanic–plutonic connection as a stage Chesner, C.A., Rose, W., Deino, A., Drake, R., Westgate, J., 1991. Eruptive history of Earth's
for understanding crustal magmatism. J. Volcanol. Geotherm. Res. 167 (1–4), 1–23. largest Quaternary caldera (Toba, Indonesia) clarified. Geology 19 (3), 200–203.
Bacon, C.R., 1983. Eruptive history of Mount Mazama and Crater Lake caldera, Cascade Chiodini, G., Todesco, M., Caliro, S., Del Gaudio, C., Macedonio, G., Russo, M., 2003. Magma
Range, U.S.A. J. Volcanol. Geotherm. Res. 18, 57–115. degassing as a trigger of bradyseismic events: the case of Phlegrean Fields (Italy).
Bacon, C.R., Druitt, T.H., 1988. Compositional evolution of the zoned calcalkaline magma Geophys. Res. Lett. 30 (8).
chamber of Mount Mazama, Crater Lake, Oregon. Contrib. Mineral. Petrol. 98 (2), Chu, R., Helmberger, D.V., Sun, D., Jackson, J.M., Zhu, L., 2010. Mushy magma beneath
224–256. Yellowstone. Geophys. Res. Lett. 37, L01306.
Bacon, C.R., Newman, S., Stolper, E., 1992. Water, CO2, Cl, and F in melt inclusions in phe- Cioni, R., Pistolesi, M., Bertagnini, A., Bonadonna, C., Hoskuldsson, A., Scateni, B., 2014. In-
nocrysts from three Holocene explosive eruptions, Crater Lake, Oregon. Am. Mineral. sights into the dynamics and evolution of the 2010 Eyjafjallajökull summit eruption
77, 1021–1030. (Iceland) provided by volcanic ash textures. Earth Planet. Sci. Lett. 394, 111–123.
Bacon, C.R., Lanphere, M.A., Champion, D.E., 1999. Late Quaternary slip rate and seismic Cole, J.W., Milner, D.M., Spinks, K.D., 2005. Calderas and caldera structures: a review.
hazards of the West Klamath Lake fault zone near Crater Lake, Oregon Cascades. Ge- Earth Sci. Rev. 69 (1–2), 1–26.
ology 27, 43–46. Conticelli, S., Boari, E., Avanzinelli, R., De Benedetti, A.A., Giordano, G., Mattei, M., et al.,
Bagdassarov, N., Dorfman, A.M., 1998. Viscoelastic behaviour of partially molten granites. 2010. Geochemistry, isotopes and mineral chemistry of the Colli Albani volcanic
Tectonophysics 290, 27–45. rocks: constraints on magma genesis and evolution. The Colli Albani Volcano. Spec.
Bea, F., 2010. Crystallization dynamics of granite magma chambers in the absence of re- Publ. IAVCE I 3, 107–139.
gional stress: multiphysics modeling with natural examples. J. Petrol. 51 (7), Cooper, G.F., Wilson, C.J.N., Millet, M.-A., Baker, J.A., Smith, E.G.C., 2012. Systematic tap-
1541–1569. ping of independent magma chambers during the 1 Ma Kidnappers supereruption.
Bear, A., Cas, R., Giordano, G., 2009. Variations in eruptive style and depositional processes Earth Planet. Sci. Lett. 313–314, 23–33.
associated with explosive, phonolitic composition, caldera-forming eruptions: the Dahren, B., Troll, V., Andersson, U., Chadwick, J., Gardner, M., Jaxybulatov, K., Koulakov, I.,
151 ka Sutri eruption, Vico Caldera, central Italy. J. Volcanol. Geotherm. Res. 184 2012. Magma plumbing beneath Anak Krakatau volcano, Indonesia: evidence for
(3), 225–255. multiple magma storage regions. Contrib. Mineral. Petrol. 163 (4), 631–651.
Beddoe-Stephens, B., Aspden, J.A., Shepherd, T.J., 1983. Glass inclusions and melt compo- de Angelis, S.H., Larsen, J., Coombs, M., 2013. Pre-eruptive magmatic conditions at Augustine
sitions of the, Toba Tuffs, northern Sumatra. Contrib. Mineral. Petrol. 83 (3–4), Volcano, Alaska, 2006: evidence from amphibole geochemistry and textures. J. Petrol.
278–287. 54, 1939–1961.
Bégué, F., Deering, C.D., Gravley, D.M., Kennedy, B.M., Chambefort, I., et al., 2014. Extrac- de Silva, S.L., Gregg, P.M., 2014. Thermomechanical feedbacks in magmatic systems: Im-
tion, storage and eruption of multiple isolated magma batches in the paired Mamaku plications for growth, longevity, and evolution of large caldera-forming magma reser-
and Ohakuri eruption, Taupo Volcanic Zone, New Zealand. J. Petrol. 55, 1653–1684. voirs and their supereruptions. J. Volcanol. and Geotherm. Res. 282, 77–91.
