Skip to content
Open Access Published by De Gruyter October 29, 2016

Silicic magma reservoirs in the Earth’s crust

  • Olivier Bachmann EMAIL logo and Christian Huber
From the journal American Mineralogist

Abstract

Magma reservoirs play a key role in controlling numerous processes in planetary evolution, including igneous differentiation and degassing, crustal construction, and volcanism. For decades, scientists have tried to understand what happens in these reservoirs, using an array of techniques such as field mapping/petrology/geochemistry/geochronology on plutonic and volcanic lithologies, geophysical imaging of active magmatic provinces, and numerical/experimental modeling. This review paper tries to follow this multi-disciplinary framework while discussing past and present ideas. We specifically focus on recent claims that magma columns within the Earth’s crust are mostly kept at high crystallinity (“mush zones”), and that the dynamics within those mush columns, albeit modulated by external factors (e.g., regional stress field, rheology of the crust, pre-existing tectonic structure), play an important role in controlling how magmas evolve, degas, and ultimately erupt. More specifically, we consider how the chemical and dynamical evolution of magma in dominantly mushy reservoirs provides a framework to understand: (1) the origin of petrological gradients within deposits from large volcanic eruptions (“ignimbrites”); (2) the link between volcanic and plutonic lithologies; (3) chemical fractionation of magmas within the upper layers of our planet, including compositional gaps noticed a century ago in volcanic series (4) volatile migration and storage within mush columns; and (5) the occurrence of petrological cycles associated with caldera-forming events in long-lived magmatic provinces. The recent advances in understanding the inner workings of silicic magmatism are paving the way to exciting future discoveries, which, we argue, will come from interdisciplinary studies involving more quantitative approaches to study the crust-reservoir thermo-mechanical coupling as well as the kinetics that govern these open systems.

Introduction

Previous attempts to summarize the state of knowledge in this field (e.g., Smith 1979; Lipman 1984; Sparks et al. 1984; Marsh 1989a; de Silva 1991) have greatly helped crystallizing ideas on the contemporaneous state of knowledge and possible alleys for future directions to explore. Since then, many more publications have attempted to draw magma reservoirs geometries and internal complexities [see Fig. 1 for a potpourri of such schematic diagrams for several examples of large caldera systems in the U.S.A., as well as the more recent reviews by Petford et al. (2000), Bachmann et al. (2007b), Lipman (2007), Cashman and Sparks (2013), Cash- man and Giordano (2014), and de Silva and Gregg (2014)]. At this stage, it is clear that magmas are complex mixtures of multiple phases (silicate melt, H2O-CO2-dominated fluid, and up to 10 different mineral phases in some systems) with very different physical properties. For example, viscosity contrast between water vapor and crystals can span more than 20 orders of magnitude, and even vary significantly within one given phase (e.g., several orders of magnitude for silicate melt), as composition-P-T-fH20-f02 vary (Mader et al. 2013). The way these magmas move and stall in the crust will therefore reflect the complex interaction between all these phases and the typically colder wall rocks.

Determining the shape, size, depth, and state of magma bodies in the Earth’s crust as well as how they evolve over time remain key issues for several Earth science communities. With a better determination of these variables, petrologists could construct more accurate chemical models of differentiation processes; magma dynamicists could postulate causes for magma migration, storage, and interaction at different levels within the crust; volcanologists could better predict the style and volume of upcoming volcanic eruptions; and geochronologists could construct an evolutionary trend with greater precision. Although our knowledge on magmatic systems has expanded greatly over the past decades, much remains to be done to constrain the multiphase, multiscale processes that are at play in those reservoirs and the rates at which they occur.

Figure 1 Recent schematic diagrams for the three big active caldera systems in the Western U.S.A. There are many other caldera systems around the world (e.g., Hughes and Mahood 2008), and the focus on examples from the U.S.A. here is solely due to personal acquaintance. (a) Valles caldera (Wilcock et al. 2010), (b) Long Valley caldera (Hildreth 2004), (c) Yellowstone, Official web site and Lowenstern and Hurvitz (2008).
Figure 1

Recent schematic diagrams for the three big active caldera systems in the Western U.S.A. There are many other caldera systems around the world (e.g., Hughes and Mahood 2008), and the focus on examples from the U.S.A. here is solely due to personal acquaintance. (a) Valles caldera (Wilcock et al. 2010), (b) Long Valley caldera (Hildreth 2004), (c) Yellowstone, Official web site and Lowenstern and Hurvitz (2008).

Over the last decades, an impressive body of work has ac-cumulated on the characterization of magma mainly through:(1) high-precision, high-resolution geochemical work [with ever more powerful instruments, including the atom-probe, e.g., Valley et al. (2014) and nanoSIMS in the recent past; e.g., Eiler et al. (2011); Till et al. (2015)], and (2) experimental petrology (e.g., phase diagrams, trace element partitioning between phases, isotopic fingerprinting), culminating in powerful codes for ther-modynamic modeling (Ghiorso and Sack 1995a; Boudreau 1999; Connolly 2009; Gualda et al. 2012). This work will continue to greatly improve the constraints that we place on these magmatic systems for years to come, but we argue that the field will have to take into account kinetic, disequilibrium processes, which are commonplace, and can be severe in the world of magmas.

Thermodynamics and kinetics are both valuable to understand magmatic processes, and kinetics is bound to become even more important in the following decades as we strive to better constrain the timescale of dynamical processes associated with magma transport and evolution in the crust. This Centennial Volume of American Mineralogist appears as an appropriate landmark to summarize the recent advances on the subject and bring the multiple facets of this problem together to better frame the future directions of research.

The most exciting scientific challenges are, we believe, those that are driven by enigmas arising by observations of Nature. Below are a few examples of key questions or long-standing observations that relate to magma reservoirs and magmatic differentiation:

  1. The fundamental dichotomy in supervolcanic deposits (also called ignimbrites or ash-flow tuffs), which show, in most cases:(a) compositionally and thermally zoned dominantly crystal-poor units [sometimes grading to crystal-rich facies at the end of the eruption (Lipman 1967; Hildreth 1979; Wolff and Storey 1984; Worner and Schmincke 1984; de Silva and Wolff 1995)], and (b) strikingly homogeneous crystal-rich dominantly dacitic units (the “Monotonous Intermediates” of Hildreth 1981), known to be some of the most viscous fluids on Earth (Whitney and Stormer 1985; Francis et al. 1989; de Silva et al. 1994; Scaillet et al. 1998b; Lindsay et al. 2001; Bachmann et al. 2002, 2007b; Huber et al. 2009; Folkes et al. 2011b; see also Bachmann and Bergantz 2008b; Huber et al. 2012a for reviews).

  2. The striking resemblance in chemical composition (for major and trace elements) between the interstitial melt in crystal-rich silicic arc magmas and the melt-rich rhyolitic magmas (Bacon and Druitt 1988b; Hildreth and Fierstein 2000; Bachmann et al. 2005).

  3. The ubiquitous observation of magma recharges from depth prior to eruptions, and their putative role on the remobilization of resident (often crystal-rich) magmas (“rejuvenation”), for large and small eruptions (e.g., Sparks et al. 1977; Pallister et al. 1992; Murphy et al. 2000; Bachmann et al. 2002; Bachmann and Bergantz 2003; Wark et al. 2007; Molloy et al. 2008; Bachmann 2010a; Cooper and Kent 2014; Klemetti and Clynne 2014; Wolff and Ramos 2014; Wolff et al. 2015).

  4. The long-noticed compositional gap in volcanic series, coined the Bunsen-Daly Gap after R. Bunsen and R. Daly’s early observations (Bunsen 1851; Daly 1925, 1933), and confirmed in many areas worldwide since (Chayes 1963; Thompson 1972; Brophy 1991; Thompson et al. 2001; Deering et al. 2011a; Szymanowski et al. 2015) even where crystal fractionation, which should lead to a continuous melt composition at the surface, dominates (Bonnefoi et al. 1995; Geist et al. 1995; Peccerillo et al. 2003; Macdonald et al. 2008; Dufek and Bachmann 2010; Sliwinski et al. 2015).

  5. The complex relationship between silicic plutonic and volcanic units with possible hypotheses (not mutually exclusive in plutonic complexes) ranging from plutons being “failed eruptions” (crystallized melts; Tappa et al. 2011; Barboni et al. 2015; Glazner et al. 2015; Keller et al. 2015) to plutons being “crystal graveyards” (crystal cumulates; Bachl et al. 2001; Deering and Bachmann 2010; Gelman et al. 2014; Putirka et al. 2014; Lee and Morton 2015). Related to this volcanic-plutonic connection is the corollary that granites sensu stricto (evolved, high-SiO2, low-Sr composition) appear to be less common than their volcanic counterparts (e.g., Navdat database, Halliday et al. 1991; Hildreth 2007; Gelman et al. 2014).

  6. The observation that silicic plutons, when fully crystallized, contain little volatile left (typically <1 wt% H2O that is bound up in the minerals; e.g., Caricchi and Blundy 2015) despite the fact they started with volatile concentrations above 6 wt%, implying significant loss of volatile at different stages of the magma bodies’ evolution [e.g., by first and second boiling, pre-, syn-, and post-eruption degassing, see for example Anderson (1974); Wallace et al. (1995); Candela (1997); Lowenstern (2003); Webster (2004); Wallace (2005); Bachmann et al. (2010); Blundy et al. (2010); Sillitoe (2010); Baker and Alletti (2012); Heinrich and Candela (2012)].

  7. The well-known “Excess S,” highlighted by the much higher S content released during eruptions (measured by remote sensing or estimated from ice-core records) that can be accounted for by the melt inclusion data [i.e., petrologic estimate, see, for example, reviews by Wallace (2001); Costa et al. (2003); Shino- hara (2008)].

  8. The remarkable cyclic activity observed large-scale vol-canic systems, which culminates in caldera-forming eruptions (“caldera cycles”) that can repeat itself several times, e.g., the multi-cyclic caldera systems such as Yellowstone (Hildreth et al. 1991; Bindeman and Valley 2001; Christiansen 2001; Lowenstern and Hurvitz 2008), Southern Rocky Mountain Volcanic field (e.g., Lipman 1984, 2007; Lipman and Bachmann 2015), Taupo Volcanic Zone (e.g., Wilson 1993; Wilson et al. 1995; Sutton et al. 2000; Deering et al. 2011a; Barker et al. 2014), High Andes (e.g., Pitcher et al. 1985; Petford et al. 1996; Lindsay et al. 2001; Schmitt et al. 2003; de Silva et al. 2006; Klemetti and Grander 2008), Campi Flegreii (e.g., Orsi et al. 1996; Civetta et al. 1997; Gebauer et al. 2014), and Aegean arc (e.g., Bachmann et al. 2012; Degruyter et al. 2015).

All those questions/observations require an overarching con-cept of magma chamber growth and evolution that ties together these seemingly disparate concepts. Before attempting such a task, first this briefly outlines the methods that are typically used to infer the state of these reservoirs, highlighting some of their strengths and limitations.

How Can We Study Magma Reservoirs?

Sampling volcanic and plutonic lithologies

The foundation of volcanology is motivated by observations of ongoing and past volcanic activity, with the purpose of understanding the underlying processes that govern the evolution of magmatic systems on Earth and other bodies of the Solar System (Wilson and Head 1994). However, large-scale eruptions are rare (Simkin 1993; Mason et al. 2004) and as volcanologists/petrologists we rely much on a forensic approach: study magmatic rocks (both plutonic and volcanic) that formed/erupted in the past, and try to reconstruct the conditions (P, T, fO2 ,fH2O, fSO2,...) that prevailed in the reservoirs at the time of formation or eruption [see, for example, a review of techniques by Putirka (2008) or Blundy and Cashman (2008)]. Of course, volcanic and plutonic lithologies do not provide the same type of information, and could/should be complementary.Volcanic rocks carry limited spatial context for the reservoir, but provide an instantaneous snapshot into the state of the magma body just before the eruption. In contrast, plutonic rocks present a time-integrated image of the magma accumulation zone, often with a history spanning multimillion years (e.g., Greene et al. 2006; Walker et al. 2007; Schoene et al. 2012; Coint et al. 2013), for which we can tease out some information about sizes and shapes of magmas bodies. Much of the geochemical data acquired over the last century is now tabulated in large databases such as Georoc (http://georoc.mpch-mainz.gwdg.de/georoc/) EarthChem (http://www.earthchem.org), and Navdat (http://www.navdat.org), providing a remarkable resource to analyze global geochemical problems (see recent papers of Keller and Schoene 2012; Chiaradia 2014; Gelman et al. 2014; Glazner et al. 2015; Keller et al. 2015).

Studying active volcanoes, gas, and geophysical measurements

Measuring gases coming out of active volcanoes provides us with some key information about the state of magmatic systems (e.g., Giggenbach 1996; Goff et al. 1998; Edmonds et al. 2003; Burton et al. 2007; Humphreys et al. 2009). Due to their low density and viscosity, volcanic gases can escape their magmatic traps and provide “a telegram from the Earth’s interior,” as laid out by one of the pioneers of gas measurements, Sadao Matsuo (e.g., Matsuo 1962). For example, the abundance of chemical species like He, CO2, or Cl can help determine the type of magma that is degassing (Shimizu et al. 2005), and the potential volume that is trapped in the crust (Lowenstern and Hurvitz 2008). As mentioned above, the amount of S released during eruptions is also key in estimating the presence of an exsolved gas phase in the magma reservoir prior to an eruption (Scaillet et al. 1998a; Wallace 2001; Shinohara 2008). As discussed by Giggenbach (1996), the composition of volcanic gases measured at the surface integrates both deep, i.e., magma reservoir processes, and shallow processes within the structure of the edifice, and possibly the interaction with a hydrothermal system. As such it is often challenging to directly relate gas composition to the conditions of shallow magma storage (e.g., Burgisser and Scaillet 2007) and a better monitoring strategy is generally to look for relative changes in gas compositions and temperature rather than focusing on absolute values.

With the exception of gas measurements, imaging active magma reservoirs mainly involves geophysical methods. Several techniques are now available, and provide different pictures of active magmas bodies. Such techniques include: (1) seismic tomography [both active and passive sources (see for example Dawson et al. 1990; Lees 2007; Waite and Moran 2009; Zandomeneghi et al. 2009; Paulatto et al. 2010), and ambient-noise tomography (e.g., Brenguier et al. 2008; Fournier et al. 2010; Jay et al. 2012)]; (2) magneto-telluric surveys (e.g., Hill et al. 2009; Heise et al. 2010); (3) deformation studies using strain-meter, GPS, and satellite data (e.g., Massonnet et al. 1995; Hooper et al. 2004; Lagios et al. 2005; Newman et al. 2006; Hautmann et al. 2014); and (4) muon tomography on steep volcanic edifices (Tanaka et al. 2007, 2014; Gibert et al. 2010; Marteau et al. 2012).

Geophysical methods are based on the inversion of signals transmitted through the crust and measured from or close to the surface. The choice of methods and, in several cases, the frequency band of the signals studied de-pends on the targeted resolution and depth of interest. Geophysical methods probe variations in physical properties caused by the presence of partially molten reservoirs in the crust and provide two- or three-dimensional images of these systems. These inversions however do not provide an unequivocal picture of the thermodynamical state of the magma storage region, because several variables (melt fraction, temperature, exsolved gas content, composition, connectivity of the various phases) affect the elastic, magnetic, and electric properties of magmas. Similarly, surface deformation signals are not only function of the state, shape, orientation, and size of a magma body, but also depend on the crustal response to stresses around the body (through anelastic deformation and movement along weakness planes such as fractures and faults; Newman et al. 2006; Gregg et al. 2013).

In summary, geophysical surveys mainly provide constraints about the depth, shape, and extent of a partially molten (or hot) region under an active volcanic edifice. The many surveys using inSAR, GPS field campaignsm or ambient-noise tomography where, as for gas monitoring, the focus lies on relative changes rather than the effective state of the reservoir have significantly increased our ability to monitor volcanic unrest (see recent review by Acocella et al. 2015). They efficiently highlight rapid (decadal or less) changes in the mechanical state of magmatic systems, which is generally difficult to achieve using petrological data sets. However, the spatial scales probed by geophysical imaging techniques are limited by the resolution of the method considered. Under most circumstances, the spatial resolution remains greater or equivalent to the kilometer scale (Waite and Moran 2009; Farrell et al. 2014; Huang et al. 2015), which is in stark contrast with the extremely high resolution of geochemical and petrological data. As such, geophysical inversions can be thought of as upscaling filters of the state of heterogeneous multiphase reservoirs. A fundamental assumption underlying these inversions is that spatial resolution is com-mensurable with a Representative Elementary Volume (REV) where the physical properties of the heterogeneous medium can be (linearly) averaged out into a single effective value (a more detailed discussion of the limitations imposed by this assumption is provided below).

Reproducing the conditions prevailing in magma chambers in the laboratory: Experimental petrology

1 Phase-equilibrium experiments

Starting with N.L. Bowen at the turn of the 20th century, many geoscientists have applied experimental studies to study the thermodynamics that control the assemblage of phases in magmas with different bulk compositions. For nearly a century, a large amount of data has been collected on crystal-melt stability and composition as a function of several variables (P, T, fugacities of volatile elements and oxygen, compositions of magmas, cooling rate...). Much of these data are now available in databases, some in web-based portals, as is the case for the geochemical data (see for example Berman 1991; Holland and Powell 2011 or LEPR, http://lepr.ofm-research.org/YUI/access_user/login.php). The data sets range from crystal-melt equilibria in the mantle melting zones (e.g., see Poli and Schmidt 1995 to lower crustal (e.g., Muntener and Ulmer 2006; Alonso-Perez et al. 2009) and upper crustal conditions in many different settings (e.g., Johnson and Rutherford 1989; Johannes and Holtz 1996; Scaillet and Evans 1999; Costa et al. 2004; Almeev et al. 2012; Gardner et al. 2014; Caricchi and Blundy 2015).

For shallow magma reservoirs in arcs, a third phase, volatile bubbles, is also commonly present. This exsolved volatile phase does not only affect the thermo-physical properties of magmas (see below) but can also impact greatly the partitioning of some volatile trace elements and therefore influence the chemical fractionation that takes place in magma reservoirs. Laboratory experiments have played a major role in determining the solubility of volatile phases (H2O, CO2, S, CL, F...) as a function of pressure, temperature, and melt composition [see the review by Baker and Alletti (2012) and references therein]. Several species, such as Cl and S compounds, have a strong influence on the fractionation and transport of precious metals out of magma reservoirs to the site of ore deposits and have therefore received a thorough attention (Scaillet et al. 1998a; Williams-Jones and Heinrich 2005; Zajacz et al. 2008; Sillitoe 2010; Fiege et al. 2014).

Chemical equilibrium between the melt and newly formed crystals or crystal rims is often a decent assumption, but this assumption breaks loose when changes taking place in the reservoir are more rapid than the equilibration timescale (see Pichavant et al. 2007). Such rapid changes in reservoirs’ conditions can include: (1) magma recharges and partial (incomplete) mixing; (2) efficient physical separation of mineral and/or bubble-melt; and (3) during rapid pressure drops associated with mass withdrawal events (eruptions, dike propagation out of the chamber). In these particular cases, kinetics controls the mass balance between the different phases; one therefore needs to consider diffusion of elements in silicate melts and minerals, as well as mineral and bubble growth or dissolution. The diffusive transport of chemical species in silicate melts is a strong function of: (1) viscosity (which is a proxy for the degree of polymerization of the melt); (2) dissolved water content; and (3) to a lesser extent, oxygen fugacity (redox conditions in the magma affect speciation). We bring the attention of the interested readers to the review of Zhang and Cherniak (2010) for additional information on this topic.

The presence of kinetics is readily observed in mineral zoning patterns because solid-state diffusion is generally orders of magnitude slower than diffusion in silicate melts. Over the last decade, several models have been developed to use the diffusion of cations in silicate minerals and retrieve timescales relevant to the dynamic evolution of magmas before and during eruptions. These models rest heavily on experiments where the kinetics of diffusion as a function of temperature and composition is measured [see review of Zhang and Cherniak (2010) and reference therein]. Diffusion modeling in minerals informs us on the interval of time between a significant change in thermodynamic conditions of the magma (mixing from recharges for example) and the time at which the minerals transition across the closure temperature (e.g., eruption) where subsequent diffusion can be ignored. To date, studies have explored the diffusion of a growing quantity of cations in hosts such as quartz, plagioclase, pyroxene, biotite, and olivines (for more mafic systems than the ones considered here; see Fig. 2 for published examples).

Figure 2 Examples of diffusional profiles in quartz crystals from Taupo, New Zealand (Matthews et al. 2012), in pyroxene crystals from the Bishop Tuff, California, U.S.A. (Chamberlain et al. 2014a), and plagioclase crystals from Cosiguina, Nicaragua (Longpre et al. 2014).
Figure 2

Examples of diffusional profiles in quartz crystals from Taupo, New Zealand (Matthews et al. 2012), in pyroxene crystals from the Bishop Tuff, California, U.S.A. (Chamberlain et al. 2014a), and plagioclase crystals from Cosiguina, Nicaragua (Longpre et al. 2014).

Stable isotopes can also provide information regarding the thermo-dynamic conditions (equilibrium fractionation between different phases) and kinetics (kinetic fractionation) in the magma. Because of improved accuracy and spatial resolution of analytical measurements, there is a growing interest in using different stable isotope systems, such as O, S, Ca, as well as other major elements, to fingerprint the oxidation state of magmas (Metrich and Mandeville 2010; Fiege et al. 2015), crustal as-similation (e.g., Friedman et al. 1974; Taylor 1980; Halliday et al. 1984; Eiler et al. 2000; Bindeman and Valley 2001; Boroughs et al. 2012), or mixing processes between magmas with different compositions (Richter et al. 2003; Watkins et al. 2009a).

2 Rheological experiments

The migration, convection, and defor-mation of magmatic mixtures in the crust have direct implications on the rate and efficiency of fractionation processes, on the cooling rate of magma reservoirs, and eruption dynamics (e.g., Gonnermann and Manga 2007; Lavallee et al. 2007; Cordonnier et al. 2012; Gonnermann and Manga 2013; Parmigiani et al. 2014; Pistone et al. 2015). The response of magmas to differential stresses is complex, owing to the multiphase nature of these systems. Several research groups have performed deformation experiments to better constrain rheological laws pertinent to magmas (Spera et al. 1988; Lejeune and Richet 1995; Caricchi et al. 2007; Ishibashi and Sato 2007; Champallier et al. 2008; Cimarelli et al. 2011; Mueller et al. 2011; Vona et al. 2011; Pistone et al. 2012; Del Gaudio et al. 2013; Laumonier et al. 2014; Moitra and Gonnermann 2015). These studies have focused on diverse aspects of the rich non-linear dynamics of the rheology of suspensions, such as: (1) the effect of the crystal volume fraction; (2) the effective shear rate; and (3) the effect of polydisperse crystal size distributions and various crystal shapes. These results clearly show that suspensions become very stiff as crystal content approach or exceed 40 to 50% of the volume of the magma, although they provide little information about the onset and magnitude of a yield stress for crystal-rich suspensions. Even more challenging is the development of physically consistent rheology laws for three-phase magmas because of the additional complexity of the capillary stresses (coupling bubbles, melt, and crystals), bubble deformation, three phases lubrication effects and possible coalescence (Pistone et al. 2012; Truby et al. 2014).

Magma rheology experiments have been synthesized to yield different empirical (Lavallee et al. 2007; Costa et al. 2009) and semi-empirical (Petford 2009; Mader et al. 2013; Faroughi and Huber 2015) models for the deformation of magmas under shear flow conditions. The mechanical behavior of magmas remains a rich field for future investigations, and several important challenges remain to constrain the rate at which magma flows in reservoirs and up to the surface. They include:

  • How to treat multiphase rheology when phase separation takes place concurrently?

  • How do bubbles interact with crystals suspended in viscous melts, and how does this behavior depend on the strain rate and volume fractions of each phases?

  • Can we constrain the yield strength of crystal-rich magmas under magma chamber conditions (imposed stresses rather than imposed strain-rate)?

  • By rheology, we often restrict ourselves to measure the shear viscosity using experiments or models that relate stress and deformation under very simple geometries. As viscosity is a tensor, is it possible to measure its anisotropy for non-linear multiphase systems with complex crystal shapes under vari-ous strain rates and conditions pertinent to magma chambers?

The rheology of a magma body should therefore be considered as a dynamical quantity, which can vary significantly (orders of magnitude) because of changes in stress distributions, variations in crystal-melt-bubble phase fractions and potentially many other factors.

Several other physical properties have a resounding impact on the evolution of magmas in the crust. Magmas are far from thermal equilib-rium in the mid to upper crust, and heat transfer in and out of these bodies controls their chemical and dynamic evolution to a great extent (Bowen 1928; McBirney et al. 1985; Nilson et al. 1985; McBirney 1995; Reiners et al. 1995; Thompson et al. 2002; Spera and Bohrson 2004). In that context, the heat capacity, latent heat, and thermal conductivity of silicate melts/minerals and CO2-H2O fluids are important to establish the timescale over which these reservoirs crystallize in the crust and by extension the vigor of convective motion in these systems. It is important to emphasize that thermal convection in the classical sense is not directly applicable to magma chambers under most conditions, because density contrasts between the different phases far exceeds those by thermal expansion (e.g., Marsh and Maxey 1985; Bergantz and Ni 1999; Dufek and Bachmann 2010). Additionally, the rate of crystallization directly impacts the cooling rate of magmas (due to non-linear release of latent heat during crystallization, e.g., Huber et al. 2009; Morse 2011), and this latent heat buffering becomes dominant as silicic magmas approach eutectic behavior near their solidus (Gelman et al. 2013b; Caricchi and Blundy 2015). Some recent studies have tested how the temperature dependence of thermal properties influence the cooling and crystallization timescales of magmas in the crust and showed that these effects are sometimes non-trivial (Whittington et al. 2009; Gelman et al. 2013b; de Silva and Gregg 2014).

Tank experiments for magma chamber processes

An elegant approach to study the complex multiphase fluid dynamics that takes place in magma reservoir is to resort to laboratory tank experiments (see for example,McBirney et al. 1985; Nilson et al. 1985; Jaupart and Brandeis 1986; Martin and Nokes 1988; Shibano et al. 2012). Experiments provide the means to investigate some aspects of the complex non-linear dynamics that prevail in these systems, however, in most cases, it is difficult or impossible to satisfy dynamical similitude with real magmatic systems. Nevertheless, experiments reveal interesting feedbacks that may have been overlooked and can serve as a qualitative assessment of how processes couple to each other. 0ne of the main directions for laboratory fluid dynamics experiments is to study how magmatic suspensions behave at low Reynolds number during convection, phase segregation, and magma recharges with mixing (e.g., McBirney 1980; Sparks et al. 1984; McBirney et al. 1985; Davaille and Jaupart 1993; Koyaguchi et al. 1993). Here, the term suspension is used loosely to describe both suspension with solid particles (crystals) and bubbly emulsions.

As discussed throughout this review [and clearly alluded to by early petrologists such as R. Daly as far back as the 1910s, e.g., Daly (1914)], magma bodies are open reservoirs that exchange mass and energy with their surroundings. Chemical heterogeneities and zonations then reflect either phase separation by gravity (e.g., McBirney 1980, 1993; Hildreth 1981; Sparks et al. 1984) or incomplete mixing between magmas with different compositions (Eichelberger 1975; Dungan et al. 1978; Blake and Ivey 1986; Eichelberger et al. 2000).

Because buoyancy forces related to compositional differences (including presence of various phases of variable densities) exceed thermal buoyancy effect by several orders of magnitude, mechanical convection and stirring are dominated by chemical variations, bubble rise, and/or crystal settling. Strong spatial contrasts in bulk viscosity because of mixing between magmas with different compositions, contrasts in crystal content or thermal gradients also tend to stabilize against hybridization or magma mixing (Sparks and Marshall 1986; Turner and Campbell 1986; Koyaguchi and Blake 1989; Davaille and Jaupart 1993; Jellinek and Kerr 1999). Mixing in magma reservoirs can also be accelerated during eruptions, where substantial stress differentials can be generated by decompression or roof collapse (Blake and Campbell 1986; Blake and Ivey 1986; Kennedy et al. 2008).

