You are on page 1of 18

Journal of Volcanology and Geothermal Research 288 (2014) 2845

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
b

University of Bristol, UK
Universit Roma Tre, Italy

a r t i c l e

i n f o

Article history:
Received 4 June 2014
Accepted 15 September 2014
Available online 6 October 2014
Keywords:
Explosive eruptions
Calderas
Magma storage
Syn-eruptive melt extraction

a b s t r a c t
Large caldera-forming eruptions have long been a focus of both petrological and volcanological studies; petrologists have used the eruptive products to probe conditions of magma storage (and thus processes that drive
magma evolution), while volcanologists have used them to study the conditions under which large volumes of
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, however, advances in analytical techniques have provided new views of magma storage regions, many of which provide evidence of multiple melt lenses feeding a single eruption, and/or rapid pre-eruptive assembly of large
volumes of melt. These new petrological views of magmatic systems have not yet been fully integrated into volcanological perspectives of caldera-forming eruptions. Here we explore the implications of complex magma reservoir congurations for eruption dynamics and caldera formation. We rst examine mac systems, where
stacked-sill models have long been invoked but which rarely produce explosive eruptions. An exception is the
2010 eruption of Eyjafjallajkull 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 mac
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 implications 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 successive melt lenses may be tapped. We conclude that emerging views of complex magma reservoir congurations
provide exciting opportunities for re-examining volcanological concepts of caldera-forming systems.
2014 Elsevier B.V. All rights reserved.

Contents
1.
2.
3.

4.

5.

Introduction why review calderas? . . . . . . . . . . . . . . . . . . . . . . . . . . .


Caldera-forming eruptions: an overview . . . . . . . . . . . . . . . . . . . . . . . . . .
Mac magmatic systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Stacked sill models of mac magmatic systems . . . . . . . . . . . . . . . . . . . .
3.2.
Caldera formation from complex mac magma reservoirs an example from Colli Albani
Storage and eruption from large silicic systems . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Pre-eruptive magma storage . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Time scales of magma accumulation . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Eruption triggers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.
Eruption dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5.
Caldera collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Implications for recognizing eruption potential of large magmatic systems . . . . . . . . . . .

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.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

29
30
31
31
32
35
35
35
36
37
38
39

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

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

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

However, there is a pleasure in recognizing old things from a new point


of view. Richard Feynman, 1948

1. Introduction why review calderas?


The importance of studying caldera-forming eruptions cannot be
under-estimated. Caldera-forming eruptions include some of the largest
volcanic events ever to affect the Earth. Many of these have produced
large volumes (N100 km3) of highly evolved crystal-poor melt. As a result, an enduring paradigm of both igneous petrology and volcanology
has been one of melt accumulation, evolution and eruption from a single
melt-dominated magma chamber. This conceptual metaphor has provided a framework for petrological models of magma evolution and differentiation, and for volcanological models of eruption initiation,
magma withdrawal and caldera collapse (Fig. 1). Key features include:
(1) development of stably zoned magma chambers by crystal fractionation, where crystal-liquid separation is driven by settling of individual

Fig. 1. The Standard Model of caldera formation. (A) Stably stratied magma chamber
forms over thousands of years by crystal settling and upward migration of volatiles.
(B) Eruption starts with Plinian activity through a single vent, driven primarily by volatile
exsolution. (C) Evacuation of magma causes under-pressurization and destabilization of
the reservoir. (D) Caldera forms by roof collapse.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

29

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

41
41
42
41
42

crystals or crystal plumes within a large batch of liquid that cools from
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
that are reversely zoned in composition and/or pre-eruptive temperature and pressure; and (4) caldera formation by collapse of an underpressured magma chamber after some fraction of magma has been
withdrawn, with the timing of collapse determined by the strength
and thickness of the overlying country rock relative to width of the
magma chamber.
Over the past few decades, however, detailed volcanological, petrological and geophysical studies of individual (intermediatesilicic) magmatic systems have shown that (1) magma storage regions are
composed primarily of crystalline mush (crystals plus interstitial liquid;
Fig. 2; e.g., Hildreth, 2004; Hildreth and Wilson, 2007; Lipman, 2007;
Bachmann and Bergantz, 2008; Reid, 2008; Bachmann, 2010; Deering
et al., 2011; Walker et al., 2013; Simon et al., 2014), (2) large melt volumes 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


(CR), and often zoned, eruptions are fed from a single melt body contained within a
much larger and more crystalline system comprising a crystal mush (~50% crystals) and
surrounding rigid sponge (N65% crystals); modied from Hildreth (2004). (B) Crystalrich (CR) eruptions occur by rejuvenation of a near-rigid crystal mush (by input of melt
and/or gas); modied from Bachmann and Bergantz (2006); Huber et al. (2011).

30

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

eruptions may be triggered either internally (chamber triggered) or


externally (roof triggered) depending on the tectonic setting, accumulated magma volume and roof aspect ratio (Lindsay et al., 2001; Jellinek
and DePaolo, 2003; de Silva et al., 2006; Gudmundsson, 2008, 2012;
Gottsmann et al., 2009; Gregg et al., 2012), and (4) large eruptions
may tap multiple and distinct melt sources and crystal populations (including phenocrysts, antecrysts and xenocrysts, e.g., Maughan et al.,
2002; Ellis et al., 2010; Wright et al., 2011; Allan et al., 2012; Cooper
et al., 2012; Ellis and Wolff, 2012; Vinkler et al., 2012; Gualda and
Ghiorso, 2013; Fig. 3). At the same time, geophysical studies of active
volcanic systems have failed to locate large volumes of crystal-poor
melt (e.g. Iyer et al., 1990; Masturyono et al., 2001; Schilling and
Partzsch, 2001; Sherburn et al., 2003; Zandt et al., 2003; Lees, 2007;
Chu et al., 2010). These observations are difcult to reconcile with a
classical magma chamber model sensu stricto, that is, a single, very
large, long-lived melt-dominated magma chamber (termed the Standard Model by Gualda and Ghiorso, 2013).
Here we review explosive caldera-forming eruptions and their products, paying particular attention to new conceptual models of complex
magma storage regions. By complex we refer to magma reservoirs
comprising multiple melt lenses within a partially to completely crystalline framework; such reservoir congurations have been advanced by
recent petrological studies of mac systems, but are consistent with
emerging views of some silicic systems (e.g., Fig. 3). We take this view
one step further by exploring the implications of complex magma reservoir geometries for syn-eruptive melt extraction, eruption dynamics
and caldera collapse. We end by examining the implications of different
magma storage models for hazard assessment, including geophysical
prospecting for magma reservoirs capable of producing very large
eruptions.
2. Caldera-forming eruptions: an overview
Excellent reviews of magmatic systems that produce large silicic
caldera-forming eruptions are provided by Hildreth (1981, 2004) and
Lipman (2007); here we summarize some pertinent points from these
and related studies. Large (100 km3 DRE) silicic eruptions can be classied by dominant magma type as either crystal-rich ( 35%) dacite
(often termed monotonous intermediates, MI) or crystal-poor ( 15%)
rhyolite (CPR). Magma within these systems is typically stored in silllike bodies, that is, they have horizontal dimensions that greatly exceed
the vertical dimension. Smaller (b 100 km3) caldera-forming eruptions

Fig. 3. Schematic view of the magmatic system that fed the very large (~1200 km3) rhyolitic Kidnappers eruption, Mangakino volcano, Taupo Volcanic Zone, New Zealand. The
eruption tapped three separate melt bodies distributed laterally along the rift.
Modied from Cooper et al. (2012).

are most commonly associated with stratovolcanoes, where storage regions tend to be vertically elongated.
The very largest eruptions almost always involve MI magma that is
(relatively) homogeneous in both composition and phenocryst content
(Fig. 4A). Resulting ignimbrites are typically composed of ash and broken crystals, with only scarce pumice clasts (b10%; e.g., Carter et al.,
1986; Bachmann et al., 2002; Gottsmann et al., 2009; Wright et al.,
2011). CPR eruptions, in contrast, initiate with pumice-bearing deposits
of crystal-poor (and often high-SiO2) rhyolite magma that may vary in
both crystallinity and bulk composition throughout the course of an
eruption (e.g., Hildreth, 1981; Deering et al., 2011; Pamukcu et al.,
2013; Fig. 4B). The high-SiO2 rhyolite melt often lacks a counterpart in
correlative plutonic sequences (although there are exceptions,
e.g., Walker et al., 2007), but overlaps the composition of matrix melts
in MI magmas (e.g., Lindsay et al., 2001; Lipman, 2007; Bachmann
et al., 2007; Fig. 4). For this reason, both magma types are inferred to
have a similar origin, with the difference being that MI eruptions evacuate the entire (rejuvenated) magma storage region, while CPR eruptions are dominated by the (segregated) melt phase (e.g., Bachmann
et al., 2007). Accumulation of rhyolitic residual melt prior to CPR eruptions is inferred to occur by crystal settling, compaction and/or lter
pressing of the crystal mush (e.g., Sisson and Bacon, 1999; Bachmann
and Bergantz, 2004; Bea, 2010; Dufek and Bachmann, 2010).
MI and CPR eruptions differ in eruption style and timing of caldera
formation. MI eruptions commonly lack an early Plinian (single vent,
high plume) phase and initiate, instead, with eruption of pyroclastic
density currents from faults along the caldera margin. Where the timing
of caldera collapse can be determined, it is coincident with the start of
the eruption (Sparks et al., 1985; Lindsay et al., 2001; de Silva et al.,
2006; Gregg et al., 2012; Willcock et al., 2013). As a result, associated
distal ash deposits derive mostly (or exclusively) from the coignimbrite plume (e.g., Chesner et al., 1991). CPR eruptions, in contrast,
typically start with a Plinian (high plume) phase, as recorded in widespread and voluminous fall deposits. With time, the vent widens
(often by propagating ring faults) and pyroclastic ows comprise an increasing proportion of the erupted volume. Caldera collapse occurs only
after withdrawal of a critical volume of magma that can be related to the
depth and geometry of the reservoir (e.g., Roche and Druitt, 2001; Geyer
et al., 2006; Geshi et al., 2014).
Smaller caldera-forming eruptions (~100 km3) encompass a wide
range of magma compositions and crystallinities (e.g., Hildreth, 1981)
and are typically associated with stratovolcanoes. When classied by
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 mac (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
DRE erupted volume. (A) CR data; yellow squares show matrix glass, orange bars show
bulk compositional range. (B) CP data; blue circles show matrix glass composition of earliest erupted magma, blue bars show bulk compositional range. Data sources are listed in
Table S1.

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

et al., 2012). By analogy with the larger systems, it is commonly assumed that evolved crystal-poor (CP) magma is segregated into a single
large body prior to eruption, and that late-erupted crystal-rich (CR)
magma is mobilized by recharge melts (e.g., Bacon and Druitt, 1988;
Pallister et al., 1992; Allen, 2001; Bachmann, 2010). When crystal-rich
magmas are erupted early (for example, Pinatubo 1991, Philippines;
Quilotoa 800 ybp, Ecuador), erupted magmas appear similar to larger
MI eruptions in their (general) homogeneity, bulk composition and
high crystallinity (e.g., Polacci et al., 2001; Rosi et al., 2004). Caldera collapse in these systems is attributed to magma withdrawal and underpressurization, and may happen at some point during the eruptive sequence (e.g., Druitt and Sparks, 1984).
Of the caldera-forming ignimbrite family, the smallest, and in many
ways the oddest, group is that of ultrapotassic (SiO2 b 55%) ignimbrites
found primarily in the Quaternary Roman Magmatic Province (QRMP;
Italy). The QRMP comprises four major caldera complexes that have
produced recurrent eruptions of tephritic to tephri-phonolitic ignimbrites with DRE (dense rock equivalent) volumes of 150 km3 (Fig. 5;
Giordano et al., 2006; Boari et al., 2009; Masotta et al., 2010; Freda
et al., 2011; Vinkler et al., 2012; Acocella et al., 2012). The recurrence
times for caldera-forming eruptions at each caldera complex are 40 to
50 ka, and caldera areas range from 30 to 300 km2. These eruptions
are curious because ultrapotassic magmas have low viscosities
(104.5 Pa s, even accounting for up to 30% syn-eruptive crystallization;
Vinkler et al., 2012; Campagnola, 2014) and are therefore susceptible
to gas escape during magma ascent. The eruption sequence of
ultrapotassic ignimbrites is, however, comparable to that of their silicic

