You are on page 1of 21

Bull Volcanol (2011) 73:1513–1533

DOI 10.1007/s00445-011-0525-5

RESEARCH ARTICLE

Heterogeneous pumice populations in the 2.08-Ma Cerro


Galán Ignimbrite: implications for magma recharge
and ascent preceding a large-volume silicic eruption
Heather M. N. Wright & Chris B. Folkes &
Raymond A. F. Cas & Katharine V. Cashman

Received: 16 May 2009 / Accepted: 14 June 2011 / Published online: 2 October 2011
# Springer-Verlag 2011

Abstract Triggering mechanisms of large silicic eruptions suggest that the two pumice populations are not related by
remain a critical unsolved problem. We address this simple fractionation. Trace element concentrations in
question for the ~2.08-Ma caldera-forming eruption of crystals mimic bulk variations between clast types, with
Cerro Galán volcano, Argentina, which produced distinct grey pumice containing elevated Ba, Cu, Pb, and Zn
pumice populations of two colors: grey (5%) and white concentrations in both bulk samples (average Cu, Pb, and
(95%) that we believe may hold clues to the onset of Zn concentrations are 27, 35, and 82 in grey pumice vs. 11,
eruptive activity. We demonstrate that the color variations 19, and 60 in white pumice) and biotite phenocrysts and
correspond to both textural and compositional variations white pumice showing elevated Li concentrations in biotite
between the clast types. Both pumice types have bulk and plagioclase phenocrysts. White and grey clasts are also
compositions of high-K, high-silica dacite to low-silica texturally distinct: White pumice clasts contain abundant
rhyolite, but there are sufficient compositional differences phenocrysts (44–57%), lack microlites, and have highly
(e.g., ~150 ppm lower Ba at equivalent SiO2 content and evolved groundmass glass compositions (76.4–79.6 wt.%
0.03 wt.% higher TiO2 in white pumice than grey) to SiO2), whereas grey pumice clasts contain a lower
percentage of phenocrysts/microphenocrysts (35–49%),
have abundant microlites, and have less evolved ground-
Editorial responsibility: J. Stix mass glass compositions (69.4–73.8 wt.% SiO2). There is
This paper constitutes part of a special issue: also evidence for crystal transfer between magma produc-
ing white and grey pumice. Thin highly evolved melt rims
Cas RAF, Cashman K (eds) The Cerro Galan Ignimbrite and Caldera: surround some fragmental crystals in grey pumice clasts
characteristics and origins of a very large-volume ignimbrite and its
magma system.
and appear to have come from magma that produced white
pumice. Furthermore, based on crystal compositions, white
Electronic supplementary material The online version of this article bands within banded pumice contain crystals originating in
(doi:10.1007/s00445-011-0525-5) contains supplementary material,
which is available to authorized users. grey magma. Finally, only grey pumice clasts form
breadcrusted surface textures. We interpret these composi-
H. M. N. Wright : C. B. Folkes : R. A. F. Cas
School of Geosciences, Monash University,
tional and textural variations to indicate distinct magma
Clayton, VIC 3800, Australia batches, where grey pumice originated from an originally
deeper, more volatile-rich dacite recharge magma that
K. V. Cashman ascended through and mingled with the volumetrically
Department of Geological Sciences, University of Oregon,
Eugene, OR 97403-1272, USA
dominant, more highly crystalline chamber that produced
white pumice. Shortly before eruption, the grey pumice
Present Address: magma stalled within shallow fractures, forming a van-
H. M. N. Wright (*) guard magma phase whose ascent may have provided a
U.S. Geological Survey,
345 Middlefield Rd,
trigger for eruption of the highly crystalline rhyodacite
Menlo Park, CA 94025, USA magma. We suggest that in the case of the Cerro Galán
e-mail: hwright@usgs.gov eruption, grey pumice provides evidence not only for
1514 Bull Volcanol (2011) 73:1513–1533

cryptic silicic recharge in a large caldera system but also a fragments in the grey pumice (Pallister et al. 1996; Rosi et
probable trigger for the eruption. al. 2004) that are interpreted to result from high strain rates
along conduit margins during periods of rapid magma
Keywords Caldera . Pumice textures . Shear . Phenocryst ascent (Polacci et al. 2001; Rosi et al. 2004). Because clast
fragments . Magma ascent . Silicic recharge color can be attributed to different processes, it is important
to identify variations between clast populations beyond
simple color observations.
Introduction The most recent, climactic eruption of Cerro Galán
caldera, Argentina, at 2.08±0.02 Ma (average 40Ar/39Ar
The eruptibility of magma is a fundamental issue in sanidine age from samples north and west of the caldera,
volcanology. The problem is particularly acute in the case Kay et al. 2011), produced both dominant coarsely
of high crystallinity magma, where magma ascent is crystalline and coarsely vesicular white pumice and
expected to be very slow, if not completely inhibited. Such proportionally subordinate finely crystalline and finely
is the case for many large-volume, so-called monotonous vesicular grey pumice. Here we relate chemical and textural
intermediate ignimbrites (Hildreth 1981), for which the characteristics of pumice to conditions of magma storage
majority of the erupted material is phenocryst-rich and the and ascent that precede a cataclysmic, caldera-forming
bulk compositions are restricted to a narrow range between eruption. We then consider these results in the context of
63 and 71 wt.% SiO2 (e.g., Fish Canyon Tuff, Lipman et al. similar textural variations in other caldera-forming pyro-
1970; Atana ignimbrite, Lindsay et al. 2000). One clastic deposits.
suggested mechanism for magma chamber destabilization
(and eruption triggering) in high crystallinity deposits is
recharge by less crystalline, less evolved, and less viscous Cerro Galán Ignimbrite
magma into the larger magma chamber, particularly when
such an intrusion releases volatiles into the overlying Cerro Galán is located in the Central Volcanic Zone of
chamber, causing convective mixing between magma northwestern Argentina (Fig. 1). The Cerro Galán volcanic
bodies and/or transfers heat into the overlying chamber system has erupted a sequence of nine major ignimbrites
(Pallister et al. 1992; Venezky and Rutherford 1997; since ~6 Ma (Folkes et al. 2011b). In addition, small
Bachmann et al. 2002). However, while some eruptions volume dacitic eruptions within the caldera produced
may contain clear evidence of mafic recharge (mafic domes, block and ash flows, and pyroclastic flows
enclaves, compositionally distinct pyroclast populations), throughout this interval. The climactic eruption of the
such evidence is not universal. An alternative suggestion is volcanic sequence produced the Cerro Galán Ignimbrite
that the recharge (and triggering) magma is silicic, although (CGI) at ~2.08±0.02 Ma. The large deposit volume,
in this case the evidence of recharge is often cryptic, ~630 km3 (dense-rock equivalent; Folkes et al. 2011b),
involving subtle textural variations within eruptive products makes this eruption one of the largest in the world (Mason
(e.g., Eichelberger et al. 2000, 2006; Smith et al. 2004, et al. 2004). Trap-door style collapse occurred along a
2009; Hildreth and Wilson 2007; Kennedy and Stix 2007; rectilinear set of faults, producing a piecemeal collapse
Wiebe et al. 2007; de Silva et al. 2008). block ~27×16 km in dimension. The CGI is generally
The most commonly observed textural variation between pumice-poor, lithic-poor (average 10% pumice and 5%
pumice clasts is clast color, which generally varies from lithics, based on field approximation), and crystal-rich. The
grey to white within a given eruption. In some cases, clast pyroclastic flow deposit lacks an associated precursory fall
color corresponds with bulk composition, particularly when deposit, forming a single depositional unit through much of
the compositional difference between the resident and its areal distribution, with the exception of distal deposi-
recharge magma is large (e.g., andesite to rhyodacite, tional localities where multiple flow units are observed,
Bacon and Druitt 1988; tephriphonolite to phonolite, Cioni separated by thin pyroclastic surge layers (Cas et al. 2011).
et al. 1995). In contrast, pumice color may not correlate We have identified three distinct juvenile clast types:
with bulk composition but instead reflect variations in syn- white pumice, grey pumice (which includes grey/white
eruptive rates of decompression and degassing. For exam- banded varieties), and crystal-rich juvenile clasts. Propor-
ple, at Mount St Helens volcano, clast color varies with tionally minor, crystal-rich dense juvenile clasts represent
vesicle size and the presence or absence of microlites, small cognate material, inferred to have formed part of a crystal
crystals that formed during ascent in response to decom- mush (Folkes et al. 2011a) and will not be discussed here.
pression and degassing (Klug and Cashman 1994). Alter- We focus instead on differences between white and grey
natively, at Pinatubo and Quilotoa volcanoes, color pumice. Grey pumice commonly contains microscopic
differences record the presence of abundant, small crystal bands of white pumice but can be macroscopically banded
Bull Volcanol (2011) 73:1513–1533 1515

Fig. 1 Location map of Cerro


Galán caldera, Argentina. Loca-
tions of samples used in this
study are shown with small
circles. Topographic caldera rim
is shown with a solid line;
inferred caldera-bounding faults
are shown with dashed lines

(Fig. 2). While white pumice clasts are sometimes found a slightly higher percentage of large-pumice clasts than of
within grey pumice, the reverse is not observed. Further- small clasts, and in rare large-pumice concentration
more, the only clasts to exhibit breadcrusted surfaces are horizons, proportions of grey pumice are as high as 75%.
grey and banded clasts. Small plagioclase-phyric dacite In some distal localities, these pumice concentration
lithic clasts, inferred to originate from shallow wall rock, horizons form the tops of topography-induced depositional
are present in many grey pumice clasts. Grey pumice forms subunits.
~5% of the total pumice population, varying between 1% Using these macroscopic textural variations as a basis for
and 10% at any given locality. However, grey pumice forms classification, we collected samples of white and grey
1516 Bull Volcanol (2011) 73:1513–1533