Convection strongly influences phase separation, which in turn controls chemical differentiation. Several groups have investigated the effect of crystals (Jaupart et al. 1984; Brandeis and Jaupart 1986; Brandeis and Marsh 1990; Koyaguchi et al. 1993; Shibano et al. 2012), bubbles (Huppert et al. 1982; Thomas et al. 1993; Cardoso and Woods 1999; Phillips and Woods 2002;Namiki et al. 2003), and shapes of reservoirs (e.g., de Silva and Wolff 1995) on convective motion. The presence of crystals or bubbles controls the gravitational energy potential that is converted to kinetic energy during convection, as experiments confirm that they impact convection even at relatively small volume fractions. Obviously, suspended phases affect convection, but convection itself also influences crystal settling or bubble migration. For example, Martin and Nokes (1988, 1989) used a series of experiments to quantify the timescale over which a convecting suspension of crystals clears out because of Stokes settling. Hydrodynamic interactions between the different phases can significantly reduce the settling velocity and affect phase separation even at low-particle (bubbles or crystals) volume fractions (Faroughi and Huber 2015, see Fig. 3).

Figure 3 Injection of negatively buoyant droplets of (low-viscosity) water dyed in green into a tank filled with (more viscous) silicon oil (from Faroughi and Huber 2015). Droplet interactions, even in the presence of droplet trains, can significantly reduce the settling velocity and decrease the rate of phase separation.
Figure 3

Injection of negatively buoyant droplets of (low-viscosity) water dyed in green into a tank filled with (more viscous) silicon oil (from Faroughi and Huber 2015). Droplet interactions, even in the presence of droplet trains, can significantly reduce the settling velocity and decrease the rate of phase separation.

Overview of Magma Reservoir Dynamics from Theoretical Models

Numerical studies enable us to isolate specific dynamical feedbacks and to provide a framework to understand the complex temporal evolution of these systems, which can often be challenging to infer from natural samples or laboratory experiments where dynamic similarity is difficult to satisfy. Numerical approaches based on different sets of starting assumptions and governing equations have been used to address questions such as:

  • The chemical evolution of open-system magma res-ervoirs undergoing assimilation, crystal fractionation and magma evacuation.

  • The thermal evolution and longevity of magma res-ervoirs in the crust.

  • The relative importance of crustal melting vs. crystal fractionation of more mafic parents in driving the evolution of magmas toward silicic compositions.

  • The pressure variations in and around magma cham-bers and stability of reservoirs following recharges.

  • Gas exsolution and how it impacts the dynamics in shallow magma reservoirs.

1 Box models (no spatial dimensions)

The first category of studies summarized below is based on box models where the reservoirs are considered homogeneous and internal structures are not explicitly accounted for. These models involve ordinary differential equations that depend solely on time (no spatial dependency) and by extension assume that homog- enization takes place over a shorter timescale than the processes considered. In other words, in these models, the different phases (melt, crystals, sometimes bubbles) are assumed to remain in thermal and chemical equilibrium with each other over the scale of the whole reservoir.

Since the seminal work of H. Taylor and D. DePaolo (Taylor 1980; DePaolo 1981), where a box model was constructed to infer the relative importance of assimilation and crystal fractionation processes on the trace elements and isotopic record, many studies have expanded on the method to relate chemical tracers to the actual dynamical processes that control the evolution of magmatic systems. The scope of these models ranges from non-linear reactor models (Bonnefoi et al. 1995) investigating the development of bimodal compositions from a single magmatic system to constraining the effect of melt extraction on the chemical signature of the residual cumulate and high-silica melts (Gelman et al. 2014; Lee and Morton 2015) as well as focusing on the development of crystal zonations in a well-mixed reservoir (Wallace and Bergantz 2002; Nishimura 2009). These models provide insights into the chemical response of reservoirs to various fractionation processes, but do not explicitly treat the source of these processes (e.g., cooling and crystallization leading to gravitational settling or the reheating of the surrounding crust and its partial melting/ assimilation).

As a first step toward answering these limitations, W. Bohrson and F. Spera developed, in a series of papers (Spera and Bohrson 2001, 2004; Bohrson and Spera 2007; Bohrson et al. 2014), a model that couples the mass balance of trace elements and isotopes in a reservoir where cooling and assimilation processes are parameterized. They also added a thermal energy equation for the reservoir. The energy balance in open magma chambers is quite complex and involves more than just sensible and latent heat contributions in response to the heat loss to the surrounding crust. Even within the assumption that the reservoir remains homogeneous and the phases remain in equilibrium at all times, enthalpy is constantly transferred from mechanical work to heat and vice versa. As an example, the injection of fresh magma into a reservoir affects the energy budget not only by providing a possible heat source but also exerts mechanical work (pressure changes) that affects the stability of the phases present in the magma. These feedbacks become even more important when shallow volatile-saturated systems are considered (Huppert and Woods 2002; Degruyter and Huber 2014; Degruyter et al. 2015). It is therefore important to extend the energy conservation statement in these models to include mechanical work and phase changes when possible [hence introducing a consistent two-way coupling between the hot magma reservoir and the surrounding crust (de Silva and Gregg 2014; Degruyter and Huber 2014)].

2 Models with spatial dimensions

The assumption of thermal and chemical equilibrium at the scale of the reservoir is useful and allows us to draw interesting and informative conclusions on the general trends that govern magma chamber evolution. However, this assumption is unten-able under most conditions. Introducing spatial heterogeneities in the magma body and therefore allowing for some degree of disequilibrium between the phases (at least down to the scale of the model’s resolution, where local equilibrium is generally still assumed between phases) is necessary to get realistic estimates of the temporal scales over which magma bodies evolve. More specifically, a spatiotemporal description of magma reservoirs provides the opportunity to test for the conditions under which efficient mixing is no longer possible (Huber et al. 2009) and chemical differentiation by mechanical separation (flow) occurs. We will divide the spectrum of numerical models geared to ad-dress these questions into: (1) thermal models on one hand and (2)coupled thermal and mechanical models on the other hand.

Thermal models. Thermal models are concerned only with one aspect of the energy balance and consider only sensible and latent heat. These models involve various levels of complexity, from equilibrium crystallization, i.e., the amount of crystallization over each time step is that predicted from equilibrium thermodynamics (e.g., Annen et al. 2008; Leeman et al. 2008; Annen 2009; Gelman et al. 2013 b; Karakas and Dufek 2015) to the model of C. Michaut and coworkers where crystallization is considered kinetically limited (Michaut and Jaupart 2006). Equilibrium crystallization models rely on a constitutive equation that relates temperature and crystallinity for a given choice of magma composition and pressure, generally constrained experimentally or using thermodynamic models such as MELTS (Ghiorso and Sack 1995b). Because no momentum conservation is considered, crystals and melt are static and bound to each other, i.e., no phase separation and by extension no chemical differentiation by crystal fractionation is possible. Heat is therefore transported only by diffusion within and out of the magma reservoir. These models have been used extensively to study the longevity of active magma bodies in the crust as a function of the rate of magma injections. While they suggest high average rates of magma injections to counteract the heat lost to the crust, they overlook the role of mechanical pro-cesses and volatile exsolution on the energy balance in response to repeated magma recharges and sporadic evacuation events (Degruyter et al. 2015).

Thermo-mechanical models. Another type of model accounts for the coupled momentum and enthalpy balance in magma reser-voirs and has been used to model the thermal evolution as well as chemical differentiation (by crystal fractionation and assimilation) in magmatic systems. In these models, the momentum balance is either parameterized (Huber et al. 2009) or requires a set of coupled continuum equations that allows the different phases to separate from one another (Dufek and Bachmann 2010; Gutierrez and Parada 2010; Lohmar et al. 2012; Simakin and Bindeman 2012; Solano et al. 2012). In those latter models, the different phases are coupled in the momentum equation through drag terms. The first numerical models applied to mixing of chemical heterogeneities (single-phase fluids) in magmas by convective stirring where developed by C. 0ldenburg and colleagues (Oldenburg et al. 1989; Spera et al. 1995), highlighted the different rates of mixing by normal and shear strains. Later, Huber et al. (2009) discussed a method to characterize the efficiency of mixing by convective stirring in time-dependent systems where cooling and crystallization dampen the efficiency of convection over time. Multiphase mixing by sinking crystal plumes initiated at thermal boundary layers has been studied extensively as the source of convection in magmas (Bergantz and Ni 1999; Dufek and Bachmann 2010; Simakin and Bindeman 2012). The focus of these models is to look at crystal fractionation (Bergantz and Ni 1999; Dufek and Bachmann 2010; Solano et al. 2012) and crustal assimilation (Simakin and Bindeman 2012) as sources of chemical differentiation for magmas emplaced in the crust.

At this stage, a model that involves a thermo-mechanical solution to the crust-magma body systems and that also solves for the internal fluid mechanics and chemical evolution of the reservoir is yet to be completed [see Gregg et al. (2013) and Degruyter and Huber (2014) for preliminary attempts to do this]. Some of the processes associated with the mechanical response to these reservoirs to mass addition or withdrawal can operate over much faster timescales than the thermal and chemical evolution of the reservoir and coupling these various processes proves to be challenging.

The mechanical and chemical interaction between the many phases that coexist in magma reservoirs is complex. It is generally parameterized with simple constitutive empirical relationships (e.g., drag terms, interphase heat transfer coefficients). Because models are only as good as the assumptions they rely on, it is important to revisit the validity of these constitutive models that introduce the dynamical coupling between the different phases in continuum models. In truth, these interaction terms involve length scales that are far beyond the resolution of continuum models, and it is the interplay between crystals, melt, and, possibly, gas bubbles that drive most of the dynamics in magma bodies. One example for such an empirical law is the mechanical separation between crystals and the melt by gravity, which controls chemical differentiation. Below a threshold crystallinity, gravitational settling laws based on Stokes settling are generally assumed valid (Martin and Nokes 1989). However, it assumes that crystals are far apart and never interact hydrodynamically and that there is no return flow of melt to conserve the mass flux over any horizontal surface in the magma body, i.e., the magma body has an infinite volume. The mechanics of multi-particles gravitational settling is quite complex though and requires an account of both short- and long-range hydrodynamic corrections even at crystal volume fractions significantly below 10% (Koyaguchi et al. 1990, 1993; Faroughi and Huber 2015). Another example lies in a proper way to account for the thermal energy balance in multiphase media (crystals+melt for example). At the continuum scale, the two general approaches commonly used assume either that. over the resolution of the numerical model, the two phases are in thermal equilibrium (local thermal equilibrium conditions or LTE) as in Simakin and Bindeman (2012) or that the contrast in thermal properties requires a thermal lag or inertia between the phases parameterized with inter-phase heat transfer coefficients (local thermal non-equilibrium conditions or LTNE) as in Dufek and Bachmann (2010). Although the latter thermal model (LTNE) is more realistic and provides better results (the current state-of- the-art), it has several issues that sometimes precludes its use for multiphase heat transfer (e.g., Karani and Huber 2015). There is therefore a need to study multiphase magmas at the discrete scale and develop new continuum-scale conservation laws that do a more careful job when filtering out heterogeneities during volume averaging (see Frontiers section).

3 Future development in modeling

A possible solution to resolve the actual thermal and me-chanical interactions in multiphase magmas is to resort to granu-lar (discrete) scale models as a complement to these continuum scale studies. The fundamental difference between continuum and granular scale models is that the latter explicitly solve for the mass (including chemical exchanges), momentum (e.g., drag), and energy (e.g., heat transfer) exchange between the different components through their interfaces, while the former involves a parameterization of these interactions or, in some cases (e.g., effective medium theory), relies on the definition of an effective medium with hybrid (parameterized) properties. The granular- scale approach offers a significant advantage by solving a more realistic model and allows us to study complex feedbacks arising from the interaction of multiple phases, but at the cost of heavy computational requirements.

Several “granular” multiphase models have been developed over the last decade to study magma dynamics. For example, multiphase fluid dynamics modeling where crystals are introduced using the discrete element method (DEM) coupled to Stokes flow solvers has been applied to magma chamber dynamics recently by Furuichi and Nishiura (2014) and Bergantz et al. (2015), considering mostly how crystal melt mechanical interactions affect settling in magma chambers. Granular-scale models, using the lattice Boltzmann technique, have also permitted to study the migration of exsolved volatile bubbles in magma chambers at high crystal content (see Fig. 4; Parmigiani et al. 2011, 2014; Huber et al. 2012b). As mentioned above, discrete models (for the dispersed phase) are valuable in that they provide an accurate description of the actual physics that governs the mass, momentum, and energy exchange between phases, but they are generally limited by computer power to small computational domains and require upscaling strategies to extrapolate the dynamics to the scale of the reservoir. Upscaling multiphase dynamics is and will remain a great challenge in the years to come to better constrain the chemical and dynamical evolution of magma bodies in the crust (see Frontiers topic at the end of this review).

Figure 4 Example of granular scale calculation for multiphase magma chamber dynamics. (a-c) modified from from Furuichi and Nishiara (2014) for crystal settling from a stratified chamber with a crystal-free region at the bottom and a denser, crystal-rich horizon, at the top, and (d) Parmigiani et al. 2011 for the migration of a buoyant volatile phase in a crystal-rich rigid magma.
Figure 4

Example of granular scale calculation for multiphase magma chamber dynamics. (a-c) modified from from Furuichi and Nishiara (2014) for crystal settling from a stratified chamber with a crystal-free region at the bottom and a denser, crystal-rich horizon, at the top, and (d) Parmigiani et al. 2011 for the migration of a buoyant volatile phase in a crystal-rich rigid magma.

Model of Magma Reservoir: Focusing on the “Mush”

After reviewing most of the techniques and tools used to study magma reservoirs in the Earth’s crust, the next section will focus on a specific model of magma reservoir evolution, the so-called “Mush Model” (Bachmann and Bergantz 2004; Hildreth 2004; Huber et al. 2009), strongly influenced by the pioneering efforts of many previous researchers (e.g., Smith 1960, 1979; Hildreth 1981; Lipman 1984; Bacon and Druitt 1988b; Brophy 1991; Sinton and Detrick 1992 to cite a few). Although we fully admit that this choice is strongly dictated by our previous experience, we believe it helps providing a framework to many of the questions we listed in the introduction. Obviously, the model needs many refinements, and we will try to bring out the caveats as we see them. We are also aware of the controversial aspects that remain to be resolved (e.g., Glazner et al. 2008, 2015; Tappa et al. 2011; Simakin and Bindeman 2012; Rivera et al. 2014; Streck 2014; Wotzlaw et al. 2014; Keller et al. 2015), but hope that such a review can provide a stepping stone to bring the discussion forward.

Since volcanic rocks are, at times, relatively crystal poor [particularly the compositional extremes, basalts and rhyolites, whereas intermediate magmas are often more crystal rich (Ewart 1982)], researchers have tended to draw magma reservoirs as dominantly liquids [big pools of liquid magmas for both mafic and silicic units, see Yellowstone reservoir depicted on Figure 1 (Fig. 9-14 of McBirney 1993; Irvine et al. 1998; Maughan et al. 2002)], which easily caught on in the general public’s views of magma reservoirs. However, as magmas are emplaced in contact with a colder crust, heat loss to the surrounding crust limits the time magmas remain in a mostly liquid state in these reservoirs. Hence, many researchers argued that solidification zones [or fronts, (Marsh 1981, 1989b, 2002; Marsh and Maxey 1985; Gutierrez et al. 2013; Fig. 5)] are likely to develop quickly, and form a crystal-rich buffer zone between the hottest part, most liquid part of the reservoir, and the subsolidus wall rocks. Since, many lines of evidence have suggested that magmas are dominantly kept as high-crystallinity bodies in the Earth’s last few tens of kilometers. These high-crystallinity bodies are often referred to as “crystal mushes.” Miller and Wark (2008) defined crystal mush as a “mixture of crystals and silicate liquid whose mobility, and hence eruptibility, is inhibited by a high fraction of solid particles.” (Miller and Wark 2008). This is the case for subvolcanic reservoirs (Hildreth 2004; Bachmann et al. 2007b) in large silicic systems, but also at Mid 0cean Ridges (Sinton and Detrick 1992). The supporting evidences for the importance of crystal mushes stem from petrological, geophysical, geochemical, and geological observations, as well as thermal modeling. These are detailed below.

Figure 5 Crystallinity variations in magmas, from solidus to liquidus, which can span up to 250 °C (modified from Marsh 1996). The typical amount of heat liberated by 1 kg of magma from liquidus to solidus is ~600 000 J, of which ~1/2 is from latent heat.
Figure 5

Crystallinity variations in magmas, from solidus to liquidus, which can span up to 250 °C (modified from Marsh 1996). The typical amount of heat liberated by 1 kg of magma from liquidus to solidus is ~600 000 J, of which ~1/2 is from latent heat.

As the temperature contrast with the surrounding crust diminishes, the cooling rate also decrease and magmas spend more time at temperature close to their solidus (Marsh 1981; Koyaguchi and Kaneko 1999; Huber et al. 2009). The waning of convection also plays a role decreasing the cooling rate with decreasing T. As stirring occurs, it steepens the temperature gradient at the edge of the body, and tends to hasten the cooling. Since convection only occurs at low crystallinity, it enhances the cooling near the liquidus (see Fig. 6 of Huber et al. 2009). Finally, latent heat released during crystallization provides energy that needs to be removed before subsequent cooling is possible. If the temperature vs. crystallinity diagram is relatively linear, latent heat is released equally over the cooling interval between the liquidus and solidus, and does not favor a prolonged state at either the low- or high-crystallinity ranges. However, the temperature vs. crystallinity relationship is quite complex and non-linear for many types of magmas, particularly those with near-eutectic points. For example, evolved high-Si02 magmas rapidly reach the haplogranite eutectic and are thus characterized by a strongly non-linear phase diagram [e.g., dacitic magmas reach near-eutectic conditions at 40-50 vol% crystals (Bachmann et al. 2002)]. The latent heat is mostly released when the magma approaches this eutectic condition and provides an additional thermal buffering effect, i.e., slows down the cooling rate significantly from that point onward [a process called “mushification” by Huber et al. (2009)]. Magma recharges can provide a sufficient amount of enthalpy to slow or even reverse the cooling trend generally observed in these reservoirs, which can prolong their existence at low-melt fraction (Annen and Sparks 2002; Annen et al. 2006; Leeman et al. 2008; Annen 2009; Gutierrez et al. 2013; Gelman et al. 2014), see Figure 6 for an attempt to illustrate the temperature-time evolution in the warmest parts of the reservoirs. The maturation of a magmatic field through repeated intrusions of magmas appears to also play an important role in priming the thermal state of the crust to host long-lived active magma reservoirs (e.g., Lipman et al. 1978; de Silva and Gosnold 2007a; Lipman 2007; Grunder et al. 2008).

Figure 6 Schematic temperature-time diagram for a part of mature magma reservoirs situated in the hottest, core zone. Note the slower cooling rates as the crystal content increases (magma approaching their solidus).
Figure 6

Schematic temperature-time diagram for a part of mature magma reservoirs situated in the hottest, core zone. Note the slower cooling rates as the crystal content increases (magma approaching their solidus).

Evidence from erupted rocks

The eruption of large, homogeneous crystal-rich units, such as the Fish Canyon Tuff (Whitney and Stormer 1985; Lipman et al 1997; Bachmann et al. 2002), the Lund Tuff (Maughan et al. 2002), or La Pacana Tuff (Lindsay et al. 2001) have also been pivotal in the development of the Mush Model (see below). Such “monotonous intermediates” (Hildreth 1981) are common in large magmatic provinces, and unambiguously document the presence of large, homogeneous zones of crystal-rich magmas in the upper crust, potentially remobilizable by eruption. Their homogeneity is striking, and requires a relatively thorough stirring prior to eruption, likely triggered by a recharge and remobilization prior to eruption [see below, Huber et al. (2012a) and Parmigiani et al. (2014) for reviews].

Evidence from geophysics

Large, low-seismic velocity zones and conductive MT areas beneath active volcanoes or volcanic provinces also suggest the presence of near-solidus magmas bodies. Although the resolution remains poor (hundreds of meters to kilometers), the anomalies (both in terms of electric conductivity and seismic velocities) and the size of those bodies are not supportive of dominantly molten bodies (e.g., Dawson et al. 1990; Lees 1992; Weiland et al. 1995; Steck et al. 1998; Zollo et al. 1998; Masturyono et al. 2001; Zandt et al. 2003; De Natale et al. 2004; Husen et al. 2004; Hill et al. 2009; Waite and Moran 2009; Heise et al. 2010; Farrell et al. 2014; Ward et al. 2014). Estimates of melt range from a few percent to nearly 50% melt in some areas of the anomalies (Heise et al. 2010; Farrell et al. 2014). In addition to those seismic and MT anomalies, Bouguer gravity anomalies under older volcanic fields also suggest the accumulation of low-density, silicic rocks in the upper crust (e.g., SRMVF, Fig. 7, Plouff and Pakiser 1972; Lipman 1984; Drenth and Keller 2004; Drenth et al. 2012; Lipman and Bachmann 2015).

Figure 7 Comparison between the SRMVF batholith model (Lipman and Bachmann 2015). (a and b) Schematized cross sections based on erupted products and gravity data and a joint ambient noise- receiver function inversion S-velocity model for the Altiplano region of the Andes (c) Ward et al. (2014).
Figure 7

Comparison between the SRMVF batholith model (Lipman and Bachmann 2015). (a and b) Schematized cross sections based on erupted products and gravity data and a joint ambient noise- receiver function inversion S-velocity model for the Altiplano region of the Andes (c) Ward et al. (2014).

Evidence from presence of cumulate lithologies in volcanic and plutonic units

The presence of crystal-rich, cumulate blocks erupted in volcanic units (Wager 1962; Arculus and Wills 1980; Heliker 1995; Ducea and Saleeby 1998), as well as large plutonic masses in crustal sections with clear crystal accumulation zones (Voshage et al. 1990; Greene et al. 2006; Jagoutz and Schmidt 2012; Otamendi et al. 2012; Coint et al. 2013) suggest the presence of mush zones within the crust in active magma provinces. Cumulate signatures are not only seen in deep crustal zones, but also in shallow plutonic bodies, both using bulk chemistry (McCarthy and Groves 1979; Bachl et al. 2001; Miller and Miller 2002; Gelman et al. 2014; Lee and Morton 2015) and/or textural features (Beane and Wiebe 2012; Graeter et al. 2015).

Over the next sections, we will try to revisit the questions asked in the introduction, and see how mush zones (or the “mush model”) can help accounting for the main observations that we have laid out.

Presence of crystal-poor rhyolites, and zoned ignimbrites

First of all, we observe many silicic crystal-poor deposits at the surface of the Earth (see Hildreth 1981; Lindsay et al. 2001; Mason et al. 2004; Huber et al. 2012a). Hence, the extraction of high-SiO2, cold (<800 °C), viscous melt from crystalline residue in large quantities (several hundreds of km3) clearly happens in many cases. In some locations, these crystal-poor pockets of erupted magmas are homogenous (Dunbar et al. 1989; Ellis and Wolff 2012; Ellis et al. 2014), in others they grade into crystal-rich, typically hotter intermediate magmas at the end of the eruption (upper parts of the deposits; Lipman 1967; Hildreth 1981; Bacon and Druitt 1988a; Wolff et al. 1990; Deering et al. 2011b; Huber et al. 2012a; Lipman and Bachmann 2015; Fig. 8).

Figure 8 aVariations in SiO2 and crystal contents for ignimbrites in western United States (LC = Lava Creek Tuff; T = Tshigere Member of Bandelier Tuff; BT = Bishop Tuff; WP = Wason Park Tuff; CR = Carpenter Ridge Tuff; RC = Rat Creek Tuff; NM = Nelson Mountain Tuff; AT = Ammonia Tanks Tuff; FC = Fish Canyon Tuff; BC = Blue Creek Tuff; SM = Snowshoe Mountain Tuf; MP = Masonic Park Tuff). Modified from Hildreth (1981) and Huber et al. (2012a); data from Hildreth (1981) and Lipman (2000, 2006). (b) REE patterns from crystal-poor pumices and crystal-rich clasts from the Carpenter Ridge Tuff, with, for reference, patterns for lamproitic magmas, indicating that mixing with such high-K, incompatible-element-enriched liquids is not an option to generate the high-Ba-Zr composition of the late-erupted crystal-rich clasts (modified from Bachmann et al. 2014).
Figure 8

aVariations in SiO2 and crystal contents for ignimbrites in western United States (LC = Lava Creek Tuff; T = Tshigere Member of Bandelier Tuff; BT = Bishop Tuff; WP = Wason Park Tuff; CR = Carpenter Ridge Tuff; RC = Rat Creek Tuff; NM = Nelson Mountain Tuff; AT = Ammonia Tanks Tuff; FC = Fish Canyon Tuff; BC = Blue Creek Tuff; SM = Snowshoe Mountain Tuf; MP = Masonic Park Tuff). Modified from Hildreth (1981) and Huber et al. (2012a); data from Hildreth (1981) and Lipman (2000, 2006). (b) REE patterns from crystal-poor pumices and crystal-rich clasts from the Carpenter Ridge Tuff, with, for reference, patterns for lamproitic magmas, indicating that mixing with such high-K, incompatible-element-enriched liquids is not an option to generate the high-Ba-Zr composition of the late-erupted crystal-rich clasts (modified from Bachmann et al. 2014).

Generating these eruptible silicic pockets is challenging, as extracting such viscous melt [up to 105—106 Pa s, even with several wt% dissolved volatiles (Scaillet et al. 1998b)] from a network of millimeter-sized crystals is, at best, sluggish (McKenzie 1985; Wickham 1987) and the process can be easily disrupted. For example, convection currents should not occur, as this will significantly stir the whole magma reservoir, and lead to re-homogenizing of the crystal-melt mixture (i.e., impeding crystal-melt separation). Hence, a possibility, particularly in long-lived, incrementally built, sill-like magma bodies, is that melt is slowly extracted by gravitational processes (hindered settling, microsettling, compac-tion) as the magma reaches the rheological lock-up [~50 vol% (Brophy 1991; Thompson et al. 2001; Bachmann and Bergantz 2004; Hildreth 2004; Solano et al. 2012)], an extraction process potentially enhanced by gas filter-pressing (Sisson and Bacon 1999; Bachmann and Bergantz 2006; Ellis et al. 2014; Pistone et al. 2015). This melt extraction following lock-up in large mush zones has the advantage of: (1) preventing stirring and mixing of the crystal-melt mixture by convective currents, and (2) keeping the crystal-poor melt pocket in a warm environment, leading to significantly reduced crystallization rate and more time for crystal-liquid separation. Moreover, a sill-like geometry (see Fig. 9) optimizes extraction as it reduces the average vertical distance traveled by the viscous melt. The interstitial melt extraction from mushes seems not to be restricted to evolved magmas, but may also occurred at more mafic, less viscous compositions, such as basalts and andesites (Dufek and Bachmann 2010).

Petrological observations in many large, zoned ignimbrites support such a magma reservoir model, with a cap of crystal-poor material immediately underlain by a more crystal-rich zone. Some well-exposed deposits such as the 7700 BP Crater Lake, or 1912 Katmai eruptions show abrupt changes in crystallinity, from nearly crystal free early in the eruption, to 40-50 vol% crystals late in the sequence, with a very similar melt composition throughout the stratigraphy (Bacon and Druitt 1988a; Hildreth and Fierstein 2000; Bacon and Lanphere 2006). Other large ignimbrites present more gradual variations, as exemplified by famous Bishop and Bandelier Tuffs (Hildreth 1979; Hildreth and Wilson 2007; Wolff and Ramos 2014) among many others (see Hildreth 1981 and Bachmann and Bergantz 2008a for lists of these deposits). Such zoned ignimbrites are best explained by erupting a melt- rich cap followed by partial remobilization and entrainment of a silicic cumulate from the mushy roots of the system, chocking the eruption (see Smith 1979; Worner and Wright 1984; Deering et al. 2011b; Bachmann et al. 2014; Sliwinski et al. 2015; Wolff et al. 2015; Evans et al. 2016). This partial cumulate remobilization is likely a consequence of hotter recharge at the base of the silicic cap, as suggested by textural, mineralogical, and geochemical features indicative of mixing and reheating (Anderson et al. 2000; Wark et al. 2007; Bachmann et al. 2014; Wolff and Ramos 2014; Forni et al. 2015).