31

counterparts, with an early fallout phase from a central single vent


followed by climactic, syn-collapse highly mobile ignimbrites with associated proximal collapse breccias (e.g. Giordano and Dobran, 1994;
Watkins et al., 2002). As is typical for other stratovolcano eruptions,
none of these caldera complexes has undergone post-collapse resurgence; instead, post-collapse subsidence, intracaldera volcanism and
sedimentation are common (Giordano et al., 2006; Acocella et al., 2012).
3. Mac magmatic systems
To explore relations between conditions of magma eruption and
storage, we rst briey review mac and mac ultrapotassic magmatic
systems, where both recent observations of eruptions (Tarasewicz et al.,
2012; Carbotte et al., 2013) and detailed petrological (e.g., Marsh, 2013;
Neave et al., 2013) and volcanological (e.g., Brown et al., 2014; Vinkler
et al., 2012) studies all point to syn-eruptive, and sometimes explosive,
tapping of multiple melt lenses stored within complex magma plumbing systems. Geochemical data suggest, moreover, that these melt
lenses may be vertically distributed throughout the crust. In fact, isotopic evidence for crustal assimilation of ood basalt magmas (e.g., Wolff
et al., 2008; Vye-Brown et al., 2013) requires that very large volumes of
mac melt accumulate within the crust prior to eruption.
3.1. Stacked sill models of mac magmatic systems
Mac magmatic systems are commonly envisioned as sequences of
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 eruption unit: black lines = isopachs of the basal scoria fall (shown in yellow); orange = Tufo Lionato pre-collapse ignimbrite; pink = Pozzolanelle climactic ignimbrite. (C) Chemical composition of VSN units: orange eld = 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) 2845

zones (e.g., Marsh, 1996; Annen et al., 2006; Gudmundsson, 2012). Sills
form when repeated magma injections are sufciently spaced in time to
allow complete cooling between injection events (e.g., Annen et al.,
2006; Burchardt, 2008); melt lenses dominate when new inputs are
added to systems that are still partially molten, or within sills that are
sufciently thick to allow internal redistribution of melt (e.g., Marsh,
2002, 2013). Variants of stacked sill models have been invoked to explain the petrologic diversity of mac magma erupted during individual,
often long-lived, eruptive episodes (e.g., Aarnes et al., 2008; Kelley and
Barton, 2008; Erlund et al., 2010; Dahren et al., 2012; Passmore et al.,

2012; Neave et al., 2013; van der Zwan et al., 2013). In these conceptual
models, each thermally zoned mac melt lens evolves by the progression of a solidication front that is multiply saturated, such that melt
composition is buffered along a eutectic/cotectic. At the same time, crystallization generates highly differentiated melt compositions that may
(1) segregate by compaction at intermediate crystallinities, (2) become
trapped within small pores in regions of high crystallinity (N 70%;
e.g., Dufek and Bachmann, 2010), or (3) segregate into lenses, pockets,
and bulbous masses (Marsh, 2002; Masotta et al., 2012).
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 downwardpropagating seismicity that accompanied the 2010 FimmvrduhlsEyjafjallajkull eruption in Iceland (Tarasewicz et al., 2012; Fig. 6). Evidence for at least two sill intrusions in the 1990s (Sigmundsson et al.,
2010) provides support for this interpretation. An important consequence was an eruption that involved several distinct explosive episodes 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 macintermediate 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 reect 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 solidication front (Marsh,
2002; Humphreys and Holness, 2010), local gas build-up caused by
volatile-saturated crystallization (Tait et al., 1989), changes in melt volume 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 mac magma reservoirs an
example from Colli Albani

Fig. 6. Time-progressive unloading of a stacked sill sequence beneath Eyjafjallajkull volcano, Iceland, 2010; illustrates response to a downward propagating decompression front
of a partially molten crustal magmatic system.
Redrafted from Tarasewicz et al. (2012).

The Eyjafjallajkull 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
mac ultrapotassic volcano Colli Albani (QRMP; Fig. 5), which has produced several large (3050 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 eruptions, 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 appearance between VSN1 and VSN2 of an intercalated co-ignimbrite breccia
with up to 30% deep-seated lithics (from ~1 to 6 km depth, including intrusives and cumulates). The breccia also contains an isotopically distinct K-foidite juvenile magma that does not appear elsewhere in the
VSN stratigraphy (Conticelli et al., 2010). The deposit characteristics,
as well as the ubiquitous vapor-phase lithication of the lower VSN1 ignimbrite, suggest that a (short?) hiatus preceded caldera collapse.

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

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 postcollapse deposits.

One unusual feature of the eruptive deposit relates to the pyroclast


vesicularities, which are generally very low (b~40% bulk, and b ~60%,
melt-referenced; Fig. 7), and contrast with typical crystal-poor silicic
pumice vesicularities of N 70% (e.g., Klug et al., 2002; Gurioli et al.,
2005; Adams et al., 2006; Houghton et al., 2010). At the same time, vesicle number densities of 108 cm3 reach those of other mac plinian
eruptions (Vinkler et al., 2012). These vesicle characteristics suggest extensive syn-eruptive outgassing, although concomitant with high rates
of magma decompression. And yet tens of cubic kilometers of low viscosity magma were erupted explosively to form two separate, largevolume ignimbrites. This apparent conundrum strongly suggests that
volatile exsolution is not necessarily the primary driving force for eruptive activity.
At the same time, geochemical examination of the eruptive products shows that extensive ( 60%) crystallization of clinopyroxene
and leucite are required to produce Colli Albani magma (Freda
et al., 1997, 2011; Peccerillo, 2005; Boari et al., 2009). This implies
that the bulk of the magma storage system was highly crystalline,
even though both early-erupted (pre-collapse) and very evolved
(syn-collapse foidite) magma are crystal-poor (Fig. 8a, b). Further
support for a largely crystalline magma reservoir can be found in

the ejection of both cognate cumulates and loose cumulate crystals


(up to 35% of mm- to cm-sized leucite, clinopyroxene and biotite;
Fig. 8c, d) during and after caldera collapse. Incorporation of cumulate crystals into the erupted melt explains the observed change in
bulk composition from tephri-phonolite to phono-tephrite, and
suggests that progressive disruption of the crystal framework
accompanied caldera collapse. Additionally, eruption of K-foidite in
association with syn-caldera breccias shows that at least one isolated melt lens was tapped at the time of caldera formation.
The data reviewed above low pyroclast vesicularity, time gaps,
multiple explosive episodes, variable melt and bulk compositions
are difcult to reconcile with magma extraction from a single, pressurized, well mixed, melt-dominated magma chamber. Instead, the
eruptive sequence has many elements that suggest involvement of a
complex magma reservoir composed of both laterally and vertically distributed melt lenses within a variably crystalline mush (Fig. 9). Most importantly, both the explosive nature of the eruptive activity and the
poorly vesicular scoria appear to require sustained overpressures to
drive the eruption. We suggest that a complex magma reservoir can explain these observations if overpressure is accommodated within individual melt lenses (e.g., Bagdassarov and Dorfman, 1998). In this

34

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

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


a sufciently high driving pressure to sustain fast withdrawal of even
volatile-poor magma. As a consequence, sustained melt extraction
would not require volatiles to provide the only driving force for eruption. Instead, as for oil or water extraction from an over-pressured reservoir, melt out-ow could be driven initially by release of stored
deformational energy within the (bubble-bearing?) magma, the (viscoelastic) crystalline parts of the reservoir, and the country rock. Vesiculation triggered by magma ascent and decompression would then
enhance magma ascent, particularly at low pressures. In this complex

magma reservoir model, time gaps reect modulation of magma withdrawal by the strength of the crystal framework (Fig. 10), and the
time required for connection of melt lenses to the main conduit
(e.g., Fig. 6).
To summarize, magma storage within a complex melt/mush reservoir helps to explain many unusual characteristics of the Villa Senni
eruption, including the stability afforded by distributing, rather than
concentrating melt, the potential for isolated lenses to develop both
high overpressures and highly evolved compositions, and the opportunity 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 ows
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 solidication 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) 2845

Fig. 10. Changes in magma strength as a function of melt volume fraction. Dashed lines are
t to experimental data on the Westerly granite (upper) and Delegate aplite (lower).
Adapted from Rosenberg et al. (2007).

throughout an eruptive sequence. Deposits from the VSN eruption further suggest that the timing and style of caldera collapse may be controlled by processes within the reservoir (specically, failure of parts
of the crystal network), in addition to processes external to the reservoir
(such as the geometry, thickness and mechanical properties of the roof
rock and caldera faults). More broadly, we highlight emerging views of
mac magma reservoirs as vertically extensive and comprising both
melt-rich and melt-poor (or melt-absent) regions. This view derives
not only from the geophysical, petrologic and volcanologic studies
reviewed above, but also from new thermal models that examine conditions required to develop complex storage regions (e.g., Annen et al.,
2006; Annen, 2011; Solano et al., 2012). More important from a
volcanological perspective, however, are the implications for conditions
leading to, and evolving during, volcanic eruptions from magma
reservoirs that contain multiple and variably connected melt lenses
(e.g., Gudmundsson, 2012).
4. Storage and eruption from large silicic systems
We now address the question of the extent to which complex (meltlens-dominated), as compared to simple (single melt body), magma
reservoirs can provide insight into processes that contribute to the
much more common eruptions of intermediate to silicic magmas. Our
goal is not to dismiss the Standard Model, but instead to evaluate the extent to which emerging, and sometimes conicting, observations about
very large explosive eruptions can be reconciled by broadening our
views of magmatic systems.
4.1. Pre-eruptive magma storage
Several recent studies of crystal-poor rhyolitic ignimbrites suggest
that multiple, rather than single, melt batches were tapped during individual caldera-forming eruptions. Evidence for multiple melt batches is
particularly common in extensional environments such as the Snake
River Plain (US; Ellis et al., 2010; Ellis and Wolff, 2012) and the Taupo
Volcanic Zone (TVZ, New Zealand; Brown et al., 1998; Charlier et al.,
2003; Gravley et al., 2007; Wilson and Charlier, 2009; Bgu et al.,
2014), and demonstrates the importance of crustal forcing on both
magma storage and eruption (e.g., Lindsay et al., 2001; Gottsmann
et al., 2009; Cooper et al., 2012; Allan et al., 2013). Two laterally
displaced (and non-communicating) melt lenses may also have been
tapped during the 600 km3 Bishop Tuff eruption from the Long Valley
caldera (Gualda and Ghiorso, 2013), which lies within a transtensional setting at the eastern edge of the Sierra Nevada and has long
been considered the iconic example of the Standard Model of a single
zoned magma chamber. In all cases, melt lenses are similar in bulk,
but distinct in trace element and isotopic, composition and were stored

35

in laterally extensive (rather than vertically elongated) reservoirs


(e.g., Fig. 3).
Monotonous (crystal-rich) ignimbrites (MI) may also preserve evidence of melt segregation. In fact, although the very name monotonous denotes homogeneity, careful examination of some large MI
deposits has shown that these systems may also be spatially heterogeneous. 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 outow sheet have distinct
compositional characteristics (Francis et al., 1989; Kay et al., 2011),
and compositionally distinct white and gray pumice clasts provide evidence for at least two magmas involved in the eruption (Folkes et al.,
2011; Wright et al., 2011). In fact, discrete evolved melt pockets are inferred even for systems that lack evidence for diverse melt compositions
(e.g., Huber et al., 2012; Willcock et al., 2013).
Compositionally zoned ignimbrites are perhaps most representative
of the Standard Model, in that they are interpreted to record top-down
evacuation of individual magma chambers. They also demonstrate the
fundamental role of mac magma in providing either heat (to partially
melt roof rocks) or evolved melt (from cooling and crystallization). Examples of the former include Gran Canaria (Freundt and Schmincke,
1995) and Iceland (Askja, Sigurdsson and Sparks, 1981), where mac
inputs are hot and water-poor. Examples of the latter are common in
water-rich environments, where (often cooler) crystal-poor silicic
magma overlies hotter (often more crystalline) mac magma
(e.g., Hildreth, 1981; Bacon and Druitt, 1988; Druitt and Bacon, 1989;
Deering et al., 2011; Pamukcu et al., 2013). As both lower temperatures
and lower PH2O promote crystallization (Blundy and Cashman, 2008;
Cashman and Blundy, 2013), the lower temperature of the dominant
crystal-poor rhyolite requires commensurate zoning in water unless
compositional differences are large. Zoned magmas are particularly
characteristic of (although not unique to) vertically elongated magmatic systems that feed arc stratovolcanoes, perhaps because vertically
elongated (and water-rich) reservoirs are less susceptible to convective
mixing than sill-like melt bodies (e.g., Maughan et al., 2002; Blundy and
Cashman, 2008).
A more unusual example of a zoned magma reservoir is provided by
the NovaruptaKatmai eruption of 1912, which involved 7.5 km3 of
crystal-poor high-silica rhyolite and 5.5 km3 of crystal-rich continuously
zoned dacite to andesite magma. The unusual aspect of the eruption is
not the compositional zonation but the observed caldera collapse at
Mt. Katmai, which lies 10 km from the eruptive vent of Novarupta. Detailed studies by Hildreth and Fierstein (2012) demonstrate that the
rhyolite was extracted from the intermediate composition magma and
that extraction and lateral transport of crystal-poor rhyolite from the
storage region beneath Mt. Katmai was necessarily rapid, perhaps occurring during the 5 days of recorded precursory activity. This raises
some interesting questions. First, to what extent does zoning observed
in ignimbrites provide direct evidence of zoning within magma reservoirs? Second, how can large amounts of rhyolite melt be extracted
from crystal mush zones both efciently (without accompanying crystals) and rapidly (in days)? An alternative suggestion is that the rhyolite
magma intruded as an ascending rhyolite dike that intersected the resident intermediate composition magma reservoir under Mount Katmai
(Eichelberger and Izbekov, 2000). Neither model provides a good explanation for the lateral displacement of the eruptive vent from the magma
storage region.
4.2. Time scales of magma accumulation
The time scale of magma accumulation prior to large eruptions has
been the subject of numerous recent studies that apply new diffusion
chronometers to observed crystal zoning patterns. A surprising result