Fig. 2 Photographs of pumice textures. a White (left) and banded (right) pumice textures in outcrop. b Breadcrusted grey pumice clasts. c Grey
pumice clast containing white pumice core

pumice (including banded, breadcrusted, and white pumice- eliminate this possibility in all cases. 29Si was used as an
cored samples) for detailed textural and chemical study. We internal standard in conjunction with SiO2 contents mea-
use the coupled textural and chemical variations to better sured by electron microprobe. The trace element composi-
constrain interpretation of the eruption trigger and ascent tions of individual phenocrysts were analyzed in one white,
history of the CGI magma. one grey, and two grey clasts with white cores, with a single
analysis performed on each crystal. Modal proportions of
crystal phases and groundmass glass were determined by
Methods point counting crystals on petrographic thin sections. A
minimum of 500 counts (usually 1,000 counts) were made
Pumice samples from each textural group were collected for on a total of 24 white and 13 grey samples. Cathodolumi-
chemical and textural analyses from sample localities nescence (CL) images of quartz crystals were taken using
distributed both around the caldera (Fig. 1) and at a range an external Gatan cathodoluminescence detector attached to
of stratigraphic levels within the ignimbrite. To determine the JEOL scanning electron microscope (SEM) at the
chemical variations between pumice clasts, we analyzed University of Melbourne, using a 30-keV beam with a spot
bulk pumice, groundmass glass, and individual crystal size of 8. The maximum imaging aperture available was
compositions. Bulk compositions were determined at ~1 mm, and where grains were >1 mm, multiple images
Washington State University using X-ray fluorescence were stitched together.
(XRF) and inductively coupled plasma mass spectrometry To determine variations in vesicle textures, we analyzed
(ICP-MS) of 22 white and 13 grey crushed pumice samples porosities (vesicularities), permeabilities, and vesicle tex-
using the protocol of Johnson et al. (1999) and Knaack et tures of selected clasts. Connected porosities were deter-
al. (1994), respectively. Groundmass glass and mineral mined using cores drilled from individual pumice clasts.
compositions were analyzed in 13 samples using the Clasts were cored with a 2.5-cm-diameter drill bit, cut to
Cameca SX100 electron microprobe at the University of 2.5 cm length, weighed, and core volume was measured.
Oregon. A 12-keV electron beam at 20 nA with a 2–10-μm- Connected porosity (ϕc) was determined by comparing the
diameter defocused beam was used to analyze groundmass He-accessible volume of the core with the bulk core
glass compositions, whereas a focused beam was used for volume, measured using a micromeritic multi-volume He-
feldspar, biotite, and oxide crystal analyses. Volatile pycnometer at the University of Oregon (e.g., Wright et al.
migration through glass due to sample bombardment (and 2007). At least three measurements were made per sample;
heating) by the electron beam was corrected using an resultant precision is ±0.2% (1σ) of porosity value. Darcian
extrapolation method in the software package Probe for permeability was measured using a top-down permeameter
Windows (Donovan 2007). Trace element concentrations in at the University of Oregon according to Rust and Cashman
individual minerals were analyzed by LA-ICP-MS (laser (2004). At least two measurements were made per sample.
ablation ICP-MS) in the W.M. Keck Collaboratory for Qualitative characterization of vesicle shapes, sizes,
Plasma Spectrometry, Oregon State University using a and crystal orientations was made using backscatter
NewWave DUV 193-mm ArF Excimer laser and VG PQ electron (BSE) images and crossed polarized light
ExCell Quadrupole ICP-MS. Plagioclase and biotite crys- (XPL) images of polished thin sections of pumice. BSE
tals were ablated for 45 s using a 40-μm spot size and 4-Hz images were obtained on the FEI Quanta 200 FEG SEM
pulse rate, following the methods of Kent et al. (2004). at the University of Oregon using a 10-keV electron
Care was taken to avoid intersecting inclusions in the beam at 5–10 nA sample current and a 10-mm working
analyses by visual inspection, although we cannot entirely distance and on the JEOL JSM 840A SEM at the Centre
Bull Volcanol (2011) 73:1513–1533 1517

for Electron Microscopy at Monash University using a Ba concentrations in grey pumice are ~150 ppm higher at
15-keV electron beam at 3–10 nA current and a 15-mm equal SiO2 content than in white pumice (Fig. 3c).
working distance on carbon-coated thin sections. Similarly, slightly higher Ti contents (0.03 wt.% TiO2) in
To determine water and CO2 concentrations in melt white pumice than grey produces an offset between the two
inclusions trapped in quartz crystals in white pumice, pumice types in a plot of FeO vs. TiO2 (Fig. 3d). Offset
inclusions were doubly intersected and polished for Fourier trends cannot be explained simply by differences in degree
transform infrared spectroscopic (FTIR) analyses. Only of crystal fractionation of a single fractionating assemblage,
white pumice samples were analyzed because of a general as illustrated by the constant Sr/Ba ratio with changing Ba
paucity of intact melt inclusions within grey pumice. In in the grey pumice, which contrasts with the Sr/Ba
addition, wafers of matrix glass from pumice clast margins fractionation signal preserved in the white pumice samples
and a breadcrust bomb rind were doubly polished according (Fig. 3b).
to the method of Wright et al. (2007). Transmission spectra
were obtained using a Thermo Nicolet Nexus 670 FTIR Crystal content, glass composition, and mineral
spectrometer interfaced with a continuum IR microscope at composition
the University of Oregon. Absorptions (peak heights) were
measured by subtracting background fits using straight line Modal proportions of phenocryst phases also vary
fits and were used to determine species concentration between pumice types (Table 2). Both white and grey
through Beer’s law (e.g., Wallace et al. 1999). We used pumice contain plagioclase, biotite, quartz, sanidine,
εCO2 =1,214 l/mol cm based on measurements of synthe- apatite, ilmenite, and magnetite, with trace amounts of
sized rhyolitic samples (Behrens et al. 2004). We iteratively zircon, monazite, and titanite. Differences between the
calculated εH2O =73 l/mol cm by minimizing error to a pumice types include the presence of trace hornblende
linear least-squares fit between cH2O from the 3,550-cm−1 (small and variably broken down) in some grey pumice
peak and combined water concentration from individual samples and textural contrasts. One obvious textural
species peaks at 5,230 and 4,520 cm−1 using the calibration difference is crystal size: The maximum dimensions of
of Zhang et al. (1997). This value is consistent with the feldspar and quartz are <8 mm in the white pumice as
range of values calculated by Newman et al. (1986). Based opposed to <1.6 mm in the grey. Similarly, biotite crystals
on a comparison between 5,230 and 4,520 cm−1 calculated are <4 mm in white pumice but <1.2 mm in the grey
water contents and 3,550 cm−1 water contents and pumice. In general, the size and proportion of crystals in
including error in thickness measurements, we calculate a grey pumice is less than in grey bands mingled with white
maximum error of ±0.4 wt.% H2O and ±10 ppm CO2 for pumice (Fig. 4). However, grey pumice clasts are not
melt inclusion data. Matrix glass water and CO2 concen- homogeneous; local bands with higher and lower crystal-
trations are calculated using the same absorption peaks and linities are present in some grey clasts. Furthermore, the
absorption coefficients as melt inclusions but are analyzed range in point-counted crystallinities between clasts is
using a wider aperture (>100 μm) and corrected for average large within both pumice types (from 35% to 49% in grey
vesicularity, as in Wright et al. (2007). At least three sets of pumice; 44% to 57% in white pumice; Fig. 5).
measurements on different glassy wafers were made per The most striking feature of grey pumice clasts is the
sample. Standard deviation was calculated between average pervasive fracturing of all microphenocrysts and rare
values for each glass wafer. phenocrysts (Fig. 6). Fractures form margin-parallel,
length-parallel, and cross-hatched sets (Fig. 6a–c). Highly
fractured quartz and feldspar crystals are also present but
Results not common in white pumice clasts; crystal fractures are
generally more widely spaced than in grey clasts (Fig. 6d).
Bulk composition Grey pumice clasts also contain microlites (crystals <30 μm
in size) and tiny (<1 μm) Fe–Ti oxides, whereas white
White and grey pumice populations are geochemically pumice clasts do not. Figure 7 shows biotite, plagioclase,
similar. Both pumice types are high-K high-silica dacites to and quartz microlites in the glassy groundmass of grey
low-silica rhyolites (hereafter called rhyodacites) with pumice. Plagioclase microlite shapes vary from angular,
broadly similar bulk compositions (Table 1; Sup. Table 1; commonly triangular 2-D shapes grouped in clusters to
Fig. 3a). Compositional variation between pumice clast irregular, multi-pointed, and rounded varieties. Quartz
types does not correlate with geographic or stratigraphic microlites are rounded and commonly surrounded by cracks
level in the ignimbrite. Variations of some minor and trace in the groundmass glass. BSE images show that ground-
element abundances demonstrate, however, that the two mass glass in the grey pumice is inhomogeneous and often
pumice populations are chemically distinct. For example, complexly mingled; thin (<5 μm) rims of evolved (white
Table 1 Selected bulk compositions of white and grey pumice; SiO2 to Cu analyses are XRF results; Nd to Eu analyses are ICPMS results
1518

White Grey

Sample number CG257 CG267 CG338 CG142a CG121a CG287 CG163 CG140 CG218 CG337 CG216 CG219 CG327 CG125 CG256 CG1a2 CG122

SiO2 69.07 69.07 69.25 68.99 69.50 69.50 70.24 69.87 70.38 70.12 69.52 68.32 69.53 70.04 70.13 69.87 69.93
TiO2 0.62 0.59 0.61 0.63 0.56 0.64 0.63 0.61 0.59 0.52 0.60 0.57 0.58 0.54 0.57 0.56 0.56
Al2O3 15.45 15.33 15.14 15.30 14.92 15.23 14.61 14.82 14.65 15.09 15.32 15.86 15.23 14.86 15.59 15.01 14.95
FeOT 3.05 2.95 3.04 3.14 2.75 3.20 3.10 3.03 2.92 2.66 3.06 2.80 3.04 2.80 2.93 2.84 2.97
MnO 0.05 0.05 0.05 0.05 0.05 0.05 0.05 0.05 0.05 0.04 0.04 0.05 0.04 0.05 0.05 0.07 0.05
MgO 1.26 1.23 1.19 1.19 1.14 1.14 1.27 1.19 1.16 0.99 0.69 1.08 0.77 1.11 1.06 1.06 1.17
CaO 2.59 2.81 2.67 2.73 3.18 2.56 2.66 2.79 2.56 2.40 2.61 2.91 2.63 2.66 2.17 2.56 2.49
Na2O 3.13 2.73 3.28 3.46 3.39 3.18 3.05 3.42 3.31 3.26 3.51 3.62 3.44 3.14 2.79 3.41 3.24
K2O 4.58 5.02 4.56 4.27 4.29 4.19 4.17 3.98 4.16 4.74 4.45 4.57 4.50 4.62 4.50 4.43 4.45
P2O5 0.20 0.21 0.21 0.22 0.22 0.29 0.22 0.22 0.21 0.18 0.21 0.23 0.24 0.20 0.21 0.19 0.19
Raw sum (XRF) 97.16 96.07 97.76 96.81 96.76 97.88 97.06 97.35 97.39 97.63 98.25 98.39 98.48 96.79 96.11 97.15 97.14
Ni 0.7 1.6 0.7 8.1 7.5 4.0 b.d. 7.1 b.d. 1.3 b.d. b.d. 1.1 4.6 b.d. 7.0 8.5
Cr 15.6 12.5 14.4 15.4 13.3 15.4 15.2 14.0 13.6 13.0 43.8 13.8 25.9 14.3 13.5 13.1 15.0
V 68 60 68 67 58 74 64 65 63 59 62 63 64 56 60 57 61
Ba 511 459 389 383 362 358 265 282 252 485 547 604 516 504 494 488 516
Rb 258 303 292 260 259 261 256 258 323 296 265 247 260 259 261 263 255
Sr 294 311 274 280 265 275 249 266 244 268 295 328 297 273 268 277 266
Zn 64 56 64 65 61 58 62 63 59 50 63 64 65 93 79 132 133
Ga 24 22 23 23 22 23 24 23 22 22 23 23 21 21 22 22 22
Cu 16.9 5.5 22.6 8.6 9.7 22.4 6.9 7.0 15.7 22.5 9.7 11.7 23.1 25.3 26.5 21.6 39.5
La 49 49 51 56 42 59 44 43 44 48 49 42 54 46 46 45 43
Ce 99 98 103 111 84 118 89 88 89 96 97 84 107 92 92 90 86
Pr 11 11 12 13 10 13 10 10 10 11 11 10 12 11 11 11 10
Nd 41 41 42 45 36 47 39 38 38 39 40 35 43 39 39 38 36
Sm 7.8 7.6 8.1 8.2 7.2 8.6 7.6 7.4 7.9 7.5 7.6 7.0 8.1 7.6 7.7 7.4 7.1
Eu 1.3 1.3 1.3 1.3 1.1 1.3 1.2 1.2 1.2 1.2 1.3 1.3 1.3 1.2 1.3 1.2 1.2
Gd 5.7 5.5 6.0 6.0 5.4 6.2 5.8 5.8 6.0 5.4 5.7 5.3 6.0 5.8 5.8 5.6 5.4
Tb 0.77 0.77 0.82 0.82 0.78 0.86 0.82 0.79 0.86 0.75 0.79 0.76 0.81 0.82 0.82 0.78 0.75
Dy 4.0 4.0 4.3 4.2 4.1 4.3 4.2 4.1 4.5 3.8 4.1 3.9 4.1 4.3 4.3 4.1 4.0
Ho 0.69 0.72 0.73 0.72 0.70 0.76 0.74 0.72 0.78 0.66 0.70 0.71 0.72 0.75 0.76 0.74 0.69
Er 1.7 1.8 1.8 1.8 1.8 1.9 1.9 1.8 2.0 1.7 1.8 1.8 1.8 1.9 1.9 1.9 1.7
Tm 0.24 0.25 0.26 0.25 0.25 0.26 0.27 0.24 0.28 0.23 0.25 0.25 0.24 0.27 0.27 0.27 0.24
Yb 1.5 1.5 1.6 1.5 1.5 1.6 1.6 1.5 1.7 1.4 1.5 1.5 1.5 1.7 1.6 1.6 1.5
Lu 0.21 0.23 0.23 0.23 0.23 0.24 0.24 0.23 0.26 0.22 0.23 0.23 0.22 0.26 0.24 0.24 0.22
Th 31 29 31 31 29 34 31 31 31 30 29 25 31 29 30 29 29
Nb 18 18 20 20 19 21 21 20 21 19 19 18 19 19 19 18 19
Y 18 19 20 19 19 20 20 19 21 18 19 19 19 21 20 20 18
Bull Volcanol (2011) 73:1513–1533
Bull Volcanol (2011) 73:1513–1533 1519