Figure 9 Schematic diagram of the polybaric mush model [modified from Lipman (1984), Hildreth (2004), and Bachmann and Bergantz (2008c). (1) Pre-existing crust, (2) upper mantle, (3) feeding zone of primitive magmas from the mantle (“basalt s.l.”), (4) lower crustal mush zone, with internal variability in melt content, (5) upper crustal mush zone, (6) melt-rich pockets in upper crust, (7) melt-rich pockets in the lower crust, (8) caldera structure, (9), stratovolcano (e.g., Mount St. Helens, Washington, U.S.A.).
Figure 9

Schematic diagram of the polybaric mush model [modified from Lipman (1984), Hildreth (2004), and Bachmann and Bergantz (2008c). (1) Pre-existing crust, (2) upper mantle, (3) feeding zone of primitive magmas from the mantle (“basalt s.l.”), (4) lower crustal mush zone, with internal variability in melt content, (5) upper crustal mush zone, (6) melt-rich pockets in upper crust, (7) melt-rich pockets in the lower crust, (8) caldera structure, (9), stratovolcano (e.g., Mount St. Helens, Washington, U.S.A.).

Other processes, such as incomplete/partial mixing between different upper crustal magma pockets and recharge (e.g., Dorais et al. 1991; Mills et al. 1997; Eichelberger et al. 2000; Bindeman and Valley 2003) or co-erupting physically separated and chemically different melt-rich lenses (Gualda and Ghiorso 2013 a), have also been suggested to explain the ubiquitous chemical zoning in ignimbrites. Although we do not question that such processes are likely playing a role in generating chemical complexities in the erupted deposit, these hypotheses fail to account for the presence of co-magmatic crystal-rich cumulate blocks that erupt concurrently with the crystal-poor parts of the reservoirs [see also Ellis et al. (2014) for a case in a hot-dry rhyolitic system, and “The unzoned, crystal-poor to intermediate deposits” section below]. Hence, we argue that magma recharge reaching the base of one or several crystal-poor caps above a cumulate zone (as depicted in Fig. 9) is the most likely reservoir geometry to explain such deposits.

The “monotonous intermediates”—Remobilized mushes

A second conspicuous group of ignimbrites are the “monotonous intermediates,” consisting of homogeneous crystal-rich, typically dacitic units, without any evidence for significant crystal accumulation (Hildreth 1981; Lindsay et al. 2001; Maughan et al. 2002; Folkes et al. 2011c; Huber et al. 2012a). Such units typically display corrosion textures on the low-temperature minerals (quartz and feldspars; Fig. 10) and pervasive reheating prior to eruption (Bachmann and Dungan 2002; Bachmann et al. 2002). Smaller eruptions of crystal-rich magmas [e.g., Montserrat (Murphy et al. 2000), Kos Plateau Tuff (Keller 1969; Bachmann 2010a), Taupo Volcanic Zone (Molloy et al. 2008), Cascade arc (Cooper and Kent 2014; Klemetti and Clynne 2014)] also show a very similar reheating and partial melting signal following recharge from deeper, hotter magmas (although melting evidence can be subtle or even nonexistent in some cases). The physical processes acting upon this remobilization process have been explored by several papers (Bacon and Druitt 1988b; Druitt and Bacon 1989; Couch et al. 2001; Huber et al. 2010a, 2011; Burgisser and Bergantz 2011; Karlstrom et al. 2012). The reheating signal and the partial melting signature in minerals favor a model whereby the reactivation results from the combination of melting and pore pressure increase in the mush in response to melting and the slow and progressive disaggregation of the weakened mush, a process coined as thermomechanical remobilization by Huber et al. (2010a). This model is also consistent with several striking characteristics (whole-rock homogeneity, high crystallinity, partial corrosion of minerals) that such deposits display (Huber et al. 2012a; Cooper and Kent 2014; Parmigiani et al. 2014).

Figure 10 Phase maps of crystal-rich ignimbrites, both remobilized crystal mushes at different stages of rejuvenation; the Masonic Park Tuff no longer has any sanidine and shows only tiny quartz microcrysts, whereas the Fish Canyon Tuff still shows large, albeit highly resorbed quartz and sanidine phenocrysts (Bachmann et al. 2002). In both cases, plagioclase and mafic crystals (hornblende in the FCT, and pyroxene in the MPT) show reverse zoning (Lipman et al. 1996; Bachmann and Dungan 2002).
Figure 10

Phase maps of crystal-rich ignimbrites, both remobilized crystal mushes at different stages of rejuvenation; the Masonic Park Tuff no longer has any sanidine and shows only tiny quartz microcrysts, whereas the Fish Canyon Tuff still shows large, albeit highly resorbed quartz and sanidine phenocrysts (Bachmann et al. 2002). In both cases, plagioclase and mafic crystals (hornblende in the FCT, and pyroxene in the MPT) show reverse zoning (Lipman et al. 1996; Bachmann and Dungan 2002).

Thermo-mechanical remobilization of locked areas following recharge is likely to be a very common process in incrementally built magma reservoirs. In most cases, such remobilization events would not lead to an eruption, but could potentially be inferred from surface deformation, change in gas outputs, and/or seismic signals related to the emplacement of the recharge. When partial melting of crystal-rich, mostly anhydrous material, occurs, the new melt released is dry, and will impact the rheological response and phase diagram of the newly produced mixture (Evans and Bachmann 2013; Caricchi and Blundy 2015; Wolff et al. 2015). As drier melt increases the temperature of mineral precipitation (e.g., Johannes and Holtz 1996), the amount of melting will tend to be limited in most cases (Huber et al. 2010a; Wotzlaw et al. 2013). Hence, the condition for eruption during remobilization might only be reached when significant heat and volatiles (mainly H2O) are injected by the incoming recharge. Exactly how much heat and volatiles need to be injected strongly depends on several factors (including relative size of recharge and mush, geometry, compositions of magmatic end-members, average crystallinity of the mush, ...), which will vary from case to case and are difficult to predict.

The unzoned, crystal-poor to intermediate deposits

There is a third type of ignimbrite that commonly appears in volcanic sequences, particularly in areas where magmatism is relatively dry (hot spot or rift environments such as Yellow-stone—Snake River Plain or Taupo Volcanic Zone). In such cases, homogeneous ignimbrites with crystal contents on the order of 10-20% are commonplace (Dunbar et al. 1989; Nash et al. 2006; Ellis and Wolff 2012; Ellis et al. 2013). Ubiquitous in such units are millimeter- to centimeter-sized glomerocrysts of pyroxene- plagioclase-oxides (Fig. 11). These glomerocrysts are much more mafic (when analyzed in bulk) than the bulk rocks they are found in, and entirely lack sanidine and quartz, while they are found as phenocrysts in the host units. However, the chemical composi-tions of their pyroxenes and plagioclase crystals are identical to the individual phenocrysts around them. These glomerocrystic pyroxene and plagioclase also show very evolved REE pattern, suggesting growth from rhyolitic melts (Ellis et al. 2014).

Figure 11 Glomerocrysts with cumulate characteristics found in ignimbrites from the Snake River Plain (modified from Ellis et al. 2014).
Figure 11

Glomerocrysts with cumulate characteristics found in ignimbrites from the Snake River Plain (modified from Ellis et al. 2014).

Within the framework of the mush model, such deposits are best understood as pockets of melt-dominated regions in relatively hot/refractory mushes built in such dry magmatic environments. Upon recharge from below, these dry mushes are not likely to melt as easily as the ones that contain sanidine and quartz (as found in wetter, colder subduction zone environments). Hence, the zoning produced by partial melting of cumulates does not happen, and the deposits remain homogenous. As gas sparging can also play an important role in mush defrosting (Bachmann and Bergantz 2006; Huber et al. 2010b), mushes in magmatically dry environments are more difficult to remobilize because exsolved volatiles fail to reach the critical volume fraction required for efficient heat transfer by advective transport (Huber et al. 2010b), and produce mostly this remarkable type of unzoned, crystal-poor to intermediate deposit (Ellis et al. 2014; Wolff et al. 2015).

Chemical fractionation in mushy reservoirs: The development of compositional gaps in volcanic sequences

Compositional gaps, or the paucity of erupted products within a range of compositions, are commonplace in volcanic sequences and called the Bunsen-Daly gaps, in homage to early work from Bunsen (1851) and Daly (1925) on Iceland and Ascension Island. Such gaps can be seen in different elements (major and trace), in all series, including tholeiitic trends (e.g., Thompson 1972) and subduction zones (Brophy 1991; see Fig. 12). Possible hypotheses to explain these gaps dominantly revolve around three main lines of reasoning: (1) two main magma types (one mafic, one felsic/ silicic) are generated in the upper parts of our planet, mostly by melting mantle and crustal components (e.g., Bunsen 1851; Chayes 1963; Reubi and Blundy 2009); (2) two magma types are generated by liquid immiscibility (e.g., Charlier et al. 2011); and (3) magma evolution is dominated by crystal fractionation from mafic parents (generating a continuum of melt composition) but due to some mechanical trapping (potentially enhanced by non-linear temperature-crystallinity relationships), magma reaching the surface are dominantly clustered at some compositions (Marsh 1981; Grove and Donnelly-Nolan 1986; Brophy 1991; Bonnefoi for nearly two centuries [discussion started by James Hutton at the 18th century, and pushed further in classical books by Lyell (1838), Daly (1914), and Bowen (1928); and papers or memoirs such as Read (1948, 1957), Buddington (1959), Hamilton and Myers (1967), and Pitcher (1987)]. Several recent review papers summarize the main issues at play (Bachmann et al. 2007b; de Silva and Gosnold 2007a; Lipman 2007; Glazner et al. 2015; Keller et al. 2015; Lipman and Bachmann 2015). Focusing on felsic/ silicic magmas (mafic plutons are clearly mostly considered cumulates for decades; e.g., Wager et al. 1960), the main hypotheses revolve around whether intermediate to silicic plutons are mostly the expression of crystallized melts (“failed eruptions,” i.e., have not undergone significant crystal-liquid separation), or cumulate leftovers (“crystal graveyards”) from volcanic eruptions (those two end-member hypotheses not being mutually exclusive in plutonic complexes containing multiple facies). Textural analysis of plutonic lithologies and trace element geochemistry (in both bulk rock and minerals) should be the two main lines of arguments to differentiate between these competing hypotheses.

  1. If crystal accumulations occurs in areas of the crust: compatible trace elements should show significant increase in their concentrations, while incompatible elements should remain low (e.g., Mohamed 1998; Bachl et al. 2001; Wiebe et al. 2002; Kamiyama et al. 2007; Deering and Bachmann 2010; Turnbull et al. 2010).

  2. Plutonic rocks should show some mineral orientations or preferential alignment/foliation (e.g., Wager et al. 1960; Arculus and Wills 1980; Shirley 1986; Seaman 2000) related to magmatic processes (e.g., hindered settling and/or compaction), which will tend to orient anisotropically shaped crystals as they accumulate.

Figure 12 (a) Compositional groups observed in the Kos-Nisyros volcanic center (Eastern Aegean; see Pe-Piper and Moulton 2008 and Francalanci et al. 1995). (b) Tri-modality in compositions of volcanic rocks from Tenerife (modified from Sliwinski et al. 2015).
Figure 12

(a) Compositional groups observed in the Kos-Nisyros volcanic center (Eastern Aegean; see Pe-Piper and Moulton 2008 and Francalanci et al. 1995). (b) Tri-modality in compositions of volcanic rocks from Tenerife (modified from Sliwinski et al. 2015).

Despite several attempts to look in detail at this issue, this topic remains controversial (see de Silva and Gregg 2014; Glazner et al. 2015; Lipman and Bachmann 2015). The root of the controversy may largely stem from the fact that the amount of crystal/ melt segregation is less important in evolved (silicic) magmas compared to more mafic compositions [i.e., more trapped melt component in the intermediate to silicic crystal cumulates, due largely to viscosity difference (Bachmann et al. 2007b)]. Hence, silicic units have more subtle geochemical or textural signatures of crystal accumulation (or melt loss) than their mafic counterparts, and can easily be overlooked (see Gelman et al. 2014, and Fig. 13). However, although recent publications involving compilations of geochemical data from large databases (e.g., Glazner et al. 2015; Keller et al. 2015) cannot clearly differentiate volcanic from plutonic compositions at the felsic/silicic end of the spectrum, other compilations do see significant variabilities with volcanic units being on average richer in SiO2 (Lipman 1984, 2007; Halliday et al. 1991; Gelman et al. 2014; Deering et al. 2016). Recent studies focusing on specific field examples show that several intermediate to silicic plutonic lithologies have high-compatible/low-incompatible element concentrations in bulk rock (although such units clearly still have significant trapped melt components; e.g., Bachl et al. 2001; Barnes et al. 2001; Deering and Bachmann 2010; Lee and Morton 2015; Fig. 13). Additionally, detailed textural data, using electron backscattered electron diffraction (EBSD) techniques, on non-metamorphosed granitoids (i.e., observed mineral orientation must be magmatic) indicate crystal alignment compatible with compaction/hindered settling, even for very viscous magma compositions (Beane and Wiebe 2012; Graeter et al. 2015).

Figure 13 Example of a cross section through a wellexposed pluton (Searchlight pluton, Nevada), showing areas rich in crystallized melts and areas with cumulate characteristics, but still containing trapped melt (modified from Bachl et al. 2001; Gelman et al. 2014).
Figure 13

Example of a cross section through a wellexposed pluton (Searchlight pluton, Nevada), showing areas rich in crystallized melts and areas with cumulate characteristics, but still containing trapped melt (modified from Bachl et al. 2001; Gelman et al. 2014).

The volcanic-plutonic connection

How compositionally and temporally related (i.e., same age range, see section on timescales below) volcanic and plutonic units fit onto a common framework has been actively debated for nearly two centuries [discussion started by James Hutton at the 18th century, and pushed further in classical books by Lyell (1838), Daly (1914), and Bowen (1928); and papers or memoirs such as Read (1948, 1957), Buddington (1959), Hamilton and Myers (1967), and Pitcher (1987)]. Several recent review papers summarize the main issues at play (Bachmann et al. 2007b; de Silva and Gosnold 2007a; Lipman 2007; Glazner et al. 2015; Keller et al. 2015; Lipman and Bachmann 2015). Focusing on felsic/ silicic magmas (mafic plutons are clearly mostly considered cumulates for decades; e.g., Wager et al. 1960), the main hypotheses revolve around whether intermediate to silicic plutons are mostly the expression of crystallized melts (“failed eruptions,” i.e., have not undergone significant crystal-liquid separation), or cumulate leftovers (“crystal graveyards”) from volcanic eruptions (those two end-member hypotheses not being mutually exclusive in plutonic complexes containing multiple facies). Textural analysis of plutonic lithologies and trace element geochemistry (in both bulk rock and minerals) should be the two main lines of arguments to differentiate between these competing hypotheses.

  1. If crystal accumulations occurs in areas of the crust: compatible trace elements should show significant increase in their concentrations, while incompatible elements should remain low (e.g., Mohamed 1998; Bachl et al. 2001; Wiebe et al. 2002; Kamiyama et al. 2007; Deering and Bachmann 2010; Turnbull et al. 2010).

  2. Plutonic rocks should show some mineral orientations or preferential alignment/foliation (e.g., Wager et al. 1960; Arculus and Wills 1980; Shirley 1986; Seaman 2000) related to magmatic processes (e.g., hindered settling and/or compaction), which will tend to orient anisotropically shaped crystals as they accumulate.

Despite several attempts to look in detail at this issue, this topic remains controversial (see de Silva and Gregg 2014; Glazner et al. 2015; Lipman and Bachmann 2015). The root of the controversy may largely stem from the fact that the amount of crystal/ melt segregation is less important in evolved (silicic) magmas compared to more mafic compositions [i.e., more trapped melt component in the intermediate to silicic crystal cumulates, due largely to viscosity difference (Bachmann et al. 2007b)]. Hence, silicic units have more subtle geochemical or textural signatures of crystal accumulation (or melt loss) than their mafic counterparts, and can easily be overlooked (see Gelman et al. 2014, and Fig. 13). However, although recent publications involving compilations of geochemical data from large databases (e.g., Glazner et al. 2015; Keller et al. 2015) cannot clearly differentiate volcanic from plutonic compositions at the felsic/silicic end of the spectrum, other compilations do see significant variabilities with volcanic units being on average richer in SiO2 (Lipman 1984, 2007; Halliday et al. 1991; Gelman et al. 2014; Deering et al. 2016). Recent studies focusing on specific field examples show that several intermediate to silicic plutonic lithologies have high-compatible/low-incompatible element concentrations in bulk rock (although such units clearly still have significant trapped melt components; e.g., Bachl et al. 2001; Barnes et al. 2001; Deering and Bachmann 2010; Lee and Morton 2015; Fig. 13). Additionally, detailed textural data, using electron backscattered electron diffraction (EBSD) techniques, on non-metamorphosed granitoids (i.e., observed mineral orientation must be magmatic) indicate crystal alignment compatible with compaction/hindered settling, even for very viscous magma compositions (Beane and Wiebe 2012; Graeter et al. 2015).

The high eruptability of rhyolitic melt pockets

While high-SiO2 rhyolites are not rare in the volcanic record, including several very large units (>100 km3; Lipman 2000; Christiansen 2001; Mason et al. 2004), granites sensu stricto appear much less commonly. The upper crustal plutonic record is dominantly granodioritic/tonalitic in composition (Taylor and McClennan 1981; Rudnick 1995). When compiling data from large geochemical databases, such as the NAVDAT (http://www.navdat.org), one notices that low-Sr rocks are, relatively speaking, much more abundant in the volcanic realm than in the plutonic record (Parmigiani et al. 2016), an observation already discussed a quarter of a century ago on a much smaller database (Halliday et al. 1991; Cashman and Giordano 2014).

A possible explanation for the unusually high volcanic/plutonic ratio of evolved magmas is the fact that rhyolites are generated and stored in the upper crust (Tuttle and Bowen 1958; Gualda and Ghiorso 2013b; Ward et al. 2014; Lipman and Bachmann 2015), where volatiles can exsolve in quantity. If exsolved magmatic volatiles can accumulate in such melt pools, it will render them highly eruptible (increasing the gravitational potential energy of the magma) and will tend to promote more voluminous eruptions (Huppert and Woods 2002). As a consequence it will also lead to a deficit in the plutonic record. Using multiphase numerical simulations of exsolved volatile transport in magma reservoir, Parmigiani et al. (2016) have shown that bubbles tend to accumulate in crystal-poor environments. In crystal-rich environments, the volume available for the exsolved volatile phase is significantly reduced (porosity) by crystal confinement and promotes bubble coalescence and the formation of buoyant fingering pathways. Once the pathways are established, the very viscous melt does not limit the migration efficiency of the vapor phase and transport is efficient. In contrast, in crystal-poor environment, fingering channels are not stable and break into bubbles under the action of capillary forces. Discrete bubbles can only rise as fast as the viscous melt can be moved out of their way, which significantly reduces the migration efficiency of exsolved volatiles in crystal- poor magmas. As a result, bubbles tend to accumulate in the convecting rhyolitic cap, providing additional potential energy to drive or sustain future eruptions.

The accumulation of a buoyant volatile phase in these high- SiO2 melt pools favors large eruption, but it is unlikely to prevent entirely the formation plutonic counterparts as it only brings these magmas closer to a critical state. In fact, zones of evolved high- SiO2 granites do seldom occur in plutonic series. They appear to be mostly present in the upper parts of plutons, near the roof. Typical examples of such evolved pockets have been described in plutons from Nevada (Bachl et al. 2001; Barnes et al. 2001; Miller and Miller 2002), Klamath Mountain, California (Coint et al. 2013), Sierra Nevada Batholith, California (Putirka et al. 2014), and Peninsular Range Batholith, Mexico (Lee and Morton 2015).

Pluton degassing and volatile cycling

The chemical composition of the atmosphere is tied to some extent to magmatic degassing (e.g., CO2 and H2O cycles). As magmas are dominantly forming plutons [(plutonic/volcanic ratios are typically at least 5:1 to 10:1, and possibly even greater in some areas (Crisp 1984; White et al. 2006; Ward et al. 2014), the cycle of degassing and outgassing in plutons exerts an important control on the chemistry of the surface reservoirs on Earth. The protracted periods of storage and slow crystallization at high crystallinity predicted by the mush model are prone to lead to particularly ef-ficient magma degassing at shallow depths. In such slowly cooled crystal-rich environments, exsolved magmatic volatile build up in the mush by second boiling (exsolution produced by the crystallization of dominantly anhydrous minerals), and progressively fill more of the constricted pore space leading to a more efficient outgassing (Parmigiani et al. 2011; Huber et al. 2012b, 2013). Once the pore space, and more specifically the pore throats, are significantly reduced in size (down to the scale of a few microm-eters), capillary stresses can even become large enough for gas pathways to deform locally the granular structure in the mush and volatile-filled “dikes” may finalize the outgassing at large crystal fractions (>80% vol; Holtzman et al. 2012; Oppenheimer et al. 2015), possibly leading to the formation of the ubiquitous aplitic dykes seen in plutons (e.g., Jahns and Tuttle 1963; Candela 1997). In either case, the combination ofhigh-crystal content and increasing exsolved volatiles volume fraction can lead to the formation of efficient degassing pathways even during quiescent periods (sometimes potentially preserved as residual porosity in granites, e.g., Dunbar et al. 1996; Edmonds et al. 2003; Lowenstern and Hurvitz 2008).

The Excess S paradox associated with the eruption of silicic arc magmas

The mass balance of volatiles released during volcanic erup-tions can provide valuable insights into the state of shallow magma reservoirs prior to the eruption. As such, the mismatch in S mass balance between petrological inferences (i.e., comparing the S content of pre-degassed melt inclusions and fully degassed matrix glass) and remote sensing spectroscopy or sulfur output estimated from ice core data can provide additional clues about the dynamics that prevail in magma chambers and more specifically about the factors that control volatile exsolution. The notion of “Excess S” refers to the underestimation of the sulfur output from petrological constraints (melt inclusion record) during an eruption. This Excess S paradox has been evident since the eruption of El Chichon in Mexico (Luhr et al. 1984) and described in detail for the eruption of Mt. Pinatubo in 1991. At Pinatubo, remote sensing methods estimated the S output to nearly 20 Mt (e.g., Soden et al. 2002), while the petrological method provided an estimate smaller by a factor of 10 (Gerlach et al. 1996). Since these two eruptions, the release of S in excess to petrologic estimates seems to be more the rule than the exception, especially for silicic magmas in arcs (see Fig. 14 and Wallace 2001; Shinohara 2008).

Figure 14 Examples of volcanic eruptions (both explosive and effusive, from arc and non-arc systems) with and without excess sulfur (modified from Shinohara 2008). Excess S is based on the mismatch between the SO2 flux measured during volcanic eruptions (y-axis; typically done by spectroscopic methods or ice core data) and petrological estimates based on the difference in S content between melt inclusions and interstitial glass, estimating the amount of S released by the eruptive decompression process (x-axis). Calculated second boiling effect (gray inclined bar) is from a bubble growth model tracking the evolution of S partitioning between melt, crystals, and an exsolved gas phase in a cooling and crystallizing magma reservoir (from 5 to 50 vol% crystals; from Su et al. 2016).
Figure 14

Examples of volcanic eruptions (both explosive and effusive, from arc and non-arc systems) with and without excess sulfur (modified from Shinohara 2008). Excess S is based on the mismatch between the SO2 flux measured during volcanic eruptions (y-axis; typically done by spectroscopic methods or ice core data) and petrological estimates based on the difference in S content between melt inclusions and interstitial glass, estimating the amount of S released by the eruptive decompression process (x-axis). Calculated second boiling effect (gray inclined bar) is from a bubble growth model tracking the evolution of S partitioning between melt, crystals, and an exsolved gas phase in a cooling and crystallizing magma reservoir (from 5 to 50 vol% crystals; from Su et al. 2016).

Although multiple hypotheses have been suggested to account for this Excess S (see Gerlach et al. 1996 for a review), the presence of a S-rich exsolved volatile phase in the magma reservoir seems to be the most commonly accepted postulate (Gerlach et al. 1996; Scaillet et al. 1998a; Wallace 2001; Shinohara 2008). Crystallization-driven (second boiling) exsolution in a shallow magma reservoir can generate significant excess S where the S trapped by vapor bubbles during crystallization is not accessible to melt trapped subsequently as inclusion in crystal hosts. Although second boiling is bound to play an important role on generating Excess S in crystal-rich magmas, in crystal-poor silicic eruptions the deficit in S in melt inclusions requires an alternative process. In light of the previous section, accumulation of exsolved (S-rich) volatile bubbles in the erupted, low-crystallinity portions of magma reservoirs would provide a closure to the problematic missing S in crystal-poor silicic magmas, These bubbles (poorly constrained, but likely up to 20-30 of percent by volume) can be extremely rich in S, due to the very high affinity of S for the exsolved volatile phase (Zajacz et al. 2008). Hence, the mush model, predicting efficiently degassed crystal-rich zones and gas-charged, highly eruptible pockets of magma, provides a consistent framework to explain the missing source of S necessary to close the volatile budget for crystal-poor units.

Caldera cycles

Caldera-forming events typically happen after long periods of maturation in long-lived magmatic provinces during which andesitic volcanism dominates for up to several millions of years (Lipman 2007; de Silva and Gosnold 2007b; Grunder et al. 2008; Deering et al. 2011a) (Fig. 15). During this andesite-dominated period, the crust initially warms up to become more amenable to host silicic magma reservoirs in the mid to upper crust (Jellinek and DePaolo 2003; de Silva and Gregg 2014). Once this mature stage is reached, it is not rare to see several caldera-forming events occurring in relatively short time spans [~2-3 Myr; multi- cyclic caldera systems (Hildreth et al. 1991; Graham et al. 1995; Christiansen 2001; Lipman and McIntosh 2008; Lipman and Bachmann 2015)]. During such caldera cycles, magmas erupting shortly (~5-30 kyr) prior to the caldera-forming event typically appear to be compositionally similar to the climactic event (e.g., Bacon and Druitt 1988b; Bacon and Lanphere 2006; Bachmann et al. 2012), whereas the post-caldera magmas appear more mafic and drier (Shane et al. 2005; Bachmann et al. 2012; Gelman et al. 2013a; Barker et al. 2014; Fig. 15). These petrological shifts can be explained by the fact that the leftover, crystal-rich (mush) portion of the magma system following the eruption will be strongly affected by the caldera-forming event (Hildreth 2004). As those eruptions typically lead to significant decompression of the underlying crustal column (potentially up to 50-100 MPa, accounting for the overpressure necessary for the eruption to start and the redistribution of mass following the removal of much of the eruptible parts of the reservoir, see Bachmann et al. 2012), some rapid and significant degassing and crystallization is expected in the mush, particularly if it is near-eutectic and volatile-saturated (Bachmann et al. 2012). For mushes that were volatile-saturated prior to the decompression, the caldera eruption will likely promote the rapid formation of gas channels resulting in deep-gas release. Hence, the leftover mush is expected to become more crystalline (decompression-driven crystallization) and more degassed than it was before the caldera eruption (Degruyter et al. 2015). It will take another long maturation stage for the system to be primed for a new caldera event with a shallow, gas-charged, evolved magma body.

Figure 15 A caldera cycle recorded by changes in temperature, oxygen fugacity, bulk-rock composition, and mineralogy in the Kos-Nisyros volcanic system, eastern Aegean. Pre-caldera units (Kefalos domes and pyroclastic units) show highly evolved magma compositions (high-SiO2 rhyolites), low temperature, oxidized, and water-rich conditions, similar to the caldera-forming event (Kos Plateau Tuff, KPT). Following the KPT, Nisyros volcano built up, generating more typically less evolved magmas, including two large rhyodacitic units (Lower Pumice and Upper Pumice),
with drier, more reduced compositions, and hotter magma temperatures. Similar cycles have been suggested for the Taupo Volcanic Zone, in New Zealand (modified from Bachmann et al. 2012).
Figure 15

A caldera cycle recorded by changes in temperature, oxygen fugacity, bulk-rock composition, and mineralogy in the Kos-Nisyros volcanic system, eastern Aegean. Pre-caldera units (Kefalos domes and pyroclastic units) show highly evolved magma compositions (high-SiO2 rhyolites), low temperature, oxidized, and water-rich conditions, similar to the caldera-forming event (Kos Plateau Tuff, KPT). Following the KPT, Nisyros volcano built up, generating more typically less evolved magmas, including two large rhyodacitic units (Lower Pumice and Upper Pumice), with drier, more reduced compositions, and hotter magma temperatures. Similar cycles have been suggested for the Taupo Volcanic Zone, in New Zealand (modified from Bachmann et al. 2012).