36

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

of these studies is the suggestion that large volumes of silicic melt may
accumulate over short (centuries or less) time scales (e.g., Charlier
et al., 2007; Gualda et al., 2012; Allan et al., 2013; Fig. 11). Other studies
describe very short time scales (decades to years or even months) of
late-stage crystallization (e.g., Wark et al., 2007; Saunders et al., 2010;
Druitt et al., 2012; Gualda et al., 2012; Matthews et al., 2012) and/or incorporation of xenocrysts (e.g., Gardner et al., 2002) prior to large eruptions. Short timescales of magma accumulation are difcult to reconcile
with long times (104105 years) required for compaction-driven melt
extraction (e.g., McKenzie, 1984; Bachmann and Bergantz, 2004; Dufek
and Bachmann, 2010). An alternative mechanism for segregating silicic
magma invokes formation of melt channels or dikes within more crystalline parts of the magma reservoir; melt extraction in this scenario is
driven by either pore pressure response to an anisotropic stress eld
(e.g., Eichelberger et al., 2006; Allan et al., 2013) or rapid interconnection of isolated melt lenses (e.g., Eichelberger and Izbekov,
2000; Fig. 6). Rapid extraction and shallow accumulation of melt may
also be driven by perturbations of local stress elds surrounding crystal
mush zones. Perturbations could be caused by the arrival of new
magma inputs, gas exsolution (Sisson and Bacon, 1999), or tectonic
stresses (particularly extension related to rifting; Allan et al., 2012).

4.3. Eruption triggers


Critical to understanding caldera-forming eruptions is the consideration of processes that cause a stable melt/mush system to destabilize.
By denition, an eruption starts when the magmatic system becomes
connected to the surface. This connection can be established by
upward-propagating dikes driven by magmatic overpressure (an internal trigger), or downward-propagating faults generated by thermomechanical instabilities in the roof rocks (an external trigger;
e.g., Gudmundsson, 2008; Gregg et al., 2012). Upward-propagating
dikes are commonly invoked when there is evidence for intrusion of
hotter recharge melt (and/or volatile phase). Evidence of magma inux
may be preserved in the form of magmatic inclusions, banded pumice or
multiple pumice populations (e.g., Hildreth, 1981; Pallister et al., 1992;
Polacci et al., 2001; Rosi et al., 2004). Although most easily recognized
when mac magma is intruded into a silicic system, the intruding
magma may be silicic, in which case it is typically hotter and less

Fig. 11. Time constraints on melt accumulation prior to the Oruanui eruption, Taupo
Volcanic Zone, New Zealand. Orange curve represents the time required to construct the
primary magma reservoir (high-Si rhyolite), using FeMg interdiffusion in orthopyroxene
(opx). Purple curve represents late-stage re-equilibration of opx incorporated into the
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
that were intersected during the eruption.
Complied from Allan et al. (2013)).

evolved than the resident magma in the uppermost part of the system
(Eichelberger et al., 2006; Hildreth and Wilson, 2007; Wright et al.,
2011). The geometry of the magma reservoir may also control both
the extent of interaction of new hotter melt inputs with cooler mush
(e.g., Humphreys et al., 2009) and the interaction of volatiles with the
mush (Wright et al., 2012).
Late-stage disturbances to magmatic systems may also be recorded
as selective crystal dissolution (e.g., feldspar and/or quartz; Bachmann
et al., 2002), phenocryst rim growth (Wark et al., 2007; Saunders
et al., 2010; Druitt et al., 2012; Matthews et al., 2012; Allan et al.,
2013) or microlite formation (Pamukcu et al., 2012). Both dissolution
and phenocryst rim growth are commonly interpreted to reect intrusion of mac magma into the system; in the latter crystal growth results
from cooling of the mac input. In the absence of evidence for mac inputs, however, selective dissolution and new crystal growth can also be
explained by changes in PH2O in response to decompression or addition
of volatiles (Bachmann et al., 2002; Wark et al., 2007; Blundy and
Cashman, 2008; Matthews et al., 2012; Cashman and Blundy, 2013). Evidence for volatile transfer underlies the concept of gas sparging, whereby sufcient heat to unlock crystal networks is transferred by an inux
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
(e.g., Gottsmann et al., 2009). Fluxing with H2O-rich uids could unlock
crystal networks more rapidly, however, by resorbing anhydrous
phases. Introduction of CO2-rich uids, in contrast, would promote crystal growth, particularly of feldspar (e.g., Cashman and Blundy, 2013),
thereby strengthening crystal networks.
Early (precursory) phases of eruptive activity provide insight into
conditions required to initiate and sustain an eruption. Interestingly,
many eruptions are preceded by leaks from the magma reservoir. Examples include the eruption of the 200 km3, largely degassed, Pagosa
Peak dacite just before the very Fish Canyon Tuff eruption (FCT;
e.g., Bachmann et al., 2000), the explosive-to-effusive Cleetwood eruption that preceded the c. 50 km3 caldera-forming (zoned) eruption of
Crater Lake, USA by weeks to months (Bacon, 1983; Kamata et al.,
1993), and the small (~ 0.3 km3) explosive eruption that preceded,
probably by months, the 530 km3 (crystal-poor) Oruanui eruption in
New Zealand (Allan et al., 2012). In all three cases, precursory eruptions
clearly tapped the primary magmatic system, and yet did not immediately trigger the climactic eruption. In both the FCT and Oruanui examples, 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 phenocrysthosted 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 necessary to fully mobilize magma from the crystal-rich reservoir. Preeruptive crystal breakage may also occur in response to migration of recharge 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 ignimbrite (Wright et al., 2011). Heating accompanying melt migration
can also cause crystal rupture by volatile expansion within melt inclusions (Gualda et al., 2004; Bindeman, 2005). Finally, it has been suggested that crystal breakage could be a response to seismic shaking
(Gottsmann et al., 2009). In all cases, physical breakage of the crystal
framework would help to mobilize magma preparatory to the climactic
event.

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

4.4. Eruption dynamics


The triggering event can also determine the nature of initial eruptive
activity. For example, very large MI eruptions inferred to be triggered by
roof collapse (e.g., Jellinek and DePaolo, 2003; Gregg et al., 2012; de
Silva and Gregg, 2014) lack an early single vent (Plinian) phase
(e.g., Druitt and Sparks, 1984; Sparks et al., 1985; de Silva et al., 2006;
Cas et al., 2011; Chesner, 2012). They initiate instead with eruption of
poorly expanded pyroclastic density currents along bounding ring faults
(Willcock et al., 2013) that are sustained by very high mass uxes (Cas
et al., 2011; Lesti et al., 2011). In contrast, (often smaller) eruptions of
crystal-poor magma typically have protracted single vent phases prior
to caldera collapse (e.g., Crater Lake, USA (Bacon, 1983); AD 161
Taupo, New Zealand (Wilson and Walker, 1985); Minoan Santorini,
Greece (Druitt and Bacon, 1989; Sparks and Wilson, 1990); 39 ka
Campanian Ignimbrite, Italy (Rosi et al., 1999)). Initial vents may be
located either on marginal ring faults when eruptions tap large sills
(e.g., Hildreth and Mahood, 1986), or the summits of stratovolcanoes.
In these eruptions, the mass eruption rate probably increases with
time, particularly after caldera collapse, and pressure changes within
the magma reservoir may be preserved in pyroclast textures
(e.g., Bacon, 1983; Klug et al., 2002; Gurioli et al., 2005).
The mode of eruption will also affect the nature of the eruptive products. Ignimbrite deposits from large MI eruptions contain mostly ash
and pervasively shattered individual crystals, with limited abundance
of (often low vesicularity) pumice (e.g., Carter et al., 1986; Bachmann
et al., 2002; Gottsmann et al., 2009; Wright et al., 2011). Shattering of
crystals provides evidence of extensive disruption of magma during extraction from the reservoir (e.g., Maughan et al., 2002). When combined
with the low vesicularity of rare pumice clasts, these textural characteristics suggest that vesiculation played a limited role in the eruption process, which was probably dominated instead by catastrophic
decompression (Gottsmann et al., 2009). This scenario is similar to
that outlined above for Colli Albani. At the same time, large shear strains
may be imposed by magma mobilization along vertically extensive and

37

horizontally propagating yield surfaces (Karlstrom et al., 2012) and


around subsiding caldera blocks (Kennedy et al., 2008). These processes
would promote extensive syn-eruptive mixing, which could explain the
apparent dichotomy between the broad homogeneity of MI pyroclast
hand samples and the extreme complexity commonly recorded within
phenocryst populations. Alternatively, mobilization of crystal-rich
magma is common in late stage (syn- or post-collapse) eruptive activity
from (inferred) vertically zoned magma reservoirs (e.g., Bacon and
Druitt, 1988; Druitt and Bacon, 1989; Deering et al., 2011; Hildreth
and Fierstein, 2012; Pamukcu et al., 2013), where eruptions are well explained by the Standard Model of top-down magma withdrawal
(e.g., Bacon and Druitt, 1988; Allen, 2001; Mandeville et al., 2009). In
these examples, however, crystals are not pervasively broken.
The scarcity of pumice clasts in MI ignimbrites makes it difcult to
establish details of magma extraction. Deposits from crystal-poor rhyolite eruptions, in contrast, often produce abundant pumice that allows
individual parcels of magma to be related directly to their preeruptive storage conditions. Most informative are detailed studies of
phenocryst-hosted melt inclusions, which preserve dissolved volatiles
that can be used to estimate entrapment pressures, and major and
trace element compositions that can be used to track magma evolution.
From a volcanological perspective, an interesting observation is that
melt inclusion studies often indicate magma extraction from a large
pressure range, even very early in the eruptive sequence (e.g., Wallace
et al., 1999; Liu et al., 2006; Mangiacapra et al., 2008; Roberge et al.,
2013). One explanation for this could be pre-eruptive mixing of crystals
from throughout the magma storage region. Alternatively, magma
could be extracted from a large depth (pressure) range syn-eruptively
by lateral melt migration to a vertically extensive feeder dike.
The most thoroughly documented example is that of the Bishop Tuff,
where detailed stratigraphic and volcanological studies of the eruptive
deposits (Wilson and Hildreth, 1997) provide exceptional constraints
on both the timing and location of magma extraction from the underlying reservoir (Fig. 12A). Here quartz-hosted melt inclusions record pressures 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 (modied 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) 2845

deeper magma level may have been tapped late in the eruption
(Fig. 12B). Phase equilibrium constraints, in contrast, allow both early
and late Bishop magma to span the entire (100250 MPa) pressure
range (Gualda and Ghiorso, 2013). Here continuous magma extraction
from a large pressure range can be explained by lateral magma supply
to vertically extensive feeder dikes located on the caldera margin
(e.g., Gardner et al., 1991). Horizontally directed melt ow is also suggested by lateral propagation of ring faults during caldera collapse
(Wilson and Hildreth, 1997). Corresponding trace element analyses
show that late (post-collapse) eruptive activity tapped magma that
was both more and less evolved than prior to collapse (Fig. 12C),
which suggests late stage involvement of both less evolved melt lenses
(as also suggested from zircon data; Chamberlain et al., 2014) and more
evolved matrix melt, the latter perhaps extracted during caldera
collapse.
4.5. Caldera collapse
Comprehensive reviews of caldera collapse are provided in Lipman
(1997); Cole et al. (2005); Acocella (2007); Marti et al. (2008). These reviews focus largely on structural controls on caldera collapse styles, a
topic that is beyond the scope of this review. Instead we explore the relation between melt storage and caldera formation, as indicated by variations in the timing of caldera formation within an eruptive sequence.
Caldera formation after evacuation of substantial magma volumes has
been interpreted as a consequence of under-pressurization of the
magma storage region (Druitt and Sparks, 1984; Scandone, 1990;
Mart, 1991; Branney, 1995; Lipman, 1997; Cole et al., 2005). Caldera
collapse coincident with the onset of eruptive activity, in contrast,
suggests that pressurization and pre-eruptive doming caused by shallow magma accumulation may trigger collapse of large calderas
(e.g., Gudmundsson, 2008; Gregg et al., 2012; de Silva and Gregg,
2014). Also important is the tectonic stress eld, which can control
the location of caldera-bounding faults (e.g., Holohan et al., 2008a).
Models of caldera collapse (e.g., Roche and Druitt, 2001; Geyer et al.,
2006; Stix and Kobayashi, 2007; Geshi et al., 2014) typically measure
the timing of collapse by f, the fraction of the total (DRE) magma volume
that is erupted prior to the onset of collapse. The value of f can be related
to the roof rock strength and magma chamber aspect ratio (or roof aspect ratio R = thickness/width; Roche and Druitt, 2001; Fig. 13). Collapse is assumed to occur when the roof can no longer support the
(under-pressured) magma chamber. Theoretical analysis suggests that
collapse begins earlier (smaller f) for magma chambers that are shallow
and wide compared to those that are deep and narrow; these models
also predict that collapse will be piston-like when R b ~ 1.4, and