Major elements are normalized to 100% on a volatile-free basis. T Total Fe is expressed as FeO. All oxides normalized to 100% on a volatile-free basis. Eu* is the ratio between chondrite
normalized Eu and the geometric mean between chondrite normalized Sm and Gd (normalized to chondrite of Anders and Ebihara 1982). Oxides are listed in weight percent and trace elements in
CG122

0.59
5.1
2.1

6.5
pumice) glass form veneers around some plagioclase and

11
35
30
170
quartz microlites, demonstrating that mingling included
CG1a2
exchange of crystals (Fig. 8).

0.56
5.0
1.8

6.5
As anticipated from BSE imaging, groundmass glass

171
8
32
24
compositions differ for the two clast types, with white pumice
glass compositions more evolved (more silica-rich and Fe-
CG256

0.56
5.6
2.2

6.4
13 poor) than those of grey pumice (Fig. 9). Among white pumice
45
29
185
and white portions of banded or mingled pumice clasts, the
groundmass glass composition ranges from 76.4 to 79.6 wt.%
CG125

0.54
5.5
2.2

6.8
11
34
18
179
SiO2. Within single clasts, heterogeneity in glass compositions
includes up to 1.8 wt.% variation in SiO2 (e.g., CG405;
CG327

0.55
Table 3). Glass compositions in grey pumice are more varied
5.2
2.2

6.4
27
14
173
11

and less evolved than in white pumice (69.4–73.8 wt.%


SiO2), except in sparse, thin high-silica bands in grey pumice
CG219

0.64

(77–78 wt.% SiO2). We interpret the latter to be microscopic


5.5
2.1

6.6
10
31
12
183

bands of white pumice within grey pumice (Table 3).


Plagioclase crystals in grey and white pumice clasts have
CG216

0.59
5.7
2.1

6.7

overlapping compositions, although grey pumice clasts


11
27
17
188

contain a more restricted compositional range than white


pumice clasts. Plagioclase core compositions in white
CG337

0.57

pumice are variable (An29–An69), whereas rims are homo-


5.2
2.2

5.5
Grey

12
27
32
169

geneous (An31–An39). In contrast, plagioclase core compo-


sitions of An34–An50 in the grey pumice overlap with
CG218

0.49
5.6
2.6

6.8

compositions of rims and microlites (Fig. 10; Table 4).


9
14

29
180

Preliminary analyses of trace elements in plagioclase for


one white, two grey and one grey clast with a white core
CG140

0.54
5.1
2.3

6.7
12
23
27
169

suggest that trace element concentrations are also broadly


similar in the two pumice types with the exception of Li,
which is substantially higher in white pumice (maximum=
CG163

0.52
5.8
2.4

7.3

62 μg/g) than in grey (maximum=12 μg/g; Fig. 11a;


13
19
20
192

Table 4).
CG287

0.50

Biotite compositions also differ between grey and white


5.9
2.4

7.4
11
24
17
196

pumice clasts. Although we have analyzed only four


samples (one grey, one white, and two grey clasts with
CG121a

0.54

white cores), there is a small contrast in both average Fe/


5.2
2.4

6.3
12
23
19
172

(Fe+Mg) (= 0.43, 0.40, 0.41, and 0.41) and average Ti


content per formula unit (= 0.50, 0.48, 0.50, and 0.49) in
CG142a

white, grey, and white and grey portions of mingled clasts,


0.54
5.8
2.3

6.9
11
23
19
192

respectively (Fig. 11b). Trace element concentration varia-


tions in biotite crystals parallel bulk compositional varia-
CG338

0.53

tions; grey pumice biotites have higher Ba, Pb, and Cu


5.6
2.4

6.4
12
27
27
185

concentrations than in white pumice (average Ba, Pb, and


Cu concentrations, respectively—891, 18, and 60 μg/g in
grey pumice vs. 418, 6, and 15 μg/g in white; Fig. 11c).
CG267

0.60
5.5
2.0

6.4
10
17
105
184

Additionally, white pumice biotites have higher Li concen-


trations than do grey pumice (maximum 713 μg/g in white
b.d. below detection limits
CG257

0.56

pumice vs. 129 μg/g in grey pumice; Fig. 11d; Table 4).
White

5.7
2.0

6.4
10
23
23
192
Table 1 (continued)

Both white and grey pumice clasts contain co-existing


magnetite and ilmenite crystals. The sizes (80–320 μm) and
parts per million
Sample number

compositions of analyzed oxides are similar for the two


Eu/Eu*

pumice types. If we assume that neighboring oxide crystals


Pb
Cs
Hf

Sc
Ta

Zr
U

(touching or in close proximity) are in equilibrium, using


the criteria of Bacon and Metz 1984 and Bacon and
1520 Bull Volcanol (2011) 73:1513–1533

Fig. 3 Bulk compositions of


grey and white pumice samples
(open symbols are white pumice;
closed symbols are grey and
banded pumice). Grey and white
pumice samples contain over-
lapping bulk concentrations of a
Na2O+K2O vs. SiO2 (composi-
tional fields as in Le Bas et al.
1992), but form distinct trends
in b Sr/Ba vs. Ba, c Ba vs. SiO2,
and d FeOT vs. TiO2; e Pb, Cu,
and Zn concentrations are higher
in grey pumice than white pum-
ice samples; f Eu anomaly var-
iation with Ba concentration; g
Ba vs. K2O—note covariation in
white pumice, but not in grey
pumice population concentra-
tions; and h Y/Ba vs. TiO2
values of grey pumice are more
geochemically similar to older
Toconquis ignimbrite units than
to white pumice clasts. Error
bars show maximum relative
error calculated from repeat
bead analyses; where error bars
are not shown, they are smaller
than the symbol size; error is
generally less than 3% (relative)
of major element and 10% (rel-
ative) of trace element concen-
trations; error is propagated for
ratios and sums

Hirschmann 1988, application of the geothermometer of 2011a). At high oxygen fugacity, calculated temperatures
Andersen and Lindsley (1988) gives crystallization temper- using the Ghiorso and Evans (2008) model can be in error
atures of 796–806°C and 796–809°C and log10 oxygen by as much as 100°C, due to analytical error. Therefore,
fugacities of −11.8 and −11.7 for white and grey pumice, actual temperatures likely like somewhere between the
respectively (Fig. 12; Table 4). However, at these two estimates, with some overlap or slight offset between
relatively oxidizing conditions (~NNO +1.5 log units), white and grey pumice (where grey pumice is slightly
these temperatures are probably overestimates of the hotter); we suggest that eruption temperature was <780°C.
actual temperature (cf. Pinatubo temperature estimates,
Geschwind and Rutherford 2002). Using the calibration of Vesicle content
Ghiorso and Evans (2008), calculated temperatures are
much lower—682–698°C and 705–709°C, for white and Bulk porosity (vesicularity) is highly varied but mostly
grey pumice, respectively, at oxygen fugacities >1.5 log overlaps in the two pumice types. For a solid density of
units above the nickel–nickel oxide buffer (cf. Folkes et al. 2,700 kg/m3, the bulk porosity of grey pumice ranges from
Bull Volcanol (2011) 73:1513–1533 1521

Table 2 Modal phase percentages of grey and white pumice clasts

Grey White

Min Max Ave Min Max Ave

Quartz 7.4 15.3 10.3 5.5 13.3 8.7


Plag 13.8 21.8 17.5 13.4 26.1 20.6
Biotite 6.5 12.3 9.3 10.7 15.6 13.2
Oxides 0.5 6.6 2.4 0.0 7.3 3.2
Sanidine 0.1 1.5 0.9 0.8 2.2 1.5
Apatite – 2.8 0.9 0.1 2.8 1.5
Hornblende – 0.20 0.03 – – –
Accessory minerals 0.5 1.1 0.7 0.2 2.2 1.0
Crystallinity 34.8 48.8 42.1 43.9 57.1 49.6

Note that phase abundances do not strictly co-vary (i.e., the sample
with the minimum percentage of quartz is not necessarily the exact
same sample as that with the minimum overall crystallinity, e.g.,
Fig. 5) Fig. 5 Plagioclase crystallinity (left axis, circles) and biotite crystal-
linity (right axis, diamonds) vs. overall phenocryst and micro-
phenocryst crystallinity for all point-counted samples; grey pumice
33% to 71%; bulk porosity of white pumice ranges from is shown with filled symbols; white pumice is shown with open
24% to 69%. The connected porosities of white and grey symbols
pumice populations cover a similar range, between 21%
and 69% (Fig. 13). The measured permeability of grey and
white pumice clasts also overlaps and ranges from 1.0× tube pumice in white clasts, although tube pumice is
10−14 to 5.3×10−12 m2. This range is typical for pumice subordinate to white pumice with isotropic structure.
clasts from pyroclastic flow deposits (e.g., Wright et al. Breadcrusted varieties exist of both tube and isotropic grey
2009) and reflects the anisotropy of vesicle textures in these pumice clasts; breadcrusting is absent in white pumice
samples (perpendicular clasts are shown with tie lines in clasts. Vesicle sizes vary between clasts, with vesicles in
Fig. 13). Anisotropy is manifested by a predominance of grey pumice generally smaller than those in white pumice
parallel, elongated vesicles in grey pumice clasts and true (e.g., Fig. 14).