Timescales associated with magma reservoirs

How long does a magma body stay above the solidus in the Earth’s crust (i.e., potentially active) remains one of the key questions in magmatic petrology and volcanology. The longevity, of course, impacts the extent of magmatic differentiation (including ore formation) and the eruptive volume that is available at any given time. As discussed above, magmatic reservoirs are mostly kept as high-crystallinity regions above the solidus when active, but some controversies remain as to how long the system remains in such a mush state (with estimates from a few thousand to more than 1000000 years). The crux of the debate lies in the fact that different analytical techniques provide different answers, and likely date different processes. Zircon dating of silicic plutons and volcanic rocks record extended crystallization histories, ranging from tens of thousands to several millions of years (Reid et al. 1997; Brown and Fletcher 1999; Lowenstern et al. 2000; Reid and Coath 2000; Charlier et al. 2003, 2005, 2007; Schmitt et al. 2003, 2010b; Coleman et al. 2004; Vazquez and Reid 2004; Bacon and Lowenstern 2005; Simon and Reid 2005; Bindeman et al. 2006; Matzel et al. 2006; Bachmann et al. 2007a; Miller et al. 2007; Walker et al. 2007; Reid 2008; Claiborne et al. 2010; Klemetti et al. 2011; Storm et al. 2012, 2014; Walker et al. 2013; Wotzlaw et al. 2013, 2014; Chamberlain et al. 2014b). In contrast, crystal size distributions (CSD; Pappalardo and Mastrolorenzo 2012), and diffusion timescales (see Druitt et al. 2012 and the compilation of data in Cooper and Kent 2014) are typically much shorter (typically months to a few hundreds of years).

As most elements used in diffusion timescale modeling move relatively fast (e.g., Fe-Mg in mafic minerals, Ti in quartz, Sr and Mg in plagioclase, Li in zircon ...), it is not surprising that preserved zoning patterns in such elements indicate relatively short timescales. Using simple scaling, one can provide a rough estimate of maximum timescale that diffusion re-equilibration would provide if one knows the size of the crystals and the dif-fusion coefficients (time ~R2/D, where R is the radius and D the diffusivity). Elements that are diffusing slower (REE, Ba) can be used to obtain longer timescales (e.g., Morgan and Blake 2006; Turner and Costa 2007), but, in such cases, periodic resorption, concurrent crystal growth, and 3D effects associated with the finite size of the host lead to significant complications in estimating time. For example, growth of feldspars in the Yellowstone magma reservoirs (Till et al. 2015) demonstrably led to “diffusion-like” profiles that look similar for elements that are diffusing at very different rates. Similarly, CSD studies typically assume linear growth rates (no pauses in crystallization, no resorption event), with values obtained from experiments (hence typically faster than what actually happens in silicic magma reservoirs). Hence, crystal size distributions provide a lower bound, and sometimes strongly underestimated, timescale.

The short timescales obtained by diffusion modeling are best interpreted as reactivation/remobilization/mixing timescales (e.g., Martin et al. 2008; Matthews et al. 2012; Till et al. 2015). They, however, do not provide a timescale of magma reservoir growth and storage in the crust. For such information, the zircon age distributions or bulk U/Th/Ra/Pb ages likely provide more reliable estimates (Reid et al. 1997; Cooper and Reid 2003; Bachmann et al. 2007a; Turner et al. 2010; Cooper and Kent 2014; Guillong et al. 2014). The difference between eruption age (estimated using the young zircon population or the Ar/Ar age when available) and the oldest co-genetic zircons (i.e., antecrysts, see Bacon and Lowenstern 2005; Miller et al. 2007 for a definition) is typically at least 104-105 years (see Costa 2008; Reid 2008; Simon et al. 2008; Bachmann 2010b for reviews on the topic). Xenocrysts (foreign crystals coming from unrelated wall rock lithologies) can also occur, but, in some cases, they can be isolated and removed from the data set (as they can be significantly older, plot off the Concordia and/or have different trace element chemistries; Miller et al. 2007; Lukács et al. 2015).Of course, the presence of older but co-genetic zircons (antecrysts) does not mean that the system remained above the solidus for the whole time; some parts of the system could have reached the solidus, and later recycled by reactivation events (Schmitt et al. 2010a; Folkes et al. 2011a). Hence, physical models are needed to address this issue.

Numerical models of the thermal state of such systems usu-ally agree with relatively long timescales of tens to hundreds of thousands of years, particularly when they include incremental recharge at relatively high fluxes (e.g., Annen et al. 2008; Gelman et al. 2013b). Moreover, it is very costly, both in terms of volatiles and energy, to reactivate sub-solidus plutons; nearly all volatile elements (including CO2 and H2O; Caricchi and Blundy 2015) and all latent heat has been lost to the surrounding. If the subsolidus material has to be melted again, the thermal budget required involves not only the sensible and latent heat loss, but also a large fraction of energy wasted on reheating solid material that will not undergo melting (Dufek and Bergantz 2005). The assimilation of these devolatilized sub-solidus portions of the body will also induce a depletion of volatile content in the reservoir. From a mechanical standpoint, entrainment/recycling of material that is not completely solid is much more likely, as it will disaggregate more easily, particularly if some local melting at grain boundaries is involved (Beard et al. 2005; Huber et al. 2011).

In summary, the periodic recharge of magma at reasonably high fluxes (>10-4—10-3 km3/yr; Annen et al. 2008; Gelman et al. 2013a, 2013b; Karakas and Dufek 2015) is likely to maintain at least parts of silicic magma reservoirs in a mush state for periods extending to at least several hundreds of thousands of years, even in the upper part of the crust (8-10 km depth). Shorter timescales, obtained with other techniques (particularly diffusion modeling), mainly refer to periods of reactivation following significant heat and mass addition into the reservoirs (Cooper and Kent 2014; Till et al. 2015). Hence, crystal/liquid separation is likely to have enough time to occur, despite being sluggish, to some extent in those reservoirs, driving differentiation toward more evolved and more eruptible crystal-poor magma pods (e.g., Bachmann and Bergantz 2004, 2008c; Hildreth 2004; Lee et al. 2015). We note that those thermal models are lacking the mechanical part of the energy budget (overpressurization during recharge, depressuriza- tion during eruption), which can have a significant impact on the stability and duration of reservoirs (Degruyter and Huber 2014; Degruyter et al. 2015) and the feedbacks with crustal rheology (Gregg et al. 2013; de Silva and Gregg 2014). For example the exsolution of volatiles affect the effective thermal properties of the magma by reducing its thermal capacity and conductivity (Huber et al. 2010b) and consuming latent heat. We note, however, that adding this mechanical part to the models is unlikely to significantly shorten the lifetimes of the reservoirs.

Frontiers in our understanding of magma chamber evolution

Over the last paragraphs, we have tried to assemble many important questions in volcanology/petrology into a coherent framework. The Polybaric Mush Model (Fig. 9) provides a context to explain several seemingly unconnected observations, such as the geophysical imaging of crystal-rich zones in magmatically active zones at different levels in the crust, the chemical differentiation of magmas dominated by fractional crystallization (with some assimilation, explaining some of the variability in isotopic data) ofmafic parents, the production of compositional gaps in volcanic series, the presence of cumulates in the plutons records, and the volatile mass balance in shallow reservoirs. However, there are many questions that remain to be answered, and the next section tries to outline some of the challenges that our community faces in the years to come.

At the most fundamental level, the goal is to continue de-veloping a more integrated picture between physical models, geophysical/geodetical inversions and petrology/geochemistry. In particular, we should tend to obtain a multiphase and multiscale (time and space) picture of magmatic systems, from the formation of the primary melts to their final resting place in the crust or at the surface (including interactions with the atmosphere). To reach such an ambitious goal, we would need, among many things:

  1. To revisit the information that can be extracted from the inversion of geophysical data sets. The best resolution that one can obtain from a magma body at depth is much lower than most structures of interest in these dynamical environments. As such, it is important to understand how the implicit filtering operates and how it impacts the set of questions that can be logically addressed with these studies. For example, if the target of the analysis is to estimate melt fraction from vp/vs tomography, even with good experimental calibration, one is left to wonder what the estimates of a melt fraction at the scale of hundreds of meters to even kilometers really mean. Is the outcome of the inversion to be understood as a volumetric average (as is commonly assumed)? The effective mechanical properties of heterogeneous media, which is what one measures through inversions, are generally quite different from a volumetric average. For complex, but structured, heterogeneous media, the effective properties inferred from elastic properties for example, result from a complex non-linear optimization process. For instance, it is quite possible that the effective medium elastic properties that one retrieves from tomography are dominantly controlled by heterogeneities that are volumetrically secondary. We certainly need more effort to relate the best-fit results of these inversions to the physical reality of complex multiphase magma bodies. This starts with a better understanding of the role of subgrid scale structures on the effective properties at the resolution of the inversion.

  2. To continue developing and improving our interpretation of tracers for rates of processes in magmatic systems. For example, it is necessary to improve precision and spatial resolution of mineral geochronology; and do more diffusion modeling of measured gradients to determine timescales, and focus on what they mean. It also becomes possible to use stable and radiogenic isotopes to constrain thermodynamics and kinetics, although the most substantial effort in this direction so far relates to mafic magmas (Richter et al. 2003; Watkins et al. 2009b; Huang et al. 2010; Savage et al. 2011). We also need better constraints on possible growth rates of minerals and bubbles. Dissolution events are transparent to the present chronometers and leaves hiatus in the temporal evolution of magma bodies that are impossible to quantify at the present time. Can we find kinetic tracers that would provide information about the duration and frequency ofresorption/dissolution events?

  3. To better constrain disequilibrium/kinetic effects in magma reservoirs. Most of the petrology and trace element partitioning among phases in magma reservoirs assumes at least a local equilibrium. The mere fact that crystallization (and sometimes dissolution) takes place requires a finite degree of disequilibrium, even over small (crystal sizes) scales. To relate dynamical processes to the rock record, a kinetic framework is necessary. There is ample information to mine from heterogeneities and disequilibrium; these clues will become crucial in testing dynamical models and establishing what set of processes controls the temporal evolution of these reservoirs over different spatial and temporal scales. Shall the high-T geochemistry and petrology community follow the low- T geochemistry community and invest time in developing reactive transport modeling for trace elements and isotopes in magma bodies? The chemistry can become very complicated, but the idea even in the context of oversimplified models is worth pursuing.

  4. To focus on eruption triggering mechanisms. One of the fundamental societal goals in volcanology is to understand the causes (triggers) that drive volcanic eruptions. There are several possible factors that can influence the state of a shallow reservoir and influence the timing of an eruption. Triggers can be external [e.g., caused by a regional earthquakes, roof collapse, or magma recharges from deeper (Sparks et al. 1977; Pallister et al. 1992; Watts et al. 1999; Eichelberger and Izbekov 2000; Gottsmann et al. 2009; Gregg et al. 2012)] or internal [e.g., buoyancy, pressure buildup during recharge or gas exsolution (e.g., Caricchi et al. 2014; Degruyter and Huber 2014; Malfait et al. 2014]. Fundamentally, the notion of trigger, which is related to an event or a process that has a causal link to a subsequent eruption, remains quite loose. Establishing a causality between a trigger and an eruption is much more challenging than establishing a correlation between the two.

  5. To continue providing better models of long-standing is-sues such as the “room problem.” Although the “room problem” is alleviated by having long-term incremental addition in a mainly ductile crust, pushing material (a) to the side, (b) down, as dense cumulates from the lower crust founder back into the mantle (Kay and Mahlburg Kay 1993; Jull and Kelemen 2001; Dufek and Bergantz 2005; Jagoutz and Schmidt 2013) and (c) upward, as upper crustal reservoir dome up during resurgence (e.g., Smith and Bailey 1968; Lipman et al. 1978; Marsh 1984; Hildreth 2004), the question continues to drive controversies (e.g., Glazner and Bartley 2006, 2008; Clarke and Erdmann 2008; Paterson et al. 2008; Yoshinobu and Barnes 2008).

Outlook

At the present time, many questions remain to be answered about the migration, emplacement, evolution, and ultimately eruption of silicic magmas. The main challenges stem from the fact that: (1) we have access to only a limited set of indirect observations, in most cases difficult to relate to the actual processes that control the chemical and dynamical evolution of these magmas, and (2) most magmas never reach the surface (trapped as plutons within the crust).

The development of a self-consistent model for silicic magmas residing in the upper crust is complex and requires a joint effort across multiple disciplines. The mush model presented in this review provides, we believe, a testable hypothesis from which better models can be developed. In spite of what remains to be done to understand the fate of magmas in the crust, the mush model offers a self-consistent paradigm that is supported by several independent observations from geophysics, geochemistry, and petrology. For instance the development of mush zones is consistent with:

  1. Thermal models of incrementally growing magma bodies and geophysical data on active systems require that such reservoirs are dominantly crystal-rich on long timescales (>100 kyr; Marsh 1981; Koyaguchi and Kaneko 1999; Bachmann and Bergantz 2004; Cooper and Kent 2014).

  2. Chemical differentiation can occur internally by melt- crystal separation and produce shorter-lived eruptible pockets (with melt content higher than 50 vol%) from time to time. Melt extraction can largely occur when convection currents wane down, i.e., when mixing by stirring becomes less efficient than gravitational settling. Phase separation should be most efficient at intermediate (40-70%) crystallinities.

  3. Melt extraction in slowly cooled mush zones (receiving periodic enthalpy additions by recharges and thermally buffered by latent heat at high crystallinity) helps explaining the longevity of magma centers/ lifetime of volcanic fields (as inferred from zircon age populations) and the comparatively short duration over which magmas are mobile and prone to erupt (Cooper and Kent 2014).

  4. The presence of cumulate plutonic lithologies in well-exposed upper crustal batholiths (e.g., Bachl et al. 2001; Barnes et al. 2001; Putirka et al. 2014) and as cognate clasts in silicic volcanic units (Gelman et al. 2014; Lee and Morton 2015; Wolff et al. 2015) confirms the close spatio-temporal link between fossil mush zones (“crystal graveyards”) and their extracted melt-rich complements.

Acknowledgments

Despite our attempt to cite the largest amount of previous work (we believe the reference list is extensive with nearly 500 citations), we are certainly not giving enough due credit to the amazing work that people have done over the years in our topic of interest. We apologize for this in advance. Most of these ideas put forward here have been nucleating in the womb of our own biased view. We have them here for discussion, and trust our community that some or most of them will be challenged. We look forward to it. We thank Mikito Furuichi, Naomi Matthews, Marc-Antoine Longpre, and Katy Chamberlain for providing figures that we use to illustrate some recent advances in our field, and Keith Putirka for motivating us to write this paper. Comments from Ben Ellis, Wim Degruyter, and Peter Lipman on a previous version of the manuscript helped better crystallize some ideas that were too tangled up in a mushy discussion. We are indebted to Charlie Bacon and Shan de Silva as formal reviewers, for the insightful and constructive comments, as well as to the very efficient editorial handling by Fidel Costa.

References cited

Acocella, V., Di Lorenzo, R., Newhall, C., and Scandone, R. (2015) An overview of recent (1988 to 2014) caldera unrest: Knowledge and perspectives. Reviews of Geophysics, 53, 896–955.10.1002/2015RG000492Search in Google Scholar

Almeev, R.R., Bolte, T., Nash, B.P., Holtz, F.O., Erdmann, M., and Cathey, H.E. (2012) High-temperature, low-H2O silicic magmas of the Yellowstone Hotspot: an experimental study of rhyolite from the Bruneau-Jarbidge Eruptive Center, Central Snake River Plain, U.S.A. Journal of Petrology, 53, 1837–1866.10.1093/petrology/egs035Search in Google Scholar

Alonso-Perez, R., Müntener, O., and Ulmer, P. (2009) Igneous garnet and amphibole fractionation in the roots of island arcs: experimental constraints on andesitic liquids. Contributions to Mineralogy and Petrology, 157, 541–558.10.1007/s00410-008-0351-8Search in Google Scholar

Anderson, A.T. (1974) Chlorine, sulfur and water in magmas and oceans. Geological Society of America Bulletin, 85, 1485–1492.10.1130/0016-7606(1974)85<1485:CSAWIM>2.0.CO;2Search in Google Scholar

Anderson, A.T., Davis, A.M., and Lu, F. (2000) Evolution of Bishop Tuff rhyolitic magma based on melt and magnetite inclusions and zoned phenocrysts. Journal of Petrology, 41, 449–473.10.1093/petrology/41.3.449Search in Google Scholar

Annen, C. (2009) From plutons to magma chambers: Thermal constraints on the accumulation of eruptible silicic magma in the upper crust. Earth and Planetary Science Letters, 284, 409–416.10.1016/j.epsl.2009.05.006Search in Google Scholar

Annen, C., and Sparks, R.S.J. (2002) Effects of repetitive emplacement of basaltic intrusions on thermal evolution and melt generation in the crust. Earth and Planetary Science Letters, 203, 937–955.10.1016/S0012-821X(02)00929-9Search in Google Scholar

Annen, C., Blundy, J.D., and Sparks, R.S.J. (2006) The genesis of intermediate and silicic magmas in deep crustal hot zones. Journal of Petrology, 47, 505–539.10.1093/petrology/egi084Search in Google Scholar

Annen, C., Pichavant, M., Bachmann, O., and Burgisser, A. (2008) Conditions for the growth of a long-lived shallow crustal magma chamber below Mount Pelee volcano (Martinique, Lesser Antilles Arc). Journal of Geophysical Research-Solid Earth, 113(B7).10.1029/2007JB005049Search in Google Scholar

Arculus, R.J., and Wills, K.J.A. (1980) The petrology of plutonic blocks and inclusions from the Lesser Antilles Island Arc. Journal of Petrology, 21, 743–799.10.1093/petrology/21.4.743Search in Google Scholar

Bachl, C.A., Miller, C.F., Miller, J.S., and Faulds, J.E. (2001) Construction of a pluton: Evidence from an exposed cross-section of the Searchlight pluton, Eldorado Mountains, Nevada. Geological Society of America Bulletin, 113, 1213–1228.10.1130/0016-7606(2001)113<1213:COAPEF>2.0.CO;2Search in Google Scholar

Bachmann, O. (2010a) The petrologic evolution and pre-eruptive conditions of the rhyolitic Kos Plateau Tuff (Aegean arc). Central European Journal of Geosciences, 2, 270–305.10.2478/v10085-010-0009-4Search in Google Scholar

——— (2010b) Timescales associated with large silicic magma bodies. In A. Dosseto, S.P. Turner, and J.A. Van Orman, Eds., Timescales of Magmatic Processes: From core to atmosphere. Kluwer.Search in Google Scholar

Bachmann, O., and Bergantz, G.W. (2003) Rejuvenation of the Fish Canyon magma body: A window into the evolution of large-volume silicic magma systems. Geology, 31, 789–792.10.1130/G19764.1Search in Google Scholar

——— (2004) On the origin of crystal-poor rhyolites: Extracted from batholithic crystal mushes. Journal of Petrology, 45, 1565–1582.10.1093/petrology/egh019Search in Google Scholar

——— (2006) Gas percolation in upper-crustal silicic crystal mushes as a mechanism for upward heat advection and rejuvenation of near-solidus magma bodies. Journal of Volcanology and Geothermal Research, 149, 85–102.10.1016/j.jvolgeores.2005.06.002Search in Google Scholar

——— (2008a) Deciphering magma chamber dynamics from styles of compositional zoning in large silicic ash flow sheets. Minerals, Inclusions and Volcanic Processes, 69, 651–674.10.2138/rmg.2008.69.17Search in Google Scholar

——— (2008b) Deciphering magma chamber dynamics from styles of compositional zoning in large silicic ash flow sheets. Reviews in Mineralogy and Geochemistry. 69, 651–674.10.2138/rmg.2008.69.17Search in Google Scholar

——— (2008c) Rhyolites and their source mushes across tectonic settings. Journal of Petrology, 49, 2277–2285.10.1093/petrology/egn068Search in Google Scholar

Bachmann, O., and Dungan, M.A. (2002) Temperature-induced Al-zoning in hornblendes of the Fish Canyon magma, Colorado. American Mineralogist, 87, 1062–1076.10.2138/am-2002-8-903Search in Google Scholar

Bachmann, O., Dungan, M.A., and Lipman, P.W. (2002) The Fish Canyon magma body, San Juan volcanic field, Colorado: Rejuvenation and eruption of an upper-crustal batholith. Journal of Petrology, 43, 1469–1503.10.1093/petrology/43.8.1469Search in Google Scholar

Bachmann, O., Dungan, M.A., and Bussy, F. (2005) Insights into shallow magmatic processes in large silicic magma bodies: The trace element record in the Fish Canyon magma body, Colorado. Contributions to Mineralogy and Petrology, 149, 338–349.10.1007/s00410-005-0653-zSearch in Google Scholar

Bachmann, O., Charlier, B.L.A., and Lowenstern, J.B. (2007a) Zircon crystallization and recycling in the magma chamber of the rhyolitic Kos Plateau Tuff (Aegean Arc). Geology, 35, 73–76.10.1130/G23151A.1Search in Google Scholar

Bachmann, O., Miller, C.F., and de Silva, S.L. (2007b) The volcanic-plutonic connection as a stage for understanding crustal magmatism. Journal of Volcanology and Geothermal Research, 167, 1–23.10.1016/j.jvolgeores.2007.08.002Search in Google Scholar

Bachmann, O., Wallace, P.J., and Bourquin, J. (2010) The melt inclusion record from the rhyolitic Kos Plateau Tuff (Aegean Arc). Contributions to Mineralogy and Petrology, 159, 187–202.10.1007/s00410-009-0423-4Search in Google Scholar

Bachmann, O., Deering, C.D., Ruprecht, J.S., Huber, C., Skopelitis, A., and Schnyder, C. (2012) Evolution of silicic magmas in the Kos-Nisyros volcanic center, Greece: A petrological cycle associated with caldera collapse. Contributions to Mineralogy and Petrology, 163, 151–166.10.1007/s00410-011-0663-ySearch in Google Scholar

Bachmann, O., Deering, C., Lipman, P., and Plummer, C. (2014) Building zoned ignimbrites by recycling silicic cumulates: Insight from the 1,000 km3 Carpenter Ridge Tuff, Colorado. Contributions to Mineralogy and Petrology, 167, 1–13.10.1007/s00410-014-1025-3Search in Google Scholar

Bacon, C.R., and Druitt, T.H. (1988a) Compositional evolution of the zoned calcalkaline magma chamber of Mount Mazama, Crater Lake, Oregon. Contributions to Mineralogy and Petrology, 98, 224–256.10.1007/BF00402114Search in Google Scholar

——— (1988b) Compositional evolution of the zoned calcalkaline magma chamber of Mount-Mazama, Crater Lake, Oregon. Contributions to Mineralogy and Petrology, 98, 224–256.10.1007/BF00402114Search in Google Scholar

Bacon, C.R., and Lanphere, M.A. (2006) Eruptive history and geochronology of Mount Mazama and the Crater Lake region, Oregon. GSA Bulletin, 118, 1331–1359.10.1130/B25906.1Search in Google Scholar

Bacon, C.R., and Lowenstern, J.B. (2005) Late Pleistocene granodiorite source for recycled zircon and phenocrysts in rhyodacite lava at Crater Lake, Oregon. Earth and Planetary Science Letters, 233, 277–293.10.1016/j.epsl.2005.02.012Search in Google Scholar

Baker, D.R., and Alletti, M. (2012) Fluid saturation and volatile partitioning between melts and hydrous fluids in crustal magmatic systems: The contribution of experimental measurements and solubility models. Earth-Science Reviews, 114, 298–324.10.1016/j.earscirev.2012.06.005Search in Google Scholar

Barboni, M., Annen, C., and Schoene, B. (2015) Evaluating the construction and evolution of upper crustal magma reservoirs with coupled U/Pb zircon geochronology and thermal modeling: A case study from the Mt. Capanne pluton (Elba, Italy). Earth and Planetary Science Letters, 432, 436–448.10.1016/j.epsl.2015.09.043Search in Google Scholar

Barker, S.J., Wilson, C.J.N., Smith, E.G.C., Charlier, B.L.A., Wooden, J.L., Hiess, J., and Ireland, T.R. (2014) Post-supereruption magmatic reconstruction of Taupo Volcano (New Zealand), as reflected in zircon ages and trace elements. Journal of Petrology, 55, 1511–1533.10.1093/petrology/egu032Search in Google Scholar

Barnes, C.G., Burton, B.R., Burling, T.C., Wright, J.E., and Karlsson, H.R. (2001) Petrology and Geochemistry of the Late Eocene Harrison Pass Pluton, Ruby Mountains Core Complex, Northeastern Nevada. Journal of Petrology, 42, 901–929.10.1093/petrology/42.5.901Search in Google Scholar

Beane, R., and Wiebe, R. (2012) Origin of quartz clusters in Vinalhaven granite and porphyry, coastal Maine. Contributions to Mineralogy and Petrology, 1–14.10.1007/s00410-011-0717-1Search in Google Scholar

Beard, J.S., Ragland, P.C., and Crawford, M.L. (2005) Reactive bulk assimilation: A model for crust-mantle mixing in silicic magmas. Geology, 33, 681–684.10.1130/G21470AR.1Search in Google Scholar

Bergantz, G.W., and Ni, J. (1999) A numerical study of sedimentation by dripping instabilities in viscous fluids. International Journal of Multiphase Flow, 25, 307–320.10.1016/S0301-9322(98)00050-0Search in Google Scholar

Bergantz, G.W., Schleicher, J.M., and Burgisser, A. (2015) Open-system dynamics and mixing in magma mushes. Nature Geoscience, 8, 793–796.10.1038/ngeo2534Search in Google Scholar

Berman, R.G. (1991) Thermobarometry using multi-equilibrium calculations: A new technique, with petrological implications. Contributions to Mineralogy and Petrology, 29, 833–855.Search in Google Scholar

Bindeman, I.N., and Valley, J.W. (2001) Low-18O rhyolites from Yellowstone: Magmatic evolution based on zircons and individual phenocrysts. Journal of Petrology, 42, 1491–1517.10.1093/petrology/42.8.1491Search in Google Scholar

——— (2003) Rapid generation of both high- and low-δ18O, large volume silicic magmas at the Timber Mountain/Oasis Valley caldera complex, Nevada. Geological Society of America Bulletin, 115, 581–595.10.1130/0016-7606(2003)115<0581:RGOBHA>2.0.CO;2Search in Google Scholar

Bindeman, I., Schmitt, A., and Valley, J. (2006) U–Pb zircon geochronology of silicic tuffs from the Timber Mountain/Oasis Valley caldera complex, Nevada: Rapid generation of large volume magmas by shallow-level remelting. Contributions to Mineralogy and Petrology, 152, 649–665.10.1007/s00410-006-0124-1Search in Google Scholar

Blake, S., and Campbell, I.H. (1986) The dynamics of magma-mixing during flow in volcanic conduits. Contributions to Mineralogy and Petrology, 94, 72–81.10.1007/BF00371228Search in Google Scholar

Blake, S., and Ivey, G.N. (1986) Magma mixing and the dynamics of withdrawal from stratified reservoirs. Journal of Volcanology and Geothermal Research, 27, 153–178.10.1016/0377-0273(86)90084-3Search in Google Scholar

Blundy, J., and Cashman, K. (2008) Petrologic Reconstruction of Magmatic System Variables and Processes. Reviews in Mineralogy and Geochemistry, 69, 179–239.10.1515/9781501508486-007Search in Google Scholar

Blundy, J., Cashman, K.V., Rust, A., and Witham, F. (2010) A case for CO2-rich arc magmas. Earth and Planetary Science Letters, 290, 289–301.10.1016/j.epsl.2009.12.013Search in Google Scholar