Fig. 13. Volume ratio f of pre-collapse to total eruption volume as a function of R, the ratio
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
reservoir was evacuated. Specic eruptions are labeled.
Modied from Roche and Druitt (2001).

incoherent when R N ~1.4. Conditions of f = 0, that is, where collapse


is synchronous with the start of an eruption, require coupling between
shallow, laterally extensive magma chambers and the overlying roof
rocks (Gregg et al., 2012). Model predictions have been tested experimentally with analogue magmas that are withdrawn steadily from
chambers. Fluids used in analogue experiments include air ( =
105 Pa s), water ( = 10 3 Pa s), and silicone ( = 104 Pa s), and
thus span a range of viscosities. All experiments, however, investigate
steady withdrawal of uid from a single (simple) reservoir; caldera formation by uid extraction from complex reservoirs has not yet been
explored.
Variations in f can be evaluated as a function of both eruption magnitude (DRE volume) and eruption type (crystal-rich [CR] or crystalpoor [CP]; Fig. 14A; Table S1 in Supplementary material). These data
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 nonzero) 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 stratovolcano eruptive sequences (e.g., Vesuvius [VS], Tambora [TMB] and
Aso [AS]), even for CR magma (e.g., Pinatubo [PN]) or substantial
erupted volumes (e.g., AS). This compilation shows that collapse
can occur over a wide range of f for similar total erupted volumes
(i.e., after very different volumes of magma extraction from the reservoir), and at similar f for erupted volumes that vary over two orders of
magnitude. Some of this variation can be explained by differences in
magma chamber geometry (R), but the extreme variability must reect
other controlling factors, including the distribution of melt within a
crystalline reservoir.
To relate f directly to conditions of magma withdrawal requires that
we know the thickness (pressure) of the melt lens evacuated during collapse. If magma is assumed to be withdrawn from a single, sill-like reservoir, and if the surface expression of the caldera can be taken as the
footprint of that reservoir, then the total thickness of magma extracted during an eruption can be estimated using the erupted (DRE) volume
and caldera area, a value commonly referred to as the collapse height.
Early studies identied a linear relation between DRE volume and collapse area (Smith, 1979; Spera and Crisp, 1981) that suggested a constant thickness of magma evacuation characterized many collapse
events. This thickness could be related to the tensile strength of the
crust (Walker, 1984; Scandone, 1990). Our re-examination of calculated
collapse heights for all sufciently well characterized eruptions (120 in
total) in the Collapse Caldera Data Base (Geyer and Marti, 2008;
Table S1; Fig. 14B) shows that many eruptions do have collapse heights
that cluster at a single value ~ 1000 m, although the data are highly
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 inferred. 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 eruption triggers. The dashed line on the diagram denes a caldera area
A = 100 km2, which has been identied as a thermomechanical boundary that separates eruptions triggered by (external) roof collapse or (internal) chamber collapse (Gregg et al., 2012; de Silva and Gregg, 2014).
CR eruptions tend to lie above this boundary (in the roof-collapse region), a placement that is consistent with the observed synchroneity between eruption initiation and caldera collapse. Data from many CP
eruptions also lie above this line, however, even for eruptions with

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

Fig. 14. Relations between erupted volume (DRE), fraction of magma erupted prior to collapse (f) and inferred collapse height (calculated as erupted volume/caldera area). (A) f vs.
volume; (B) volume vs. inferred collapse height. Symbols are the same in both, with blue
circles CP magma and yellow squares CR magma. Lines in (A) show pre-eruptive DRE volumes (labeled, in km3). Dashed line in (B) shows the contour for a caldera area of 100 km2,
which Gregg et al. (2012) suggest as the bounding limit between chamber-trigged and
roof-triggered (yellow shading) eruptions. Labeled eruptions are as follows: AR = Aira,
AS = Aso, CEB = Ceboruco, CG = Cerro Galan, CL = Crater Lake, CMP = Campanian
(Campi Flegrei), KT = Katmai, KOS = Kos, KRA = Krakatau, LG = La Garita (Fish Canyon
Tuff), LV = Long Valley (Bishop Tuff), NYT = Neapolitan Yellow Tuff (Campi Flegrei),
PN = Pinatubo, SAN = Santorini (Minoan), TMB = Tambora, TP = Taupo (181 AD),
VICO = Vico, VS = Vesuvius (79 AD), VSN = Villa Senni (Colli Albani). All data from
Table S1.

very large f values (e.g., Campanian [CMP] and Aso [AS]) that denote late
stage collapse (which is not consistent with a roof trigger). These discrepancies indicate that caldera area is not the sole control on eruption
triggering, which will also be affected by magma storage depth, reservoir conguration, magma input and tectonic triggering (e.g. Lindsay
et al., 2001; Allan et al., 2012). Support for the latter includes the relative
placement of small to moderate (b100 km3) stratovolcano eruptions
(e.g., Tambora [TM], Krakatau [KR] Crater Lake [CL] and Santorini [SN];
A b 100 km2) compared with eruptions of similar sizes in extensional
settings (e.g., Taupo [TP]; A N 100 km2).
Ideally, the entire vertical extent of magma extraction (the drainage
height) should be recorded by the volatile contents of phenocrysthosted melt inclusions (Wallace et al., 1999) and/or the stability of phenocryst assemblages (e.g., Hammer et al., 2002; Gualda and Ghiorso,
2013). There are very few caldera-forming eruptions, however, for
which the drainage height is well constrained. An exception is the Bishop Tuff, where both melt inclusion and phase equilibria data suggest
magma withdrawal over 130 MPa (~ 5000 m; Fig. 12B). This value

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 23 km indicated by analysis of drill core samples taken from
within the caldera (Hildreth and Mahood, 1986). At face value, this discrepancy implies that on average, melt was extracted from less than 1/
21/3 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 largescale 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 discrepancies 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 drainage 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
34 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 eruptions are not all fed by single magma bodies, and that accumulations
of eruptible melt do not necessarily require assembly over long time periods. Instead, some systems may store melt within multiple sills (within a rigid framework) or lenses (within a crystal mush) that may either
amalgamate into a single melt body shortly before eruption, or may be
tapped syn-eruptively, particularly in extensional environments. In
these complex magma reservoirs, pulsed interconnection of isolated
melt lenses promoted by syn-eruptive depressurization could both prolong explosive activity and explain commonly observed hiatuses in
eruptive sequences (e.g., Aramaki, 1984; Allen, 2004; Palladino and
Simei, 2005; Bear et al., 2009; Vinkler et al., 2012), particularly if adjustments 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
may also bear an unexplored inuence on the development of ring
faults (e.g., Kennedy et al., 2004; Holohan et al., 2008b; Burchardt and
Walter, 2010) and extent of collapse in different sectors of a caldera.
In fact, it seems likely that the spatial distribution of melt within a complex reservoir will contribute to the identied spectrum of piston,
downsag, trapdoor and piecemeal collapse styles (e.g., Cole et al.,
2005). Taken together, we suggest that the full range of eruptive conditions (precursors, triggers and eruption dynamics) created by tapping
complex storage regions has yet to be explored, and represents exciting
opportunities for future research.
5. Implications for recognizing eruption potential of large
magmatic systems
Effective volcano monitoring requires identication of systems
capable of producing very large eruptions. For this reason, several
potentially active volcanic regions have been the recent targets of
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) 2845

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 sufcient 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


geophysical images assess melt contents at ~30% (e.g., Schilling
and Partzsch, 2001; Zandt et al., 2003; Chu et al., 2010; Luttrell
et al., 2013), a number that appears safely at odds with the 50%
melt considered necessary for melt to be eruptible (e.g., Bachmann
and Bergantz, 2008). New views of magmatic systems, however,
show that large volumes of melt may accumulate rapidly, and that
multiple magma lenses may be tapped during a single eruptive episode. Syn-eruption melt extraction from a largely crystalline reservoir seems likely for eruptions such as Aso, where a large volume

(~ 200 km3) of magma was erupted under conditions of f N 0.9 but


R b 1 (estimated from reported caldera area of 200 km3 and H2O contents ~ 57 wt.% (Kaneko et al., 2007), which places the magma at
depths no greater than 810 km, R ~ 0.50.67). Also suggestive is evidence (from both melt inclusion and phase equilibria) for magma
extraction over much larger pressure (depth) ranges than the collapse height calculated from erupted volume and caldera area. Within this framework, we suggest that magma reservoirs such as those
imaged beneath Yellowstone may actually be capable of producing
a large eruption (e.g., Wotzlaw et al., 2014).

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

Another challenge relates to monitoring complex magmatic systems. An increasingly important and effective volcano monitoring technique is measurement of surface deformation, particularly using
satellite-based Interferometric Synthetic Aperture Radar (InSAR;
e.g., Sparks et al., 2012; Pyle et al., 2013). Surface deformation over
large, shallow and sill-like magma bodies (that is, those susceptible to
roof triggers) should be substantial; this makes them particularly good
targets for monitoring by InSAR. One complication, however, is that
large magmatic reservoirs often have well-developed active hydrothermal systems that may show extensive deformation related to shallow
changes in pore-pressure and water levels (e.g., Chiodini et al., 2003;
Husen et al., 2004; Chang et al., 2007). Pore-pressure-generated deformation signals may either mask or mimic magmatic activity. Surface deformation can also provide evidence of deep intrusive activity, such as
that currently ongoing at Uturuncu volcano, Bolivia (e.g., Sparks et al.,
2008), and thus provides a potential tool for tracking long-term migration of magma inputs at different crustal levels.
Magmatic activity that precedes internally triggered eruptions (that
is, triggers involving intrusion of gas or hot melt from below, or overpressurization caused by crystallization and associated gas exsolution)
may be more difcult to recognize. One potential precursor is early
magma leakage from large reservoirs (e.g., Bacon, 1983; Dufeld,
1990; Dufeld et al., 1990; Hildreth, 2004; Fabbro et al., 2013). An interesting observation is that precursory leaks from large magmatic systems
are often sourced from shallow levels and may produce either unusual
low energy fountains (e.g., Dufeld, 1990; Bachmann et al., 2000) or
lava ows (e.g., Bacon, 1983), despite tapping volatile-rich components
of the magmatic system (e.g., Dufeld and Dalrymple, 1990; Bacon et al.,
1992; Mandeville et al., 2009). A good illustration of this phenomenon,
and a cautionary tale for event-tree-based hazard analysis, is provided
by the eruptive sequence at Crater Lake, OR (Bacon, 1983). Here a composite eruption (Llao Rock; 1.7 km3 DRE pumice fall and 0.5 km3 DRE
lava ow) tapped the main magma reservoir about 200 years before
caldera formation. Another composite eruption (Cleetwood; 1.5 km3
DRE pumice fall and 0.6 km3 DRE lava ow) preceded the main (calderaforming) phase of the eruption by only weeks to months (Kamata
et al., 1993). In both cases, the magma apparently came from the
climactic reservoir (Bacon et al., 1992; Mandeville et al., 2009) and yet
each eruption transitioned from explosive to effusive. Why, then, was
the Cleetwood eruption followed so promptly by a very large
(~ 50 km3 DRE) explosive eruption from the same magmatic system?
One possible explanation is that the precursory eruptions tapped relatively shallow and partially to fully isolated melt lenses within a larger
reservoir. Magma withdrawal from the Cleetwood melt lens could
then have triggered the climactic event by either downward or lateral
propagation of a decompression wave capable of connecting the isolated lens to the larger reservoir. This scenario illustrates the importance of
developing methods to monitor processes internal to magma reservoirs
(e.g., induced seismicity; Catalli et al., 2013) during, as well as prior to
eruptions, and to distinguish between precursors and the main event
(e.g., Allan et al., 2012).
Summary
It has long been known that caldera-forming eruptions may evacuate large volumes of either crystal-rich or crystal-poor magma
(e.g., Hildreth, 1981). The past ten years have seen a growing number
of studies that relate the chemical and physical conditions in magma
storage regions to the conditions under which different parts of the system 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;
Gregg et al., 2012; Hildreth and Fierstein, 2012; Huber et al., 2012;
Karlstrom et al., 2012; Gualda and Ghiorso, 2013). Key observations
arising from these studies include: (1) many large eruptions tap multiple melt sources, (2) large melt bodies are probably transient features,