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 43

de Silva, S., Zandt, G., Trumbull, R., Viramonte, J.G., Salas, G., Jimenez, N., 2006. In: Troise, C., Gravley, D.M., Wilson, C.J.N., Leonard, G.S., Cole, J.W., 2007. Double trouble: Paired ignim-
De Natale, G., Kilburn, C.R.J. (Eds.), Large ignimbrite eruptions and volcano-tectonic brite eruptions and collateral subsidence in the Taupo Volcanic Zone, New Zealand.
depressions in the Central Andes: a thermomechanical perspectiveMechanisms of Ac- Geol. Soc. Am. Bull. 119, 18–30.
tivity and Unrest at Large Calderas vol. 269. Geological Society of London, London, pp. Gregg, P.M., de Silva, S.L., Grosfils, E.B., Parmigiani, J.P., 2012. Catastrophic caldera-forming
47–63. eruptions: thermomechanics and implications for eruption triggering and maximum
Deering, C.D., Bachmann, O., Vogel, T.A., 2011. The Ammonia Tanks Tuff: erupting a melt- caldera dimensions on Earth. J. Volcanol. Geotherm. Res. 241–242, 1–12.
rich rhyolite cap and its remobilized crystal cumulate. Earth Planet. Sci. Lett. 310 Gualda, G.A.R., Ghiorso, M.S., 2013. The Bishop Tuff giant magma body: an alternative to
(2011), 518–525. the Standard Model. Contrib. Mineral. Petrol. 166, 755–775.
Druitt, T., Bacon, C., 1989. Petrology of the zoned calcalkaline magma chamber of Mount Gualda, G.A.R., Cook, D.L., Chopra, R., Qin, L.P., Anderson Jr., A.T., Rivers, M., 2004. Frag-
Mazama, Crater Lake, Oregon. Contrib. Mineral. Petrol. 101 (2), 245–259. mentation, nucleation and migration of crystals and bubbles in the Bishop Tuff rhyo-
Druitt, T.H., Sparks, R.S.J., 1984. On the formation of calderas during ignimbrite eruptions. litic magma. Trans. R. Soc. Edinb. Earth Sci. 95, 375–390.
Nature 310 (5979), 679–681. Gualda, G.A.R., Ghiorso, M.S., Lemons, R.V., Carley, T.L., 2012. Rhyolite-MELTS: a modified
Druitt, T.H., Costa, F., Deloule, E., Dungan, M., Scaillet, B., 2012. Decadal to monthly time- calibration of MELTS optimised for silica-rich, fluid-bearing magmatic systems. J. of
scales of magma transfer and reservoir growth at a caldera volcano. Nature 482 Petrol. 53, 875–890.
(7383), 77–80. Gudmundsson, A., 2008. Magma-chamber geometry, fluid transport, local stresses and
Dufek, J., Bachmann, O., 2010. Quantum magmatism magmatic compositional gaps gener- rock behaviour during collapse caldera formation. Dev. Volcanol. 10, 313–349.
ated by melt-crystal dynamics. Geology 38 (8), 687–690. Gudmundsson, A., 2012. Magma chambers: formation, local stresses, excess pressures,
Duffield, W.A., 1990. Eruptive fountains of silicic magma and their possible effects on the and compartments. J. Volcanol. Geotherm. Res. 237–238, 19–41.
tin content of fountain-fed lavas, Taylor Creek Rhyolite, New Mexico. Geol. Soc. Am. Gurioli, L., Houghton, B.F., Cashman, K.V., Cioni, R., 2005. Complex changes in eruption dy-
Spec. Pap. 246, 251–262. namics during the 79 AD eruption of Vesuvius. Bull. Volcanol. 67 (2), 144–159.
Duffield, W., Dalrymple, G., 1990. The Taylor Creek Rhyolite of New Mexico: a rapidly Hammer, J.E., Rutherford, M.J., Hildreth, W., 2002. Magma storage prior to the 1912 erup-
emplaced field of lava domes and flows. Bull. Volcanol. 52 (no. 6), 475–487. tion at Novarupta, Alaska. Contrib. Mineral. Petrol. 144, 144–162.
Eichelberger, J.C., Izbekov, P.E., 2000. Eruption of andesite triggered by dyke injection: Hildreth, W., 1979. The Bishop Tuff: evidence for the origin of compositional zonation in
contrasting cases at Karymsky Volcano, Kamchatka and Mt Katmai, Alaska: philo- silicic magma chambers. Geol. Soc. Am. Spec. Pap. 180, 43–75.
sophical transactions of the Royal Society of London. Ser. A: Math. Phys. Eng. Sci. Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithospheric
358 (1770), 1465–1485. magmatism. J. Geophys. Res. 86, 10153–10192.
Eichelberger, J.C., Izbekov, P.E., Browne, B.L., 2006. Bulk chemical trends at arc volcanoes Hildreth, W., 2004. Volcanological perspectives on Long Valley Mammoth Mountain, and
are not liquid lines of descent. Lithos 87 (1–2), 135–154. Mono Craters: several contiguous but discrete systems. J. Volcanol. Geotherm. Res.