Fig. 4 a Plane polarized and b


crossed polarized light image of
banded pumice clast. Note the
predominance of broken quartz
and feldspar crystals in grey
bands (grey phases in b) as
opposed to large quartz and
feldspar phenocrysts and higher
percentage of biotite phenoc-
rysts (dark phases in a) in white
bands
1522 Bull Volcanol (2011) 73:1513–1533

Fig. 6 Plane polarized light


images of crystal textures, in-
cluding a two of the largest
quartz crystals in a grey pumice
sample; both are fractured, but
the crystal on the right is only
fractured around its outer mar-
gins; b parallel fractures in
plagioclase; and c common,
pervasively fractured quartz
crystals in grey pumice. d Typ-
ical crystal fractures in quartz in
white pumice. Fractures are not
due to thin section preparation;
fracture surfaces are commonly
slightly altered and can be cov-
ered by a thin glassy film

Volatile content melt inclusions in white pumice clasts contain up to one co-
existing vapor bubble with a total volume less than 20% of
Melt inclusions are common in quartz and plagioclase the total inclusion volume. Measured volatile contents
phenocrysts in white pumice but are extremely rare in grey range between 2.4 and 6.6 wt.% H2O and 0 and 296 ppm
pumice due largely to the smaller crystal size and CO2 (Fig. 15; Sup. Fig. 1). These values correspond to
fragmental nature of crystals in grey pumice. Measured between 36 and 245 MPa at 750°C (Newman and Low-

Fig. 7 a, c SEM BSE images of


microlite phases in grey pumice:
Black areas are vesicles, dark
grey is quartz, light grey is
feldspar, white is biotite (or
oxide phases where <1 μm in
diameter), and medium grey is
glass; b, d sketches of crystal
phases: Quartz crystals are filled
in dark grey, feldspar crystals
are filled in light grey; biotite is
outlined in black and filled with
white. Note angular and irregu-
lar morphologies of feldspar
fragments and rounded, resorbed
morphologies of quartz
microlites
Bull Volcanol (2011) 73:1513–1533 1523

Fig. 8 SEM BSE image of grey


pumice clast. Sketches fill and
outline colors as in Fig. 7. Note
the presence of a dark (lower
average atomic number) veneer
of glass around the feldspar
crystal shown in the inset panel

enstern 2002). Interestingly, groundmass glass volatile CL zonation (Fig. 15b; cf. Renou 2009). Volatile content
contents reach within 0.4 wt.% H2O of the lower limit of does not systematically increase or decrease with distance
melt inclusions (0.85–2 wt.% H2O; 5–26 MPa at 750 C). from cores; furthermore, while melt inclusions within
The lowest water content was measured in the dense bomb recycled cores commonly do not contain CO2, not all
rind of a breadcrusted grey pumice clast; the highest value CO2-absent inclusions are within the cores of crystals.
is from the margin of an expanded white pumice clast.
These elevated values are supported by elevated volatiles-
by-difference estimates of water content from electron Discussion
microprobe analyses of groundmass glass compositions
(using the iterative water-by-difference program in Probe Here we use the physical and chemical data presented
for Windows software, e.g., Roman et al. 2006). The above to address the origin of the two prominent pumice
evolution of volatile contents in the melt is indicated by CL populations (grey and white) and their relationship to the
images of quartz crystals. Many crystals show concentric large CGI eruption. In particular, we examine ways in
oscillatory CL zoning in quartz crystals, where bright zones which characterization of the two pumice types can inform
often have flame-like outer margins (Fig. 15b). Melt our understanding of possible triggers for this eruption, as
inclusions are common in these zones. In addition, some well as for the eruption of crystal-rich “monotonous
crystals contain cores separated from rims by truncations in intermediate” (e.g., Hildreth 1981) magmas, in general.

Fig. 9 Bulk and groundmass


glass compositions of pumice
clasts in the Cerro Galán
Ignimbrite. Each symbol repre-
sents a distinct clast; error bars
represent standard deviation of
analyses for each clast. Grey
pumice glass compositions plot
between bulk and white pumice
glass compositions
Table 3 Selected glass compositions of white and grey pumice
1524

Sample Number SiO2 σ TiO2 Al2O3 σ FeO σ MgO CaO σ Na2O σ K2O σ P2O5 Cl Raw total

CG128a White 13 77.2 0.25 0.10 13.4 0.26 0.57 0.03 0.06 0.91 0.02 2.4 0.19 5.2 0.13 0.01 0.16 95.91
CG131b White 24 79.5 0.42 0.10 13.9 0.20 0.38 0.13 0.03 0.75 0.07 1.1 0.19 4.1 0.26 b.d. 0.20 95.15
CG140b White 6 77.8 0.25 0.11 13.2 0.17 0.51 0.16 0.06 0.51 0.16 1.9 0.10 5.3 0.22 0.02 0.19 97.64
CG142b-b White 9 78.4 0.81 0.11 13.1 0.28 0.55 0.08 0.07 0.89 0.04 2.0 0.26 4.7 0.35 b.d. 0.19 92.91
CG142b-b White 4 77.4 0.30 0.09 12.6 0.20 0.51 0.12 0.07 0.94 0.02 2.7 0.20 5.5 0.20 b.d. 0.18 94.01
CG405 White 10 76.6 0.34 0.12 13.0 0.20 0.32 0.14 0.04 0.96 0.04 3.1 0.13 5.7 0.30 b.d. b.d. 98.33
CG405 White 4 78.4 0.21 0.03 12.6 0.50 0.55 0.21 0.09 0.88 0.02 1.9 0.16 5.3 0.46 b.d. b.d. 96.80

CG1a1a Grey 2 75.8 0.60 0.23 13.7 0.04 0.49 0.15 0.05 1.42 0.17 2.2 0.64 5.9 0.32 0.02 b.d. 94.30
CG62a Grey 2 69.6 0.17 0.66 17.7 0.03 1.11 1.03 0.82 1.84 0.28 3.9 0.40 4.9 1.31 0.06 0.03 94.08
CG122a Grey 2 71.7 1.68 0.38 15.9 0.34 0.98 0.59 0.34 2.62 0.97 3.3 0.23 4.7 1.60 0.05 0.13 93.33
CG256 g Grey 9 70.1 0.72 0.53 16.1 0.43 2.25 0.16 1.03 1.84 0.30 2.8 0.27 5.2 0.24 0.41 b.d. 97.59

CG2a-a White band 3 77.9 0.74 0.17 14.6 0.19 0.20 0.06 0.02 1.17 0.27 1.5 0.27 4.3 0.30 0.01 0.13 95.79
CG2a-a Grey band 5 73.8 0.66 0.44 16.2 0.63 1.24 0.53 0.37 1.65 0.17 1.7 0.37 4.2 0.31 0.06 0.24 96.58
CG119b White band 3 78.7 0.93 0.13 13.9 0.25 0.19 0.01 0.03 0.91 0.05 1.4 0.34 4.6 0.30 b.d. 0.05 95.89
CG119b White band 3 77.5 1.48 0.11 14.3 0.69 0.20 0.04 0.03 1.19 0.33 2.0 0.79 4.6 0.40 b.d. 0.03 97.87
CG119b Grey band 2 73.8 0.30 0.31 15.9 0.11 0.64 0.57 0.06 1.61 0.18 2.6 1.07 4.9 0.02 0.07 0.06 97.68
CG119b Grey band 6 72.5 1.38 0.51 16.3 0.38 2.23 1.11 0.20 1.90 0.14 1.8 0.23 4.3 0.27 0.11 0.09 96.12
CG125a White band 5 77.3 0.36 0.17 13.2 0.14 0.51 0.07 0.10 0.96 0.04 1.9 0.26 5.9 0.22 b.d. 0.02 95.93
CG125a Grey band 3 71.1 1.74 0.60 16.6 0.28 1.11 0.59 0.44 2.04 0.19 2.3 0.45 5.7 0.78 0.08 0.03 95.38
CG136aa White band 7 78.6 1.23 0.12 13.6 0.43 0.25 0.03 0.03 1.00 0.48 1.7 0.45 4.5 0.23 b.d. 0.12 96.78
CG136aa Grey band 13 71.1 0.68 0.53 16.3 0.18 2.52 0.15 1.14 2.15 0.11 2.0 0.29 4.1 0.26 0.11 0.10 96.28

CG476 LM white 15 77.9 0.58 0.11 12.2 0.42 0.68 0.07 0.08 0.92 0.11 2.6 0.31 5.3 0.16 b.d. 98.08
CG476 LM white 4 78.1 0.28 0.05 12.8 0.29 0.67 0.05 0.08 0.92 0.03 2.1 0.22 5.1 0.20 b.d. 97.36
CG87a-b MM white 7 77.7 0.72 0.15 13.1 0.24 0.34 0.14 0.06 0.94 0.02 2.3 0.48 5.4 0.35 b.d. 96.52
CG481 MM grey 4 78.9 0.29 0.11 12.4 0.13 0.31 0.09 0.04 0.77 0.07 1.7 0.32 5.7 0.10 b.d. 96.59
CG483 UM grey 2 77.3 0.12 0.11 13.4 0.23 0.64 0.08 0.14 0.82 0.12 1.9 0.12 5.4 0.24 b.d. 96.71
CG98a-a Pitas white 13 77.4 0.25 0.12 12.8 0.21 0.24 0.04 0.04 0.85 0.07 2.6 0.31 5.9 0.14 b.d. 96.57
CG137a-b RG white 17 77.2 0.69 0.15 12.5 0.43 0.75 0.09 0.11 0.85 0.04 2.3 0.32 5.8 0.19 b.d. 95.35
CG68bb-b RG banded 10 78.0 0.60 0.16 12.8 0.44 0.31 0.15 0.06 0.75 0.04 1.9 0.31 5.9 0.28 0.01 96.59