41

(3) crystals carried by the transporting melt have been stored at a


range of pressures and temperatures, and (4) eruptions of crystal-rich
magma are probably driven by roof collapse and fragmented by sudden
decompression (with a limited role for volatile exsolution and expansion). These observations require new models to explain and anticipate
triggering and eruption of magma from complex storage reservoirs.
We address this question by rst considering eruptions from mac
systems, which are commonly modeled as stacked sills. Here both direct
observations and petrologic studies of recently active systems show that
single eruptions may tap multiple melt lenses. One consequence is
pulsatory (and often protracted) eruptive activity and alternation between explosion and lava effusion; another is eruption of a range of
magma compositions. We extend these modern examples to larger,
caldera-forming eruptions using an example from the ultrapotassic
Quaternary Roman Magmatic Province (QRMP), where we suggest
that apparently contradictory observations of protracted explosivity
and low pyroclast vesicularity can be reconciled if the eruption tapped
a complex reservoir containing multiple melt lenses.
We then turn to the more common, and larger, eruptions of silicic
crystal-poor (CP) and crystal-rich (CR) magma, and review conditions
of magma storage, time scales of melt accumulation, eruption triggers,
eruption dynamics and conditions of caldera collapse. We show where
the Standard Model (single magma chamber) appears consistent with
the nature and stratigraphy of the eruptive products, as well as examples where a complex (melt-lens-dominated) magma reservoir may
better explain both petrological and volcanological observations. Most
important is the growing evidence that in many systems that erupt (initially) crystal-poor silicic magma, particularly those in extensional environments, melt may be stored within multiple isolated and/or partially
connected lenses. This type of complex storage geometry could provide
a mechanism for rapid assembly of large melt bodies, as well as a local
source of triggering (recharge) magma. In fact, Eichelberger et al.
(2006) extend this concept to suggest that a combination of porous
media and triggered dike ow could allow sufciently rapid syneruptive melt extraction to feed a silicic Plinian eruption, as we have inferred for the QRMP example. From a hazard perspective, eruptions fed
from complex magma reservoirs may show abrupt changes in eruptive
processes (such as pauses and transitions between explosive and effusive activity) that pose challenges for volcano monitoring and forecasting efforts.
From a heuristic perspective, magma storage in complex, rather than
simple, magma reservoirs resolves several existing paradoxes about
conditions of both pre-eruptive magma storage and syn-eruptive
magma withdrawal and caldera collapse. This perspective also represents a logical extension of recent attempts to reconcile petrological
views of incremental assembly of plutons with volcanological requirements of instantaneous availability of very large melt volumes
(Hildreth, 2004; Bachmann et al., 2007; Lipman, 2007; Walker et al.,
2007). Melt accumulation within, and syn-eruptive extraction from, interconnected melt lenses also alleviates problems related to maintaining 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
within a single coherent framework apparently disparate observations
related to magma extraction and caldera collapse. Finally, such a
model is consistent with geophysical observations that even active
magma storage reservoirs often contain 30% melt (Zandt et al.,
2003; Chu et al., 2010); the latter observation has important
implications for recognizing the eruption potential of large magma
reservoirs.

Acknowledgments
This work was supported by the AXA Research Fund through a
Research Professorship to KC and by Regione Lazio (818000-2009-RM-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) 2845

by G. Gualda and S. Burchardt and for the encouragement of the editor


(L. Wilson) to write this review.
Appendix A. Supplementary data
Supplementary data to this article can be found online at http://dx.
doi.org/10.1016/j.jvolgeores.2014.09.007.
References
Aarnes, I., Podladchikov, Y.Y., Neumann, E.-R., 2008. Post-emplacement melt ow induced
by thermal stresses: implications for differentiation in sills. Earth Planet. Sci. Lett. 276
(12), 152166.
Acocella, V., 2007. Understanding caldera structure and development: an overview of analogue models compared to natural calderas. Earth Sci. Rev. 85 (34), 125160.
Acocella, V., Palladino, D.M., Cioni, R., Russo, P., Simei, S., 2012. Caldera structure, amount
of collapse, and erupted volumes: the case of Bolsena caldera, Italy. Geol. Soc. Am.
Bull. 124 (910), 15621576.
Adams, N.K., Houghton, B.F., Hildreth, W., 2006. Abrupt transitions during sustained explosive eruptions: examples from the 1912 eruption of Novarupta: Alaska. Bull.
Volcanol. 69 (2), 189206.
Allan, A.S.R., Wilson, C.J.N., Millet, M.-A., Wysoczanski, R.J., 2012. The invisible hand: tectonic triggering and modulation of a rhyolitic supereruption. Geology 40 (6),
563566.
Allan, A.R., Morgan, D., Wilson, C.N., Millet, M.-A., 2013. From mush to eruption in centuries: assembly of the super-sized: Oruanui magma body. Contrib. Mineral. Petrol. 166
(1), 143164.
Allen, S.R., 2001. Reconstruction of a major caldera-forming eruption from pyroclastic deposit characteristics: Kos Plateau Tuff, eastern Aegean Sea. J. Volcanol. Geotherm. Res.
105 (12), 141162.
Allen, S.R., 2004. Complex spatter- and pumice-rich pyroclastic deposits from an andesitic
caldera-forming eruption: the Siwi pyroclastic sequence, Tanna, Vanuatu. Bull.
Volcanol. 67, 2741.
Annen, C., 2011. Implications of incremental emplacement of magma bodies for magma
differentiation, thermal aureole dimensions and plutonismvolcanism relationships.
Tectonophysics 500, 310.
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicic magmas
in deep crustal hot zones. J. Petrol. 47, 505539.
Aramaki, S., 1984. Formation of the Aira Caldera, Southern Kyushu, 22,000 years ago. J.
Geophys. Res. 89, 84858501.
Bachmann, O., 2010. The petrologic evolution and pre-eruptive conditions of the rhyolitic
Kos Plateau Tuff (Aegean arc). Cent. Eur. J. Geosci. 2 (3), 270305.
Bachmann, O., Bergantz, G.W., 2004. On the origin of crystal-poor rhyolites: extracted
from batholithic crystal mushes. J. Petrol. 45 (8), 15651582.
Bachmann, O., Bergantz, G.W., 2006. Gas percolation in upper-crustal magma bodies as a
mechanism for upward heat advection and rejuvenation of silicic crystal mushes. J.
Volcanol. Geotherm. Res. 149, 85102.
Bachmann, O., Bergantz, G.W., 2008. Rhyolites and their source mushes across tectonic
settings. J. Petrol. 49, 22772285.
Bachmann, O., Dungan, M.A., Lipman, P.W., 2000. Voluminous lava-like precursor to a
major ash-ow tuff: low-column pyroclastic eruption of the Pagosa Peak Dacite,
San Juan volcanic eld, Colorado. J. Volcanol. Geotherm. Res. 98 (14), 153171.
Bachmann, O., Dungan, M.A., Lipman, P.W., 2002. The Fish canyon magma body, San Juan
Volcanic Field, Colorado: rejuvenation and eruption of an upper-crustal batholith. J.
Petrol. 43 (8), 14691503.
Bachmann, O., Miller, C.F., de Silva, S.L., 2007. The volcanicplutonic connection as a stage
for understanding crustal magmatism. J. Volcanol. Geotherm. Res. 167 (14), 123.
Bacon, C.R., 1983. Eruptive history of Mount Mazama and Crater Lake caldera, Cascade
Range, U.S.A. J. Volcanol. Geotherm. Res. 18, 57115.
Bacon, C.R., Druitt, T.H., 1988. Compositional evolution of the zoned calcalkaline magma
chamber of Mount Mazama, Crater Lake, Oregon. Contrib. Mineral. Petrol. 98 (2),
224256.
Bacon, C.R., Newman, S., Stolper, E., 1992. Water, CO2, Cl, and F in melt inclusions in phenocrysts from three Holocene explosive eruptions, Crater Lake, Oregon. Am. Mineral.
77, 10211030.
Bacon, C.R., Lanphere, M.A., Champion, D.E., 1999. Late Quaternary slip rate and seismic
hazards of the West Klamath Lake fault zone near Crater Lake, Oregon Cascades. Geology 27, 4346.
Bagdassarov, N., Dorfman, A.M., 1998. Viscoelastic behaviour of partially molten granites.
Tectonophysics 290, 2745.
Bea, F., 2010. Crystallization dynamics of granite magma chambers in the absence of regional stress: multiphysics modeling with natural examples. J. Petrol. 51 (7),
15411569.
Bear, A., Cas, R., Giordano, G., 2009. Variations in eruptive style and depositional processes
associated with explosive, phonolitic composition, caldera-forming eruptions: the
151 ka Sutri eruption, Vico Caldera, central Italy. J. Volcanol. Geotherm. Res. 184
(3), 225255.
Beddoe-Stephens, B., Aspden, J.A., Shepherd, T.J., 1983. Glass inclusions and melt compositions of the, Toba Tuffs, northern Sumatra. Contrib. Mineral. Petrol. 83 (34),
278287.
Bgu, F., Deering, C.D., Gravley, D.M., Kennedy, B.M., Chambefort, I., et al., 2014. Extraction, storage and eruption of multiple isolated magma batches in the paired Mamaku
and Ohakuri eruption, Taupo Volcanic Zone, New Zealand. J. Petrol. 55, 16531684.

Bindeman, I.N., 2005. Fragmentation phenomena in populations of magmatic crystals.