Ellis, B.S., Wolff, J.A., 2012. Complex storage of rhyolite in the central Snake River Plain. J. 136, 169–198.
Volcanol. Geotherm. Res. 211–212, 1–11. Hildreth, W., Fierstein, J., 2012. The Novarupta-Katmai eruption of 1912 — largest erup-
Ellis, B.S., Barry, T., Branney, M.J., Wolff, J.A., Bindeman, I., Wilson, R., Bonnichsen, B., 2010. tion of the twentieth century; centennial perspectives. U.S.G.S. Prof. Paper 1791.
Petrologic constraints on the development of a large-volume, high temperature, silic- Hildreth, W., Mahood, G.A., 1986. Ring-fracture eruption of the Bishop Tuff. Geol. Soc. Am.
ic magma system: the Twin Falls eruptive centre, central Snake River Plain. Lithos 120 Bull. 97, 396–403.
(3–4), 475–489. Hildreth, W., Wilson, C.J.N., 2007. Compositional zoning of the Bishop Tuff. J. Petrol. 48,
Erlund, E.J., Cashman, K.V., Wallace, P.J., Pioli, L., Rosi, M., Johnson, E., Granados, H.D., 2010. 951–999.
Compositional evolution of magma from Parícutin Volcano, Mexico: the tephra re- Holohan, E., van Wyk de Vries, B., Troll, V., 2008a. Analogue models of caldera collapse in
cord. J. Volcanol. Geotherm. Res. 197 (1–4), 167–187. strike-slip tectonic regimes. Bull. Volcanol. 70 (no. 7), 773–796.
Fabbro, G., Druitt, T., Scaillet, S., 2013. Evolution of the crustal magma plumbing system Holohan, E., Troll, V., van Wyk de Vries, B., Walsh, J.J., Walter, T.R., 2008b. Unzipping Long
during the build-up to the 22-ka caldera-forming eruption of Santorini (Greece). Valley: an explanation for vent migration patterns during elliptical ring fracture erup-
Bull. Volcanol. 75 (12), 1–22. tion. Geology 36, 323–326.
Folkes, C.B., Wright, H.M., Cas, R.A., de Silva, S.L., Lesti, C., Viramonte, J.G., 2011. A re- Houghton, B.F., Carey, R.J., Cashman, K.V., Wilson, C.J., Hobden, B.J., Hammer, J.E., 2010. Di-
appraisal of the stratigraphy and volcanology of the Cerro Galán volcanic system, verse patterns of ascent, degassing, and eruption of rhyolite magma during the 1.8 ka
NW Argentina. Bull. Volcanol. 73, 1427–1454. Taupo eruption, New Zealand: evidence from clast vesicularity. J. Volcanol. Geotherm.
Fowler, S.J., Spera, F.J., Bohrson, W.A., Belkin, H.E., De Vivo, B., 2007. Phase equilibria con- Res. 195 (1), 31–47.
straints on the chemical and physical evolution of the Campanian Ignimbrite. J. Petrol. Huber, C., Bachmann, O., Dufek, J., 2011. Thermo-mechanical reactivation of locked crystal
48, 459–493. mushes: melting-induced internal fracturing and assimilation processes in magmas.
Francis, P.W., Sparks, R., Hawkesworth, C., Thorpe, R., Pyle, D., Tait, S., Mantovani, M., Earth Planet. Sci. Lett. 3–4, 443–454.
McDermott, F., 1989. Petrology and geochemistry of volcanic rocks of the Cerro Huber, C., Bachmann, O., Dufek, J., 2012. Crystal-poor versus crystal-rich ignimbrites; a
Galan caldera, northwest Argentina. Geol. Mag. 126 (05), 515–547. competition between stirring and reactivation. Geology 40 (2), 115–118.
Freda, C., Gaeta, M., Palladino, D.M., Trigila, R., 1997. The Villa Senni Eruption (Alban Hills, Humphreys, M.C.S., Holness, M.B., 2010. Melt-rich segregations in the Skaergaard Margin-
central Italy): the role of H2O and CO2 on the magma chamber evolution and on the al Border Series: tearing of a vertical silicate mush. Lithos 119 (3–4), 181–192.
eruptive scenario. J. Volcanol. Geotherm. Res. 78 (no. 1–2), 103–120. Humphreys, M., Christopher, T., Hards, V., 2009. Microlite transfer by disaggregation of
Freda, C., Gaeta, M., Giaccio, B., Marra, F., Palladino, D., Scarlato, P., Sottili, G., 2011. CO2- mafic inclusions following magma mixing at Soufrière Hills volcano, Montserrat.
driven large mafic explosive eruptions: the Pozzolane Rosse case study from the Contrib. Mineral. Petrol. 157 (5), 609–624.