Analyses of Toconquis white and grey pumice are also included


LM Lower Merihuaca, MM Middle Merihuaca, UM Upper Merihuaca, RG Real Grande ignimbrites of Folkes et al. (2011b), b.d. below detection limits
Bull Volcanol (2011) 73:1513–1533
Bull Volcanol (2011) 73:1513–1533 1525

and phenocrysts that preserve melt inclusions. Composi-


tional and modal crystal phase variations within the white
pumice population can be adequately explained by sanidine
fractionation in a large magma chamber. In particular,
correlative variation between decreasing Ba, K, Sr, and Eu
and increasing SiO2 support the interpretation of variable
separation of sanidine from the melt (sanidine fractionation;
Fig. 3b, f, g; Table 1). The presence of sanidine and the
absence of hornblende can be used to infer magma storage
at relatively low pressures. In comparison, experimentally
determined phase relations for the geochemically similar
Fig. 10 Plagioclase core compositions (a) and rim or microlite Fish Canyon rhyodacitic magma show that sanidine is
compositions (b) in white and grey pumice. An anorthite, Ab albite
present and hornblende is absent at <100 MPa at 800°C
(Johnson and Rutherford 1989). Crystallization at these
pressures at Cerro Galán is confirmed by the presence of
We pay particular attention to the origin of the grey pumice, melt inclusions trapped at pressures <100 MPa (Fig. 15)
as subordinate grey pumice is a common feature in many and the range of white pumice glass compositions (e.g.,
silicic eruptions. Generation of grey pumice has been Blundy and Cashman 2001, 2008). If the >630 km3 of
attributed to a variety of processes that are either syn- erupted magma (D.R.E.) resided in a sill-like body with
eruptive or active during precursory and/or intra-eruptive areal dimensions of the resulting caldera (~450 km2), then
periods (i.e., associated with eruptive processes). the sill thickness would have been 1.4 km (ΔP>35 MPa,
depending on magma density). This thickness corresponds
Geochemical distinction between pumice types well with inferred subsidence, based on the maximum
observed intracaldera thickness of the CGI (1.4 km; Sparks
Grey and white pumices are found as both discrete clasts et al. 1985). However, the wider observed range of melt
and as complexly mingled clasts (e.g., Fig. 2a). It is not inclusion pressures (~200 MPa) suggests a more vertically
uncommon to see either white pumice clasts or mingled extensive magma body (~8.5 km), the presence of a mush
pumice clasts as inclusions with subangular and in some column or a feeder dike system below the large sill-like
cases slightly concave margins in grey pumice (e.g., body or a distinct origin for higher volatile content
Fig. 2c). The presence of subangular, vesicular inclusions inclusions.
requires vesiculation followed by brittle breakage. Further- High volatile content melt inclusions could have
more, the existence of mingled clasts within grey pumice formed in a deeper portion of a more vertically
requires formation of grey magma prior to mingling; these extensive magma chamber that produced white pumice
observations would seem to rule out formation of grey (hereafter called white magma) and could be antecrysts
pumice solely by high strain rates created by rapid magma from older Galán ignimbrite-forming magma chambers,
ascent and decompression (e.g., Polacci et al. 2001; Rosi et or could be xenocrysts. Truncated CL zonation patterns
al. 2004). Geochemical data support this interpretation, as in crystal interiors support the interpretation that some
variations in minor and trace elements between clast types degree of quartz recycling has taken place, such that
require that grey and white pumice samples formed from melt inclusion entrapment pressures should be viewed
distinct, though geochemically similar, magma batches. For with caution (Fig. 14). High An cores in plagioclase
example, the observed co-variations between Ba and SiO2, crystals in white pumice may similarly indicate either
and FeO and TiO2 (Fig. 3) contents in each suite, in vertically extensive crystallization or xenocrysts (Fig. 10).
addition to trace element concentration variations in Xenocrystic zircons of basement age in the CGI (517–
phenocrysts, cannot be explained by simple fractionation, 539 Ma) and antecrystic zircons of Tonconquis age (5.60±
by simple mixing of two end-member compositions, or by 0.20 to 4.51±0.11 Ma 40Ar/39Ar biotite age) in the Cueva
shear-induced melting during eruption. Instead, it appears Negra ignimbrite (3.77±0.08 Ma 40Ar/39Ar biotite age)
that grey and white pumice were produced from distinct provide evidence for zircon recycling in the Galán system
magma bodies that mingled shortly prior to eruption. (C. Folkes, unpublished data). Therefore, it seems plausi-
ble that at least some of the inclusions with high H2O
Dominant chamber storage region contents may not have been trapped in the climactic
magma chamber. However, the dichotomy between inclu-
Questions of magma storage are most easily addressed for sions that contain CO2 and those that do not, in addition to
the white pumice, which has a limited compositional range the large range in entrapment pressures, present the
Table 4 Selected plagioclase, biotite, and oxide compositions in white and grey pumice
1526

Plagioclase Sanidine Biotite


CG122 mic CG140 core CG140 rim CG483 mic CG462 rim CG140 rim
Grey White White Grey Grey White
SiO2 59.7 59.6 58.8 64.0 SiO2 36.81 36.43
Al2O3 25.1 25.3 25.7 18.7 TiO2 4.13 4.33
FeO 0.24 0.18 0.19 0.09 Al2O3 14.35 14.11
CaO 7.05 6.46 7.36 0.16 MnO 0.21 0.25
Na2O 7.20 6.78 6.52 2.45 FeO 16.55 17.49
K2O 0.74 0.73 0.62 13.2 MgO 14.07 13.38
Total 100.04 99.08 99.23 98.7 Na2O 0.37 0.39
Ab 62.1 62.6 59.3 21.8 K2O 9.07 9.31
An 33.6 32.9 37.0 0.8 Total 95.56 95.68
Or 4.2 4.5 3.7 77.4
Plagioclase Biotite
CG428 CG140 CG428 CG140
Grey Grey White White Grey Grey White White
SiO2 calib 59.5 59.5 58.0 58.0 SiO2 calib 36.5 36.5 37.0 37.0
Li 3.82 9.19 7.58 62.13 Li 48 62 634 53
Ti 89.2 34.6 34.8 29.8 Ti 23363 13523 19724 21803
Sr 134 753 567 549 Cu 76.0 31.1 23.7 4.8
Zn 2.44 5.95 19.93 6.28 Zn 417 300 496 494
Rb 21 3 2 2 Y 3.20 1.60 0.86 0.12
Ba 157 93 61 75 Ba 2147 429 307 459
Nd 1.42 4.03 4.63 1.67 Nd 5.60 2.16 0.92 0.15
Pb 5.72 15.95 16.77 15.12 Pb 56.17 11.27 6.91 3.61
Oxides
CG428 mt CG428 ilm CG140 mt CG140 ilm
Grey Grey White White
SiO2 0.04 0.00 0.04 0.02
TiO2 4.01 36.57 3.74 35.97
Al2O3 1.31 0.06 1.33 0.15
MnO 0.26 0.75 0.90 1.23
FeO 83.31 54.87 83.68 54.58
MgO 0.51 1.29 1.15 1.34
Total 89.50 93.56 91.51 93.73
size (um) 90 190 120 120
log fO2 -11.62 -11.80
Temp (°C) 809 798

Values listed in percent of each component. Oxides are listed in weight percent; trace elements are listed in micrograms per gram. Trace element concentrations calibrated based on electron
microprobe SiO2 contents (listed). External errors (2σ) for ICP-MS trace element concentrations are ±5%, except for Pb, which is ±10%
Ab albite, An anorthite, Or orthoclase, Mic microlite, core phenocryst core, rim phenocryst rim
Bull Volcanol (2011) 73:1513–1533
Bull Volcanol (2011) 73:1513–1533 1527

Fig. 11 Plagioclase and biotite


compositions of grey (filled
symbols), white (open symbols),
and grey pumice clasts with
white cores (grey portions filled
in dark grey, white portions
filled in light grey). a Li con-
centrations are higher in plagio-
clase crystals in white pumice, b
Fe/(Fe+Mg) and Ti concentra-
tions are lower in biotites in
grey pumice, and c Li concen-
trations are higher in biotite in
white pumice, but Ba, d Pb and
Cu are higher in biotite in grey
pumice than in white

possibility that some inclusions have leaked. Although although we do not see inclusions with variable CO2
inclusions were carefully selected to be free from cracks or contents at the highest water content that record this CO2
thin glass septa connected to the matrix, we cannot rule loss. Alternatively, these inclusions could indicate lack of
out the possibility of leakage in any of the samples. CO2 in the original melt. Elevated CO2 contents across a
range of H2O contents, although common (e.g., linear H2O–
Volatile fluxing CO2 trends, Newman et al. 1988; Spilliaert et al. 2006;
Wright et al. 2007; Vigouroux et al. 2008), are more
Volatile and fluid-mobile trace element data may indicate difficult to explain. Suggested processes for preserving
spatially variable volatile fluxing prior to eruption. Quartz- elevated CO2 contents during degassing include (1) very
hosted melt inclusions contain two distinct populations, one high and variable CO2 in the original magma (Newman et
with CO2 and one without CO2 (Fig. 15). Although some al. 1988), (2) non-equilibrium vesiculation (Gonnermann
CO2-absent inclusions are found within recycled crystal and Manga 2005), or (3) gas fluxing (Rust et al. 2004;
cores, this is not universally the case. If the variation in Blundy et al. 2010). Here, we favor pre-eruptive gas
CO2 is real (i.e., the CO2-absent values are not due to post- fluxing of selected portions of the white magma reservoir,
entrapment volatile loss), the CO2-absent inclusions can be particularly given both CO2-absent and CO2-present
explained by open-system degassing and the resulting rapid volatile content (combined H2O–CO2) trends (which seem
loss of CO2 (Newman and Lowenstern 2002; Papale 1999), to rule out disequilibrium degassing).