Bohrson, W.A., and Spera, F.J. (2007) Energy-constrained recharge, assimilation, and fractional crystallization (EC-RAFC): A Visual Basic computer code for calculating trace element and isotope variations of open-system magmatic systems. Geochemistry, Geophysics, Geosystems, 8.10.1029/2007GC001781Search in Google Scholar

Bohrson, W.A., Spera, F.J., Ghiorso, M.S., Brown, G.A., Creamer, J.B., and Mayfield, A. (2014) Thermodynamic model for energy-constrained open-system evolution of crustal magma bodies undergoing simultaneous recharge, assimilation and crystallization: The magma chamber simulator. Journal of Petrology, 55, 1685–1717.10.1093/petrology/egu036Search in Google Scholar

Bonnefoi, C.C., Provost, A., and Albarede, F. (1995) The “Daly gap” as a magmatic catastrophe. Nature, 378, 270–272.10.1038/378270a0Search in Google Scholar

Boroughs, S., Wolff, J.A., Ellis, B.S., Bonnichsen, B., and Larson, P.B. (2012) Evaluation of models for the origin of Miocene low-d18O rhyolites of the Yellowstone/Columbia River Large Igneous Province. Earth and Planetary Science Letters, 313–314, 45–55.10.1016/j.epsl.2011.10.039Search in Google Scholar

Boudreau, A.E. (1999) PELE—a version of the MELTS software program for the PC platform. Computers and Geosciences, 25, 201–203.10.1016/S0098-3004(98)00117-4Search in Google Scholar

Bowen, N.L. (1928) The Evolution of Igneous Rocks, 332 p. Dover, New York.Search in Google Scholar

Brandeis, G., and Jaupart, C. (1986) On the interaction between convection and crystallization in cooling magma chambers. Earth and Planetary Science Letters, 77, 345–361.10.1016/0012-821X(86)90145-7Search in Google Scholar

Brandeis, G., and Marsh, B.D. (1990) Transient magmatic convection prolonged by solidification. Geophysical Research Letters, 17, 1125–1128.10.1029/GL017i008p01125Search in Google Scholar

Brenguier, F., Shapiro, N.M., Campillo, M., Ferrazzini, V., Duputel, Z., Coutant, O., and Nercessian, A. (2008) Towards forecasting volcanic eruptions using seismic noise. Nature Geoscience, 1, 126–130.10.1038/ngeo104Search in Google Scholar

Brophy, J.G. (1991) Composition gaps, critical crystallinity, and fractional crystallization in orogenic (calc-alkaline) magmatic systems. Contributions to Mineralogy and Petrology, 109, 173–182.10.1007/BF00306477Search in Google Scholar

Brown, S.J.A., and Fletcher, I.R. (1999) SHRIMP U-Pb dating of the preeruption growth history of zircons from the 340 ka Whakamaru Ignimbrite, New Zealand: Evidence for >250 k.y. magma residence times. Geology, 27, 1035–1038.10.1130/0091-7613(1999)027<1035:SUPDOT>2.3.CO;2Search in Google Scholar

Buddington, A.F. (1959) Granite emplacement with special reference to North America. Geological Society of America Bulletin, 70, 671–748.10.1130/0016-7606(1959)70[671:GEWSRT]2.0.CO;2Search in Google Scholar

Bunsen, R. (1851) Ueber die prozesse der vulkanishen Gesteinsbildungen Islands. Annalen der Physics (Leipzig), 83, 197–272.10.1002/andp.18511590602Search in Google Scholar

Burgisser, A., and Bergantz, G.W. (2011) A rapid mechanism to remobilize and homogenize highly crystalline magma bodies. Nature, 471, 212–215.10.1038/nature09799Search in Google Scholar

Burgisser, A., and Scaillet, B. (2007) Redox evolution of a degassing magma rising to the surface. Nature, 194–197.10.1038/nature05509Search in Google Scholar

Burton, M., Allard, P., Muré, F., and La Spina, A. (2007) Magmatic gas composition reveals the source depth of slug-driven strombolian explosive activity. Science, 317, 227–230.10.1126/science.1141900Search in Google Scholar

Candela, P.A. (1997) A review of shallow, ore-related granites: Textures, volatiles, and ore metals. Journal of Petrology, 38, 1619–1633.10.1093/petroj/38.12.1619Search in Google Scholar

Cardoso, S.S.S., and Woods, A.W. (1999) On convection in a volatile-saturated magma. Earth and Planetary Science Letters, 168, 301–310.10.1016/S0012-821X(99)00057-6Search in Google Scholar

Caricchi, L., and Blundy, J. (2015) Experimental petrology of monotonous intermediate magmas. Geological Society, London, Special Publications, 422.10.1144/SP422.9Search in Google Scholar

Caricchi, L., Burlini, L., Ulmer, P., Gerya, T., Vassali, M., and Papale, P. (2007) Non-Newtonian rheology of crystal-bearing magmas and implications for magma ascent dynamics. Earth and Planetary Science Letters, 264, 402–419.10.1016/j.epsl.2007.09.032Search in Google Scholar

Caricchi, L., Annen, C., Blundy, J., Simpson, G., and Pinel, V. (2014) Frequency and magnitude of volcanic eruptions controlled by magma injection and buoyancy. Nature Geoscience, 7, 126–130.10.1038/ngeo2041Search in Google Scholar

Cashman, K.V., and Giordano, D. (2014) Calderas and magma reservoirs. Journal of Volcanology and Geothermal Research, 288, 28–45.10.1016/j.jvolgeores.2014.09.007Search in Google Scholar

Cashman, K.V., and Sparks, R.S.J. (2013) How volcanoes work: A 25 year perspective. Geological Society of America Bulletin.10.1130/B30720.1Search in Google Scholar

Chamberlain, K.J., Morgan, D.J., and Wilson, C.J.N. (2014a) Timescales of mixing and mobilisation in the Bishop Tuff magma body: Perspectives from diffusion chronometry. Contributions to Mineralogy and Petrology, 168, 1–24.10.1007/s00410-014-1034-2Search in Google Scholar

Chamberlain, K.J., Wilson, C.J.N., Wooden, J.L., Charlier, B.L.A., and Ireland, T.R. (2014b) New perspectives on the Bishop Tuff from zircon textures, ages and trace elements. Journal of Petrology, 55, 395–426.10.1093/petrology/egt072Search in Google Scholar

Champallier, R., Bystricky, M., and Arbaret, L. (2008) Experimental investigation of magma rheology at 300 MPa: From pure hydrous melt to 76 vol% of crystals. Earth and Planetary Science Letters, 267, 571–583.10.1016/j.epsl.2007.11.065Search in Google Scholar

Charlier, B., Namur, O., Toplis, M.J., Schiano, P., Cluzel, N., Higgins, M.D., and Auwera, J.V. (2011) Large-scale silicate liquid immiscibility during differentiation of tholeiitic basalt to granite and the origin of the Daly gap. Geology, 39, 907–910.10.1130/G32091.1Search in Google Scholar

Charlier, B.L.A., Peate, D.W., Wilson, C.J.N., Lowenstern, J.B., Storey, M., and Brown, S.J.A. (2003) Crystallisation ages in coeval silicic magma bodies: 238U-230Th disequilibrium evidence from the Rotoiti and Earthquake Flat eruption deposits, Taupo Volcanic Zone, New Zealand. Earth and Planetary Science Letters, 206, 441–457.10.1016/S0012-821X(02)01109-3Search in Google Scholar

Charlier, B.L.A., Wilson, C.J.N., Lowenstern, J.B., Blake, S., Van Calsteren, P.W., and Davidson, J.P. (2005) Magma generation at a large, hyperactive silicic volcano (Taupo, New Zealand) revealed by U–Th and U–Pb systematics in zircons. Journal of Petrology, 46, 3–32.10.1093/petrology/egh060Search in Google Scholar

Charlier, B.L.A., Bachmann, O., Davidson, J.P., Dungan, M.A., and Morgan, D. (2007) The upper crustal evolution of a large silicic magma body: Evidence from crystal-scale Rb/Sr isotopic heterogeneities in the Fish Canyon magmatic system, Colorado. Journal of Petrology, 48, 1875–1894.10.1093/petrology/egm043Search in Google Scholar

Chayes, F. (1963) Relative abundance of intermediate members of the oceanic basalt-trachyte association. Journal of Geophysical Research, 68, 1519–1534.10.1029/JZ068i005p01519Search in Google Scholar

Chiaradia, M. (2014) Copper enrichment in arc magmas controlled by overriding plate thickness. Nature Geoscience, 7, 43–46.10.1038/ngeo2028Search in Google Scholar

Christiansen, R.L. (2001) The quaternary and Pliocene Yellowstone Plateau Volcanic Field of Wyoming, Idaho, and Montana. U.S. Geological Survey Professional Paper, 729-G, 145.10.3133/pp729GSearch in Google Scholar

Cimarelli, C., Costa, A., Mueller, S., and Mader, H.M. (2011) Rheology of magmas with bimodal crystal size and shape distributions: Insights from analog experiments. Geochemistry, Geophysics, Geosystems, 12.10.1029/2011GC003606Search in Google Scholar

Civetta, L., Orsi, G., Pappalardo, L., Fisher, R.V., Heiken, G., and Ort, M. (1997) Geochemical zoning, mingling, eruptive dynamics and depositional processes-the Campanian Ignimbrite, Campi Flegrei, Italy. Journal of Volcanology and Geothermal Research, 75, 183–219.10.1016/S0377-0273(96)00027-3Search in Google Scholar

Claiborne, L.L., Miller, C.F., Flanagan, D.M., Clynne, M.A., and Wooden, J.L. (2010) Zircon reveals protracted magma storage and recycling beneath Mount St. Helens. Geology, 38, 1011–1014.10.1130/G31285.1Search in Google Scholar

Clarke, D.B., and Erdmann, S. (2008) Is stoping a volumetrically significant pluton emplacement process?: Comment. Geological Society of America Bulletin, 120, 1072–1074.10.1130/B26147.1Search in Google Scholar

Coint, N., Barnes, C.G., Yoshinobu, A.S., Barnes, M.A., and Buck, S. (2013) Use of trace element abundances in augite and hornblende to determine the size, connectivity, timing, and evolution of magma batches in a tilted batholith. Geosphere, 9, 1747–1765.10.1130/GES00931.1Search in Google Scholar

Coleman, D.S., Gray, W., and Glazner, A.F. (2004) Rethinking the emplacement and evolution of zoned plutons; geochronologic evidence for incremental assembly of the Tuolumne Intrusive Suite, California. Geology, 32, 433–436.10.1130/G20220.1Search in Google Scholar

Connolly, J.A.D. (2009) The geodynamic equation of state: What and how. Geochemistry, Geophysics, Geosystems, 10.10.1029/2009GC002540Search in Google Scholar

Cooper, K.M., and Kent, A.J.R. (2014) Rapid remobilization of magmatic crystals kept in cold storage. Nature, 506, 480–483.10.1038/nature12991Search in Google Scholar

Cooper, K.M., and Reid, M.R. (2003) Re-examination of crystal ages in recent Mount St. Helens lavas: Implications for magma reservoir processes. Earth and Planetary Science Letters, 213, 149–167.10.1016/S0012-821X(03)00262-0Search in Google Scholar

Cordonnier, B., Caricchi, L., Pistone, M., Castro, J., Hess, K.-U., Gottschaller, S., Manga, M., Dingwell, D.B., and Burlini, L. (2012) The viscous-brittle transition of crystal-bearing silicic melt: Direct observation of magma rupture and healing. Geology, 40, 611–614.10.1130/G3914.1Search in Google Scholar

Costa, A., Caricchi, L., and Bagdassarov, N. (2009) A model for the rheology of particle-bearing suspensions and partially molten rocks. Geochemistry, Geophysics, Geosystems, 10.10.1029/2008GC002138Search in Google Scholar

Costa, F. (2008) Residence times of silicic magmas associated with calderas. In J. Gottsmann, and J. Martì, Eds., Developments in Volcanology, 10, 1–55. Elsevier, Amsterdam.10.1016/S1871-644X(07)00001-0Search in Google Scholar

Costa, F., Scaillet, B., and Gourgaud, A. (2003) Massive atmospheric sulfur loading of the AD 1600 Huaynaputina eruption and implications for petrologic sulfur estimates. Geophysical Research Letters, 30.10.1029/2002GL016402Search in Google Scholar

Costa, F., Scaillet, B., and Pichavant, M. (2004) Petrological and experimental constraints on the pre-eruption conditions of Holocene dacite from Volcan San Pedro (36° S, Chilean Andes) and the importance of sulphur in silicic subduction-related magmas. Journal of Petrology, 45, 855–881.10.1093/petrology/egg114Search in Google Scholar

Couch, S., Sparks, R.S.J., and Carroll, M.R. (2001) Mineral disequilibrium in lavas explained by convective self-mixing in open magma chambers. Nature, 411, 1037–1039.10.1038/35082540Search in Google Scholar

Crisp, J.A. (1984) Rates of magma emplacement and volcanic output. Journal of Volcanology and Geothermal Research, 20, 177–211.10.1016/0377-0273(84)90039-8Search in Google Scholar

Daly, R.A. (1914) Igneous Rocks and Their Origin. 563 p. McGraw-Hill, London.Search in Google Scholar

——— (1925) The geology of Ascension Island. Proceedings of the American Academy of Arts and Sciences, 60, 1–80.Search in Google Scholar

——— (1933) Igneous Rocks and the Depths of the Earth. McGraw Hill, New York.Search in Google Scholar

Davaille, A., and Jaupart, C. (1993) Transient high-Rayleigh number thermal convection with large viscosity variations. Journal of Fluid Mechanics, 253, 141–166.10.1017/S0022112093001740Search in Google Scholar

Dawson, P.B., Evans, J.R., and Iyer, H.M. (1990) Teleseismic tomography of the compressional wave velocity structure beneath the Long Valley region. Journal of Geophysical Research, 95(B7), 11021–11050.10.1029/JB095iB07p11021Search in Google Scholar

De Natale, G., Troise, C., Trigila, R., Dolfi, D., and Chiarabba, C. (2004) Seismicity and 3-D substructure at Somma-Vesuvius volcano: evidence for magma quenching. Earth and Planetary Science Letters, 221, 181–196.10.1016/S0012-821X(04)00093-7Search in Google Scholar

De Silva, S.L. (1991) Styles of zoning in the central Andean ignimbrites: Insights into magma chamber processes. In R.S. Harmon and C.W. Rapela, Eds., Andean Magmatism and its Tectonic Setting. Geological Society of America Special Paper, 265, 233–243.Search in Google Scholar

De Silva, S.L., and Gosnold, W. (2007a) Cordilleran Batholith development: Insights from an ignimbrite flare-up. Journal of Volcanology and Geothermal Research, 167, 320–335.10.1016/j.jvolgeores.2007.07.015Search in Google Scholar

——— (2007b) Episodic construction of batholiths: Insights from the spatiotemporal development of an ignimbrite flare-up. Journal of Volcanology and Geothermal Research, 167, 320–335.10.1016/j.jvolgeores.2007.07.015Search in Google Scholar

De Silva, S.L., and Gregg, P.M. (2014) Thermomechanical feedbacks in magmatic systems: Implications for growth, longevity, and evolution of large caldera-forming magma reservoirs and their supereruptions. Journal of Volcanology and Geothermal Research, 282, 77–91.10.1016/j.jvolgeores.2014.06.001Search in Google Scholar

De Silva, S.L., and Wolff, J.A. (1995) Zoned magma chambers; the influence of magma chamber geometry on sidewall convective fractionation. Journal of Volcanology and Geothermal Research, 65, 111–118.10.1016/0377-0273(94)00105-PSearch in Google Scholar

De Silva, S.L., Self, S., Francis, P.W., Drake, R.E., and Carlos Ramirez, R. (1994) Effusive silicic volcanism in the Central Andes: The Chao dacite and other young lavas of the Altiplano-Puna Volcanic Complex. Journal of Geophysical Research, 99(B9), 17805–17825.10.1029/94JB00652Search in Google Scholar

De Silva, S.L., Zandt, G., Trumbull, R., and Viramonte, J. (2006) Large scale silicic volcanism—the result of thermal maturation of the crust. In Y.-t. Chen, Ed., Advances in Geosciences, p. 215–230. World Scientific Press.10.1142/9789812707178_0021Search in Google Scholar

Deering, C.D., and Bachmann, O. (2010) Trace element indicators of crystal accumulation in silicic igneous rocks. Earth and Planetary Science Letters, 297, 324–331.10.1016/j.epsl.2010.06.034Search in Google Scholar

Deering, C.D., Bachmann, O., Dufek, J., and Gravley, D.M. (2011a) Rift-related transition from andesite to rhyolite volcanism in the Taupo Volcanic Zone (New Zealand) controlled by crystal-melt dynamics in mush zones with variable mineral assemblages. Journal of Petrology, 52, 2243–2263.10.1093/petrology/egr046Search in Google Scholar

Deering, C.D., Bachmann, O., and Vogel, T.A. (2011b) The Ammonia Tanks Tuff: Erupting a melt-rich rhyolite cap and its remobilized crystal cumulate. Earth and Planetary Science Letters, 310, 518–525.10.1016/j.epsl.2011.08.032Search in Google Scholar

Deering, C.D., Keller, B., Schoene, B., Bachmann, O., Beane, R., and Ovtcharova, M. (2016) Zircon record of the plutonic-volcanic connection and protracted rhyolite melt evolution. Geology, 10.1130/G37539.1.Search in Google Scholar

Degruyter, W., and Huber, C. (2014) A model for eruption frequency of upper crustal silicic magma chambers. Earth and Planetary Science Letters, 403, 117–130.10.1016/j.epsl.2014.06.047Search in Google Scholar

Degruyter, W., Huber, C., Bachmann, O., Cooper, K.M., and Kent, A.J.R. (2015) Magma reservoir response to transient recharge events: The case of Santorini volcano (Greece). Geology, 44, 23–26.10.1130/G37333.1Search in Google Scholar

Del Gaudio, P., Ventura, G., and Taddeucci, J. (2013) The effect of particle size on the rheology of liquid-solid mixtures with application to lava flows: Results from analogue experiments. Geochemistry, Geophysics, Geosystems, 14, 2661–2669.10.1002/ggge.20172Search in Google Scholar

DePaolo, D.J. (1981) Trace element and isotopic effects of combined wallrock assimilation and fractional crystallization. Earth and Planetary Science Letters, 53, 189–202.10.1016/0012-821X(81)90153-9Search in Google Scholar

Dorais, M.J., Whitney, J.A., and Stormer, J.C. (1991) Mineralogical constraints on the petrogenesis of trachytic inclusions, Carpenter Ridge Tuff, Central San Juan volcanic field, Colorado. Contributions to Mineralogy and Petrology, 107, 219–230.10.1007/BF00310708Search in Google Scholar

Drenth, B.J., and Keller, G.R. (2004) New gravity and magnetic maps of the San Juan volcanic field, southwestern Colorado. EOS Transaction, American Geophysical Union, 85(F623).Search in Google Scholar

Drenth, B.J., Keller, G.R., and Thompson, R.A. (2012) Geophysical study of the San Juan Mountains batholith complex, southwestern Colorado. Geosphere, 8, 669–684.10.1130/GES00723.1Search in Google Scholar

Druitt, T.H., and Bacon, C.R. (1989) Petrology of the zoned calcalkaline magma chamber of Mount Mazama, Crater Lake, Oregon. Contributions to Mineralogy and Petrology, 101, 245–259.10.1007/BF00375310Search in Google Scholar

Druitt, T.H., Costa, F., Deloule, E., Dungan, M., and Scaillet, B. (2012) Decadal to monthly timescales of magma transfer and reservoir growth at a caldera volcano. Nature, 482, 77–80.10.1038/nature10706Search in Google Scholar

Ducea, M., and Saleeby, J.B. (1998) Crustal recycling beneath continental arcs; silica-rich glass inclusions in ultramafic xenoliths from the Sierra Nevada, California. Earth and Planetary Science Letters, 156, 101–116.10.1016/S0012-821X(98)00021-1Search in Google Scholar

Dufek, J., and Bachmann, O. (2010) Quantum magmatism: Magmatic compositional gaps generated by melt-crystal dynamics. Geology, 38, 687–690.10.1130/G30831.1Search in Google Scholar

Dufek, J., and Bergantz, G.W. (2005) Lower crustal magma genesis and preservation: a stochastic framework for the evaluation of basalt–crust interaction. Journal of Petrology, 46, 2167–2195.10.1093/petrology/egi049Search in Google Scholar

Dunbar, N.W., Kyle, P.R., and Wilson, C.J.N. (1989) Evidence for limited zonation in silicic magma systems, Taupo volcanic zone, New Zealand. Geology, 17, 234–236.10.1130/0091-7613(1989)017<0234:EFLZIS>2.3.CO;2Search in Google Scholar

Dunbar, N.W., Campbell, A.R., and Candela, P.A. (1996) Physical, chemical, and mineralogical evidence for magmatic fluid migration within the Capitan pluton, southeastern New Mexico. Geological Society of America Bulletin, 108, 318–333.10.1130/0016-7606(1996)108<0318:PCAMEF>2.3.CO;2Search in Google Scholar

Dungan, M.A., Long, P.E., and Rhodes, J.M. (1978) Magma mixing at mid-ocean ridges: Evidence from legs 45 and 46 DSDP. Geophysical Research Letters, 5, 423–425.10.1029/GL005i006p00423Search in Google Scholar

Edmonds, M., Oppenheimer, C., Pyle, D.M., Herd, R.A., and Thompson, G. (2003) SO2 emissions from Soufriere Hills Volcano and their relationship to conduit permeability, hydrothermal interaction and degassing regime. Journal of Volcanology and Geothermal Research, 124, 23–43.10.1016/S0377-0273(03)00041-6Search in Google Scholar

Eichelberger, J.C., and Izbekov, P.E. (2000) Eruption of andesite triggered by dyke injection; contrasting cases at Karymsky Volcano, Kamchatka and Mt. Katmai, Alaska. Philosophical Transactions–Royal Society. Mathematical, Physical and Engineering Sciences, 358, 1465–1485.10.1098/rsta.2000.0599Search in Google Scholar

Eichelberger, J.C., Chertkoff, D.G., Dreher, S.T., and Nye, C.J. (2000) Magmas in collision; Rethinking chemical zonation in silicic magmas. Geology, 28, 603–606.10.1130/0091-7613(2000)28<603:MICRCZ>2.0.CO;2Search in Google Scholar

Eiler, J., Stolper, E.M., and McCanta, M.C. (2011) Intra- and intercrystalline oxygen isotope variations in minerals from basalts and peridotites. Journal of Petrology. Eichelberger, J.C. (1975) Origin of andesite and dacite; evidence of mixing at Glass Mountain in California and at other Circum-Pacific volcanoes. Geological Society of America Bulletin, 86, 1381–1391.Search in Google Scholar

Eiler, J.M., Crawford, A., Elliott, T.I.M., Farley, K.A., Valley, J.W., and Stolper, E.M. (2000) Oxygen isotope geochemistry of oceanic-arc lavas. Journal of Petrology, 41, 229–256.10.1093/petrology/41.2.229Search in Google Scholar

Ellis, B.S., and Wolff, J.A. (2012) Complex storage of rhyolite in the central Snake River Plain. Journal of Volcanology and Geothermal Research, 211–212, 1–11.10.1016/j.jvolgeores.2011.10.002Search in Google Scholar

Ellis, B.S., Wolff, J.A., Boroughs, S., Mark, D.F., Starkel, W.A., and Bonnichsen, B. (2013) Rhyolitic volcanism of the central Snake River Plain: A review. Bulletin of Volcanology, 75, 745.10.1007/s00445-013-0745-ySearch in Google Scholar

Ellis, B.S., Bachmann, O., and Wolff, J.A. (2014) Cumulate fragments in silicic ignimbrites: The case of the Snake River Plain. Geology, 42, 431–434.10.1130/G35399.1Search in Google Scholar

Evans, B.W., and Bachmann, O. (2013) Implications of equilibrium and disequilibrium among crystal phases in the Bishop Tuff. American Mineralogist, 98, 271–274.10.2138/am.2013.4280Search in Google Scholar

Evans, B.W., Hildreth, E., Bachmann, O., and Scaillet, B. (2016) In defense of Magnetite- Ilmenite Thermometry in the Bishop Tuff and its implication for gradients in silicic magma reservoirs. American Mineralogist, 101, 469–482.10.2138/am-2016-5367Search in Google Scholar

Ewart, A. (1982) The mineralogy and petrology of Tertiary-Recent orogenic volcanic rocks: With special reference to the andesitic-basaltic compositional range. In R.S. Thorpe, Ed., Andesites: Orogenic Andesites and Related Rocks, p. 25–95. Wiley, New York.Search in Google Scholar

Faroughi, S.A., and Huber, C. (2015) Unifying the relative hindered velocity in suspensions and emulsions of nondeformable particles. Geophysical Research Letters, 42, 2014GL062570.10.1002/2014GL062570Search in Google Scholar

Farrell, J., Smith, R.B., Husen, S., and Diehl, T. (2014) Tomography from 26 years of seismicity revealing that the spatial extent of the Yellowstone crustal magma reservoir extends well beyond the Yellowstone caldera. Geophysical Research Letters, 41, 2014GL059588.10.1002/2014GL059588Search in Google Scholar

Fiege, A., Behrens, H., Holtz, F., and Adams, F. (2014) Kinetic vs. thermodynamic control of degassing of H2O-S ± Cl-bearing andesitic melts. Geochimica et Cosmochimica Acta, 125, 241–264.10.1016/j.gca.2013.10.012Search in Google Scholar

Fiege, A., Holtz, F., Behrens, H., Mandeville, C.W., Shimizu, N., Crede, L.S., and Göttlicher, J.R. (2015) Experimental investigation of the S and S-isotope distribution between H2O–S ± Cl fluids and basaltic melts during decompression. Chemical Geology, 393–394, 36–54.10.1016/j.chemgeo.2014.11.012Search in Google Scholar

Folkes, C.B., de Silva, S.L., Schmitt, A.K., and Cas, R.A.F. (2011a) A reconnaissance of U-Pb zircon ages in the Cerro Galán system, NW Argentina: Prolonged magma residence, crystal recycling, and crustal assimilation. Journal of Volcanology and Geothermal Research, 206, 136–147.10.1016/j.jvolgeores.2011.06.001Search in Google Scholar

Folkes, C.B., de Silva, S.L., Wright, H.M., and Cas, R.A.F. (2011b) Geochemical homogeneity of a long-lived, large silicic system; evidence from the Cerro Galán caldera, NW Argentina. Bulletin of Volcanology.10.1007/s00445-011-0511-ySearch in Google Scholar

Folkes, C.B., Silva, S.L., Wright, H.M., and Cas, R.A.F. (2011c) Geochemical homogeneity of a long-lived, large silicic system; evidence from the Cerro Galán caldera, NW Argentina. Bulletin of Volcanology, 73, 1455–1486.10.1007/s00445-011-0511-ySearch in Google Scholar

Forni, F., Ellis, B.S., Bachmann, O., Lucchi, F., Tranne, C.A., Agostini, S., and Dallai, L. (2015) Erupted cumulate fragments in rhyolites from Lipari (Aeolian Islands). Contributions to Mineralogy and Petrology, 170, 1–18.10.1007/s00410-015-1201-0Search in Google Scholar

Fournier, T.J., Pritchard, M.E., and Riddick, S.N. (2010) Duration, magnitude, and frequency of subaerial volcano deformation events: New results from Latin America using InSAR and a global synthesis. Geochemistry, Geophysics, Geosystems, 11.10.1029/2009GC002558Search in Google Scholar

Francalanci, L., Varekamp, J.C., Vougioukalakis, G., Defant, M.J., Innocenti, F., and Manetti, P. (1995) Crystal retention, fractionation and crustal assimilation in a convecting magma chamber, Nisyros Volcano, Greece. Bulletin of Volcanology, 56, 601–620.10.1007/BF00301465Search in Google Scholar

Francis, P.W., Sparks, R.S.J., Hawkesworth, C.J., Thorpe, R.S., Pyle, D.M., Tait, S.R., Mantovani, M.S., and McDermott, F. (1989) Petrology and geochemistry of the Cerro Galan caldera, northwest Argentina. Geological Magazine, 126, 515–547.10.1017/S0016756800022834Search in Google Scholar