Colli Albani Volcanic District (Italy). Bull. Volcanol. 73 (3), 241–256. Husen, S., Smith, R.B., Waite, G.P., 2004. Evidence for gas and magmatic sources beneath
Freundt, A., Schmincke, H.-U., 1995. Eruption and emplacement of a basaltic welded ig- the Yellowstone volcanic field from seismic tomographic imaging. J. Volcanol.
nimbrite during caldera formation on Gran Canaria. Bull. Volcanol. 56 (8), 640–659. Geotherm. Res. 131 (3–4), 397–410.
Gardner, J.E., Sigurdsson, H., Carey, S.N., 1991. Eruption dynamics and magma withdrawal Iyer, H., Evans, J., Dawson, P., Stauber, D., Achauer, U., 1990. Differences in magma storage
during the Plinian Phase of the Bishop Tuff Eruption, Long Valley Caldera. J. Geophy. in different volcanic environments as revealed by seismic tomography; silicic volca-
Res. Solid Earth 96 (B5), 8097–8111. nic centers and subduction-related volcanoes. Magma Transport and Storage,pp.
Gardner, J.E., Layer, P.W., Rutherford, M.J., 2002. Phenocrysts versus xenocrysts in the 293–316.
youngest Toba Tuff: implications for the petrogenesis of 2800 km3 of magma. Geolo- Jellinek, A.M., DePaolo, D.J., 2003. A model for the origin of large silicic magma chambers:
gy 30 (4), 347–350. precursors of caldera-forming eruptions. Bull. Volcanol. 65 (5), 363–381.
Gebauer, S., Schmitt, A., Pappalardo, L., Stockli, D., Lovera, O., 2014. Crystallization Johnson, D., 1987. Elastic and inelastic magma storage at Kilauea volcano. US Geol. Surv.
and eruption ages of Breccia Museo (Campi Flegrei caldera, Italy) plutonic Prof. Pap. 1350, 1297–1306.
clasts and their relation to the Campanian ignimbrite. Contrib. Mineral. Petrol. 167 Kamata, H., Suzuki-Kamata, K., Bacon, C.R., 1993. Deformation of the Wineglass Welded
(no. 1), 1–18. Tuff and the timing of caldera collapse at Crater Lake, Oregon. J. Volcanol. Geotherm.
Geshi, N., Ruch, J., Acocella, V., 2014. Evaluating volumes for magma chambers and Res. 56, 253–265.
magma withdrawn for caldera collapse. Earth Planet. Sci. Lett. 396, 107–115. Kaneko, K., Kamata, H., Koyaguchi, T., Yoshikawa, M., Furukawa, K., 2007. Repeated large-
Geyer, A., Marti, J., 2008. The new worldwide collapse caldera database (CCDB): a tool for scale eruptions from a single compositionally stratified magma chamber: an example
studying and understanding caldera processes. J. Volcanol. Geotherm. Res. 175, from Aso volcano, Southwest Japan. J. Volcanol. Geotherm. Res. 167 (1–4), 160–180.
334–354. Karlstrom, L., Rudolph, M.L., Manga, M., 2012. Caldera size modulated by the yield stress
Geyer, A., Folch, A., Martì, J., 2006. Relationship between caldera collapse and magma within a crystal-rich magma reservoir. Nat. Geosci. 5, 402–405.
chamber withdrawal: An experimental approach. J. Volcanol. Geotherm. Res. 157, Kay, S., Coira, B., Wörner, G., Kay, R., Singer, B., 2011. Geochemical, isotopic and single
375–386. crystal 40Ar/39Ar age constraints on the evolution of the Cerro Galán ignimbrites.
Giordano, G., Dobran, F., 1994. Computer simulations of the Tuscolano Artemisio's IInd Bull. Volcanol. 73 (10), 1487–1511.
Pyroclastic Flow Unit (Alban Hills, Central Italy). J. Volcanol. Geotherm. Res. 61, Kelley, D.F., Barton, M., 2008. Pressures of crystallization of icelandic magmas. J. Petrol. 49
69–94. (3), 465–492.
Giordano, G., the CARG team, 2010. Stratigraphy and volcano-tectonic structures of the Kennedy, B., Stix, J., Vallance, J.W., Lavallée, Y., Longpré, M.-A., 2004. Controls on caldera
Colli Albani volcanic field. In: Funiciello, R., Giordano, G. (Eds.), The Colli Albani structure: results from analogue sandbox modeling. Geol. Soc. Am. Bull. 116 (5–6),
VolcanoSpecial Publication of IAVCEI 3. The Geological Society, London, pp. 43–97. 515–524.