Am. Mineral. 90, 18011815.
Blundy, J., Cashman, K., 2008. Petrologic reconstruction of magmatic system variables and
processes. Rev. Mineral. Geochem. 69, 179239.
Boari, E., Avanzinelli, R., Melluso, L., Giordano, G., Mattei, M., De Benedetti, A.A., Morra, V.,
Conticelli, S., 2009. Isotope geochemistry (SrNdPb) and petrogenesis of leucitebearing rocks from Colli Albani volcano, Roman Magmatic Province, Central Italy:
inferences on volcanic evolution. Bull. Volcanol. 71 (9), 9771005.
Branney, M.J., 1995. Downsag and extension at calderas: new perspectives on collapse geometries from ice-melt, mining, and volcanic subsidence. Bull. Volcanol. 57 (5), 303318.
Brown, S.J.A., Burt, R.M., Cole, J.W., Krippner, S.J.P., Price, R.C., Cartwright, I., 1998. Plutonic
lithics in ignimbrites of Taupo Volcanic Zone, New Zealand; sources and conditions of
crystallisation. Chem. Geol. 148 (12), 2141.
Brown, R.J., Blake, S., Thordarson, T., Self, S., 2014. Pyroclastic edices record vigorous lava
fountains during the emplacement of a ood basalt ow eld, Roza Member, Columbia River Basalt Province, USA. Geological Society of America Bulletin 126, 875891.
Burchardt, S., 2008. New insights into the mechanics of sill emplacement provided by
eld observations of the Njardvik Sill, Northeast Iceland. J. Volcanol. Geotherm. Res.
173, 280288.
Burchardt, S., Walter, T.R., 2010. Propagation, linkage, and interaction of caldera
ringfaults: comparison between analogue experiments and caldera collapse at
Miyakejima, Japan, in 2000. Bull. Volcanol. 72 (3), 297308.
Campagnola, S., 2014. Large scale Plinian eruptions of the Colli Albani and the Campi
Flegrei volcanoes: insights from textural and rheological studies. Universit Roma
Tre, (Unpublished PhD thesis).
Carbotte, S.M., Marjanovi, M., Carton, H., Mutter, J.C., Canales, J.P., Nedimovi, M.R., et al.,
2013. Fine-scale segmentation of the crustal magma reservoir beneath the East Pacific Rise. Nat. Geosci. 6, 866870.
Carter, N.L., Ofcer, C.B., Chesner, C.A., Rose, W.I., 1986. Dynamic deformation of volcanic
ejecta from the Toba caldera: possible relevance to Cretaceous/Tertiary boundary
phenomena. Geology 14 (5), 380383.
Cas, R.A., Wright, H.M., Folkes, C.B., Lesti, C., Porreca, M., Giordano, G., Viramonte, J.G.,
2011. The ow dynamics of an extremely large volume pyroclastic ow, the 2.08Ma Cerro Galn Ignimbrite, NW Argentina, and comparison with other ow types.
Bull. Volcanol. 73 (10), 15831609.
Cashman, K., Blundy, J., 2013. Petrological cannibalism: the chemical and textural consequences of incremental magma body growth. Contrib. Mineral. Petrol. 166 (3),
703729.
Catalli, F., Meier, M., Wiemer, S., 2013. The role of Coulomb stress changes for injectioninduced seismicity: the Basel enhanced geothermal system. Geophys. Res. Lett. 40, 7277.
Chamberlain, K.J., Wilson, C.J.N., Wooden, J.L., Charlier, B.L.A., Ireland, T.R., 2014. New perspectives on the Bishop Tuff from zircon textures, ages and trace elements. J. Petrol.
55 (2), 395426.
Chang, W.L., Smith, R.B., Wicks, C., Farrell, J., Puskas, C.M., 2007. Accelerated uplift
and magmatic intrusion of the Yellowstone caldera, 2004 to 2006. Science 318,
952956.
Charlier, B.L.A., Peate, D.W., Wilson, C.J.N., Lowenstern, J.B., Storey, M., Brown, S.J.A., 2003.
Crystallisation ages in coeval silicic magma bodies: 238U230Th disequilibrium evidence from the Rotoiti and Earthquake Flat eruption deposits, Taupo Volcanic Zone,
New Zealand. Earth Planet. Sci. Lett. 206 (34), 441457.
Charlier, B.L.A., Wilson, C.J.N., Lowenstern, J.B., Blake, S., Van Calsteren, P.W., Davidson, J.P.,
2005. Magma generation at a large, hyperactive silicic volcano (Taupo, New Zealand)
revealed by UTh and UPb systematics in zircons. J. Petrol. 46 (1), 332.
Charlier, B.L.A., Bachmann, O., Davidson, J.P., Dungan, M.A., Morgan, D.J., 2007. The upper
crustal evolution of a large silicic magma body: evidence from crystal-scale RbSr isotopic heterogeneities in the Fish Canyon magmatic system, Colorado. J. Petrol. 48
(10), 18751894.
Chesner, C.A., 2012. The Toba caldera complex. Quat. Int. 258, 518.
Chesner, C.A., Rose, W., Deino, A., Drake, R., Westgate, J., 1991. Eruptive history of Earth's
largest Quaternary caldera (Toba, Indonesia) claried. Geology 19 (3), 200203.
Chiodini, G., Todesco, M., Caliro, S., Del Gaudio, C., Macedonio, G., Russo, M., 2003. Magma
degassing as a trigger of bradyseismic events: the case of Phlegrean Fields (Italy).
Geophys. Res. Lett. 30 (8).
Chu, R., Helmberger, D.V., Sun, D., Jackson, J.M., Zhu, L., 2010. Mushy magma beneath
Yellowstone. Geophys. Res. Lett. 37, L01306.
Cioni, R., Pistolesi, M., Bertagnini, A., Bonadonna, C., Hoskuldsson, A., Scateni, B., 2014. Insights into the dynamics and evolution of the 2010 Eyjafjallajkull summit eruption
(Iceland) provided by volcanic ash textures. Earth Planet. Sci. Lett. 394, 111123.
Cole, J.W., Milner, D.M., Spinks, K.D., 2005. Calderas and caldera structures: a review.
Earth Sci. Rev. 69 (12), 126.
Conticelli, S., Boari, E., Avanzinelli, R., De Benedetti, A.A., Giordano, G., Mattei, M., et al.,
2010. Geochemistry, isotopes and mineral chemistry of the Colli Albani volcanic
rocks: constraints on magma genesis and evolution. The Colli Albani Volcano. Spec.
Publ. IAVCE I 3, 107139.
Cooper, G.F., Wilson, C.J.N., Millet, M.-A., Baker, J.A., Smith, E.G.C., 2012. Systematic tapping of independent magma chambers during the 1 Ma Kidnappers supereruption.
Earth Planet. Sci. Lett. 313314, 2333.
Dahren, B., Troll, V., Andersson, U., Chadwick, J., Gardner, M., Jaxybulatov, K., Koulakov, I.,
2012. Magma plumbing beneath Anak Krakatau volcano, Indonesia: evidence for
multiple magma storage regions. Contrib. Mineral. Petrol. 163 (4), 631651.
de Angelis, S.H., Larsen, J., Coombs, M., 2013. Pre-eruptive magmatic conditions at Augustine
Volcano, Alaska, 2006: evidence from amphibole geochemistry and textures. J. Petrol.
54, 19391961.
de Silva, S.L., Gregg, P.M., 2014. Thermomechanical feedbacks in magmatic systems: Implications for growth, longevity, and evolution of large caldera-forming magma reservoirs and their supereruptions. J. Volcanol. and Geotherm. Res. 282, 7791.

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845
de Silva, S., Zandt, G., Trumbull, R., Viramonte, J.G., Salas, G., Jimenez, N., 2006. In: Troise, C.,
De Natale, G., Kilburn, C.R.J. (Eds.), Large ignimbrite eruptions and volcano-tectonic
depressions in the Central Andes: a thermomechanical perspectiveMechanisms of Activity and Unrest at Large Calderas vol. 269. Geological Society of London, London, pp.
4763.
Deering, C.D., Bachmann, O., Vogel, T.A., 2011. The Ammonia Tanks Tuff: erupting a meltrich rhyolite cap and its remobilized crystal cumulate. Earth Planet. Sci. Lett. 310
(2011), 518525.
Druitt, T., Bacon, C., 1989. Petrology of the zoned calcalkaline magma chamber of Mount
Mazama, Crater Lake, Oregon. Contrib. Mineral. Petrol. 101 (2), 245259.
Druitt, T.H., Sparks, R.S.J., 1984. On the formation of calderas during ignimbrite eruptions.
Nature 310 (5979), 679681.
Druitt, T.H., Costa, F., Deloule, E., Dungan, M., Scaillet, B., 2012. Decadal to monthly timescales of magma transfer and reservoir growth at a caldera volcano. Nature 482
(7383), 7780.
Dufek, J., Bachmann, O., 2010. Quantum magmatism magmatic compositional gaps generated by melt-crystal dynamics. Geology 38 (8), 687690.
Dufeld, W.A., 1990. Eruptive fountains of silicic magma and their possible effects on the
tin content of fountain-fed lavas, Taylor Creek Rhyolite, New Mexico. Geol. Soc. Am.
Spec. Pap. 246, 251262.
Dufeld, W., Dalrymple, G., 1990. The Taylor Creek Rhyolite of New Mexico: a rapidly
emplaced eld of lava domes and ows. Bull. Volcanol. 52 (no. 6), 475487.
Eichelberger, J.C., Izbekov, P.E., 2000. Eruption of andesite triggered by dyke injection:
contrasting cases at Karymsky Volcano, Kamchatka and Mt Katmai, Alaska: philosophical transactions of the Royal Society of London. Ser. A: Math. Phys. Eng. Sci.
358 (1770), 14651485.
Eichelberger, J.C., Izbekov, P.E., Browne, B.L., 2006. Bulk chemical trends at arc volcanoes
are not liquid lines of descent. Lithos 87 (12), 135154.
Ellis, B.S., Wolff, J.A., 2012. Complex storage of rhyolite in the central Snake River Plain. J.
Volcanol. Geotherm. Res. 211212, 111.
Ellis, B.S., Barry, T., Branney, M.J., Wolff, J.A., Bindeman, I., Wilson, R., Bonnichsen, B., 2010.
Petrologic constraints on the development of a large-volume, high temperature, silicic magma system: the Twin Falls eruptive centre, central Snake River Plain. Lithos 120
(34), 475489.
Erlund, E.J., Cashman, K.V., Wallace, P.J., Pioli, L., Rosi, M., Johnson, E., Granados, H.D., 2010.
Compositional evolution of magma from Parcutin Volcano, Mexico: the tephra record. J. Volcanol. Geotherm. Res. 197 (14), 167187.
Fabbro, G., Druitt, T., Scaillet, S., 2013. Evolution of the crustal magma plumbing system
during the build-up to the 22-ka caldera-forming eruption of Santorini (Greece).
Bull. Volcanol. 75 (12), 122.
Folkes, C.B., Wright, H.M., Cas, R.A., de Silva, S.L., Lesti, C., Viramonte, J.G., 2011. A reappraisal of the stratigraphy and volcanology of the Cerro Galn volcanic system,
NW Argentina. Bull. Volcanol. 73, 14271454.
Fowler, S.J., Spera, F.J., Bohrson, W.A., Belkin, H.E., De Vivo, B., 2007. Phase equilibria constraints on the chemical and physical evolution of the Campanian Ignimbrite. J. Petrol.
48, 459493.
Francis, P.W., Sparks, R., Hawkesworth, C., Thorpe, R., Pyle, D., Tait, S., Mantovani, M.,
McDermott, F., 1989. Petrology and geochemistry of volcanic rocks of the Cerro
Galan caldera, northwest Argentina. Geol. Mag. 126 (05), 515547.
Freda, C., Gaeta, M., Palladino, D.M., Trigila, R., 1997. The Villa Senni Eruption (Alban Hills,
central Italy): the role of H2O and CO2 on the magma chamber evolution and on the
eruptive scenario. J. Volcanol. Geotherm. Res. 78 (no. 12), 103120.
Freda, C., Gaeta, M., Giaccio, B., Marra, F., Palladino, D., Scarlato, P., Sottili, G., 2011. CO2driven large mac explosive eruptions: the Pozzolane Rosse case study from the
Colli Albani Volcanic District (Italy). Bull. Volcanol. 73 (3), 241256.
Freundt, A., Schmincke, H.-U., 1995. Eruption and emplacement of a basaltic welded ignimbrite during caldera formation on Gran Canaria. Bull. Volcanol. 56 (8), 640659.
Gardner, J.E., Sigurdsson, H., Carey, S.N., 1991. Eruption dynamics and magma withdrawal
during the Plinian Phase of the Bishop Tuff Eruption, Long Valley Caldera. J. Geophy.
Res. Solid Earth 96 (B5), 80978111.
Gardner, J.E., Layer, P.W., Rutherford, M.J., 2002. Phenocrysts versus xenocrysts in the
youngest Toba Tuff: implications for the petrogenesis of 2800 km3 of magma. Geology 30 (4), 347350.
Gebauer, S., Schmitt, A., Pappalardo, L., Stockli, D., Lovera, O., 2014. Crystallization
and eruption ages of Breccia Museo (Campi Flegrei caldera, Italy) plutonic
clasts and their relation to the Campanian ignimbrite. Contrib. Mineral. Petrol. 167
(no. 1), 118.
Geshi, N., Ruch, J., Acocella, V., 2014. Evaluating volumes for magma chambers and
magma withdrawn for caldera collapse. Earth Planet. Sci. Lett. 396, 107115.
Geyer, A., Marti, J., 2008. The new worldwide collapse caldera database (CCDB): a tool for
studying and understanding caldera processes. J. Volcanol. Geotherm. Res. 175,
334354.
Geyer, A., Folch, A., Mart, J., 2006. Relationship between caldera collapse and magma
chamber withdrawal: An experimental approach. J. Volcanol. Geotherm. Res. 157,
375386.
Giordano, G., Dobran, F., 1994. Computer simulations of the Tuscolano Artemisio's IInd
Pyroclastic Flow Unit (Alban Hills, Central Italy). J. Volcanol. Geotherm. Res. 61,
6994.
Giordano, G., the CARG team, 2010. Stratigraphy and volcano-tectonic structures of the
Colli Albani volcanic eld. In: Funiciello, R., Giordano, G. (Eds.), The Colli Albani
VolcanoSpecial Publication of IAVCEI 3. The Geological Society, London, pp. 4397.
Giordano, G., De Benedetti, A., Diana, A., Diano, G., Gaudioso, F., Marasco, F., Miceli, M.,
Mollo, S., Cas, R., Funiciello, R., 2006. The Colli Albani mac caldera (Roma, Italy): stratigraphy, structure and petrology. J. Volcanol. Geotherm. Res. 155 (1), 4980.
Gottsmann, J., Lavalle, Y., Mart, J., Aguirre-Daz, G., 2009. Magmatectonic interaction
and the eruption of silicic batholiths. Earth Planet. Sci. Lett. 284 (34), 426434.