Fig. 12 fO2 vs. temperature of co-existing Fe–Ti oxide pairs in white Fig. 13 Connected porosity vs. permeability of Cerro Galán pumice
(open symbols) and grey (filled symbols) pumice, calculated using the (grey pumice in filled symbols; white pumice in open symbols; small
geothermometers of Andersen and Lindsley (1988) and Ghiorso and symbols are data for other rhyodacitic to rhyolitic pyroclastic flow
Evans (2008). The nickel–nickel oxide buffer is shown for compar- samples from Klug and Cashman 1996, Nakamura et al. 2008, de
ison, calculated at 200 bars from Frost (1991) Maisonneuve et al. 2009, and Wright et al. 2009)
1528 Bull Volcanol (2011) 73:1513–1533

Fig. 14 SEM BSE image of


grey and white pumice clasts,
demonstrating variations in
bubble textures; white pumice
commonly has larger bubbles
(a) than grey pumice (b)

Boyce and Hervig (2008) inferred that a volatile magmatic system through time (the “recharge magma”).
recharge event preceded the climactic eruption of the CGI Such “cryptic” recharge has been identified elsewhere by
by days to months, based on OH and Cl zonation of apatite examination of trace elements in plagioclase (Smith et al.
rims (which would not be preserved over long timescales 2009). We see evidence in both bulk and mineral chemistry
due to rapid diffusion rates). Variable, elevated lithium of the grey and white magma, which are sufficiently
concentrations in plagioclase and biotite in white pumice different that they cannot be related by any simple physical
may support spatially variable volatile fluxing, if Li was process such as fractionation, assimilation, or shear-induced
partitioned from a deeper source into the vapor phase (e.g., melting.
as inferred for Mt. St. Helens; Berlo et al. 2004; Blundy et Determination of absolute depths for grey magma
al. 2008). However, we have not measured lithium residence is hindered by the absence of melt-inclusion-
concentration profiles in crystals to verify that lithium bearing phenocrysts and large errors on Al-in-hornblende
concentrations are higher in crystal rims than in cores, as geobarometers (e.g., Blundy and Cashman 2008). A deeper
expected for diffusive Li gain. origin for the grey magma relative to the white is suggested
by the less evolved glass composition of grey magma (and
Depth of origin of the grey pumice is inferred based on Al-in-hornblende geobarometry for
older Toconquis grey magma as well; Folkes et al. 2011a).
Geochemical data suggest that the magma that produced The position of the cotectic in the haplogranitic ternary
grey pumice (hereafter called grey magma) represents system Qz–Ab–Or moves progressively Qz-ward with
rhyodacitic magma that fed the growing Cerro Galán decreasing pressure (Tuttle and Bowen 1958). Rare horn-

Fig. 15 Water vs. CO2 content


of melt inclusions and ground-
mass glass from Cerro Galán
Ignimbrite determined using
FTIR spectroscopy. Open circles
are white pumice groundmass
glass, and filled circles are grey
pumice groundmass glass. Dia-
monds are melt inclusions in
quartz phenocrysts hosted in
white pumice. Error bars are
smaller than symbols where not
shown. Saturation pressures are
calculated using VolatileCalc
at 750°C (Newman and
Lowenstern 2002). Cathodolu-
minescence images of represen-
tative quartz crystals are shown
with arrows linking melt inclu-
sions and their respective mea-
sured H2O and CO2
concentrations
Bull Volcanol (2011) 73:1513–1533 1529

blende and modest amounts of sanidine in grey pumice,


which could potentially have originated in white pumice
(positive and negative Clapeyron slope to hornblende and
sanidine, respectively; e.g., Johnson and Rutherford 1989),
are consistent with a deeper origin to grey pumice than
white. However, taking into account the errors associated
with the geobarometers and geothermometers applied to the
CGI pumice clasts, we cannot rule out that the grey magma
could have been stored at depths equivalent to portions of
the white chamber.

Migration of grey magma through white magma

The textural association between white and grey pumice


indicate that the two magmas having different compositions
mingled prior to eruption, although the physical conditions
of mingling appear to have been highly variable. White Fig. 16 Eruption sequence schematic: white pumice originated from a
pumice occurs in macroscopic bands, as cores to larger grey volumetrically dominant, large magma chamber, which was fluxed by
pumice clasts and as bands of white pumice glass volatiles; grey pumice originated from originally deeper recharge
magma, with higher Cu, Zn, and Pb contents and lower Li contents in
compositions within grey pumice clasts (Table 3); however, crystals; less viscous grey magma ascended through the larger magma
the opposite of all of these cases is not true. The white chamber, fracturing the overlying country rock, and stalling in
pumice cores can be subangular in shape, which suggests vanguard fractures at shallow levels, priming the system for eruption
brittle behavior during interaction of the two magma types.
In contrast, thin rinds of evolved melt that surround some
crystals in grey pumice provide evidence for transfer of from a more viscous, degassed melt (i.e., with lower
crystals from white magma into grey magma when the two diffusivity, e.g., Zhang and Behrens 2000), particularly
magmas were intimately (and ductilely) mingled (Fig. 8). when there are abundant microlites to limit bubble
Evidence for additional transfer of material from grey coalescence (Klug and Cashman 1994).
magma to white can also be found: White bands in grey Grey pumice clasts contain microlites that appear to be the
pumice contain plagioclase and biotite crystals with high products of both phenocryst fragmentation and degassing-
Ba and low Li, similar to uniformly grey pumice (Fig. 11) induced crystallization (i.e., they exhibit both breakage and
and white pumice cores, have wispy and thinly banded growth textures). To form microlites as a result of degassing,
margins (Fig. 2c). We therefore suggest that the grey the grey magma must have resided at shallow depths long
magma must have traveled through the volumetrically enough to degas, nucleate, and grow crystals, suggesting
dominant (and higher viscosity) white magma chamber shallow residence for at least hours (e.g., Hammer and
(Fig. 16). Rutherford 2002). A timescale of hours to days would also
have been sufficient to partially resorb the margins of small
Shallow residence of grey magma quartz fragments, where resorption occurs due to a shift in
the position of the cotectic with decreasing pressure (Tuttle
Both textural and compositional characteristics of grey and Bowen 1958; e.g., Fig. 7). Luhr and Melson (1996)
pumice suggest that the grey magma resided at shallower suggested that timescales of hours to months were necessary
levels than the white magma immediately prior to eruption. to produce similar resorption features in Pinatubo pumice.
Macroscopic evidence of shallow residence is provided by Together, these textural data suggest that the grey magma
breadcrusting of some grey clasts and associated low water formed a vanguard phase that stalled temporarily at shallow
contents (<1 wt.% H2O) of breadcrust bomb rinds (e.g., levels prior to the onset of the caldera-forming eruption.
Wright et al. 2007). Microtextural evidence for shallow Additional evidence for the grey magma as a vanguard phase
magma residence includes both vesicle and crystal charac- includes the presence of dacite lithics (wall rock) within
teristics of grey pumice. some grey pumice clasts.
Vesicles are smaller and have a higher number density in There is also clear evidence, however, that grey magma
the grey pumice than in the white (compare a and b of batches resided at different depths and stalled for different
Fig. 14). Although these characteristics could result from amounts of time. Most striking is the range in grey magma
much higher supersaturations in the grey relative to the matrix glass compositions, which span nearly the full range
white magma, they are also anticipated for vesiculation from the most evolved bulk compositions (of both grey and
1530 Bull Volcanol (2011) 73:1513–1533

white magma) to the lowest silica white pumice glass tion due to shock wave propagation during eruption (e.g.,
composition and require variable degrees of crystallization Pallister et al. 1996). Textures are consistent with any of
that we infer to have taken place at different depths of these fracture mechanisms (Figs. 4, 6, and 7). The intrusion
storage, though some of this spread in composition could of grey magma through crystal-rich white magma and into
be due to diffusive mixing during mingling. This variation vanguard shallow fractures is equally likely to cause shallow
in equilibration depths would have allowed variable shear-induced fracture, to trigger rupture of trapped melt
amounts of shallow decompression-induced crystallization inclusions and shock-induced fracture. We suggest, therefore,
(of plagioclase and oxides) and crystal breakdown and/or that all of these mechanisms may have contributed to crystal
resorption (of less stable phases, such as hornblende and disruption in the grey magma as it ascended to shallow
quartz; e.g., Blundy and Cashman 2001; Fig. 9). Moreover, levels.
if elevated Li concentrations at Cerro Galán reflect late Eruption conditions can be inferred from the range of
stage volatile enrichment, we can infer from its absence that connected porosities and permeabilities of the two clast
the shallow grey pumice magma was not fluxed by vapor types, as well as from the overall characteristics of the
prior to eruption; this was also true of vanguard crypto- deposit. Connected porosities of both clast types range from
dome magma at Mount St. Helens (Berlo et al. 2004, 2006). 40% to 70%. Although these porosities appear low, when
corrected for the high observed crystallinities of these
Triggering of the Cerro Galán eruption samples (up to 45% for the grey, up to 55% for the white
pumice), the range of melt vesicularities is ~60–80%, as
A heuristic model of grey magma traversing and interacting observed in other silicic eruptions. Cerro Galán deposits are
with white magma rising at upper crustal levels raises pumice-poor (<10% by volume), which suggests that only
questions about the physical mechanism by which this pumice with sufficient permeability was preserved. Pumice
occurred. Critical to answering this question are observa- preservation will be determined by the overpressure
tions of both mingling textures and phenocryst morphology. experienced during fragmentation and the clast permeability
Entrainment of brittlely broken white pumice clasts within (e.g., Klug and Cashman 1996). Our data show that this
grey requires that the grey magma intruded into already critical permeability was ~10−12 m2, as seen in other silicic
vesiculated white magma, or that brittle behavior preceded eruptions (e.g., Klug and Cashman 1996). Grey magma
simultaneous vesiculation in both. In contrast, intimate stored at the shallowest levels would have experienced
mingling of white and grey glass compositions and crystal smaller pressure drops than in white magma, which
cargo demands interaction when both magmas were fluid explains breadcrusting of these clasts (from delayed
and not fully vesiculated. Furthermore, the fine scale of vesiculation; Wright et al. 2007).
compositional banding in glasses (Figs. 2 and 8) and large
variation in Li concentrations in crystals (Fig. 11) argue for Cryptic silicic recharge
relatively short post-mingling timescales based on diffusiv-
ities at magmatic temperatures (e.g., Li diffusivity in We can also use our observations of the most recent
plagioclase and biotite, Raussel-Colom et al. 1965; Giletti caldera-forming eruptive deposits of Cerro Galán to
and Shanahan 1997). Therefore, we infer that interactions speculate on longer-term processes related to generation
between the two magma types occurred over a large vertical and storage of magma in that system. It seems reasonable to
extent, prior to eruption, with the shallowest interactions suggest that grey pumice may provide evidence of silicic
being at low pressures (to account for vesicular angular recharge in the Cerro Galán system. Textural support for the
pumice). recharge hypothesis includes evidence for crystal transfer
Thin section observations show that quartz and feldspar from white into grey magma. Such transfer may indicate
crystals in the grey pumice are both marginally and that measured crystallinities in grey pumice are much
pervasively fractured. Some fractured crystals have not been higher than the crystallinity of the original grey magma.
disaggregated (Figs. 6 and 7), others occur in elongate bands Furthermore, coupled resorption and growth (quartz resorp-
(Fig. 4), and others form isotropic crystal clusters (Fig. 7). tion, microlite growth) is not surprising in rejuvenated
Possible breakage mechanisms include shear created by the magma (e.g., Couch et al. 2001; Bachmann et al. 2002).
movement of grey magma through the white magma body Therefore, we interpret the grey magma as a higher volatile
and into vanguard fractures above the white magma chamber content, lower viscosity magma (due to lower crystallinities
(e.g., high strain rate fragmentation of highly viscous magma, prior to mingling-induced crystal transfer) that ascended
Gonnermann and Manga 2003; Tuffen et al. 2003), rupture of through the rigid white magma, propagating into newly
volatile-rich, phenocryst-hosted melt inclusions (as is com- opened fractures that became the Cerro Galán collapse
mon for volatile-rich melt inclusions, Tait 1992; Best and caldera block-bounding fractures (Fig. 16). These grey
Christiansen 1997; Bindeman 2005), and brittle fragmenta- magma-lined shallow fractures became the network of
Bull Volcanol (2011) 73:1513–1533 1531