Giordano, G., De Benedetti, A., Diana, A., Diano, G., Gaudioso, F., Marasco, F., Miceli, M., Kennedy, B.M., Jellinek, M.A., Stix, J., 2008. Coupled caldera subsidence and stirring in-
Mollo, S., Cas, R., Funiciello, R., 2006. The Colli Albani mafic caldera (Roma, Italy): stra- ferred from analogue models. Nat. Geosci. 1 (6), 385–389.
tigraphy, structure and petrology. J. Volcanol. Geotherm. Res. 155 (1), 49–80. Klug, C., Cashman, K., Bacon, C., 2002. Structure and physical characteristics of pumice
Gottsmann, J., Lavallée, Y., Martí, J., Aguirre-Díaz, G., 2009. Magma–tectonic interaction from the climactic eruption of Mount Mazama (Crater Lake), Oregon. Bull. Volcanol.
and the eruption of silicic batholiths. Earth Planet. Sci. Lett. 284 (3–4), 426–434. 64 (7), 486–501.
44 K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45

Lees, J.M., 2007. Seismic tomography of magmatic systems. J. Volcanol. Geotherm. Res. Roberge, J., Wallace, P.J., Kent, A.J.R., 2013. Magmatic processes in the Bishop Tuff rhyolitic
167 (1–4), 37–56. magma based on trace elements in melt inclusions and pumice matrix glass. Contrib.
Lesti, C., Porreca, M., Giordano, G., Mattei, M., Cas, R.A., Wright, H.M., Folkes, C.B., Mineral. Petrol. 165, 237–257.
Viramonte, J., 2011. High-temperature emplacement of the Cerro Galán and Roche, O., Druitt, T.H., 2001. Onset of caldera collapse during ignimbrite eruptions. Earth
Toconquis Group ignimbrites (Puna plateau, NW Argentina) determined by TRM Planet. Sci. Lett. 191, 191–202.
analyses. Bull. Volcanol. 73 (10), 1535–1565. Roman, D.C., Cashman, K.V., Gardner, C.A., Wallace, P.J., Donovan, J.J., 2006. Storage and
Lindsay, J.M., Schmitt, A.K., Trumbull, R.B., de Silva, S.L., Siebel, W., Emmermann, R., 2001. Mag- interaction of compositionally heterogeneous magmas from the 1986 eruption of
matic evolution of the La Pacana Caldera System, Central Andes, Chile: compositional var- Augustine Volcano, Alaska. Bull. Volcanol. 68, 240–254.
iation of two cogenetic, large-volume felsic ignimbrites. J. Petrol. 42 (3), 459–486. Rosenberg, C.L., Medvedev, S., Handy, M.R., 2007. Effects of melting on faulting and con-
Lipman, P.W., 1997. Subsidence of ash-flow calderas: relation to caldera size and magma- tinental deformation. Dahlem Workshop Report vol. 95, pp. 357–401.
chamber geometry. Bull. Volcanol. 59 (3), 198–218. Rosi, M., Vezzoli, L., Castelmenzano, A., Grieco, G., 1999. Plinian pumice fall deposit of the
Lipman, P.W., 2007. Incremental assembly and prolonged consolidation of Cordilleran Campanian Ignimbrite eruption (Phlegraean Fields, Italy). J. Volcanol. Geotherm. Res.
magma chambers: evidence from the Southern Rocky Mountain volcanic field. 91, 179–198.
Geosphere 3. http://dx.doi.org/10.1130/GES00061.1. Rosi, M., Landi, P., Polacci, M., Di Muro, A., Zandomeneghi, D., 2004. Role of conduit shear
Lipman, P., Dungan, M., Bachmann, O., 1997. Comagmatic granophyric granite in the Fish on ascent of the crystal-rich magma feeding the 800-year-b.p. Plinian eruption of
Canyon Tuff, Colorado: implications for magma-chamber processes during a large Quilotoa Volcano (Ecuador). Bull. Volcanol. 66 (4), 307–321.
ash-flow eruption. Geology 25 (10), 915–918. Rust, A.C., Cashman, K.V., 2011. Permeability controls on expansion and size distributions
Liu, Y., Anderson, A.T., Wilson, C.J.N., Davis, A.M., Steele, I.M., 2006. Mixing and differenti- of pyroclasts. J. Geophys. Res. 116. http://dx.doi.org/10.1029/2011JB008494.
ation in the Oruanui rhyolitic magma, Taupo, New Zealand: evidence from volatiles Saunders, K.E., Morgan, D.J., Baker, J.A., Wysoczanski, R.J., 2010. The magmatic evolution of
and trace elements in melt inclusions. Contrib. Mineral. Petrol. 151, 71–87. the Whakamaru Supereruption, New Zealand, constrained by a microanalytical study
Luttrell, K., Mencin, D., Francis, O., Hurwitz, S., 2013. Constraints on the upper crustal of plagioclase and quartz. J. Petrol. 51 (12), 2465–2488.
magma reservoir beneath Yellowstone Caldera inferred from lake-seiche induced Scandone, R., 1990. Chaotic collapse of calderas. J. Volcanol. Geotherm. Res. 42 (3),
strain observations. Geophys. Res. Lett. 40 (3), 501–506. 285–302.
