You are on page 1of 7

Earth and Planetary Science Letters 333–334 (2012) 63–69

Contents lists available at SciVerse ScienceDirect

Earth and Planetary Science Letters


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

The role of melt-fracture degassing in defusing explosive rhyolite eruptions


at volcán Chaitén
Jonathan M. Castro a,b,n, Benoit Cordonnier c,d, Hugh Tuffen e, Mark J. Tobin f, Ljiljana Puskar f,
Michael C. Martin g, Hans A. Bechtel g
a
Institute of Geosciences, Becherweg 21, University of Mainz, 55099 Mainz, Germany
b
School of Geosciences, Monash University, Clayton, Victoria, Australia
c
Earth and Planetary Sciences, University of California, Berkeley, CA 94720, USA
d
Geological Institute, Swiss Federal Institute of Technology, Zurich, Switzerland
e
Lancaster Environment Centre, Lancaster University, LA1 4YQ, UK
f
Australian Synchrotron, 800 Blackburn Rd., Clayton, Victoria, Australia
g
Advanced Light Source, 1 Cyclotron Rd., Berkeley, CA 94720, USA

a r t i c l e i n f o abstract

Article history: Explosive volcanic eruptions of silicic magma often evolve towards non-explosive emissions of lava.
Received 18 January 2012 The mechanisms underlying this transition remain unclear, however, a widely cited idea holds that
Received in revised form shear-induced magma fragmentation plays a critical role by fostering volatile loss from fragmentary
2 April 2012
magma and through ash-filled cracks termed tuffisite. We test this hypothesis by measuring H2O
Accepted 17 April 2012
concentrations within glassy tuffisite from the 2008–2011 rhyolitic eruption at volcán Chaitén, Chile.
Editor: B. Marty
We show that while H2O concentrations decrease next to tuffisite veins and at the margins of obsidian
fragments, the depletions cannot account for the disparity in H2O between explosively and effusively
Keywords: erupted rhyolite. Tuffisite vein lifetimes derived from diffusion modeling (min to h) imply degassing
obsidian
rates that are too slow to effectively degas magma, unless the magma was entirely fragmented to mm
eruption
or smaller particles. This level of brecciation may locally develop near conduit margins, but is
explosive–effusive transition
Chaitén unrealistic for entire conduits. The primary role of melt fracturing may therefore be to provide gas-
FTIR escape pathways for more efficient degassing of permeable vesicular magma in the conduit interior.
volcanology & 2012 Elsevier B.V. All rights reserved.

1. Introduction part of this degassing mechanism is that coherent magma behaves as


a chemically open system (Newman et al., 1988), and the ‘‘opening’’
The broad range of volcanic eruption styles, bound by explosive of the system occurs through the development of bubble perme-
and effusive eruptions, is underpinned by magma’s ability to shed its ability (e.g., Westrich and Eichelberger, 1994; Klug and Cashman,
volatiles whilst rising to the surface. Magma loses its dissolved 1996; Okumura et al., 2009) and flow of exsolved volatiles upward
volatiles via degassing, a two-part process involving firstly, short- and out of permeable conduit walls (Jaupart and Alle gre, 1991;
path-length (10’s to 100’s mm) diffusive transfer of volatile compo- Woods and Koyaguchi, 1994).
nents into an area of lower chemical potential (e.g., a melt–bubble A recent hypothesis (Gonnermann and Manga, 2003) posits
interface) and then advective transport of those exsolved volatiles to that effective, open-system degassing of magma can happen in
the surroundings, such as the conduit margins and atmosphere response to shear-induced fracturing of melt (Stasiuk et al., 1996;
(Eichelberger, 1995). The common progression of explosive to Rust et al., 2004; Rust and Cashman, 2007; Tuffen et al., 2003;
effusive eruption requires that magma degassing become more Cabrera et al., 2011; Yoshimura and Nakamura, 2010). According
effective with time such that magma loses most of its volatiles to this hypothesis, strain-rate dependent brittle failure of the
during transit from storage to the surface. This can happen if, for magma prompts diffusive degassing of cm-sized fragments com-
example, syneruptive gas escape occurs through permeable mag- bined with gas flow through cracks, thereby causing the ‘‘inhibi-
matic foam that after releasing its vapor, collapses to form dense lava tion of explosive behavior’’ (Gonnermann and Manga, 2003).
(Eichelberger et al., 1986; Taylor et al., 1983; Westrich et al., 1988; Cabrera et al. (2011) discovered evidence of this degassing
Westrich and Eichelberger, 1994; Eichelberger, 1995). The critical mechanism in the form of H2O-concentration depletions around
a healed microfault in an obsidian bomb and suggested that these
signatures reflect an efficient degassing mechanism that could
n
explain the transition from explosive to effusive eruption in silicic
Corresponding author at: Institute of Geosciences, Becherweg 21, University of
Mainz, 55099 Mainz, Germany. Tel.: þ 49 6131 392 4146. volcanoes. Does shear fracturing of silicic magma really play such
E-mail address: castroj@uni-mainz.de (J.M. Castro). an important role in syn-eruptive magma degassing and forcing

0012-821X/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.epsl.2012.04.024
64 J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69

Fig. 1. Tuffisite-bearing bombs collected from volcán Chaitén. Bombs are typically dense and glassy like (a) and (b), but some are vesicular as shown in (c) (sample ChR-FT-1).
Frames d–f give photomicrographic views of three superjacent frames and show the tuffisite-vein-filling material and fine structure of the fracture margin. Note the lobate
wall structures in (d); these indicate flowage of melt into the vein after it had formed from brittle deformation, and therefore signals the formation of the vein above the
calorimetric glass transition possibly due to shear-induced fracturing. Note also the microvesicular grain in (e), which also indicates viscous deformation following initial
fragmentation. Frame f shows a thin tuffisite vein transecting a pumiceous bomb (sample ChR-FT-1) containing roughly 32 percent irregular shaped vesicles by volume.