Freundt-Malecha, B., Schmicke, H.-U., and Freundt, A. (2001) Plutonic rocks of intermediate compositions on Gran Canaria: The missing link of the bimodal volcanic suite. Contributions to Mineralogy and Petrology, 141, 430–445.10.1007/s004100100250Search in Google Scholar

Friedman, I., Lipman, P.W., Obradovich, J.D., Gleason, J.D., and Christiansen, R.L. (1974) Meteoric water in magmas. Science, 184, 1069–1072.10.1126/science.184.4141.1069Search in Google Scholar PubMed

Furuichi, M., and Nishiura, D. (2014) Robust coupled fluid-particle simulation scheme in Stokes-flow regime: Toward the geodynamic simulation including granular media. Geochemistry, Geophysics, Geosystems, 15, 2865–2882.10.1002/2014GC005281Search in Google Scholar

Gardner, J., Befus, K., Gualda, G.R., and Ghiorso, M. (2014) Experimental constraints on rhyolite-MELTS and the Late Bishop Tuff magma body. Contributions to Mineralogy and Petrology, 168, 1–14.10.1007/s00410-014-1051-1Search in Google Scholar

Gebauer, S., Schmitt, A., Pappalardo, L., Stockli, D., and Lovera, O. (2014) Crystallization and eruption ages of Breccia Museo (Campi Flegrei caldera, Italy) plutonic clasts and their relation to the Campanian ignimbrite. Contributions to Mineralogy and Petrology, 167, 1–18.10.1007/s00410-013-0953-7Search in Google Scholar

Geist, D., Howard, K.A., and Larson, P. (1995) The Generation of oceanic rhyolites by crystal fractionation: The basalt-rhyolite association at Volcan Alcedo, Galapagos Archipelago. Journal of Petrology, 36, 965–982.10.1093/petrology/36.4.965Search in Google Scholar

Gelman, S.E., Deering, C.D., Gutierrez, F.J., and Bachmann, O. (2013a) Evolution of the Taupo Volcanic Center, New Zealand: Petrological and thermal constraints from the Omega dacite. Contributions to Mineralogy and Petrology, 166, 1355–1374.10.1007/s00410-013-0932-zSearch in Google Scholar

Gelman, S.E., Gutierrez, F.J., and Bachmann, O. (2013b) On the longevity of large upper crustal silicic magma reservoirs. Geology, 41, 759–762.10.1130/G34241.1Search in Google Scholar

Gelman, S.E., Deering, C.D., Bachmann, O., Huber, C., and Gutierrez, F.J. (2014) Identifying the crystal graveyards remaining after large silicic eruptions. Earth and Planetary Science Letters, 403, 299–306.10.1016/j.epsl.2014.07.005Search in Google Scholar

Gerlach, T.M., Westrich, H.R., and Symonds, R.B. (1996) Preeruption vapor in magma of the climactic Mount Pinatubo eruption: Source of the giant stratospheric sulfur dioxide cloud. In C. Newhall and R.S. Punongbayan, Eds., Fire and Mud: Eruptions and Lahars of Mount Pinatubo, Philippines. University of Washington Press, Seattle.Search in Google Scholar

Ghiorso, M.S., and Sack, R.O. (1995a) Chemical mass transfer in magmatic processes IV: A revised and internally consistent thermodynamic model for the interpolation and extrapolation of liquid-solid equilibria in magmatic systems at elevated temperatures and pressures. Contributions to Mineralogy and Petrology, 119, 197–212.10.1007/BF00307281Search in Google Scholar

——— (1995b) Chemical mass-transfer in magmatic processes, 4. A revised and internally consistent thermodynamic model for the interpolation and extrapolation of liquid-solid equilibria in magmatic systems at elevated-temperatures and pressures. Contributions to Mineralogy and Petrology, 119, 197–212.10.1007/BF00307281Search in Google Scholar

Gibert, D., Beauducel, F., Declais, Y., Lesparre, N., Marteau, J., Nicollin, F., and Tarantola, A. (2010) Muon tomography: Plans for observations in the Lesser Antilles. Earth, Planets and Space, 62, 153–165.10.5047/eps.2009.07.003Search in Google Scholar

Giggenbach, W.F. (1996) Chemical composition of volcanic gases. In R. Scarpa and R.I. Tilling, Eds., Monitoring and Mitigation of Volcanic Hazards, p. 221–256. Springer, Berlin.10.1007/978-3-642-80087-0_7Search in Google Scholar

Glazner, A., Coleman, D., and Mills, R. (2015) The Volcanic-Plutonic Connection, 1–22. Springer, Berlin.10.1007/11157_2015_11Search in Google Scholar

Glazner, A.F., and Bartley, J.M. (2006) Is stoping a volumetrically significant pluton emplacement process? Geological Society of America Bulletin, 118, 1185–1195.10.1130/B25738.1Search in Google Scholar

——— (2008) Reply to comments on “Is stoping a volumetrically significant pluton emplacement process?” Geological Society of America Bulletin, 120, 1082–1087.10.1130/B26312.1Search in Google Scholar

Glazner, A.F., Coleman, D.S., and Bartley, J.M. (2008) The tenuous connection between high-silica rhyolites and granodiorite plutons. Geology, 36.10.1130/G24496A.1Search in Google Scholar

Goff, F., Janik, C.J., Delgado, H., Werner, C., Counce, D., Stimac, J.A., Siebe, C., Love, S.P., Williams, S.N., Fischer, T.P., and Johnson, L. (1998) Geochemical surveillance of magmatic volatiles at Popocatepetl Volcano, Mexico. Geological Society of America Bulletin, 110, 695–710.10.1130/0016-7606(1998)110<0695:GSOMVA>2.3.CO;2Search in Google Scholar

Gonnermann, H.M., and Manga, M. (2007) The fluid mechanics inside a volcano. Annual Reviews of Fluid Mechanics, 39, 321–356.10.1146/annurev.fluid.39.050905.110207Search in Google Scholar

——— (2013) Dynamics of magma ascent in the volcanic conduit. In S.A. Fagents, T. Gregg, and R. Lopes, Eds., Modeling Volcanic Processes, p. 55–84. Cambridge University Press.10.1017/CBO9781139021562.004Search in Google Scholar

Gottsmann, J., Lavallee, Y., Marti, J., and Aguirre-Diaz, G. (2009) Magma-tectonic interaction and the eruption of silicic batholiths. Earth and Planetary Science Letters, 284, 426–434.10.1016/j.epsl.2009.05.008Search in Google Scholar

Graeter, K.A., Beane, R.J., Deering, C.D., Gravley, D., and Bachmann, O. (2015) Formation of rhyolite at the Okataina Volcanic Complex, New Zealand: New insights from analysis of quartz clusters in plutonic lithics. American Mineralogist, 100, 1778–1789.10.2138/am-2015-5135Search in Google Scholar

Graham, I.J., Cole, J.W., Briggs, R.M., Gamble, J.A., and Smith, I.E.M. (1995) Petrology and petrogenesis of volcanic rocks from the Taupo Volcanic Zone: a review. Journal of Volcanology and Geothermal Research, 68, 59–87.10.1016/0377-0273(95)00008-ISearch in Google Scholar

Greene, A.R., Debari, S.M., Kelemen, P.B., Blusztajn, J., and Clift, P.D. (2006) A detailed geochemical study of Island Arc Crust: The Talkeetna Arc Section, South-Central Alaska. Journal of Petrology, 47, 1051–1093.10.1093/petrology/egl002Search in Google Scholar

Gregg, P.M., de Silva, S.L., Grosfils, E.B., and Parmigiani, J.P. (2012) Catastrophic caldera-forming eruptions: Thermomechanics and implications for eruption triggering and maximum caldera dimensions on Earth. Journal of Volcanology and Geothermal Research, 241–242, 1–12.10.1016/j.jvolgeores.2012.06.009Search in Google Scholar

Gregg, P.M., de Silva, S.L., and Grosfils, E.B. (2013) Thermomechanics of shallow magma chamber pressurization: Implications for the assessment of ground deformation data at active volcanoes. Earth and Planetary Science Letters, 384, 100–108.10.1016/j.epsl.2013.09.040Search in Google Scholar

Grove, T.L., and Donnelly-Nolan, J.M. (1986) The evolution of young silicic lavas at Medicine Lake Volcano, California: Implications for the origin of compositional gaps in calc-alkaline series lavas. Contributions to Mineralogy and Petrology, 92, 281–302.10.1007/BF00572157Search in Google Scholar

Grove, T.L., Donnelly-Nolan, J.M., and Housh, T. (1997) Magmatic processes that generated the rhyolite of Glass Mountain, Medicine Lake volcano, N. California. Contributions to Mineralogy and Petrology, 127, 205–223.10.1007/s004100050276Search in Google Scholar

Grunder, A.L., Klemetti, E.W., Feeley, T.C., and McKee, C.L. (2008) Eleven million years of arc volcanism at the Aucanquilcha Volcanic Cluster, northern Chilean Andes: Implications for the lifespan and emplacement of batholiths. Transactions of the Royal Society of Edinburgh: Earth Sciences, 97, 415–436.10.1017/S0263593300001541Search in Google Scholar

Gualda, G.A.R., and Ghiorso, M.S. (2013a) The Bishop Tuff giant magma body: An alternative to the Standard Model. Contributions to Mineralogy and Petrology, 166, 755–775.10.1007/s00410-013-0901-6Search in Google Scholar

——— (2013b) Low-pressure origin of high-silica rhyolites and granites. The Journal of Geology, 121, 537–545.10.1086/671395Search in Google Scholar

Gualda, G.A.R., Ghiorso, M.S., Lemons, R.V., and Carley, T.L. (2012) Rhyolite-MELTS: A modified calibration of MELTS optimized for silica-rich, fluid-bearing magmatic systems. Journal of Petrology, 53, 875–890.10.1093/petrology/egr080Search in Google Scholar

Guillong, M., von Quadt, A., Sakata, S., Peytcheva, I., and Bachmann, O. (2014) LA-ICP-MS Pb-U dating of young zircons from the Kos-Nisyros volcanic centre, SE Aegean arc. Journal of Analytical Atomic Spectrometry, 29, 963–970.10.1039/C4JA00009ASearch in Google Scholar

Gutierrez, F., and Parada, M.A. (2010) Numerical modeling of time-dependent fluid dynamics and differentiation of a shallow basaltic magma chamber. Journal of Petrology, 51, 731–762.10.1093/petrology/egp101Search in Google Scholar

Gutierrez, F., Payacan, I., Gelman, S.E., Bachmann, O., and Parada, M.A. (2013) Late-stage magma flow in a shallow felsic reservoir: Merging the anisotropy of magnetic susceptibility record with numerical simulations in La Gloria Pluton, central Chile. Journal of Geophysical Research-Solid Earth, 118, 1984–1998.10.1002/jgrb.50164Search in Google Scholar

Halliday, A.N., Fallick, A.E., Hutchinson, J., and Hildreth, W. (1984) A Nd, Sr, O isotopic investigation into the causes of chemical and isotopic zonation in the Bishop Tuff, California. Earth and Planetary Science Letters, 68, 378–391.10.1016/0012-821X(84)90123-7Search in Google Scholar

Halliday, A.N., Davidson, J.P., Hildreth, W., and Holden, P. (1991) Modeling the petrogenesis of high Rb/Sr silicic magmas. Chemical Geology, 92, 107–114.10.1016/0009-2541(91)90051-RSearch in Google Scholar

Hamilton, W., and Myers, W.B. (1967) The nature of batholiths. U.S. Geological Survey Professional Paper, 554-C, 30 p.10.3133/pp554CSearch in Google Scholar

Hautmann, S., Witham, F., Christoper, T., Cole, P., Linde, A.T., Sacks, S., and Sparks, R.S.J. (2014) Strain field analysis on Montserrat (W.I.) as tool for assessing permeable flow paths in the magmatic system of Soufrière Hills Volcano. Geochemistry, Geophysics, Geosystems, 15.10.1002/2013GC005087Search in Google Scholar

Heinrich, C.A., and Candela, P.A. (2012) Fluids and Ore Formation in the Earth's Crust. Treatise on Geochemistry, 13, 1–28.10.1016/B978-0-08-095975-7.01101-3Search in Google Scholar

Heise, W., Caldwell, T.G., Bibby, H.M., and Bennie, S.L. (2010) Three-dimensional electrical resistivity image of magma beneath an active continental rift, Taupo Volcanic Zone, New Zealand. Geophysical Research Letters, 37, L10301.10.1029/2010GL043110Search in Google Scholar

Heliker, C. (1995) Inclusions in Mount St. Helens dacite erupted from 1980 through 1983. Journal of Volcanology and Geothermal Research, 66, 115–135.10.1016/0377-0273(94)00074-QSearch in Google Scholar

Hildreth, W. (1979) The Bishop Tuff: Evidence for the origin of the compositional zonation in silicic magma chambers. In C.E. Chapin and W.E. Elston, Eds., Ash-Flow Tuffs, 180, 43–76. Geological Society of America, Special Paper 180.10.1130/SPE180-p43Search in Google Scholar

——— (1981) Gradients in silicic magma chambers: Implications for lithospheric magmatism. Journal of Geophysical Research, 86(B11), 10153–10192.Search in Google Scholar

Hildreth, W., and Fierstein, J. (2000) Katmai volcanic cluster and the great eruption of 1912. Geological Society of America Bulletin, 112, 1594–1620.10.1130/0016-7606(2000)112<1594:KVCATG>2.0.CO;2Search in Google Scholar

Hildreth, W., Halliday, A.N., and Christiansen, R.L. (1991) Isotopic and chemical evidence concerning the genesis and contamination of basaltic and rhyolitic magma beneath the Yellowstone Plateau volcanic field. Journal of Petrology, 32, 63–137.10.1093/petrology/32.1.63Search in Google Scholar

Hildreth, W.S. (2004) Volcanological perspectives on Long Valley, Mammoth Mountain, and Mono Craters: Several contiguous but discrete systems. Journal of Volcanology and Geothermal Research, 136, 169–198.10.1016/j.jvolgeores.2004.05.019Search in Google Scholar

——— (2007) Quaternary magmatism in the cascades. Geological Perspectives. U.S. Geological Survey Professional Paper, 1744, 125 p.Search in Google Scholar

Hildreth, W.S., and Wilson, C.J.N. (2007) Compositional zoning in the Bishop Tuff. Journal of Petrology, 48, 951–999.10.1093/petrology/egm007Search in Google Scholar

Hill, G.J., Caldwell, T.G., Chertkoff, D.G., Bibby, H.M., Burgess, M.K., Cull, J.P., and Cas, R.A.F. (2009) Distribution of melt beneath Mount St. Helens and Mount Adams inferred from magnetotelluric data. Nature Geoscience, 2, 785–789.10.1038/ngeo661Search in Google Scholar

Holland, T.J.B., and Powell, R. (2011) An improved and extended internally consistent thermodynamic dataset for phases of petrological interest, involving a new equation of state for solids. Journal of Metamorphic Geology, 29, 333–383.10.1111/j.1525-1314.2010.00923.xSearch in Google Scholar

Holtzman, R., Szulczewski, M.L., and Juanes, R. (2012) Capillary fracturing in granular media. Physical Review Letters, 108, 264504.10.1103/PhysRevLett.108.264504Search in Google Scholar

Hooper, A., Zebker, H., Segall, P., and Kampes, B. (2004) A new method for measuring deformation on volcanoes and other natural terrains using InSAR persistent scatterers. Geophysical Research Letters, 31.10.1029/2004GL021737Search in Google Scholar

Huang, F., Chakraborty, P., Lundstrom, C.C., Holmden, C., Glessner, J.J.G., Kieffer, S.W., and Lesher, C.E. (2010) Isotope fractionation in silicate melts by thermal diffusion. Nature, 464, 396–400.10.1038/nature08840Search in Google Scholar PubMed

Huang, H.-H., Lin, F.-C., Schmandt, B., Farrell, J., Smith, R.B., and Tsai, V.C. (2015) The Yellowstone magmatic system from the mantle plume to the upper crust. Science, 348, 773–776.10.1126/science.aaa5648Search in Google Scholar PubMed

Huber, C., Bachmann, O., and Manga, M. (2009) Homogenization processes in silicic magma chambers by stirring and mushification (latent heat buffering). Earth and Planetary Science Letters, 283, 38–47.10.1016/j.epsl.2009.03.029Search in Google Scholar

Huber, C., Bachmann, O., and Dufek, J. (2010a) The limitations of melting on the reactivation of silicic mushes. Journal of Volcanology and Geothermal Research, 195, 97–105.10.1016/j.jvolgeores.2010.06.006Search in Google Scholar

Huber, C., Bachmann, O., and Manga, M. (2010b) Two competing effects of volatiles on heat transfer in crystal-rich magmas: Thermal insulation vs. defrosting. Journal of Petrology, 51, 847–867.10.1093/petrology/egq003Search in Google Scholar

Huber, C., Bachmann, O., and Dufek, J. (2011) Thermo-mechanical reactivation of locked crystal mushes: Melting-induced internal fracturing and assimilation processes in magmas. Earth and Planetary Science Letters, 304, 443–454.10.1016/j.epsl.2011.02.022Search in Google Scholar

——— (2012a) Crystal-poor versus crystal-rich ignimbrites: A competition between stirring and reactivation. Geology, 40, 115–118.10.1130/G32425.1Search in Google Scholar

Huber, C., Bachmann, O., Vigneresse, J.L., Dufek, J., and Parmigiani, A. (2012b) A physical model for metal extraction and transport in shallow magmatic systems. Geochemistry Geophysics Geosystems, 13, Q08003, 10.1029/2012GC004042.Search in Google Scholar

Huber, C., Parmigiani, A., Latt, J., and Dufek, J. (2013) Channelization of buoyant nonwetting fluids in saturated porous media. Water Resources Research, 49, 6371–6380.10.1002/wrcr.20514Search in Google Scholar

Hughes, G.R., and Mahood, G.A. (2008) Tectonic controls on the nature of large silicic calderas in volcanic arcs. Geology, 36, 627–630.10.1130/G24796A.1Search in Google Scholar

Humphreys, M.C.S., Edmonds, M., Christopher, T., and Hards, V. (2009) Chlorine variations in the magma of Soufriere Hills Volcano, Montserrat: Insights from Cl in hornblende and melt inclusions. Geochimica et Cosmochimica Acta, 73, 5693–5708.10.1016/j.gca.2009.06.014Search in Google Scholar

Huppert, H.E., and Woods, A.W. (2002) The role of volatiles in magma chamber dynamics. Nature, 420, 493–495.10.1038/nature01211Search in Google Scholar

Huppert, H.E., Sparks, R.S.J., and Turner, J.S. (1982) Effects of volatiles on mixing in calc-alkaline magma systems. Nature, 332, 554–557.10.1038/297554a0Search in Google Scholar

Husen, S., Smith, R.B., and Waite, G.P. (2004) Evidence for gas and magmatic sources beneath the Yellowstone volcanic field from seismic tomographic imaging. Journal of Volcanology and Geothermal Research, 131, 397–410.10.1016/S0377-0273(03)00416-5Search in Google Scholar

Irvine, T.N., Andersen, J.C.ø., and Brooks, C.K. (1998) Included blocks (and blocks within blocks) in the Skaergaard intrusion: Geologic relations and the origins of rhythmic modally graded layers. Geological Society of America Bulletin, 110, 1398–1447.10.1130/0016-7606(1998)110<1398:IBABWB>2.3.CO;2Search in Google Scholar

Ishibashi, H., and Sato, H. (2007) Viscosity measurements of subliquidus magmas: Alkali olivine basalt from the Higashi-Matsuura district, Southwest Japan. Journal of Volcanology and Geothermal Research, 160, 223–238.10.1016/j.jvolgeores.2006.10.001Search in Google Scholar

Jagoutz, O., and Schmidt, M.W. (2012) The formation and bulk composition of modern juvenile continental crust: The Kohistan arc. Chemical Geology, 298–299, 79–96.10.1016/j.chemgeo.2011.10.022Search in Google Scholar

——— (2013) The composition of the foundered complement to the continental crust and a re-evaluation of fluxes in arcs. Earth and Planetary Science Letters, 371–372, 177–190.Search in Google Scholar

Jahns, R.H., and Tuttle, O.F. (1963) Layered pegmatites-aplites intrusives. Mineralogical Society of America Special Paper, 1, 78–92.Search in Google Scholar

Jaupart, C., and Brandeis, G. (1986) The stagnant bottom layer of convecting magma chambers. Earth and Planetary Science Letters, 80, 183–199.10.1016/0012-821X(86)90032-4Search in Google Scholar

Jaupart, C., Brandeis, G., and Alegre, C.J. (1984) Stagnant layers at the bottom of convecting magma chambers. Nature, 308, 535–538.10.1038/308535a0Search in Google Scholar

Jay, J., Pritchard, M., West, M., Christensen, D., Haney, M., Minaya, E., Sunagua, M., McNutt, S., and Zabala, M. (2012) Shallow seismicity, triggered seismicity, and ambient noise tomography at the long-dormant Uturuncu Volcano, Bolivia. Bulletin of Volcanology, 74, 817–837.10.1007/s00445-011-0568-7Search in Google Scholar

Jellinek, A.M., and DePaolo, D.J. (2003) A model for the origin of large silicic magma chambers: precursors of caldera-forming eruptions. Bulletin of Volcanology, 65, 363–381.10.1007/s00445-003-0277-ySearch in Google Scholar

Jellinek, A.M., and Kerr, R.C. (1999) Mixing and compositional stratification produced by natural convection: 2. Applications to the differentiation of basaltic and silicic magma chambers and komatiite lava flows. Journal of Geophysical Research, 104(B4), 7203–7218.10.1029/1998JB900117Search in Google Scholar

Johannes, W., and Holtz, F. (1996) Petrogenesis and Experimental Petrology of Granitic Rocks, 335 p. Springer-Verlag, Berlin.10.1007/978-3-642-61049-3Search in Google Scholar

Johnson, M., and Rutherford, M. (1989) Experimentally determined conditions in the Fish Canyon Tuff, Colorado, magma chamber. Journal of Petrology, 30, 711–737.10.1093/petrology/30.3.711Search in Google Scholar

Jull, M., and Kelemen, P.B. (2001) On the conditions for lower crustal convective instability. Journal of Geophysical Research, 106, 6423–6446.10.1029/2000JB900357Search in Google Scholar

Kamiyama, H., Nakajima, T., and Kamioka, H. (2007) Magmatic stratigraphy of the Tilted Tottabetsu Plutonic Complex, Hokkaido, North Japan: Magma chamber dynamics and pluton construction. Journal of Geology, 115, 295–314.10.1086/512754Search in Google Scholar

Karakas, O., and Dufek, J. (2015) Melt evolution and residence in extending crust: Thermal modeling of the crust and crustal magmas. Earth and Planetary Science Letters, 425, 131–144.10.1016/j.epsl.2015.06.001Search in Google Scholar

Karani, H., and Huber, C. (2015) Pore-scale study of Horton-Rogers-Lapwood convection in porous media: Effect of microscale thermophysical heterogeneity on the onset of convection. APS Division of Fluid Dynamics Meeting Abstracts, http://meetings. aps.org/Meeting/DFD15/Session/D1.4Search in Google Scholar

Karlstrom, L., Rudolph, M.L., and Manga, M. (2012) Caldera size modulated by the yield stress within a crystal-rich magma reservoir. Nature Geoscience, 5, 402–405.10.1038/ngeo1453Search in Google Scholar

Kay, R.W., and Mahlburg Kay, S. (1993) Delamination and delamination magmatism. Tectonophysics, 219, 177–189.10.1016/0040-1951(93)90295-USearch in Google Scholar

Keller, C.B., and Schoene, B. (2012) Statistical geochemistry reveals disruption in secular lithospheric evolution about 2.5 Gyr ago. Nature, 485, 490–493.10.1038/nature11024Search in Google Scholar PubMed

Keller, C.B., Schoene, B., Barboni, M., Samperton, K.M., and Husson, J.M. (2015) Volcanic-plutonic parity and the differentiation of the continental crust. Nature, 523, 301–307.10.1038/nature14584Search in Google Scholar

Keller, J. (1969) Origin of rhyolites by anatectic melting of granitic crustal rocks; the example of rhyolitic pumice from the island of Kos (Aegean Sea). Bulletin Volcanologique, 33, 942–959.10.1007/BF02596758Search in Google Scholar

Kennedy, B.M., Mark Jellinek, A., and Stix, J. (2008) Coupled caldera subsidence and stirring inferred from analogue models. Nature Geoscience, 1, 385–389.10.1038/ngeo206Search in Google Scholar

Klemetti, E., and Grunder, A. (2008) Volcanic evolution of Volcán Aucanquilcha: A long-lived dacite volcano in the Central Andes of northern Chile. Bulletin of Volcanology, 70, 633–650.10.1007/s00445-007-0158-xSearch in Google Scholar

Klemetti, E.W., and Clynne, M.A. (2014) Localized rejuvenation of a crystal mush recorded in zircon temporal and compositional variation at the Lassen Volcanic Center, Northern California. PLoS ONE, 9, e113157.10.1371/journal.pone.0113157Search in Google Scholar

Klemetti, E.W., Deering, C.D., Cooper, K.M., and Roeske, S.M. (2011) Magmatic perturbations in the Okataina Volcanic Complex, New Zealand at thousand-year timescales recorded in single zircon crystals. Earth and Planetary Science Letters, 305, 185–194.10.1016/j.epsl.2011.02.054Search in Google Scholar

Koyaguchi, T., and Blake, S. (1989) The dynamics of magma mixing in a rising magma batch. Bulletin of Volcanology, 52, 127–137.10.1007/BF00301552Search in Google Scholar

Koyaguchi, T., and Kaneko, K. (1999) A two-stage thermal evolution model of magmas in continental crust. Journal of Petrology, 40, 241–254.10.1093/petroj/40.2.241Search in Google Scholar

Koyaguchi, T., Hallworth, M.A., Huppert, H.E., and Sparks, R.S.J. (1990) Sedimentation of particles from a convecting fluid. Nature, 343, 447–450.10.1038/343447a0Search in Google Scholar

Koyaguchi, T., Hallworth, M.A., and Huppert, H.E. (1993) An experimental study on the effects of phenocrysts on convection in magmas. Journal of Volcanology and Geothermal Research, 55, 15–32.10.1016/0377-0273(93)90087-8Search in Google Scholar

Lagios, E., Sakkas, V., Parcharidis, I., and Dietrich, V. (2005) Ground deformation of Nisyros Volcano (Greece) for the period 1995–2002: Results from DInSAR and DGPS observations. Bulletin of Volcanology, 68, 201–214.10.1007/s00445-005-0004-ySearch in Google Scholar

Laumonier, M., Scaillet, B., Pichavant, M., Champallier, R.m., Andujar, J., and Arbaret, L. (2014) On the conditions of magma mixing and its bearing on andesite production in the crust. Nature Communications, 5.10.1038/ncomms6607Search in Google Scholar

Lavallée, Y., Hess, K.-U., Cordonnier, B., and Dingwell, D.B. (2007) Non-Newtonian rheological law for highly crystalline dome lavas. Geology, 35, 843–846.10.1130/G23594A.1Search in Google Scholar

Lee, C.-T.A., and Morton, D.M. (2015) High silica granites: Terminal porosity and crystal settling in shallow magma chambers. Earth and Planetary Science Letters, 409, 23–31.10.1016/j.epsl.2014.10.040Search in Google Scholar

Lee, C.-T.A., Morton, D.M., Farner, M.J., and Moitra, P. (2015) Field and model constraints on silicic melt segregation by compaction/hindered settling: The role of water and its effect on latent heat release. American MIneralogist, 100, 1762–1777.10.2138/am-2015-5121Search in Google Scholar

Leeman, W.P., Annen, C., and Dufek, J. (2008) Snake River Plain—Yellowstone silicic volcanism: implications for magma genesis and magma fluxes. In C. Annen and G. Zellmer, Eds., Dynamics of Crustal Magma Transfer, Storage and Differentiation, 304, 235–259. Geological Society, London.10.1144/SP304.12Search in Google Scholar