43

Gravley, D.M., Wilson, C.J.N., Leonard, G.S., Cole, J.W., 2007. Double trouble: Paired ignimbrite eruptions and collateral subsidence in the Taupo Volcanic Zone, New Zealand.
Geol. Soc. Am. Bull. 119, 1830.
Gregg, P.M., de Silva, S.L., Grosls, E.B., Parmigiani, J.P., 2012. Catastrophic caldera-forming
eruptions: thermomechanics and implications for eruption triggering and maximum
caldera dimensions on Earth. J. Volcanol. Geotherm. Res. 241242, 112.
Gualda, G.A.R., Ghiorso, M.S., 2013. The Bishop Tuff giant magma body: an alternative to
the Standard Model. Contrib. Mineral. Petrol. 166, 755775.
Gualda, G.A.R., Cook, D.L., Chopra, R., Qin, L.P., Anderson Jr., A.T., Rivers, M., 2004. Fragmentation, nucleation and migration of crystals and bubbles in the Bishop Tuff rhyolitic magma. Trans. R. Soc. Edinb. Earth Sci. 95, 375390.
Gualda, G.A.R., Ghiorso, M.S., Lemons, R.V., Carley, T.L., 2012. Rhyolite-MELTS: a modied
calibration of MELTS optimised for silica-rich, uid-bearing magmatic systems. J. of
Petrol. 53, 875890.
Gudmundsson, A., 2008. Magma-chamber geometry, uid transport, local stresses and
rock behaviour during collapse caldera formation. Dev. Volcanol. 10, 313349.
Gudmundsson, A., 2012. Magma chambers: formation, local stresses, excess pressures,
and compartments. J. Volcanol. Geotherm. Res. 237238, 1941.
Gurioli, L., Houghton, B.F., Cashman, K.V., Cioni, R., 2005. Complex changes in eruption dynamics during the 79 AD eruption of Vesuvius. Bull. Volcanol. 67 (2), 144159.
Hammer, J.E., Rutherford, M.J., Hildreth, W., 2002. Magma storage prior to the 1912 eruption at Novarupta, Alaska. Contrib. Mineral. Petrol. 144, 144162.
Hildreth, W., 1979. The Bishop Tuff: evidence for the origin of compositional zonation in
silicic magma chambers. Geol. Soc. Am. Spec. Pap. 180, 4375.
Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithospheric
magmatism. J. Geophys. Res. 86, 1015310192.
Hildreth, W., 2004. Volcanological perspectives on Long Valley Mammoth Mountain, and
Mono Craters: several contiguous but discrete systems. J. Volcanol. Geotherm. Res.
136, 169198.
Hildreth, W., Fierstein, J., 2012. The Novarupta-Katmai eruption of 1912 largest eruption of the twentieth century; centennial perspectives. U.S.G.S. Prof. Paper 1791.
Hildreth, W., Mahood, G.A., 1986. Ring-fracture eruption of the Bishop Tuff. Geol. Soc. Am.
Bull. 97, 396403.
Hildreth, W., Wilson, C.J.N., 2007. Compositional zoning of the Bishop Tuff. J. Petrol. 48,
951999.
Holohan, E., van Wyk de Vries, B., Troll, V., 2008a. Analogue models of caldera collapse in
strike-slip tectonic regimes. Bull. Volcanol. 70 (no. 7), 773796.
Holohan, E., Troll, V., van Wyk de Vries, B., Walsh, J.J., Walter, T.R., 2008b. Unzipping Long
Valley: an explanation for vent migration patterns during elliptical ring fracture eruption. Geology 36, 323326.
Houghton, B.F., Carey, R.J., Cashman, K.V., Wilson, C.J., Hobden, B.J., Hammer, J.E., 2010. Diverse patterns of ascent, degassing, and eruption of rhyolite magma during the 1.8 ka
Taupo eruption, New Zealand: evidence from clast vesicularity. J. Volcanol. Geotherm.
Res. 195 (1), 3147.
Huber, C., Bachmann, O., Dufek, J., 2011. Thermo-mechanical reactivation of locked crystal
mushes: melting-induced internal fracturing and assimilation processes in magmas.
Earth Planet. Sci. Lett. 34, 443454.
Huber, C., Bachmann, O., Dufek, J., 2012. Crystal-poor versus crystal-rich ignimbrites; a
competition between stirring and reactivation. Geology 40 (2), 115118.
Humphreys, M.C.S., Holness, M.B., 2010. Melt-rich segregations in the Skaergaard Marginal Border Series: tearing of a vertical silicate mush. Lithos 119 (34), 181192.
Humphreys, M., Christopher, T., Hards, V., 2009. Microlite transfer by disaggregation of
mac inclusions following magma mixing at Soufrire Hills volcano, Montserrat.
Contrib. Mineral. Petrol. 157 (5), 609624.
Husen, S., Smith, R.B., Waite, G.P., 2004. Evidence for gas and magmatic sources beneath
the Yellowstone volcanic eld from seismic tomographic imaging. J. Volcanol.
Geotherm. Res. 131 (34), 397410.
Iyer, H., Evans, J., Dawson, P., Stauber, D., Achauer, U., 1990. Differences in magma storage
in different volcanic environments as revealed by seismic tomography; silicic volcanic centers and subduction-related volcanoes. Magma Transport and Storage,pp.
293316.
Jellinek, A.M., DePaolo, D.J., 2003. A model for the origin of large silicic magma chambers:
precursors of caldera-forming eruptions. Bull. Volcanol. 65 (5), 363381.
Johnson, D., 1987. Elastic and inelastic magma storage at Kilauea volcano. US Geol. Surv.
Prof. Pap. 1350, 12971306.
Kamata, H., Suzuki-Kamata, K., Bacon, C.R., 1993. Deformation of the Wineglass Welded
Tuff and the timing of caldera collapse at Crater Lake, Oregon. J. Volcanol. Geotherm.
Res. 56, 253265.
Kaneko, K., Kamata, H., Koyaguchi, T., Yoshikawa, M., Furukawa, K., 2007. Repeated largescale eruptions from a single compositionally stratied magma chamber: an example
from Aso volcano, Southwest Japan. J. Volcanol. Geotherm. Res. 167 (14), 160180.
Karlstrom, L., Rudolph, M.L., Manga, M., 2012. Caldera size modulated by the yield stress
within a crystal-rich magma reservoir. Nat. Geosci. 5, 402405.
Kay, S., Coira, B., Wrner, G., Kay, R., Singer, B., 2011. Geochemical, isotopic and single
crystal 40Ar/39Ar age constraints on the evolution of the Cerro Galn ignimbrites.
Bull. Volcanol. 73 (10), 14871511.
Kelley, D.F., Barton, M., 2008. Pressures of crystallization of icelandic magmas. J. Petrol. 49
(3), 465492.
Kennedy, B., Stix, J., Vallance, J.W., Lavalle, Y., Longpr, M.-A., 2004. Controls on caldera
structure: results from analogue sandbox modeling. Geol. Soc. Am. Bull. 116 (56),
515524.
Kennedy, B.M., Jellinek, M.A., Stix, J., 2008. Coupled caldera subsidence and stirring inferred from analogue models. Nat. Geosci. 1 (6), 385389.
Klug, C., Cashman, K., Bacon, C., 2002. Structure and physical characteristics of pumice
from the climactic eruption of Mount Mazama (Crater Lake), Oregon. Bull. Volcanol.
64 (7), 486501.

44

K.V. Cashman, G. Giordano / Journal of Volcanology and Geothermal Research 288 (2014) 2845

Lees, J.M., 2007. Seismic tomography of magmatic systems. J. Volcanol. Geotherm. Res.
167 (14), 3756.
Lesti, C., Porreca, M., Giordano, G., Mattei, M., Cas, R.A., Wright, H.M., Folkes, C.B.,
Viramonte, J., 2011. High-temperature emplacement of the Cerro Galn and
Toconquis Group ignimbrites (Puna plateau, NW Argentina) determined by TRM
analyses. Bull. Volcanol. 73 (10), 15351565.
Lindsay, J.M., Schmitt, A.K., Trumbull, R.B., de Silva, S.L., Siebel, W., Emmermann, R., 2001. Magmatic evolution of the La Pacana Caldera System, Central Andes, Chile: compositional variation of two cogenetic, large-volume felsic ignimbrites. J. Petrol. 42 (3), 459486.
Lipman, P.W., 1997. Subsidence of ash-ow calderas: relation to caldera size and magmachamber geometry. Bull. Volcanol. 59 (3), 198218.
Lipman, P.W., 2007. Incremental assembly and prolonged consolidation of Cordilleran
magma chambers: evidence from the Southern Rocky Mountain volcanic eld.
Geosphere 3. http://dx.doi.org/10.1130/GES00061.1.
Lipman, P., Dungan, M., Bachmann, O., 1997. Comagmatic granophyric granite in the Fish
Canyon Tuff, Colorado: implications for magma-chamber processes during a large
ash-ow eruption. Geology 25 (10), 915918.
Liu, Y., Anderson, A.T., Wilson, C.J.N., Davis, A.M., Steele, I.M., 2006. Mixing and differentiation in the Oruanui rhyolitic magma, Taupo, New Zealand: evidence from volatiles
and trace elements in melt inclusions. Contrib. Mineral. Petrol. 151, 7187.
Luttrell, K., Mencin, D., Francis, O., Hurwitz, S., 2013. Constraints on the upper crustal
magma reservoir beneath Yellowstone Caldera inferred from lake-seiche induced
strain observations. Geophys. Res. Lett. 40 (3), 501506.
Mandeville, C.W., Webster, J.D., Tappen, C., Taylor, B.E., Timbal, A., Sasaki, A., Hauri, E.,
Bacon, C.R., 2009. Stable isotope and petrologic evidence for open-system degassing
during the climactic and pre-climactic eruptions of Mt. Mazama, Crater Lake, Oregon.
Geochim. Cosmochim. Acta 73 (10), 29783012.
Mangiacapra, A., Moretti, R., Rutherford, M., Civetta, L., Orsi, G., Papale, P., 2008. The deep
magmatic system of the Campi Flegrei caldera (Italy). Geophys. Res. Lett. 35 (21),
L21304.
Marsh, B.D., 1996. Solidication fronts and magmatic evolution. Mineral. Mag. 60 (398),
540.
Marsh, B.D., 2002. On bimodal differentiation by solidication front instability in basaltic
magmas, part 1: basic mechanics. Geochim. Cosmochim. Acta 66 (12), 22112229.
Marsh, B.D., 2013. On some fundamentals of igneous petrology. Contrib. Mineral. Petrol.
166, 665690.
Mart, J., 1991. Caldera-like structures related to Permo-Carboniferous volcanism of the
Catalan Pyrenees (NE Spain). J. Volcanol. Geotherm. Res. 45 (34), 173186.
Marti, J., Geyer, A., Folch, A., Gottsmann, J., 2008. A review on collapse caldera modelling.
Dev. Volcanol. 10, 233283.
Masotta, M., Gaeta, M., Gozzi, F., Marra, F., Palladino, D., Sottili, G., 2010. H2O-and
temperature-zoning in magma chambers: the example of the Tufo Giallo della Via
Tiberina eruptions (Sabatini Volcanic District, central Italy). Lithos 118 (1), 119130.
Masotta, M., Freda, C., Gaeta, M., 2012. Origin of crystal-poor, differentiated magmas: insights from thermal gradient experiments. Contrib. Mineral. Petrol. 163 (1), 4965.
Masturyono, R.M., Wark, D., Roecker, S., Fauzi, G.I., 2001. Distribution of magma beneath
Toba Caldera, North Sumatra, Indonesia, Constrained by 3-dimensional P-wave velocities, seismicity, and gravity data. Geochem. Geophys. Geosyst. 2.
Matthews, N.E., Huber, C., Pyle, D.M., Smith, V.C., 2012. Timescales of magma recharge and
reactivation of large silicic systems from Ti diffusion in quartz. J. Petrol. 53 (7),
13851416.
Maughan, L.L., Christiansen, E.H., Best, M.G., Gromm, C.S., Deino, A.L., Tingey, D.G., 2002.
The Oligocene Lund Tuff, Great Basin, USA: a very large volume monotonous intermediate. J. Volcanol. Geotherm. Res. 113 (12), 129157.
McKenzie, D., 1984. The generation and compaction of partially molten rock. J. Petrol. 25
(3), 713765.
Melnik, O., Barmin, A.A., Sparks, R.S.J., 2005. Dynamics of magma ow inside volcanic conduits with bubble overpressure buildup and gas loss through permeable magma. J.
Volcanol. Geotherm. Res. 143, 5368.
Neave, D.A., Passmore, E., Maclennan, J., Fitton, G., Thordarson, T., 2013. Crystalmelt relationships and the record of deep mixing and crystallization in the ad 1783 Laki
Eruption, Iceland. J. Petrol. 54 (8), 16611690.
Palladino, D.M., Simei, S., 2005. Eruptive dynamics and caldera collapse during the Onano
eruption, Vulsini, Italy. Bull. Volcanol. 67, 423440.
Pallister, J.S., Hoblitt, R.P., Reyes, A.G., 1992. A basalt trigger for the 1991 eruptions of
Pinatubo volcano? Nature 356, 426428.
Pamukcu, A.S., Gualda, G.A.R., Anderson Jr., A.T., 2012. Crystallization stages of the
Bishop Tuff magma body recorded in crystal textures in pumice clasts. J. Petrol.
53, 589609.
Pamukcu, A.S., Carley, T.L., Gualda, G.A.R., Miller, C.F., Ferguson, C.A., 2013. The evolution
of the Peach Spring giant magma body: evidence from accessory mineral textures
and compositions, bulk pumice and glass geochemistry, and rhyolite-MELTS modeling. J. Petrol. 54, 11091148.
Passmore, E., Maclennan, J., Fitton, G., Thordarson, T., 2012. Mush disaggregation in basaltic magma chambers: evidence from the ad 1783 Laki Eruption. J. Petrol. 53 (12),
25932623.
Peccerillo, A., 2005. The roman province: plio-quaternary volcanism in Italy. Petrol.
Geochem. Geodyn. 69107.
Polacci, M., Papale, P., Rosi, M., 2001. Textural heterogeneities in pumices from the climactic eruption of Mount Pinatubo, 15 June 1991, and implications for magma ascent dynamics. Bull. Volcanol. 63 (23), 8397.
Pyle, D.M., Mather, T.A., Biggs, J., 2013. Remote sensing of volcanoes and volcanic processes: integrating observation and modelling introduction. Geol. Soc. Lond., Spec. Publ.
380 (1), 113.
Reid, M.R., 2008. How long does it take to supersize an eruption? Elements 4 (1),
2328.