Mandeville, C.W., Webster, J.D., Tappen, C., Taylor, B.E., Timbal, A., Sasaki, A., Hauri, E., Schilling, F.R., Partzsch, G.M., 2001. Quantifying partial melt fraction in the crust beneath
Bacon, C.R., 2009. Stable isotope and petrologic evidence for open-system degassing the Central Andes and the Tibetan plateau: physics and chemistry of the earth. Earth
during the climactic and pre-climactic eruptions of Mt. Mazama, Crater Lake, Oregon. Solid Earth Geod. 26 (4–5), 239–246.
Geochim. Cosmochim. Acta 73 (10), 2978–3012. Sherburn, S., Bannister, S., Bibby, H., 2003. Seismic velocity structure of the central Taupo
Mangiacapra, A., Moretti, R., Rutherford, M., Civetta, L., Orsi, G., Papale, P., 2008. The deep Volcanic Zone, New Zealand, from local earthquake tomography. J. Volcanol.
magmatic system of the Campi Flegrei caldera (Italy). Geophys. Res. Lett. 35 (21), Geotherm. Res. 122 (1–2), 69–88.
L21304. Sigmundsson, F., Hreinsdóttir, S., Hooper, A., Árnadóttir, T., Pedersen, R., Roberts, M.J., et
Marsh, B.D., 1996. Solidification fronts and magmatic evolution. Mineral. Mag. 60 (398), al., 2010. Intrusion triggering of the 2010 Eyjafjallajokull explosive eruption. Nature
5–40. 468, 426–430.
Marsh, B.D., 2002. On bimodal differentiation by solidification front instability in basaltic Sigurdsson, H., Sparks, R., 1981. Petrology of rhyolitic and mixed magma ejecta from the
magmas, part 1: basic mechanics. Geochim. Cosmochim. Acta 66 (12), 2211–2229. 1875 eruption of Askja, Iceland. J. Petrol. 22 (1), 41–84.
Marsh, B.D., 2013. On some fundamentals of igneous petrology. Contrib. Mineral. Petrol. Simon, J., Weis, D., DePaolo, D., Renne, P., Mundil, R., Schmitt, A., 2014. Assimilation of
166, 665–690. preexisting Pleistocene intrusions at Long Valley by periodic magma recharge acceler-
Martí, J., 1991. Caldera-like structures related to Permo-Carboniferous volcanism of the ates rhyolite generation: rethinking the remelting model. Contrib. Mineral. Petrol. 167
Catalan Pyrenees (NE Spain). J. Volcanol. Geotherm. Res. 45 (3–4), 173–186. (1), 1–34.
Marti, J., Geyer, A., Folch, A., Gottsmann, J., 2008. A review on collapse caldera modelling. Sisson, T.W., Bacon, C.R., 1999. Gas-driven filter pressing in magmas. Geology 27 (7),
Dev. Volcanol. 10, 233–283. 613–616.
Masotta, M., Gaeta, M., Gozzi, F., Marra, F., Palladino, D., Sottili, G., 2010. H2O-and Smith, R.L., 1979. Ash-flow magmatism: ash-flow tuffs. Geol. Soc. Am. Spec. Pap. 180,
temperature-zoning in magma chambers: the example of the Tufo Giallo della Via 5–27.
Tiberina eruptions (Sabatini Volcanic District, central Italy). Lithos 118 (1), 119–130. Solano, J.M.S., Jackson, M.D., Sparks, R.S.J., Blundy, J.D., Annen, C., 2012. Melt segregation in
Masotta, M., Freda, C., Gaeta, M., 2012. Origin of crystal-poor, differentiated magmas: in- deep crustal hot zones: a mechanism for chemical differentiation, crustal assimilation
sights from thermal gradient experiments. Contrib. Mineral. Petrol. 163 (1), 49–65. and the formation of evolved magmas. J. Petrol. 53, 1999–2026.
Masturyono, R.M., Wark, D., Roecker, S., Fauzi, G.I., 2001. Distribution of magma beneath Sparks, R.S.J., Wilson, C.J.N., 1990. The Minoan deposits: a review of their characteristics
Toba Caldera, North Sumatra, Indonesia, Constrained by 3-dimensional P-wave veloc- and interpretation. Thera and the Aegean world III 2, pp. 89–99.
ities, seismicity, and gravity data. Geochem. Geophys. Geosyst. 2. Sparks, R., Francis, P., Hamer, R., Pankhurst, R., O'callaghan, L., Thorpe, R., Page, R., 1985.
Matthews, N.E., Huber, C., Pyle, D.M., Smith, V.C., 2012. Timescales of magma recharge and Ignimbrites of the Cerro Galan Caldera, NW Argentina. J. Volcanol. Geotherm. Res.
reactivation of large silicic systems from Ti diffusion in quartz. J. Petrol. 53 (7), 24 (3), 205–248.