faults that bound individual piecemeal blocks in the Cerro protracted, upper crustal chamber growth leading to large
Galán system (Folkes et al. 2011b). silicic eruption (Hildreth and Wilson 2007). Furthermore,
Evidence for silicic recharge has recently been identified our results suggest that abundant broken crystal fragments
at a number of volcanoes (Eichelberger et al. 2006; Hildreth and variations in groundmass glass compositions are not
and Wilson 2007; Kennedy and Stix 2007; Wiebe et al. necessarily indicative of syn-eruptive shear along conduit
2007; de Silva et al. 2008; Smith et al. 2009). Our model, boundaries (e.g., Polacci et al. 2001; Rosi et al. 2004), but
whereby silica-rich recharge magma may also ascend may instead represent recharge magma (e.g., at Pinatubo
through a large crystal-rich chamber to trigger an eruption, and Quilotoa volcanoes).
is a logical extension of these recharge models. Importantly,
we suggest that triggering of eruptions by recharge magma
does not require the intrusion of mafic (e.g., basaltic) Conclusions
material into the shallow chamber (as at Pinatubo, Pallister
et al. 1992; Soufriere Hills, Murphy et al. 1998; or Mt. Using a combination of bulk, glass, and crystal chemistries
Rainier, Venezky and Rutherford 1997). Instead, we infer and textural variations, we interpret the grey and white
that the recharge batch is buoyant (e.g., volatile-rich) and pumice populations of the CGI to represent distinct magma
mobile. Grey magma potentially had higher temperature (in batches that mingled prior to and during eruption. Grey
older Toconquis magma as well, cf. Folkes et al. 2011a); pumice is interpreted as originally deeper, more volatile-
however, we do not see the signature of this temperature rich, recharge magma, whereas white pumice is interpreted
difference at Cerro Galán, probably because the grey to represent a volumetrically dominant, overlying magma.
magma had thermally equilibrated by the time it traversed Grey magma then ascended through white magma, filling
the magma reservoir or because oxides in the grey magma shallow vanguard fractures and stalling at shallow depths
were transferred from white magma. Such silicic recharge prior to eruption. As a combined result of shear-induced
likely occurs multiple times over the period of chamber fracture, melt inclusion rupture, and shock-induced frag-
growth. The repeated CL bright growth zones in quartz mentation, most crystals in grey magma were broken, to
(with coincident preferential melt inclusion entrapment) varying degrees. Furthermore, a spatially variable volatile-
provide textural evidence of repeated recharge (Wark et al. fluxing event caused elevated volatile (CO2) and fluid-
2007). Importantly, we do not dismiss the possible role of mobile (Li) concentrations in portions of the large magma
mafic recharge anywhere in the system; instead, we suggest chamber and may have contributed volatiles critical to
that mafic melt is likely to remain at depth (lower to mid- triggering the eruption.
crustal levels). Furthermore, we suggest that this process of
silicic recharge, likely originating from a mid-crustal Acknowledgments This work was funded by ARC grant
reservoir, has occurred within pre-CGI magma chambers DP0663560 to Cas. The authors would like to thank Jake Lowenstern,
as well, producing the same textural variations (white and editor John Stix, and an anonymous reviewer for helpful suggestions
to improve the manuscript.
grey pumice) in Toconquis ignimbrites, although with
different chemical differences (cf. Folkes et al. 2011a).
Parallel textural variations in pumice clasts are found References
from several other caldera-forming eruptions worldwide.
We note the remarkable similarities between grey and white
Anders E, Ebihara M (1982) Solar-system abundances of the
pumice in Cerro Galán and the dark/swirly and normal elements. Geochim Cosmochim Acta 46:2363–2380. doi:10.
pumice in the Bishop Tuff (Hildreth and Wilson 2007), 1016/0016-7037(82)90208-3
although absolute crystallinities in the Bishop Tuff are Andersen DJ, Lindsley DH (1988) Internally consistent solution
much lower. In the Bishop Tuff, volumetrically minor, less models for Fe–Mg–Mn-Ti oxides: Fe–Ti oxides. Am Mineral
73:714–726
evolved rhyolite clasts (dark/swirly) were found to be less Bachmann O, Dungan MA, Lipman P (2002) The Fish Canyon
crystalline and richer in Ba, Sr, and Ti than normal pumice magma body, San Juan Volcanic Field, Colorado: rejuvenation
and are also interpreted as silicic recharge products. Dark and eruption of an upper-crustal batholith. J Petrol 43:1469–1503
pumice in the Bishop Tuff contains the only hornblende Bacon CR, Druitt TH (1988) Compositional evolution of the zoned
calcalkaline magma chamber of Mount Mazama, Crater Lake,
crystals present, and crystallization in dark pumice was Oregon. Contrib Mineral Petrol 98:224–256
deeper than in the dominant chamber; Fe–Ti oxide temper- Bacon CR, Hirschmann MM (1988) Mg/Mn partitioning as a test for
atures overlap with those measured in normal pumice (at equilibrium between coexisting Fe–Ti oxides. Am Mineral
the high T end), and only dark pumice forms breadcrusted 73:57–61
Bacon CR, Metz J (1984) Magmatic inclusions in rhyolites,
surface textures. The textural and compositional similarity contaminated basalts, and compositional zonation beneath the
is striking; we suggest that the formation mechanism is Cosa volcanic field, California. Contrib Mineral Petrol 85:346–
similar and the hallmark of recharge processes during 365
1532 Bull Volcanol (2011) 73:1513–1533

Behrens H, Tamic N, Holtz F (2004) Determination of the molar and origins of a very large volume ignimbrite and its magma
absorption coefficient for the infrared absorption band of CO2 in system. Bull Volcanol. doi:10.1007/s00445-011-0511-y
rhyolitic glasses. Am Mineral 89:301–306 Folkes CB, Wright HMN, Cas RAF, de Silva SL, Lesti C, Viramonte JG
Berlo K, Blundy J, Turner S, Black S, Cashman KV, Hawkesworth C (2011) A re-appraisal of the stratigraphy and volcanology of the
(2004) Chemical tracers of conduit processes preceding the 1980 Cerro Galán volcanic system, NWArgentina. In: Cas RAF, Cashman
eruption of Mount St. Helens. IAVCEI, Pucon K (eds) The Cerro Galán Ignimbrite and Caldera: characteristics and
Berlo K, Turner S, Blundy J, Black S, Hawkesworth C (2006) Tracing origins of a very large volume ignimbrite and its magma system.
pre-eruptive magma degassing using (210Pb/226Ra) disequilibria Bull Volcanol. doi:10.1007/s00445-011-0459-y
in the volcanic deposits of the 1980-1986 eruption of Mount St. Frost BR (1991) Introduction to oxygen fugacity and its petrologic
Helens. Earth Planet Sci Lett 249:337–349 importance. Rev Mineral Geochem 25:1–9
Best MG, Christiansen EH (1997) Origin of broken phenocrysts in Geschwind CH, Rutherford MJ (2002) Cummingtonite and the
ash-flow tuffs. Geol Soc Am Bull 109:63–73 evolution of the Mount St. Helens magma system: an experi-
Bindeman IN (2005) Fragmentation phenomena in populations of mental study. Geology 20:1011–1014
magmatic crystals. Am Mineral 90:1801–1815 Ghiorso MS, Evans BW (2008) Thermodynamics of rhombohedral
Blundy J, Cashman K (2008) Petrologic reconstruction of magmatic oxide solid solutions and a revision of the Fe–Ti two-oxide
system variables and processes. Reviews in Mineralogy and geothermometer and oxygen-barometer. American Journal of
Geochemistry 69:179–239. doi:10.2138/rmg.2008.69.6 Science 308:957–1039. doi:10.2475/09.2008.01
Blundy JD, Cashman KV (2001) Magma ascent and crystallisation at Giletti BJ, Shanahan TM (1997) Alkali diffusion in plagioclase
Mount St. Helens, 1980–1986. Contrib Mineral Petrol 140:631– feldspar. Chem Geol 139:3–20
650 Gonnermann HM, Manga M (2003) Explosive volcanism may not be
Blundy J, Cashman KV, Berlo K (2008) Evolving magma storage an inevitable consequence of magma fragmentation. Nature
conditions beneath Mount St. Helens inferred from chemical 426:432–435
variations in melt inclusions from the 1980–1986 and current Gonnermann HM, Manga M (2005) Non-equilibrium magma degass-
(2004–2006) eruptions. In: Sherrod DR, Scott WE, Stauffer ing: results from modeling of the ca. 1340 A.D. eruption of
PH (eds) A volcano rekindled: the renewed eruption of Mono Craters, California. Earth Planet Sci Lett 238:1–16
Mount St. Helen, 2004–2006. US Geological Survey Profes- Hammer JE, Rutherford MJ (2002) An experimental study of the
sional Paper 1750. US Geological Survey, Reston, pp 755– kinetics of decompression-induced crystallization in silicic melt.
790 J Geophys Res 107. doi:10.1029/2001JB000281
Blundy J, Cashman KV, Rust A, Witham F (2010) A case for CO2-rich Hildreth W (1981) Gradients in silicic magma chambers: implications
arc magma. Earth Planet Sci Lett 290:289–301. doi:10.1016/j. for lithospheric magmatism. J Geophys Res 86:10153–10192
epsl.2009.12.013 Hildreth W, Wilson CJN (2007) Compositional zoning of the Bishop
Boyce JW, Hervig RL (2008) Magmatic degassing histories from Tuff. J Petrol 48:951–999. doi:10.1093/petrology/egm007
apatite volatile stratigraphy. Geology 36:63–66 Johnson DM, Hooper PR, Conrey RM (1999) XRF analysis of rocks
Cas RAF, Wright HMN, Lesti C, Porreca M, Folkes C, Giordano G, and minerals for major and trace element analysis on a single low
Viramonte J (2011) The flow dynamics of an extremely large dilution Li-tetraborate fused bead. Adv X Ray Anal 41:843–867
volume pyroclastic flow, the 2.08 Ma Cerro Galán Ignimbrite, Johnson MC, Rutherford MJ (1989) Experimentally determined
NW Argentina, and comparison with other flow types. In: Cas conditions in the Fish Canyon Tuff, Colorado magma chamber.
RAF, Cashman K (eds) The Cerro Galán ignimbrite and Caldera: J Petrol 30:711–737
characteristics and origins of a very large volume ignimbrite and Kay SM, Coira B, Wörner G, Kay RW, Singer BS (2011) Geochemical,
its magma system. Bull Volcanol (in press) isotopic, and single crystal 40Ar/39Ar age constraints on the
Cioni R, Civetta L, Marianelli P, Metrich N, Santacroce R, Sbrana A evolution of the Cerro Galán Ignimbrites. In: Cas RAF, Cashman
(1995) Compositional layering and syn-eruptive mixing of a K (eds) The Cerro Galán Ignimbrite and Caldera: characteristics
periodically refilled shallow magma chamber: the AD 79 Pinian and origins of a very large volume ignimbrite and its magma
eruption of Vesuvius. J Petrol 36:739–776 system. Bull Volcanol. doi:10.1007/s00445-010-0410-7
Couch S, Sparks RSJ, Carroll MR (2001) Mineral disequilibrium in Kennedy B, Stix J (2007) Magmatic processes associated with caldera
lavas explained by convective self-mixing in magma chambers. collapse at Ossipee ring dyke, New Hampshire. Geol Soc Am
Nature 411:1037–1039 Bull 119:3–17. doi:10.1130/B25980.1
de Maisonneuve CB, Bachmann O, Burgisser A (2009) Characteriza- Kent AJR, Stolper EM, Rancis D, Woodhead J, Frei R, Eiler J (2004)
tion of juvenile pyroclasts from the Kos Plateau Tuff (Aegean Mantle heterogeneity during formation of the North Atlantic
Arc): insights into the eruptive dynamics of a large rhyolitic Igneous Province: constraints from trace element and Sr–Nd–Os–
eruptions. Bull Volcanol 71:643–658. doi:10.1007/s00445-008- O isotope systematics of Baffin Island picrites. Geochem Geophy
0250-x Geosy 5:Q11004. doi:10.1029/2004GC000743
de Silva SL, Salas G, Schubring S (2008) Triggering explosive Klug C, Cashman KV (1994) Vesiculation of May 18, 1980, Mount St.
eruption—the case for silicic magma recharge at Huaynaputina, Helens magma. Geology 22:468–472
southern Peru. Geology 36:387–390. doi:10.1130/G24380A.1 Klug C, Cashman KV (1996) Permeability development in vesiculat-
Donovan JJ (2007) Probewin (Probe for Windows) 6.11 ing magmas: implications for fragmentation. Bull Volcanol
Eichelberger JC, Chertkoff DC, Dreher ST, Nye CJ (2000) Magmas in 58:87–100. doi:10.1007/s004450050128
collision: rethinking chemical zonation in silicic magmas. Knaack C, Cornelius SB, Hooper PR (1994) Trace element analyses of
Geology 28:603–606 rocks and minerals by ICP-MS. Technical Notes for WSU
Eichelberger JC, Izbekov PE, Browne BL (2006) Bulk chemical GeoAnalytical Lab
trends at arc volcanoes are not liquid lines of descent. Lithos Le Bas MJ, Le Maitre RW, Woolley AR (1992) The construction of the
87:135–154. doi:10.1016/j.lithos.2005.05.006 total alkali-silica chemical classification of volcanic rocks.
Folkes CB, de Silva SL, Wright HMN, Cas RAF (2011) Geochemical Mineral Petrol 46:1–22
homogeneity of a long-lived, large silicic system: evidence from Lindsay JM, Schmitt AK, Trumbull RB, De Silva SL, Siebel W,
the Cerro Galán caldera, NW Argentina. In: Cas RAF, Cashman K Emmermann R (2000) Magmatic evolution of the La Pacana
(eds) The Cerro Galán Ignimbrite and Caldera: characteristics caldera system, Central Andes, Chile: compositional variation of
Bull Volcanol (2011) 73:1513–1533 1533