Roberge, J., Wallace, P.J., Kent, A.J.R., 2013. Magmatic processes in the Bishop Tuff rhyolitic
magma based on trace elements in melt inclusions and pumice matrix glass. Contrib.
Mineral. Petrol. 165, 237257.
Roche, O., Druitt, T.H., 2001. Onset of caldera collapse during ignimbrite eruptions. Earth
Planet. Sci. Lett. 191, 191202.
Roman, D.C., Cashman, K.V., Gardner, C.A., Wallace, P.J., Donovan, J.J., 2006. Storage and
interaction of compositionally heterogeneous magmas from the 1986 eruption of
Augustine Volcano, Alaska. Bull. Volcanol. 68, 240254.
Rosenberg, C.L., Medvedev, S., Handy, M.R., 2007. Effects of melting on faulting and continental deformation. Dahlem Workshop Report vol. 95, pp. 357401.
Rosi, M., Vezzoli, L., Castelmenzano, A., Grieco, G., 1999. Plinian pumice fall deposit of the
Campanian Ignimbrite eruption (Phlegraean Fields, Italy). J. Volcanol. Geotherm. Res.
91, 179198.
Rosi, M., Landi, P., Polacci, M., Di Muro, A., Zandomeneghi, D., 2004. Role of conduit shear
on ascent of the crystal-rich magma feeding the 800-year-b.p. Plinian eruption of
Quilotoa Volcano (Ecuador). Bull. Volcanol. 66 (4), 307321.
Rust, A.C., Cashman, K.V., 2011. Permeability controls on expansion and size distributions
of pyroclasts. J. Geophys. Res. 116. http://dx.doi.org/10.1029/2011JB008494.
Saunders, K.E., Morgan, D.J., Baker, J.A., Wysoczanski, R.J., 2010. The magmatic evolution of
the Whakamaru Supereruption, New Zealand, constrained by a microanalytical study
of plagioclase and quartz. J. Petrol. 51 (12), 24652488.
Scandone, R., 1990. Chaotic collapse of calderas. J. Volcanol. Geotherm. Res. 42 (3),
285302.
Schilling, F.R., Partzsch, G.M., 2001. Quantifying partial melt fraction in the crust beneath
the Central Andes and the Tibetan plateau: physics and chemistry of the earth. Earth
Solid Earth Geod. 26 (45), 239246.
Sherburn, S., Bannister, S., Bibby, H., 2003. Seismic velocity structure of the central Taupo
Volcanic Zone, New Zealand, from local earthquake tomography. J. Volcanol.
Geotherm. Res. 122 (12), 6988.
Sigmundsson, F., Hreinsdttir, S., Hooper, A., rnadttir, T., Pedersen, R., Roberts, M.J., et
al., 2010. Intrusion triggering of the 2010 Eyjafjallajokull explosive eruption. Nature
468, 426430.
Sigurdsson, H., Sparks, R., 1981. Petrology of rhyolitic and mixed magma ejecta from the
1875 eruption of Askja, Iceland. J. Petrol. 22 (1), 4184.
Simon, J., Weis, D., DePaolo, D., Renne, P., Mundil, R., Schmitt, A., 2014. Assimilation of
preexisting Pleistocene intrusions at Long Valley by periodic magma recharge accelerates rhyolite generation: rethinking the remelting model. Contrib. Mineral. Petrol. 167
(1), 134.
Sisson, T.W., Bacon, C.R., 1999. Gas-driven lter pressing in magmas. Geology 27 (7),
613616.
Smith, R.L., 1979. Ash-ow magmatism: ash-ow tuffs. Geol. Soc. Am. Spec. Pap. 180,
527.
Solano, J.M.S., Jackson, M.D., Sparks, R.S.J., Blundy, J.D., Annen, C., 2012. Melt segregation in
deep crustal hot zones: a mechanism for chemical differentiation, crustal assimilation
and the formation of evolved magmas. J. Petrol. 53, 19992026.
Sparks, R.S.J., Wilson, C.J.N., 1990. The Minoan deposits: a review of their characteristics
and interpretation. Thera and the Aegean world III 2, pp. 8999.
Sparks, R., Francis, P., Hamer, R., Pankhurst, R., O'callaghan, L., Thorpe, R., Page, R., 1985.
Ignimbrites of the Cerro Galan Caldera, NW Argentina. J. Volcanol. Geotherm. Res.
24 (3), 205248.
Sparks, R.S.J., Folkes, C.B., Humphreys, M.C., Barfod, D.N., Clavero, J., Sunagua, M.C.,
McNutt, S.R., Pritchard, M.E., 2008. Uturuncu volcano, Bolivia: volcanic unrest due
to mid-crustal magma intrusion. Am. J. Sci. 308 (6), 727769.
Sparks, R., Biggs, J., Neuberg, J., 2012. Monitoring volcanoes. Science 335 (6074),
13101311.
Spera, F.J., Crisp, J.A., 1981. Eruption volume, periodicity, and caldera area: relationships
and inferences on development of compositional zonation in silicic magma chambers. J. Volcanol. Geotherm. Res. 11, 169187.
Stix, J., Kobayashi, T., 2007. Magma dynamics and collapse mechanisms during four historic caldera-forming events. J. Geophys. Res. 113, B09205.
Tait, S., Jaupart, C., Vergniolle, S., 1989. Pressure, gas content and eruption periodicity of a
shallow, crystallising magma chamber. Earth Planet. Sci. Lett. 92, 107123.
Tarasewicz, J., White, R.S., Woods, A.W., Brandsdttir, B., Gudmundsson, M.T., 2012.
Magma mobilization by downward-propagating decompression of the Eyjafjallajkull
volcanic plumbing system. Geophys. Res. Lett. 39 (19), L19309.
van der Zwan, F., Chadwick, J., Troll, V., 2013. Textural history of recent basalticandesites and plutonic inclusions from Merapi volcano. Contrib. Mineral. Petrol.
166 (1), 4363.
Vinkler, A.P., Cashman, K., Giordano, G., Groppelli, G., 2012. Evolution of the mac Villa
Senni caldera-forming eruption at Colli Albani volcano, Italy, indicated by textural
analysis of juvenile fragments. J. Volcanol. Geotherm. Res. 235236, 3754.
Voight, B., Widiwijayanti, C., Mattioli, G., Elsworth, D., Hidayat, D., Strutt, M., 2010.
Magma-sponge hypothesis and stratovolcanoes: case for a compressible reservoir
and quasi-steady deep inux at Soufrire Hills Volcano, Montserrat. Geophys. Res.
Lett. 37 (19).
Vye-Brown, C., Gannoun, A., Barry, T.L., Self, S., Burton, K.W., 2013. Osmium isotope variations accompanying the eruption of a single lava ow eld in the Columbia River
Flood Basalt Province. Earth Planet. Sci. Lett. 368, 183194.
Walker, G.P., 1984. Downsag calderas, ring faults, caldera sizes, and incremental caldera
growth. J.Geophys. Res. Solid Earth (19782012) 89 (B10), 84078416.
Walker, B.A., Miller, C.F., Claiborne, L.L., Wooden, J.L., Miller, J.S., 2007. Geology and geochronology of the Spirit Mountain batholith, southern Nevada; implications for timescales and physical processes of batholith construction. J. Volcanol. Geotherm. Res.
167, 239262.
Walker Jr., B., Klemetti, E., Grunder, A., Dilles, J., Tepley, F., Giles, D., 2013. Crystal reaming
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) 2845
magma cycle: thermobarometry of the Aucanquilcha Volcanic Cluster. Contrib. Mineral. Petrol. 165 (4), 663682.
Wallace, P.J., Anderson Jr., A.T., Davis, A.M., 1999. Gradients in H2O, CO2, and exsolved gas
in a large-volume silicic magma system: interpreting the record preserved in melt inclusions from the Bishop Tuff. J. Geophys. Res. 104 (B9), 2009720122.
Wark, D.A., Hildreth, W., Spear, F.S., Cherniak, D.J., Watson, E.B., 2007. Pre-eruption recharge of the Bishop magma system. Geology 35, 235238.
Watkins, S., Giordano, G., Cas, R., De Rita, D., 2002. Emplacement processes of the mac
Villa Senni Eruption Unit (VSEU) ignimbrite succession, Colli Albani volcano, Italy.
J. Volcanol. Geotherm. Res. 118 (1), 173203.
Willcock, M.A.W., Cas, R.A.F., Giordano, G., Morelli, C., 2013. The eruption, pyroclastic ow
behaviour, and caldera in-lling processes of the extremely large volume (1290 km3),
intra- to extra-caldera, Permian Ora (Ignimbrite) Formation, Southern Alps, Italy. J.
Volcanol. Geotherm. Res. 265 (2013), 102126.
Wilson, C.J.N., Charlier, B.L.A., 2009. Rapid rates of magma generation at contemporaneous magma systems, Taupo Volcano, New Zealand: insights from UTh model-age
spectra in zircons. J. Petrol. 50 (5), 875907.
Wilson, C., Hildreth, W., 1997. The Bishop Tuff: new insights from eruptive stratigraphy. J.
Geol. 105, 407440.
Wilson, C.J.N., Walker, G.P.L., 1985. The Taupo eruption, New Zealand I. General aspects:
philosophical transactions of the Royal Society of London. Ser. A Math. Phys. Sci.
314, 199228.

45

Wolff, J.A., Ramos, F.C., Davidson, J.P., 1999. Sr isotope disequilibrium during differentiation of the Bandelier Tuff: constraints on the crystallization of a large rhyolitic
magma chamber. Geology 27, 495498.
Wolff, J.A., Ramos, F.C., Hart, G.L., Patterson, J.D., Brandon, A.D., 2008. Columbia River ood
basalts from a centralized crustal magmatic system. Nat. Geosci. 1 (3), 177180.
Wotzlaw, J.-F., Bindeman, I.N., Watts, K.E., Schmitt, A.K., Caricchi, L., Schaltegger, U., 2014.
Linking rapid magma reservoir assembly and eruption trigger mechanisms at evolved
Yellowstone-type supervolcanoes. Geology 42, 807810.
Wright, H.N., Folkes, C., Cas, R.F., Cashman, K., 2011. Heterogeneous pumice populations
in the 2.08-Ma Cerro Galn Ignimbrite: implications for magma recharge and ascent
preceding a large-volume silicic eruption. Bull. Volcanol. 73 (10), 15131533.
Wright, H., Bacon, C., Vazquez, J., Sisson, T., 2012. Sixty thousand years of magmatic
volatile history before the caldera-forming eruption of Mount Mazama, Crater Lake,
Oregon. Contrib. Mineral. Petrol. 164 (6), 10271052.
Yoshimoto, M., Fujii, T., Kaneko, T., Yasuda, A., Nakada, S., 2004. Multiple magma reservoirs for the 1707 eruption of Fuji volcano, Japan. Proc. Jpn. Acad. Ser. B 80, 103106.
Zandt, G., Leidig, M., Chmielowski, J., Baumont, D., Yuan, X., 2003. Seismic detection and
characterization of the AltiplanoPuna magma body, Central Andes. Pure Appl.
Geophys. 160 (34), 789807.
Zollo, A., Maercklin, N., Vassallo, M., Dello Iacono, D., Virieux, J., Gasparini, P., 2008. Seismic
reections reveal a massive melt layer feeding Campi Flegrei caldera. Geophys. Res.
Lett. 35. http://dx.doi.org/10.1029/2008GL034242.

You might also like