1385–1416. Sparks, R.S.J., Folkes, C.B., Humphreys, M.C., Barfod, D.N., Clavero, J., Sunagua, M.C.,
Maughan, L.L., Christiansen, E.H., Best, M.G., Grommé, C.S., Deino, A.L., Tingey, D.G., 2002. McNutt, S.R., Pritchard, M.E., 2008. Uturuncu volcano, Bolivia: volcanic unrest due
The Oligocene Lund Tuff, Great Basin, USA: a very large volume monotonous interme- to mid-crustal magma intrusion. Am. J. Sci. 308 (6), 727–769.
diate. J. Volcanol. Geotherm. Res. 113 (1–2), 129–157. Sparks, R., Biggs, J., Neuberg, J., 2012. Monitoring volcanoes. Science 335 (6074),
McKenzie, D., 1984. The generation and compaction of partially molten rock. J. Petrol. 25 1310–1311.
(3), 713–765. Spera, F.J., Crisp, J.A., 1981. Eruption volume, periodicity, and caldera area: relationships
Melnik, O., Barmin, A.A., Sparks, R.S.J., 2005. Dynamics of magma flow inside volcanic con- and inferences on development of compositional zonation in silicic magma cham-
duits with bubble overpressure buildup and gas loss through permeable magma. J. bers. J. Volcanol. Geotherm. Res. 11, 169–187.
Volcanol. Geotherm. Res. 143, 53–68. Stix, J., Kobayashi, T., 2007. Magma dynamics and collapse mechanisms during four his-
Neave, D.A., Passmore, E., Maclennan, J., Fitton, G., Thordarson, T., 2013. Crystal–melt re- toric caldera-forming events. J. Geophys. Res. 113, B09205.
lationships and the record of deep mixing and crystallization in the ad 1783 Laki Tait, S., Jaupart, C., Vergniolle, S., 1989. Pressure, gas content and eruption periodicity of a
Eruption, Iceland. J. Petrol. 54 (8), 1661–1690. shallow, crystallising magma chamber. Earth Planet. Sci. Lett. 92, 107–123.
Palladino, D.M., Simei, S., 2005. Eruptive dynamics and caldera collapse during the Onano Tarasewicz, J., White, R.S., Woods, A.W., Brandsdóttir, B., Gudmundsson, M.T., 2012.
eruption, Vulsini, Italy. Bull. Volcanol. 67, 423–440. Magma mobilization by downward-propagating decompression of the Eyjafjallajökull
Pallister, J.S., Hoblitt, R.P., Reyes, A.G., 1992. A basalt trigger for the 1991 eruptions of volcanic plumbing system. Geophys. Res. Lett. 39 (19), L19309.
Pinatubo volcano? Nature 356, 426–428. van der Zwan, F., Chadwick, J., Troll, V., 2013. Textural history of recent basaltic-
Pamukcu, A.S., Gualda, G.A.R., Anderson Jr., A.T., 2012. Crystallization stages of the andesites and plutonic inclusions from Merapi volcano. Contrib. Mineral. Petrol.
Bishop Tuff magma body recorded in crystal textures in pumice clasts. J. Petrol. 166 (1), 43–63.
53, 589–609. Vinkler, A.P., Cashman, K., Giordano, G., Groppelli, G., 2012. Evolution of the mafic Villa
Pamukcu, A.S., Carley, T.L., Gualda, G.A.R., Miller, C.F., Ferguson, C.A., 2013. The evolution Senni caldera-forming eruption at Colli Albani volcano, Italy, indicated by textural
of the Peach Spring giant magma body: evidence from accessory mineral textures analysis of juvenile fragments. J. Volcanol. Geotherm. Res. 235–236, 37–54.
and compositions, bulk pumice and glass geochemistry, and rhyolite-MELTS model- Voight, B., Widiwijayanti, C., Mattioli, G., Elsworth, D., Hidayat, D., Strutt, M., 2010.
ing. J. Petrol. 54, 1109–1148. Magma-sponge hypothesis and stratovolcanoes: case for a compressible reservoir
Passmore, E., Maclennan, J., Fitton, G., Thordarson, T., 2012. Mush disaggregation in basal- and quasi-steady deep influx at Soufrière Hills Volcano, Montserrat. Geophys. Res.
tic magma chambers: evidence from the ad 1783 Laki Eruption. J. Petrol. 53 (12), Lett. 37 (19).
2593–2623. Vye-Brown, C., Gannoun, A., Barry, T.L., Self, S., Burton, K.W., 2013. Osmium isotope vari-
Peccerillo, A., 2005. The roman province: plio-quaternary volcanism in Italy. Petrol. ations accompanying the eruption of a single lava flow field in the Columbia River
Geochem. Geodyn. 69–107. Flood Basalt Province. Earth Planet. Sci. Lett. 368, 183–194.
Polacci, M., Papale, P., Rosi, M., 2001. Textural heterogeneities in pumices from the climac- Walker, G.P., 1984. Downsag calderas, ring faults, caldera sizes, and incremental caldera
tic eruption of Mount Pinatubo, 15 June 1991, and implications for magma ascent dy- growth. J.Geophys. Res. Solid Earth (1978–2012) 89 (B10), 8407–8416.
namics. Bull. Volcanol. 63 (2–3), 83–97. Walker, B.A., Miller, C.F., Claiborne, L.L., Wooden, J.L., Miller, J.S., 2007. Geology and geo-
Pyle, D.M., Mather, T.A., Biggs, J., 2013. Remote sensing of volcanoes and volcanic process- chronology of the Spirit Mountain batholith, southern Nevada; implications for time-
es: integrating observation and modelling — introduction. Geol. Soc. Lond., Spec. Publ. scales and physical processes of batholith construction. J. Volcanol. Geotherm. Res.