Lees, J.M. (1992) The magma system of Mount St. Helens: Non-linear high-resolution P-wave tomography. Journal of Volcanology and Geothermal Research, 53, 103–116.10.1016/0377-0273(92)90077-QSearch in Google Scholar

——— (2007) Seismic tomography of magmatic systems. Journal of Volcanology and Geothermal Research, 167, 37–56.10.1016/j.jvolgeores.2007.06.008Search in Google Scholar

Lejeune, A.-M., and Richet, P. (1995) Rheology of crystal-bearing silicate melts; An experimental study at high viscosities. Journal of Geophysical Research, 100, 4215–4229.10.1029/94JB02985Search in Google Scholar

Lindsay, J.M., Schmitt, A.K., Trumbull, R.B., De Silva, S.L., Siebel, W., and Emmermann, R. (2001) Magmatic evolution of the La Pacana caldera system, Central Andes, Chile: Compositional variation of two cogenetic, large-volume felsic ignimbrites. Journal of Petrology, 42, 459–486.10.1093/petrology/42.3.459Search in Google Scholar

Lipman, P.W. (1967) Mineral and chemical variations within an ash-flow sheet from Aso caldera, South Western Japan. Contributions to Mineralogy and Petrology, 16, 300–327.10.1007/BF00371528Search in Google Scholar

——— (1984) The roots of ash-flow calderas in western North America: Windows into the tops of granitic batholiths. Journal of Geophysical Research, 89 (B10), 8801–8841.Search in Google Scholar

——— (2000) The central San Juan caldera cluster: Regional volcanic framework. In P.M. Bethke and R.L. Hay, Eds., Ancient Lake Creede: Its Volcano-tectonic setting, history of sedimentation, and relation of mineralization in the Creede Mining District. Geological Society of America Special Paper, 346, 9–69.Search in Google Scholar

——— (2006) Chemical analyses of tertiary volcanic rocks, Central San Juan Caldera Complex, Southwestern Colorado. Open-File Report 2004-1194.Search in Google Scholar

——— (2007) Incremental assembly and prolonged consolidation of Cordilleran magma chambers: Evidence from the Southern Rocky Mountain Volcanic Field. Geosphere, 3, 1–29.Search in Google Scholar

Lipman, P.W., and Bachmann, O. (2015) Ignimbrites to batholiths: Integrating perspectives from geological, geophysical, and geochronological data. Geosphere, 11.10.1130/GES01091.1Search in Google Scholar

Lipman, P.W., and McIntosh, W.C. (2008) Eruptive and noneruptive calderas, northeastern San Juan Mountains, Colorado: Where did the ignimbrites come from? Geological Society of America Bulletin, 120, 771–795.10.1130/B26330.1Search in Google Scholar

Lipman, P.W., Doe, B., and Hedge, C. (1978) Petrologic evolution of the San Juan volcanic field, Southwestern Colorado: Pb and Sr isotope evidence. Geological Society of America Bulletin, 89, 59–82.10.1130/0016-7606(1978)89<59:PEOTSJ>2.0.CO;2Search in Google Scholar

Lipman, P.W., Dungan, M.A., Brown, L.L., and Deino, A.L. (1996) Recurrent eruption and subsidence at the Platoro Caldera complex, southeastern San Juan volcanic field, Colorado; new tales from old tuffs. Geological Society of America Bulletin, 108, 1039–1055.10.1130/0016-7606(1996)108<1039:REASAT>2.3.CO;2Search in Google Scholar

Lipman, P.W., Dungan, M.A., and Bachmann, O. (1997) Comagmatic granophyric granite in the Fish Canyon Tuff, Colorado: Implications for magma-chamber processes during a large ash-flow eruption. Geology, 25, 915–918.10.1130/0091-7613(1997)025<0915:CGGITF>2.3.CO;2Search in Google Scholar

Lohmar, S., Parada, M., Gutierrez, F., Robin, C., and Gerbe, M.C. (2012) Mineralogical and numerical approaches to establish the pre-eruptive conditions of the mafic Lican Ignimbrite, Villarrica Volcano (Chilean Southern Andes). Journal of Volcanology and Geothermal Research, 235–236, 55–69.10.1016/j.jvolgeores.2012.05.006Search in Google Scholar

Longpré, M.-A., Klügel, A., Diehl, A., and Stix, J. (2014) Mixing in mantle magma reservoirs prior to and during the 2011–2012 eruption at El Hierro, Canary Islands. Geology, 42.4, 315–318.10.1130/G35165.1Search in Google Scholar

Lowenstern, J.B. (2003) Melt inclusions come of age: Volatiles, Volcanoes, and Sorby's Legacy. In B. De Vivo and R.J. Bodnar, Eds., Melt Inclusions in Volcanic Systems: Methods, applications and problems. Developments in Volcanology 5, 1–22. Elsevier, Amsterdam.10.1016/S1871-644X(03)80021-9Search in Google Scholar

Lowenstern, J.B., and Hurvitz, S. (2008) Monitoring a Supervolcano in repose: Heat and volatile flux at the Yellowstone Caldera. Elements, 4, 35–40.10.2113/GSELEMENTS.4.1.35Search in Google Scholar

Lowenstern, J.B., Persing, H.M., Wooden, J.L., Lanphere, M., Donnelly-Nolan, J., and Grove, T.L. (2000) U-Th dating of single zircons from young granitoid xenoliths: New tools for understanding volcanic processes. Earth and Planetary Science Letters, 183, 291–302.10.1016/S0012-821X(00)00273-9Search in Google Scholar

Luhr, J.F., Carmichael, I.S.E., and Varekamp, J.C. (1984) The 1982 eruptions of El Chichûn Volcano, Chiapas, Mexico: Mineralogy and petrology of the anhydritebearing pumices. Journal of Volcanology and Geothermal Research, 23, 69–108.10.1016/0377-0273(84)90057-XSearch in Google Scholar

Lukács, R., Harangi, S., Bachmann, O., Guillong, M., Danišík, M., von Quadt, A., Dunkl, I., Fodor, L., Sliwinski, J., Soós, I., and Szepesi, J. (2015) Zircon geochronology and geochemistry to constrain the youngest eruption events and magma evolution of the Mid-Miocene ignimbrite flare-up in the Pannonian Basin, eastern-central Europe. Contributions to Mineralogy and Petrology, 170, 52 (26 p.), 10.1007/s00410-015-1206-8.10.1007/s00410-015-1206-8Search in Google Scholar

Lyell, C.S. (1838) Elements of Geology. John Murray, London.10.5962/bhl.title.83999Search in Google Scholar

Macdonald, R., Belkin, H.E., Fitton, J.G., Rogers, N.W., Nejbert, K., Tindle, A.G., and Marshall, A.S. (2008) The roles of fractional crystallization, magma mixing, crystal mush remobilization and volatile-melt interactions in the genesis of a young basalt-peralkaline rhyolite suite, the Greater Olkaria Volcanic Complex, Kenya Rift Valley. Journal of Petrology, 49, 1515–1547.10.1093/petrology/egn036Search in Google Scholar

Mader, H.M., Llewellin, E.W., and Mueller, S.P. (2013) The rheology of two-phase magmas: A review and analysis. Journal of Volcanology and Geothermal Research, 257, 135–158.10.1016/j.jvolgeores.2013.02.014Search in Google Scholar

Malfait, W.J., Seifert, R., Petitgirard, S., Perrillat, J.-P., Mezouar, M., Ota, T., Nakamura, E., Lerch, P., and Sanchez-Valle, C. (2014) Supervolcano eruptions driven by melt buoyancy in large silicic magma chambers. Nature Geoscience, 7, 122–125.10.1038/ngeo2042Search in Google Scholar

Marsh, B.D. (1981) On the crystallinity, probability of occurrence, and rheology of lava and magma. Contributions to Mineralogy and Petrology, 78, 85–98.10.1007/BF00371146Search in Google Scholar

——— (1984) On the mechanics of caldera resurgence. Calderas and Associate Igneous Rocks, 8245–8251. American Geophysical Union.Search in Google Scholar

——— (1989a) Magma chambers. Annual Review of Earth and Planetary Sciences, 17, 439–474.10.1146/annurev.ea.17.050189.002255Search in Google Scholar

——— (1989b) On convective style and vigor in sheet-like magma chambers. Journal of Petrology, 30, 479–530.10.1093/petrology/30.3.479Search in Google Scholar

——— (1996) Solidification fronts and magmatic evolution. Mineralogical Magazine, 60, 5–40.10.1180/minmag.1996.060.398.03Search in Google Scholar

——— (2002) On bimodal differentiation by solidification front instability in basaltic magmas, part 1: basic mechanics. Geochimica et Cosmochimica Acta, 66, 2211–2229.10.1016/S0016-7037(02)00905-5Search in Google Scholar

Marsh, B.D., and Maxey, M.R. (1985) On the distribution and separation of crystals in convecting magma. Journal of Volcanology and Geothermal Research, 24, 95–150.10.1016/0377-0273(85)90030-7Search in Google Scholar

Marteau, J., Gibert, D., Lesparre, N., Nicollin, F., Noli, P., and Giacoppo, F. (2012) Muons tomography applied to geosciences and volcanology. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 695, 23–28.10.1016/j.nima.2011.11.061Search in Google Scholar

Martin, D., and Nokes, R. (1988) Crystal settling in vigorously convecting magma chamber. Nature, 332, 534–536.10.1038/332534a0Search in Google Scholar

——— (1989) A fluid-dynamical study of crystal settling in convecting magmas. Journal of Petrology, 30, 1471–1500.10.1093/petrology/30.6.1471Search in Google Scholar

Martin, V.M., Morgan, D.J., Jerram, D.A., Caddick, M.J., Prior, D.J., and Davidson, J.P. (2008) Bang! Month-scale eruption triggering at Santorini Volcano. Science, 321, 1178.10.1126/science.1159584Search in Google Scholar PubMed

Mason, B.G., Pyle, D.M., and Oppenheimer, C. (2004) The size and frequency of the largest explosive eruptions on Earth. Bulletin of Volcanology, 66, 735–748.10.1007/s00445-004-0355-9Search in Google Scholar

Massonnet, D., Briole, P., and Arnaud, A. (1995) Deflation of Mount Etna monitored by spaceborne radar interferometry. Nature, 375, 567–570.10.1038/375567a0Search in Google Scholar

Masturyono, McCaffrey, R., Wark, D.A., and Roecker, S.W. (2001) Distribution of magma beneath the Toba caldera complex, north Sumatra, Indonesia, constrained by three-dimensional P wave velocities, seismicity, and gravity data. Geochemistry, Geophysics, Geosystems, 2.10.1029/2000GC000096Search in Google Scholar

Matsuo, S. (1962) Establishment of chemical equilibrium in the volcanic gas obtained from the lava lake of Kilauea, Hawaii. Bulletin Volcanologique, 24, 59–71.10.1007/BF02599329Search in Google Scholar

Matthews, N., Pyle, D., Smith, V., Wilson, C., Huber, C., and van Hinsberg, V. (2012) Quartz zoning and the pre-eruptive evolution of the ∼340-ka Whakamaru magma systems, New Zealand. Contributions to Mineralogy and Petrology, 163, 87–107.10.1007/s00410-011-0660-1Search in Google Scholar

Matzel, J.E.P., Bowring, S.A., and Miller, R.B. (2006) Time scales of pluton construction at differing crustal levels: Examples from the Mount Stuart and Tenpeak intrusions, North Cascades, Washington. Geological Society of America Bulletin, 118, 1412–1430.10.1130/B25923.1Search in Google Scholar

Maughan, L.L., Christiansen, E.H., Best, M.G., Gromme, C.S., Deino, A.L., and Tingey, D.G. (2002) The Oligocene Lund Tuff, Great Basin, U.S.A.: A very large volume monotonous intermediate. Journal of Volcanology and Geothermal Research, 113, 129–157.10.1016/S0377-0273(01)00256-6Search in Google Scholar

McBirney, A.R. (1980) Mixing and unmixing of magmas. Journal of Volcanology and Geothermal Research, 7, 357–371.10.1016/0377-0273(80)90038-4Search in Google Scholar

——— (1993) Igneous Petrology, 508 p. Jones and Barlett, London.Search in Google Scholar

——— (1995) Mechanisms of differentiation in the Skaergaard Intrusion. Journal of the Geological Society, 152, 421–435.10.1144/gsjgs.152.3.0421Search in Google Scholar

McBirney, A.R., Baker, B.H., and Nilson, R.H. (1985) Liquid fractionation. Part 1: Basic principles and experimental simulations. Journal of Volcanology and Geothermal Research, 24, 1–24.10.1016/0377-0273(85)90026-5Search in Google Scholar

McCarthy, T.S., and Groves, D.I. (1979) The Blue Tier Batholith, Northeastern Tasmania. Contributions to Mineralogy and Petrology, 71, 193–209.10.1007/BF00375436Search in Google Scholar

McKenzie, D.P. (1985) The extraction of magma from the crust and mantle. Earth and Planetary Science Letters, 74, 81–91.10.1016/0012-821X(85)90168-2Search in Google Scholar

Melekhova, E., Annen, C., and Blundy, J. (2013) Compositional gaps in igneous rock suites controlled by magma system heat and water content. Nature Geoscience, 6, 385–390.10.1038/ngeo1781Search in Google Scholar

Metrich, N., and Mandeville, C.W. (2010) Sulfur in magmas. Elements, 6, 81–86.10.2113/gselements.6.2.81Search in Google Scholar

Michaut, C., and Jaupart, C. (2006) Ultra-rapid formation of large volumes of evolved magma. Earth and Planetary Science Letters, 250, 38–52.10.1016/j.epsl.2006.07.019Search in Google Scholar

Miller, C.F., and Miller, J.S. (2002) Contrasting stratified plutons exposed in tilt blocks, Eldorado Mountains, Colorado River Rift, Nevada, U.S.A. Lithos, 61, 209–224.10.1016/S0024-4937(02)00080-4Search in Google Scholar

Miller, C.F., and Wark, D.A. (2008) Supervolcanoes and their explosive supereruptions. Elements, 4, 11–16.10.2113/GSELEMENTS.4.1.11Search in Google Scholar

Miller, J.S., Matzel, J.E.P., Miller, C.F., Burgess, S.D., and Miller, R.B. (2007) Zircon growth and recycling during the assembly of large, composite arc plutons. Journal of Volcanology and Geothermal Research, 167, 282–299.10.1016/j.jvolgeores.2007.04.019Search in Google Scholar

Mills, J.G., Saltoun, B.W., and Vogel, T.A. (1997) Magma batches in the timber mountain magmatic system, southwestern Nevada volcanic field, Nevada, U.S.A. Journal of Volcanology and Geothermal Research, 78, 185–208.10.1016/S0377-0273(97)00015-2Search in Google Scholar

Mohamed, F.H. (1998) Geochemistry and petrogenesis of El Gezira ring complex, Egypt: A monzosyenite cumulate derived from fractional crystallization of trachyandesitic magma. Journal of Volcanology and Geothermal Research, 84, 103–123.10.1016/S0377-0273(98)00034-1Search in Google Scholar

Moitra, P., and Gonnermann, H.M. (2015) Effects of crystal shape- and size-modality on magma rheology. Geochemistry, Geophysics, Geosystems, 16, 1–26.10.1002/2014GC005554Search in Google Scholar

Molloy, C., Shane, P., and Nairn, I. (2008) Pre-eruption thermal rejuvenation and stirring of a partly crystalline rhyolite pluton revealed by the Earthquake Flat Pyroclastics deposits, New Zealand. Journal of the Geological Society, London, 165, 435–447.10.1144/0016-76492007-071Search in Google Scholar

Morgan, D.J., and Blake, S. (2006) Magmatic residence times of zoned phenocrysts: Introduction and application of the binary element diffusion modelling (BEDM) technique. Contributions to Mineralogy and Petrology, 151, 58–70.10.1007/s00410-005-0045-4Search in Google Scholar

Morse, S.A. (2011) The fractional latent heat of crystallizing magmas. American Mineralogist, 96, 682–689.10.2138/am.2011.3613Search in Google Scholar

Mueller, S., Llewellin, E.W., and Mader, H.M. (2011) The effect of particle shape on suspension viscosity and implications for magmatic flows. Geophysical Research Letters, 38.10.1029/2011GL047167Search in Google Scholar

Muntener, O., and Ulmer, P. (2006) Experimentally derived high-pressure cumulates from hydrous arc magmas and consequences for the seismic velocity structure of lower arc crust. Geophysical Research Letters, 33, 10.1029/2006GL02762910.1029/2006GL027629Search in Google Scholar

Murphy, M.D., Sparks, R.S.J., Barclay, J., Carroll, M.R., and Brewer, T.S. (2000) Remobilization of andesitic magma by intrusion of mafic magma at the Soufrière Hills Volcano, Montserrat, West Indies. Journal of Petrology, 41, 21–42.10.1093/petrology/41.1.21Search in Google Scholar

Namiki, A., Hatakeyama, T., Toramaru, A., Kurita, K., and Sumita, I. (2003) Bubble size distributions in a convecting layer. Geophysical Research Letters, 30, 2–4.10.1029/2003GL017156Search in Google Scholar

Nash, B.P., Perkins, M.E., Christensen, J.N., Lee, D.-C., and Halliday, A.N. (2006) The Yellowstone hotspot in space and time: Nd and Hf isotopes in silicic magmas. Earth and Planetary Science Letters, 247, 143–156.10.1016/j.epsl.2006.04.030Search in Google Scholar

Newman, A.V., Dixon, T.H., and Gourmelen, N. (2006) A four-dimensional viscoelastic deformation model for Long Valley Caldera, California, between 1995 and 2000. Journal of Volcanology and Geothermal Research, 150, 244–269.10.1016/j.jvolgeores.2005.07.017Search in Google Scholar

Nilson, R.H., McBirney, A.R., and Baker, B.H. (1985) Liquid fractionation. Part 2: Fluid dynamics and quantitative implications for magmatic systems. Journal of Volcanology and Geothermal Research, 24, 25–54.10.1016/0377-0273(85)90027-7Search in Google Scholar

Nishimura, K. (2009) A trace-element geochemical model for imperfect fractional crystallization associated with the development of crystal zoning. Geochimica et Cosmochimica Acta, 73, 2142–2149.10.1016/j.gca.2009.01.011Search in Google Scholar

Oldenburg, C.M., Spera, F.J., Yuen, D.A., and Sewell, G. (1989) Dynamic mixing in magma bodies: Theory, simulations, and implications. Journal of Geophysical Research, 94(B7), 9215–9236.10.1029/JB094iB07p09215Search in Google Scholar

Oppenheimer, J., Rust, A.C., Cashman, K.V., and Sandnes, B. (2015) Gas migration regimes and outgassing in particle-rich suspensions. Frontiers in Physics, 3.10.3389/fphy.2015.00060Search in Google Scholar

Orsi, G., de Vita, S., and Di Vito, M. (1996) The restless, resurgent Campi Flegrei nested caldera (Italy): Constraints on its evolution and configuration. Journal of Volcanology and Geothermal Research, 74, 179–214.10.1016/S0377-0273(96)00063-7Search in Google Scholar

Otamendi, J.E., Ducea, M.N., and Bergantz, G.W. (2012) Geological, petrological and geochemical evidence for progressive construction of an Arc Crustal Section, Sierra de Valle Fertil, Famatinian Arc, Argentina. Journal of Petrology, 53, 761–800.10.1093/petrology/egr079Search in Google Scholar

Pallister, J.S., Hoblitt, R.P., and Reyes, A.G. (1992) A basalt trigger for the 1991 eruptions of Pinatubo volcano? Nature, 356, 426–428.10.1038/356426a0Search in Google Scholar

Pappalardo, L., and Mastrolorenzo, G. (2012) Rapid differentiation in a sill-like magma reservoir: a case study from the Campi Flegrei caldera. Scientific Reports, 2.10.1038/srep00712Search in Google Scholar PubMed PubMed Central

Parmigiani, A., Huber, C., Bachmann, O., and Chopard, B. (2011) Pore-scale mass and reactant transport in multiphase porous media flows. Journal of Fluid Mechanics, 686, 40–76.10.1017/jfm.2011.268Search in Google Scholar

Parmigiani, A., Huber, C., and Bachmann, O. (2014) Mush microphysics and the reactivation of crystal-rich magma reservoirs. Journal of Geophysical Research: Solid Earth, 119, 6308–6322.10.1002/2014JB011124Search in Google Scholar

Parmigiani, A., Faroughi, S.A., Huber, C., Bachmann, O., and Su, Y. (2016) Bubble accumulation and its role on the evolution of upper crustal magma reservoirs. Nature, 532, 492–495.10.1038/nature17401Search in Google Scholar PubMed

Paterson, S.R., Pignotta, G.S., Farris, D., Memeti, V., Miller, R.B., Vernon, R.H., and Žák, J. (2008) Is stoping a volumetrically significant pluton emplacement process?: Discussion. Geological Society of America Bulletin, 120, 1075–1079.10.1130/B26148.1Search in Google Scholar

Paulatto, M., Minshull, T.A., Baptie, B., Dean, S., Hammond, J.O.S., Henstock, T., Kenedi, C.L., Kiddle, E.J., Malin, P., Peirce, C., Ryan, G., Shalev, E., Sparks, R.S.J., and Voight, B. (2010) Upper crustal structure of an active volcano from refraction/reflection tomography, Montserrat, Lesser Antilles. Geophysical Journal International, 180, 685–696.10.1111/j.1365-246X.2009.04445.xSearch in Google Scholar

Pe-Piper, G., and Moulton, B. (2008) Magma evolution in the Pliocene-Pleistocene succession of Kos, South Aegean arc (Greece). Lithos, 106, 110–124.10.1016/j.lithos.2008.07.002Search in Google Scholar

Peccerillo, A., Barberio, M.R., Yirgu, G., Ayalew, D., Barbieri, M., and Wu, T.W. (2003) Relationships between Mafic and Peralkaline Silicic Magmatism in Continental Rift Settings: A petrological, geochemical and isotopic study of the Gedemsa Volcano, Central Ethiopian Rift. Journal of Petrology, 44, 2003–2032.10.1093/petrology/egg068Search in Google Scholar

Petford, N. (2009) Which effective viscosity? Mineralogical Magazine, 73, 167–191.10.1180/minmag.2009.073.2.167Search in Google Scholar

Petford, N., Atherton, M.P., and Halliday, A.N. (1996) Rapid magma production rates, underplating and remelting in the Andes: Isotopic evidence from northern-central Peru (9-11 °S). Journal of South American Earth Sciences, 9, 69–78.10.1016/0895-9811(96)00028-4Search in Google Scholar

Petford, N., Cruden, A.R., McCaffrey, K.J.W., and Vigneresse, J.-L. (2000) Granite magma formation, transport and emplacement in the Earth's crust. Nature, 408, 669–673.10.1038/35047000Search in Google Scholar

Phillips, J.C., and Woods, A.W. (2002) Suppression of large-scale magma mixing by melt-volatile separation. Earth and Planetary Science Letters, 204, 47–60.10.1016/S0012-821X(02)00978-0Search in Google Scholar

Pichavant, M., Costa, F., Burgisser, A., Scaillet, B., Martel, C., and Poussineau, S. (2007) Equilibration scales in silicic to intermediate magmas implications for experimental studies. Journal of Petrology, 1955–1972.10.1093/petrology/egm045Search in Google Scholar

Pistone, M., Caricchi, L., Ulmer, P., Burlini, L., Ardia, P., Reusser, E., Marone, F., and Arbaret, L. (2012) Deformation experiments of bubble- and crystal-bearing magmas: Rheological and microstructural analysis. Journal of Geophysical Research: Solid Earth, 117(B5), B05208.10.1029/2011JB008986Search in Google Scholar

Pistone, M., Arzilli, F., Dobson, K.J., Cordonnier, B., Reusser, E., Ulmer, P., Marone, F., Whittington, A.G., Mancini, L., Fife, J.L., and Blundy, J.D. (2015) Gas-driven filter pressing in magmas: Insights into in-situ melt segregation from crystal mushes. Geology, 43, 699–70210.1130/G36766.1Search in Google Scholar

Pitcher, W.S. (1987) Granites and yet more granites forty years on. Geologische Rundschau, 76, 51–79.10.1007/BF01820573Search in Google Scholar

Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D. (1985) Magmatism at a plate edge: The Peruvian Andes, 328 p. Blackie, Halstead Press, Glasgow.10.1007/978-1-4899-5820-4Search in Google Scholar

Plouff, D., and Pakiser, L.C. (1972) Gravity study in the San Juan Mountains, Colorado. U.S. Geological Survey Professional Paper, 800, B183–B190.Search in Google Scholar

Poli, S., and Schmidt, M.W. (1995) H2O transport and release in subduction zones— Experimental constraints on basaltic and andesitic systems. Journal of Geophysical Research-Solid Earth, 100(B11), 22,299–22,314.10.1029/95JB01570Search in Google Scholar

Putirka, K. (2008) Thermometers and barometers for volcanic systems. In K. Putirka and F. Tepley, Eds., Minerals, Inclusions and Volcanic Processes. Reviews in Mineralogy and Geochemistry, 69, 61–120.10.1515/9781501508486-004Search in Google Scholar

Putirka, K.D., Canchola, J., Rash, J., Smith, O., Torrez, G., Paterson, S.R., and Ducea, M.N. (2014) Pluton assembly and the genesis of granitic magmas: Insights from the GIC pluton in cross section, Sierra Nevada Batholith, California. American Mineralogist, 99, 1284–1303.10.2138/am.2014.4564Search in Google Scholar

Read, H.H. (1948) Granites and granites. Origine of granites. Geological Society of America Memoir, 28, 1–19.10.1130/MEM28-p1Search in Google Scholar

——— (1957) The Granite Controversy, 430 p. Thomas Murby and Co., London.Search in Google Scholar

Reid, M.R. (2008) How long to achieve supersize? Elements, 4, 23–28.10.2113/GSELEMENTS.4.1.23Search in Google Scholar

Reid, M.R., and Coath, C.D. (2000) In situ U-Pb ages of zircons from the Bishop Tuff: No evidence for long crystal residence times. Geology, 28, 443–446.10.1130/0091-7613(2000)28<443:ISUAOZ>2.0.CO;2Search in Google Scholar

Reid, M.R., Coath, C.D., Harrison, T.M., and McKeegan, K.D. (1997) Prolonged residence times for the youngest rhyolites associated with Long Valley Caldera; 230Th-238U ion microprobe dating of young zircons. Earth and Planetary Science Letters, 150, 27–39.10.1016/S0012-821X(97)00077-0Search in Google Scholar

Reiners, P.W., Nelson, B.K., and Ghiorso, M.S. (1995) Assimilation of felsic crust and its partial melt by basaltic magma: thermal limits and extens of crustal contamination. Geology, 23, 563–566.10.1130/0091-7613(1995)023<0563:AOFCBB>2.3.CO;2Search in Google Scholar

Reubi, O., and Blundy, J. (2009) A dearth of intermediate melts at subduction zone volcanoes and the petrogenesis of arc andesites. Nature, 461, 1269–1273.10.1038/nature08510Search in Google Scholar

Richter, F.M., Davis, A.M., DePaolo, D.J., and Watson, E.B. (2003) Isotope fractionation by chemical diffusion between molten basalt and rhyolite. Geochimica et Cosmochimica Acta, 67, 3905–3923.10.1016/S0016-7037(03)00174-1Search in Google Scholar

Rivera, T.A., Schmitz, M.D., Crowley, J.L., and Storey, M. (2014) Rapid magma evolution constrained by zircon petrochronology and 40Ar/39Ar sanidine ages for the Huckleberry Ridge Tuff, Yellowstone, U.S.A. Geology, 42, 643–646.10.1130/G35808.1Search in Google Scholar

Rudnick, R.L. (1995) Making continental crust. Nature, 378, 571–578.10.1038/378571a0Search in Google Scholar

Savage, P.S., Georg, R.B., Williams, H.M., Burton, K.W., and Halliday, A.N. (2011) Silicon isotope fractionation during magmatic differentiation. Geochimica et Cosmochimica Acta, 75, 6124–6139.10.1016/j.gca.2011.07.043Search in Google Scholar