two cogenetic large-volume felsic ignimbrites. J Petrol 42:459– Rust AC, Cashman KV, Wallace PJ (2004) Magma degassing buffered
486 by vapor flow through brecciated conduit margins. Geology
Lipman PW, Steven TA, Mehnert HH (1970) Volcanic hitory of the 32:349–352, 10.1130/G20388.2
San Juan Mountains, Colorado, as indicated by potassium-argon Smith VC, Blundy JD, Arce JL (2009) A temporal record of magma
dating. Geol Soc Amer Bull 81:2329–2352 accumulation and evolution beneath Nevado de Toluca, Mexico,
Luhr, JF, Melson WG (1996) Mineral and glass compositions in June preserved in plagioclase phenocrysts. J Petrol 50:405–426
15, 1991, pumices: evidence for dynamic disequilibrium in the Smith VC, Shane P, Nairn IA (2004) Reactivation of a rhyolitic
dacite of Mount Pinatubo. In: Newhall CG, Punongbayan RS magma body by new rhyolitic intrusion before the 15.8 ka
(eds) Fire and mud: eruptions and lahars of Mount Pinatubo, Rotorua eruptive episode: implications for magma storage in the
Philippines. University of Washington Press, Seattle, pp 733–750 Okataina Volcanic Centre, New Zealand. Journal of the Geolog-
Mason BG, Pyle DM, Oppenheimer C (2004) The size and frequency of ical Society of London 161:757–772. doi:10.1144/0016-764903-
the largest explosive eruptions on Earth. Bull Volcanol 66:735–748 092
Murphy MD, Sparks RSJ, Barclay J, Carroll MR, Lejeune A-M, Brewer Sparks RSJ, Francis PW, Hamer RD, Pankhurst RJ, O'Callaghan LO,
TS, Macdonald R, Black S, Young S (1998) The role of magma Thorpe RS, Page R (1985) Ignimbrites of the Cerro Galan
mixing in triggering the current eruption at the Soufriere Hills Caldera, NW Argentina. J Volcanol Geotherm Res 24:205–248
volcano, Montserrat, West Indies. Geophys Res Lett 25:3433–3436 Spilliaert N, Allard P, Métrich N, Sobolev AV (2006) Melt inclusion
Nakamura M, Otaki K, Takeuchi S (2008) Permeability and pore- record of the conditions of ascent, degassing, and extrusion of
connectivity variation of pumices from a single pyroclastic flow volatile-rich alkali basalt during the powerful 2002 flank eruption
eruption: implications for partial fragmentation. J Volcanol of Mount Etna (Italy). J Geophys Res 111:B04203. doi:10.1029/
Geotherm Res 176:302–314 2005JB003934
Newman S, Epstein S, Stolper E (1988) Water, carbon dioxide, and Tait S (1992) Selective preservation of melt inclusions in igneous
hydrogen isotopes in glasses from the CA. 1340 A.D. eruption of the phenocrysts. Am Mineral 77:146–155
Mono Craters, California: constraints on degassing phenomena and Tuffen H, Dingwell DB, Pinkerton H (2003) Repeated fracture and
initial volatile content. J Volcanol Geotherm Res 35:75–96 healing of silicic magma generates flow banding and earth-
Newman S, Lowenstern JB (2002) VolatileCalc: a silicate melt-H2O- quakes? Geology 31:1089–1092. doi:10.1130/G19777.1
CO2 solution model written in Visual Basic for EXCEL. Comput Tuttle OF, Bowen NL (1958) Origin of granite in the light of
Geosci 28:597–604 experimental studies in the system NaAlSi3O8-KAlSi3O8-SiO2-
Newman S, Stolper EM, Epstein S (1986) Measurement of water in H2O. Geological Society of America memoir. Geological Society
rhyolitic glasses: calibration of an infrared spectroscopic tech- of America, New York, p 153
nique. Am Mineral 71:1527–1541 Venezky DY, Rutherford MJ (1997) Preeruption conditions and timing
Pallister JS, Hoblitt RP, Meeker GP, Knight RJ, Siems DF (1996) of dacite-andesite magma mixing in the 2.2 ka eruption at Mount
Magma mixing at Mount Pinatubo: petrographic and chemical Rainier. J Geophys Res 102:20069–20086
evidence from the 1991 deposits. In: Newhall CG, Punongbayan Vigouroux N, Wallace PJ, Kent AJR (2008) Volatiles in high-K
RS (eds) Fire and mud: eruptions and lahars of Mount Pinatubo, magmas from the Western Trans-Mexican volcanic belt: evidence
Philippines. University of Washington Press, Seattle for fluid fluxing and extreme enrichment of the mantle wedge by
Pallister JS, Hoblitt RP, Reyes AG (1992) A basalt trigger for the 1991 subduction processes. J Petrol 49:1589–1618
eruptions of Pinatubo Volcano. Nature 356:426–428 Wallace PJ, Anderson ATJ, Davis AM (1999) Gradients in H2O, CO2,
Papale P (1999) Modeling of the solubility of a two-component H2O+ and exsolved gas in a large-volume silicic magma system:
CO2 fluid in silicate liquids. Am Mineral 84:477–492 interpreting the record preserved in melt inclusions from the
Polacci M, Papale P, Rosi M (2001) Textural heterogeneities in Bishop Tuff. J Geophys Res 104:20097–20122
pumices from the climactic eruption of Mount Pinatubo, 15 June Wark DA, Hildreth W, Spear FS, Cherniak DJ, Watson EB (2007) Pre-
1991, and implications for magma ascent dynamics. Bull eruption recharge of the Bishop magma system. Geology
Volcanol 63:83–97 35:235–238. doi:10.1130/G23316A.1
Raussel-Colom JA, Seatman TR, Wells CB, Norrish K (1965) In: Wiebe RA, Wark DA, Hawkins DP (2007) Insights from quartz
Hallsworth EG, Crawford DV (eds) Experimental pedology. cathodoluminescence zoning into crystallization of the Vinal-
Butterworths Scientific, London, pp 40–72 haven granite, coastal Maine. Contrib Mineral Petrol 154:439–
Renou H (2009) Volatiles in magma: evidence preserved in melt 453. doi:10.1007/s00410-007-0202-z
inclusions and the analysis of their host phenocrysts from the Wright HMN, Cashman KV, Gottesfeld EH, Roberts JJ (2009) Pore
Cerro Galan stratigraphy, NW Argentina. Honours thesis, School structure of volcanic clasts: measurements of permeability and
of Geosciences, Monash University, Melbourne, 109 p electrical conductivity. Earth Planet Sci Lett 280:93–104.
Roman DC, Cashman KV, Gardner CA, Wallace PJ, Donovan JJ doi:10.1016/j.epsl.2009.01.023
(2006) Storage and interaction of compositionally heterogeneous Wright HMN, Cashman KV, Rosi M, Cioni R (2007) Breadcrust
magmas from the 1986 eruption of Augustine Volcano, Alaska. bombs as indicators of Vulcanian eruption dynamics at Guagua
Bull Volcanol 68:240–254. doi:10.1007/s00445-005-0003-z Pichincha volcano, Ecuador. Bull Volcanol 69:281–300.
Rosi M, Landi P, Polacci M, di Muro A, Zandomeneghi D (2004) Role doi:10.1007/s00445-006-0073-6
of conduit shear on ascent of the crystal-rich magma feeding the Zhang Y, Behrens H (2000) H2O diffusion in rhyolitic melts and
800-year-(B.P.) Pinian eruption of Quilotoa volcano (Ecuador). glasses. Chem Geol 169:243–262
Bull Volcanol 66:307–321 Zhang Y, Jenkins J, Xu Z (1997) Kinetics of the reaction H2O+O-2OH
Rust AC, Cashman KV (2004) Permeability of vesicular silicic in rhyolitic glasses upon cooling: geospeedometry and compar-
magma: inertial and hysteresis effects. Earth Planet Sci Lett ison with glass transition. Geochim Cosmochim Acta 61:2167–
228:93–107 2173

You might also like