380 (1), 1–13. 167, 239–262.
Reid, M.R., 2008. How long does it take to supersize an eruption? Elements 4 (1), Walker Jr., B., Klemetti, E., Grunder, A., Dilles, J., Tepley, F., Giles, D., 2013. Crystal reaming
23–28. during the assembly, maturation, and waning of an eleven-million-year crustal
K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 28–45 45

magma cycle: thermobarometry of the Aucanquilcha Volcanic Cluster. Contrib. Min- Wolff, J.A., Ramos, F.C., Davidson, J.P., 1999. Sr isotope disequilibrium during differentia-
eral. Petrol. 165 (4), 663–682. tion of the Bandelier Tuff: constraints on the crystallization of a large rhyolitic
Wallace, P.J., Anderson Jr., A.T., Davis, A.M., 1999. Gradients in H2O, CO2, and exsolved gas magma chamber. Geology 27, 495–498.
in a large-volume silicic magma system: interpreting the record preserved in melt in- Wolff, J.A., Ramos, F.C., Hart, G.L., Patterson, J.D., Brandon, A.D., 2008. Columbia River flood
clusions from the Bishop Tuff. J. Geophys. Res. 104 (B9), 20097–20122. basalts from a centralized crustal magmatic system. Nat. Geosci. 1 (3), 177–180.
Wark, D.A., Hildreth, W., Spear, F.S., Cherniak, D.J., Watson, E.B., 2007. Pre-eruption re- Wotzlaw, J.-F., Bindeman, I.N., Watts, K.E., Schmitt, A.K., Caricchi, L., Schaltegger, U., 2014.
charge of the Bishop magma system. Geology 35, 235–238. Linking rapid magma reservoir assembly and eruption trigger mechanisms at evolved
Watkins, S., Giordano, G., Cas, R., De Rita, D., 2002. Emplacement processes of the mafic Yellowstone-type supervolcanoes. Geology 42, 807–810.
Villa Senni Eruption Unit (VSEU) ignimbrite succession, Colli Albani volcano, Italy. Wright, H.N., Folkes, C., Cas, R.F., Cashman, K., 2011. Heterogeneous pumice populations
J. Volcanol. Geotherm. Res. 118 (1), 173–203. in the 2.08-Ma Cerro Galán Ignimbrite: implications for magma recharge and ascent
Willcock, M.A.W., Cas, R.A.F., Giordano, G., Morelli, C., 2013. The eruption, pyroclastic flow preceding a large-volume silicic eruption. Bull. Volcanol. 73 (10), 1513–1533.
behaviour, and caldera in-filling processes of the extremely large volume (1290 km3), Wright, H., Bacon, C., Vazquez, J., Sisson, T., 2012. Sixty thousand years of magmatic
intra- to extra-caldera, Permian Ora (Ignimbrite) Formation, Southern Alps, Italy. J. volatile history before the caldera-forming eruption of Mount Mazama, Crater Lake,
Volcanol. Geotherm. Res. 265 (2013), 102–126. Oregon. Contrib. Mineral. Petrol. 164 (6), 1027–1052.
Wilson, C.J.N., Charlier, B.L.A., 2009. Rapid rates of magma generation at contemporane- Yoshimoto, M., Fujii, T., Kaneko, T., Yasuda, A., Nakada, S., 2004. Multiple magma reser-
ous magma systems, Taupo Volcano, New Zealand: insights from U–Th model-age voirs for the 1707 eruption of Fuji volcano, Japan. Proc. Jpn. Acad. Ser. B 80, 103–106.
spectra in zircons. J. Petrol. 50 (5), 875–907. Zandt, G., Leidig, M., Chmielowski, J., Baumont, D., Yuan, X., 2003. Seismic detection and
Wilson, C., Hildreth, W., 1997. The Bishop Tuff: new insights from eruptive stratigraphy. J. characterization of the Altiplano–Puna magma body, Central Andes. Pure Appl.
Geol. 105, 407–440. Geophys. 160 (3–4), 789–807.
Wilson, C.J.N., Walker, G.P.L., 1985. The Taupo eruption, New Zealand I. General aspects: Zollo, A., Maercklin, N., Vassallo, M., Dello Iacono, D., Virieux, J., Gasparini, P., 2008. Seismic
philosophical transactions of the Royal Society of London. Ser. A Math. Phys. Sci. reflections reveal a massive melt layer feeding Campi Flegrei caldera. Geophys. Res.
314, 199–228. Lett. 35. http://dx.doi.org/10.1029/2008GL034242.

You might also like