Scaillet, B., and Evans, B.W. (1999) The 15 June 1991 eruption of Mount Pinatubo. I. Phase equilibria and pre-eruption P-T-f O2-f H2O conditions of the dacite magma. Journal of Petrology, 40, 381–411.10.1093/petroj/40.3.381Search in Google Scholar

Scaillet, B., Clemente, B., Evans, B.W., and Pichavant, M. (1998a) Redox control of sulfur degassing in silicic magmas. Journal of Geophysical Research, 103(B10), 23937–23949.10.1029/98JB02301Search in Google Scholar

Scaillet, B., Holtz, F., and Pichavant, M. (1998b) Phase equilibrium constraints on the viscosity of silicic magmas 1. Volcanic-plutonic comparison. Journal of Geophysical Research, 103(B11), 27257–27266.10.1029/98JB02469Search in Google Scholar

Schmitt, A.K., Lindsay, J.M., de Silva, S., and Trumbull, R.B. (2003) U-Pb zircon chronostratigraphy of early-Pliocene ignimbrites from La Pacana, north Chile: implications for the formation of stratified magma chambers. Journal of Volcanology and Geothermal Research, 120, 43–53.10.1016/S0377-0273(02)00359-1Search in Google Scholar

Schmitt, A.K., Stockli, D.F., Lindsay, J.M., Robertson, R., Lovera, O.M., and Kislitsyn, R. (2010a) Episodic growth and homogenization of plutonic roots in arc volcanoes from combined U–Th and (U–Th)/He zircon dating. Earth and Planetary Science Letters, 295, 91–103.10.1016/j.epsl.2010.03.028Search in Google Scholar

Schmitt, A.K., Wetzel, F., Cooper, K.M., Zou, H., and Worner, G. (2010b) Magmatic longevity of Laacher See Volcano (Eifel, Germany) indicated by U–Th dating of intrusive carbonatites. Journal of Petrology, 51, 1053–1085.10.1093/petrology/egq011Search in Google Scholar

Schoene, B., Schaltegger, U., Brack, P., Latkoczy, C., Stracke, A., and Gunther, D. (2012) Rates of magma differentiation and emplacement in a ballooning pluton recorded by U-Pb TIMS-TEA, Adamello batholith, Italy. Earth and Planetary Science Letters, 355–356, 162–173.10.1016/j.epsl.2012.08.019Search in Google Scholar

Seaman, S.J. (2000) Crystal clusters, feldspar glomerocrysts, and magma envelopes in the Atascosa Lookout Lava Flow, Southern Arizona, U.S.A.: Recorders of Magmatic Events. Journal of Petrology, 41, 693–716.10.1093/petrology/41.5.693Search in Google Scholar

Shane, P., Smith, V., and Nairn, I. (2005) High temperature rhyodacites of the 36 ka Hauparu pyroclastic eruption, Okataina Volcanic Centre, New Zealand: Change in a silicic magmatic system following caldera collapse. Journal of Volcanology and Geothermal Research, 147, 357–376.10.1016/j.jvolgeores.2005.04.015Search in Google Scholar

Shibano, Y., Namiki, A., and Sumita, I. (2012) Experiments on upward migration of a liquid-rich layer in a granular medium: Implications for a crystalline magma chamber. Geochemistry, Geophysics, Geosystems, 13, Q03007.10.1029/2011GC003994Search in Google Scholar

Shimizu, A., Sumino, H., Nagao, K., Notsu, K., and Mitropoulos, P. (2005) Variation in noble gas isotopic composition of gas samples from the Aegean arc, Greece. Journal of Volcanology and Geothermal Research, 140, 321–339.10.1016/j.jvolgeores.2004.08.016Search in Google Scholar

Shinohara, H. (2008) Excess degassing from volcanoes and its role on eruptive and intrusive activity. Reviews of Geophysics, 46, RG4005.10.1029/2007RG000244Search in Google Scholar

Shirley, D.N. (1986) Compaction in igneous cumulates. Journal of Geology, 94, 795–809.10.1086/629088Search in Google Scholar

Sillitoe, R.H. (2010) Porphyry copper systems. Economic Geology, 105, 3–41.10.2113/gsecongeo.105.1.3Search in Google Scholar

Simakin, A.G., and Bindeman, I.N. (2012) Remelting in caldera and rift environments and the genesis of hot, recycled, rhyolites. Earth and Planetary Science Letters, 337–338, 224–235.10.1016/j.epsl.2012.04.011Search in Google Scholar

Simkin, T. (1993) Terrestrial Volcanism in Space and Time. Annual Review of Earth and Planetary Sciences, 21, 427–452.10.1146/annurev.ea.21.050193.002235Search in Google Scholar

Simon, J.I., and Reid, M.R. (2005) The pace of rhyolite differentiation and storage in an “archetypical” silicic magma system, Long Valley, California. Earth and Planetary Science Letters, 235, 123–140.10.1016/j.epsl.2005.03.013Search in Google Scholar

Simon, J.I., Renne, P.R., and Mundil, R. (2008) Implications of pre-eruptive magmatic histories of zircons for U-Pb geochronology of silicic extrusions. Earth and Planetary Science Letters, 266, 182–194.10.1016/j.epsl.2007.11.014Search in Google Scholar

Sinton, J.M., and Detrick, R.S. (1992) Mid-Ocean Ridge Magma Chambers. Journal of Geophysical Research, 97(B1), 197–216.10.1029/91JB02508Search in Google Scholar

Sisson, T.W., and Bacon, C.R. (1999) Gas-driven filter pressing in magmas. Geology, 27, 613–616.10.1130/0091-7613(1999)027<0613:GDFPIM>2.3.CO;2Search in Google Scholar

Sliwinski, J.T., Bachmann, O., Ellis, B.S., Dávila-Harris, P., Nelson, B.K., and Dufek, J. (2015) Eruption of shallow crystal cumulates during caldera-forming events on Tenerife, Canary Islands. Journal of Petrology, 1–22, 10.1093/petrology/egv068Search in Google Scholar

Smith, R.L. (1960) Ash flows. Geological Society of America Bulletin, 71, 795–842.10.1130/0016-7606(1960)71[795:AF]2.0.CO;2Search in Google Scholar

——— (1979) Ash-flow magmatism. Geological Society of America Special Paper, 180, 5–25.10.1130/SPE180-p5Search in Google Scholar

Smith, R.L., and Bailey, R.A. (1968) Stratigraphy, structure, and volcanic evolution of the Jemez Mountains, New Mexico. Special Paper—Geological Society of America, 447–448.Search in Google Scholar

Soden, B.J., Wetherald, R.T., Stenchikov, G.L., and Robock, A. (2002) Global cooling after the eruption of Mount Pinatubo: A test of climate feedback by water vapor. Science, 296, 727–730.10.1126/science.296.5568.727Search in Google Scholar

Solano, J.M.S., Jackson, M.D., Sparks, R.S.J., Blundy, J.D., and Annen, C. (2012) Melt segregation in deep crustal hot zones: A mechanism for chemical differentiation, crustal assimilation and the formation of evolved magmas. Journal of Petrology, 53, 1999–2026.10.1093/petrology/egs041Search in Google Scholar

Sparks, R.S.J., and Marshall, L.A. (1986) Thermal and mechanical constraints on mixing between mafic and silicic magmas. Journal of Volcanology and Geothermal Research, 29, 99–124.10.1016/0377-0273(86)90041-7Search in Google Scholar

Sparks, R.S.J., Sigurdsson, H., and Wilson, L. (1977) Magma mixing: A mechanism for triggering explosive eruptions. Nature, 267, 315–318.10.1038/267315a0Search in Google Scholar

Sparks, R.S.J., Huppert, H.E., and Turner, J.S. (1984) The fluid dynamics of evolving magma chambers. Philosophical Transactions of the Royal Society of London, 310, 511–534.10.1098/rsta.1984.0006Search in Google Scholar

Spera, F.J., and Bohrson, W.A. (2001) Energy-constrained open-system magmatic processes I: General model and energy-constrained assimilation and fractional crystallization (EC-AFC) formulation. Journal of Petrology, 42, 999–1018.10.1093/petrology/42.5.999Search in Google Scholar

——— (2004) Open-system magma chamber evolution: An energy-constrained geochemical model incorporating the effects of concurrent eruption, recharge, variable assimilation and fractional crystallization (EC-RAXFC). Journal of Petrology, 45, 2459–2480.10.1093/petrology/egh072Search in Google Scholar

Spera, F.J., Borgia, A., Strimple, J., and Feigenson, M. (1988) Rheology of melts and magmatic suspensions: 1. Design and calibration of concentric cylinder viscometer with application to rhyolitic magma. Journal of Geophysical Research-Solid Earth, 93(B9), 10273–10294.10.1029/JB093iB09p10273Search in Google Scholar

Spera, F.J., Oldenburg, C.M., Christiensen, C., and Todesco, M. (1995) Simulations of convection with crystallization in the system KAlSi2O6-CaMgSi2O6: Implications for compositionally zoned magma bodies. American Mineralogist, 40, 1188–1207.10.2138/am-1995-11-1210Search in Google Scholar

Steck, L.K., Thurber, C.H., Fehler, M.C., Lutter, W.J., Robbert, P.M., Baldrige, W.S., Stafford, D.G., and Sessions, R. (1998) Crustal and mantle P wave velocity structure beneath Valles caldera, New Mexico: Results from the Jemez teleseismic tomography experiment. Journal of Geophysical Research, 103, 24301–24320.10.1029/98JB00750Search in Google Scholar

Storm, S., Shane, P., Schmitt, A., and Lindsay, J. (2012) Decoupled crystallization and eruption histories of the rhyolite magmatic system at Tarawera volcano revealed by zircon ages and growth rates. Contributions to Mineralogy and Petrology, 1–15.10.1007/s00410-011-0682-8Search in Google Scholar

Storm, S., Schmitt, A., Shane, P., and Lindsay, J. (2014) Zircon trace element chemistry at sub-micrometer resolution for Tarawera volcano, New Zealand, and implications for rhyolite magma evolution. Contributions to Mineralogy and Petrology, 167, 1–19.10.1007/s00410-014-1000-zSearch in Google Scholar

Streck, M.J. (2014) Evaluation of crystal mush extraction models to explain crystal-poor rhyolites. Journal of Volcanology and Geothermal Research, 284, 79–94.10.1016/j.jvolgeores.2014.07.005Search in Google Scholar

Su, Y., Huber, C., Bachmann, O., Zajacz, Z., Wright, H.M., and Vazquez, J.A. (2016) The role of crystallization-driven exsolution on the sulfur mass balance in volcanic arc magmas. Journal of Geophysical Research, in press.10.1002/2016JB013184Search in Google Scholar

Sutton, A.N., Blake, S., Wilson, C.J.N., and Charlier, B.L.A. (2000) Late Quaternary evolution of a hyperactive rhyolite magmatic system: Taupo volcanic center, New Zealand. Journal of the Geological Society, London, 157, 537–552.10.1144/jgs.157.3.537Search in Google Scholar

Szymanowski, D., Ellis, B.S., Bachmann, O., Guillong, M., and Phillips, W.M. (2015) Bridging basalts and rhyolites in the Yellowstone-Snake River Plain volcanic province: The elusive intermediate step. Earth and Planetary Science Letters, 415, 80–89.10.1016/j.epsl.2015.01.041Search in Google Scholar

Tanaka, H.K.M., Nakano, T., Takahashi, S., Yoshida, J., Takeo, M., Oikawa, J., Ohminato, T., Aoki, Y., Koyama, E., Tsuji, H., and Niwa, K. (2007) High resolution imaging in the inhomogeneous crust with cosmic-ray muon radiography: The density structure below the volcanic crater floor of Mt. Asama, Japan. Earth and Planetary Science Letters, 263, 104–113.10.1016/j.epsl.2007.09.001Search in Google Scholar

Tanaka, H.K.M., Kusagaya, T., and Shinohara, H. (2014) Radiographic visualization of magma dynamics in an erupting volcano. Nature Communications, 5.10.1038/ncomms4381Search in Google Scholar

Tappa, M.J., Coleman, D.S., Mills, R.D., and Samperton, K.M. (2011) The plutonic record of a silicic ignimbrite from the Latir volcanic field, New Mexico. Geochemistry, Geophysics, Geosystems, 12, Q10011.10.1029/2011GC003700Search in Google Scholar

Taylor, H.P. (1980) The effects of assimilation of country rocks by magmas on 18O/16O and 87Sr/86Sr systematics in igneous rocks. Earth and Planetary Science Letters, 47, 243–254.10.1016/0012-821X(80)90040-0Search in Google Scholar

Taylor, S.R., and McClennan, S.M. (1981) The composition and evolution of the continental crust: Rare earth element evidence from sedimentary rocks. Philosophical Transactions of the Royal Society, A, 301, 381–399.Search in Google Scholar

Thomas, N., Tait, S., and Koyaguchi, T. (1993) Mixing of stratified liquids by the motion of gas bubbles: application to magma mixing. Earth and Planetary Science Letters, 115, 161–175.10.1016/0012-821X(93)90220-4Search in Google Scholar

Thompson, A.B., Matile, L., and Ulmer, P. (2002) Some thermal constraints on crustal assimilation during fractionation of hydrous, mantle-derived magmas with examples from central alpine batholiths. Journal of Petrology, 43, 403–422.10.1093/petrology/43.3.403Search in Google Scholar

Thompson, G.M., Smith, I.E.M., and Malpas, J.G. (2001) Origin of oceanic phonolites by crystal fractionation and the problem of the Daly gap: An example from Rarotonga. Contributions to Mineralogy and Petrology, 142, 336–346.10.1007/s004100100294Search in Google Scholar

Thompson, R.N. (1972) Evidence for a chemical discontinuity near the basalt-andesite transition in many anorogenic volcanic suites. Nature, 236, 106–110.10.1038/236106a0Search in Google Scholar

Till, C.B., Vazquez, J.A., and Boyce, J.W. (2015) Months between rejuvenation and volcanic eruption at Yellowstone caldera, Wyoming. Geology, 43, 695–698.10.1130/G36862.1Search in Google Scholar

Truby, J.M., Mueller, S.P., Llewellin, E.W., and Mader, H.M. (2014) The rheology of three-phase suspensions at low bubble capillary number. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 471(2173).Search in Google Scholar

Turnbull, R., Weaver, S., Tulloch, A., Cole, J., Handler, M., and Ireland, T. (2010) Field and geochemical constraints on mafic-felsic interactions, and processes in high-level arc magma chambers: An example from the Halfmoon Pluton, New Zealand. Journal of Petrology, 51, 1477–1505.10.1093/petrology/egq026Search in Google Scholar

Turner, J.S., and Campbell, I.H. (1986) Convection and mixing in magma chambers. Earth Science Reviews, 23, 255–352.10.1016/0012-8252(86)90015-2Search in Google Scholar

Turner, S., and Costa, F. (2007) Measuring timescales of magmatic evolution. Elements, 3, 267–272.10.2113/gselements.3.4.267Search in Google Scholar

Turner, S., Sandiford, M., Reagan, M., Hawkesworth, C., and Hildreth, W. (2010) Origins of large-volume, compositionally zoned volcanic eruptions: New constraints from U-series isotopes and numerical thermal modeling for the 1912 Katmai-Novarupta eruption. Journal of Geophysical Research: Solid Earth, 115(B12).10.1029/2009JB007195Search in Google Scholar

Tuttle, O.F., and Bowen, N.L. (1958) Origin of granite in light of experimental studies in the system NaAlSi3O8-KAlSi3O8-SiO2. Geological Society of America Memoir, 74, 153 pp.10.1130/MEM74-p1Search in Google Scholar

Valley, J.W., Cavosie, A.J., Ushikubo, T., Reinhard, D.A., Lawrence, D.F., Larson, D.J., Clifton, P.H., Kelly, T.F., Wilde, S.A., Moser, D.E., and Spicuzza, M.J. (2014) Hadean age for a post-magma-ocean zircon confirmed by atom-probe tomography. Nature Geoscience, 7, 219–223.10.1038/ngeo2075Search in Google Scholar

Vazquez, J.A., and Reid, M.R. (2004) Probing the Accumulation history of the voluminous toba magma. Science, 305, 991–994.10.1126/science.1096994Search in Google Scholar PubMed

Vona, A., Romano, C., Dingwell, D.B., and Giordano, D. (2011) The rheology of crystal-bearing basaltic magmas from Stromboli and Etna. Geochimica et Cosmochimica Acta, 75, 3214–3236.10.1016/j.gca.2011.03.031Search in Google Scholar

Voshage, H., Hofmann, A.W., Mazzucchelli, M., Rivalenti, G., Sinigoi, S., Raczek, I., and Demarchi, G. (1990) Isotopic evidence from the Ivrea Zone for a hybrid lower crust formed by magmatic underplating. Nature, 347(6295), 731–736.Search in Google Scholar

Wager, L. (1962) Igneous cumulates from the 1902 eruption of Soufriere, St. Vincent. Bulletin of Volcanology, 24, 93–99.10.1007/BF02599333Search in Google Scholar

Wager, L.R., Brown, G.M., and Wadsworth, W.J. (1960) Types of igneous cumulates. Journal of Petrology, 1, 73–85.10.1093/petrology/1.1.73Search in Google Scholar

Waite, G.P., and Moran, S.C. (2009) VP Structure of Mount St. Helens, Washington, U.S.A., imaged with local earthquake tomography. Journal of Volcanology and Geothermal Research, 182, 113–122.10.1016/j.jvolgeores.2009.02.009Search in Google Scholar

Walker, B. Jr., Klemetti, E., Grunder, A., Dilles, J., Tepley, F., and Giles, D. (2013) Crystal reaming during the assembly, maturation, and waning of an eleven-million-year crustal magma cycle: thermobarometry of the Aucanquilcha Volcanic Cluster. Contributions to Mineralogy and Petrology, 165, 663–682.10.1007/s00410-012-0829-2Search in Google Scholar

Walker, B.A. Jr., Miller, C.F., Lowery, L.E., Wooden, J.L., and Miller, J.S. (2007) Geology and geochronology of the Spirit Mountain batholith, southern Nevada: implications for timescales and physical processes of batholith construction. Journal of Volcanology and Geothermal Research, 167, 239–262.10.1016/j.jvolgeores.2006.12.008Search in Google Scholar

Wallace, G.S., and Bergantz, G.W. (2002) Wavelet-based correlation (WBC) of zoned crystal populations and magma mixing. Earth and Planetary Science Letters, 202, 133–145.10.1016/S0012-821X(02)00762-8Search in Google Scholar

Wallace, P.J. (2001) Volcanic SO2 emissions and the abundance and distribution of exsolved gas in magma bodies. Journal of Volcanology and Geothermal Research, 108, 85–106.10.1016/S0377-0273(00)00279-1Search in Google Scholar

——— (2005) Volatiles in subduction zone magmas: concentrations and fluxes based on melt inclusion and volcanic gas data. Journal of Volcanology and Geothermal Research, 140, 217–240.10.1016/j.jvolgeores.2004.07.023Search in Google Scholar

Wallace, P.J., Anderson, A.T., and Davis, A.M. (1995) Quantification of pre-eruptive exsolved gas contents in silicic magmas. Nature, 377, 612–615.10.1038/377612a0Search in Google Scholar

Ward, K.M., Zandt, G., Beck, S.L., Christensen, D.H., and McFarlin, H. (2014) Seismic imaging of the magmatic underpinnings beneath the Altiplano-Puna volcanic complex from the joint inversion of surface wave dispersion and receiver functions. Earth and Planetary Science Letters, 404, 43–53.10.1016/j.epsl.2014.07.022Search in Google Scholar

Wark, D.A., Hildreth, W., Spear, F.S., Cherniak, D.J., and Watson, E.B. (2007) Pre-eruption recharge of the Bishop magma system. Geology, 35, 235–238.10.1130/G23316A.1Search in Google Scholar

Watkins, J., DePaolo, D., Huber, C., and Ryerson, F. (2009a) Liquid composition-dependence of isotope fractionation during diffusion in molten silicates. Geochimica et Cosmochimica Acta, 73, 7341–7359.10.1016/j.gca.2009.09.004Search in Google Scholar

——— (2009b) Liquid composition-dependence of calcium isotope fractionation during diffusion in molten silicates. Geochimica et Cosmochimica Acta, 73, 7341–7359.10.1016/j.gca.2009.09.004Search in Google Scholar

Watts, R.B., de Silva, S.L., Jimenez de Rios, G., and Croudace, I. (1999) Effusive eruption of viscous silicic magma triggered and driven by recharge: A case study of the Cerro Chascon-Runtu Jarita Dome Complex in Southwest Bolivia. Bulletin of Volcanology, 61, 241–264.10.1007/s004450050274Search in Google Scholar

Webster, J.D. (2004) The exsolution of magmatic hydrosaline chloride liquids. Chemical Geology, 210, 33–48.10.1016/j.chemgeo.2004.06.003Search in Google Scholar

Weiland, C., Steck, L.K., Dawson, P., and Korneev, V. (1995) Crustal structure under Long Valley caldera from nonlinear teleseismic travel time tomography using three-dimensional ray. Journal of Geophysical Research, 100, 20379–20390.10.1029/95JB01147Search in Google Scholar

White, S.M., Crisp, J.A., and Spera, F.A. (2006) Long-term volumetric eruption rates and magma budgets. Geochemistry, Geophysics, Geosystems, 7, 10.1029/2005GC00100210.1029/2005GC001002Search in Google Scholar

Whitney, J.A., and Stormer, J.C. Jr. (1985) Mineralogy, petrology, and magmatic conditions from the Fish Canyon Tuff, central San Juan volcanic field, Colorado. Journal of Petrology, 26, 726–762.10.1093/petrology/26.3.726Search in Google Scholar

Whittington, A.G., Hofmeister, A.M., and Nabelek, P.I. (2009) Temperature-dependent thermal diffusivity of the Earth's crust and implications for magmatism. Nature, 458, 319–321.10.1038/nature07818Search in Google Scholar PubMed

Wickham, S.M. (1987) The segregation and emplacement of granitic magmas. Journal of the Geological Society, London, 144, 281–297.10.1144/gsjgs.144.2.0281Search in Google Scholar

Wiebe, R.A., Blair, K.D., Hawkins, D.P., and Sabine, C.P. (2002) Mafic injections, in situ hybridization, and crystal accumulation in the Pyramid Peak granite, California. Geological Society of America Bulletin, 114, 909–920.10.1130/0016-7606(2002)114<0909:MIISHA>2.0.CO;2Search in Google Scholar

Wilcock, J., Longpré, M.-A., De Moor, J.M., Ross, J., and Zimmerer, M. (2010) Calderas bottom-to-top: An online seminar and field trip. Eos, Transactions American Geophysical Union, 91, 1–2.10.1029/2010EO010002Search in Google Scholar

Williams-Jones, A.E., and Heinrich, C.A. (2005) Vapor transport and the formation of magmatic-hydrothermal ore deposits. Economic Geology, 100, 1287–1312.10.2113/gsecongeo.100.7.1287Search in Google Scholar

Wilson, C.J.N. (1993) Stratigraphy, chronology, styles and dynamics of late Quaternary eruptions from Taupo Volcano, New Zealand. Philosophical Transactions–Royal Society of London, Physical Sciences and Engineering, 343, 205–306.10.1098/rsta.1993.0050Search in Google Scholar

Wilson, C.J.N., Houghton, B.F., McWilliams, M.O., Lanphere, M.A., Weaver, S.D., and Briggs, R.M. (1995) Volcanic and structural evolution of Taupo volcanic zone, New Zealand; a review. Journal of Volcanology and Geothermal Research, 68, 1–28.10.1016/0377-0273(95)00006-GSearch in Google Scholar

Wilson, L., and Head, J.W. (1994) Mars: Review and analysis of volcanic eruption theory and relationships to observed landforms. Reviews of Geophysics, 32, 221–263.10.1029/94RG01113Search in Google Scholar

Wolff, J.A., and Ramos, F.C. (2014) Processes in caldera-forming high-silica rhyolite magma: Rb/Sr and Pb isotope systematics of the otowi member of the Bandelier Tuff, Valles Caldera, New Mexico, U.S.A. Journal of Petrology, 55, 345–375.10.1093/petrology/egt070Search in Google Scholar

Wolff, J.A., and Storey, M. (1984) Zoning in highly alkaline magma bodies. Geological Magazine, 121, 563–575.10.1017/S0016756800030715Search in Google Scholar

Wolff, J.A., Worner, G., and Blake, S. (1990) Gradients in physical parameters in zoned felsic magma bodies: implications for evolution and eruptive withdrawal. Journal of Volcanology and Geothermal Research, 43, 37–55.10.1016/0377-0273(90)90043-FSearch in Google Scholar

Wolff, J.A., Ellis, B.S., Ramos, F.C., Starkel, W.A., Boroughs, S., Olin, P.H., and Bachmann, O. (2015) Remelting of cumulates as a process for producing chemical zoning in silicic tuffs: A comparison of cool, wet and hot, dry rhyolitic magma systems. Lithos, 236–237, 275–286.10.1016/j.lithos.2015.09.002Search in Google Scholar

Worner, G., and Schmincke, H.U. (1984) mineralogical and chemical zonation of the laacher see tephra Sequence (East Eifel, W. Germany). Journal of Petrology, 25, 805–835.10.1093/petrology/25.4.805Search in Google Scholar

Worner, G., and Wright, T.L. (1984) Evidence for magma mixing within the Laacher See magma chamber (East Eifel, Germany). Journal of Volcanology and Geothermal Research, 22, 301–327.10.1016/0377-0273(84)90007-6Search in Google Scholar

Wotzlaw, J.-F., Schaltegger, U., Frick, D.A., Dungan, M.A., Gerdes, A., and Gunther, D. (2013) Tracking the evolution of large-volume silicic magma reservoirs from assembly to supereruption. Geology, 41, 867–870.10.1130/G34366.1Search in Google Scholar

Wotzlaw, J.-F., Bindeman, I.N., Watts, K.E., Schmitt, A.K., Caricchi, L., and Schaltegger, U. (2014) Linking rapid magma reservoir assembly and eruption trigger mechanisms at evolved Yellowstone-type supervolcanoes. Geology, 10.1130/G35979.1Search in Google Scholar

Yoshinobu, A.S., and Barnes, C.G. (2008) Is stoping a volumetrically significant pluton emplacement process? Discussion. Geological Society of America Bulletin, 120, 1080–1081.10.1130/B26141.1Search in Google Scholar

Zajacz, Z., Halter, W.E., Pettke, T., and Guillong, M. (2008) Determination of fluid/ melt partition coefficients by LA-ICPMS analysis of co-existing fluid and silicate melt inclusions: Controls on element partitioning. Geochimica et Cosmochimica Acta, 72, 2169–2197.10.1016/j.gca.2008.01.034Search in Google Scholar

Zandomeneghi, D., Barclay, A., Almendros, J., Ibanez Godoy, J.M., Wilcock, W.S.D., and Ben-Zvi, T. (2009) Crustal structure of Deception Island volcano from P wave seismic tomography: Tectonic and volcanic implications. Journal of Geophysical Research, 114(B6), B06310.10.1029/2008JB006119Search in Google Scholar

Zandt, G., Leidig, M., Chmielowski, J., Baumont, D., and Yuan, X. (2003) Seismic detection and characterization of the Altiplano-Puna magma body, central Andes. Pure and Applied Geophysics, 160, 789–807.10.1007/978-3-0348-8010-7_14Search in Google Scholar

Zhang, Y., and Cherniak, D.J. (2010) Diffusion in minerals and melts: Theoretical background. Reviews in Mineralogy and Geochemistry, 72, 5–60.10.1515/9781501508394-003Search in Google Scholar

Zollo, A., Gasparini, P., Virieux, J., Biella, G., Boschi, E., Capuano, P., de Franco, R., Dell’Aversana, P., de Matteis, R., De Natale, G., and others. (1998) An image of Mt. Vesuvius obtained by 2D seismic tomography. Journal of Volcanology and Geothermal Research, 82, 161–173.10.1016/S0377-0273(97)00063-2Search in Google Scholar

Received: 2016-1-6
Accepted: 2016-4-14
Published Online: 2016-10-29
Published in Print: 2016-11-1

© 2016 by Walter de Gruyter Berlin/Boston

This work is licensed under the MSA License.

Downloaded on 1.5.2024 from https://www.degruyter.com/document/doi/10.2138/am-2016-5675/html
Scroll to top button