eruptive transitions? If so, the fractures and fragments should and in the conduit-filling lavas themselves (Stasiuk et al., 1996;
show evidence of degassing at their margins and should have Tuffen et al., 2003; Heiken et al., 1988; Tuffen and Dingwell,
‘‘lifetimes’’ that allow significant release of vapor on the timescale 2005). Juvenile tuffisites are common at Chaitén volcano, where
of magma ascent and eruption. The fracturing process, further- they occur as ubiquitous blocks (5–50 cm) on the crater rim and
more, would have to repeat through cycles that allow progres- flanks (Fig. 1). Their position atop both the Plinian fall and
sively more gas to bleed off the rising magma (Chouet et al., column-collapse surge deposits suggests the bombs were ejected
2005), either through upward flow or migration through perme- as early as the waning stages of Plinian activity in the second
able conduit walls (Jaupart and Alle gre, 1991). week of May 2008 (Pallister et al., 2010; Lara, 2008). At this time,
Until now, testing the melt-fracture degassing hypothesis has lava extrusion had begun and was punctuated by large explosions
been hampered by a lack of observed eruptions that could provide from the same vent that fueled effusive and ash-venting activity
fresh samples and a timeline for the explosive-to-effusive transi- (La Penna, 2009, pers. comm.); these explosions likely reflect
tion. The 2008 eruption of volcán Chaitén provides both, the vent-clearing blasts that mined out parcels of dense tuffisite-
stringent explosive-to-effusive timeline, and fresh glassy samples bearing magma within the conduit. Tuffisite veins in these bombs
spanning the entire eruption interval. On early 2 May 2008, (cm’s to dm’s in length) contain moderately welded fine ash and
volcán Chaitén explosively erupted hydrous pumice, and 10 days fine glassy lapilli. The vein walls often have lobate structures
later, began to rapidly effuse (  20–100 m3 s  1; Pallister et al., along their margins signaling secondary flowage of melt into the
2010) chemically identical lava concurrently with tephra emis- vein after it had formed (Fig. 1d). Such textural evidence supports
sions (Lara, 2008) and sporadic large explosions (Smithsonian the interpretation that the veins formed in response to shear-
Insitution, 2008). The lava shed more than 90% of its water by the induced failure of magma (Tuffen et al., 2003).
time it emerged (initial H2O  4 wt%; Castro and Dingwell, 2009). Of 35 bombs examined on Chaitén’s rim, 28 were bore tuffisite
Moreover, the large effusive flux (Pallister et al., 2010) suggests a veins. Of these, five were selected for H2O analysis using synchro-
conduit-scale process was acting to degas magma, rather than a tron FTIR. We also analyzed H2O in obsidian from the following
localized mechanism near the conduit margins (Gonnermann and deposits so that tuffisite degassing could be related to the early
Manga, 2003; Rust and Cashman, 2007). and later eruptive phases: (1) Plinian-fall, (2) pyroclastic cone
Here we evaluate the role shear fracturing played in fostering formed during simultaneous sub-Plinian explosive and effusive
magma degassing and the explosive–effusive transition at Chai- activity, and (3) lava formed in early May, 2008.
tén. Shear fracturing is manifested as tuffisite veins in obsidian We measured the water concentrations in and around tuffisite
(Fig. 1), which form when the stress exceeds a critical threshold veins by SFTIR at the Infrared Microspectroscopy Beamline at the
for viscous flow (Gonnermann and Manga, 2003; Tuffen et al., Australian Synchrotron (AS) and Beamline 1.4.3 at the Advanced
2003). We have analyzed H2O-concentrations on tuffisite veins in Light Source (ALS), USA. Measurements were made on spots (3–
order to assess the amount and rates of degassing of magma that 10 mm diameter), along line traverses, and as 2D maps. Experiments
would erupt effusively at volcán Chaitén. were performed in transmission mode on doubly polished wafers
ranging in thickness from 30 to 570 mm with IR microscopes (Bruker
at the AS and NicPlan at ALS) having MCT detectors. Wafers were cut
2. Samples and methods perpendicular to the orientation of fracture walls and the tuffisite
vein was held intact by impregnation in epoxy. Background spectra
Tuffisites are non-welded to completely welded clastic mate- comprising the signal of a sample-free aperture were collected every
rials filling veins in the country rock surrounding silicic conduits 5 min. All experiments were collected with 16 to 128 scans and at
J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69 65

4 cm  1 spectral resolution. Spot-mode measurements were made in


groups of several (5–15) closely spaced ( 10 mm) spots positioned
in discrete grains. This procedure allowed us to assess grain-scale
H2O heterogeneity, which turned out to be very little, and provided a
number of separate points to average. Sample spectra were pro-
cessed using flexi-curve and linear baseline fits to the hydrous-
component peaks. Differences in the absorbance values of hydrous
species bands were typically less than 2% between the two baseline
methods. Magmatic CO2 was not detected in any of the samples.
Absorbance values in conjunction with absorption coefficients
(Ihinger et al., 1994; Newman et al., 1986) and the Beer–Lambert
law were used to determine the concentration of each hydrous
species (H2O or OH  ) in weight fraction. Where possible, we used
the absorbance of the OH  peak at 4512 cm  1 and molecular H2O
peak at 5200 cm  1 to quantify individual and total hydrous
components. The total-H2O content inferred from these bands
was checked against the amount given by the peak at 3570 cm  1.
In most cases, there was good agreement ( o5% relative differ-
Fig. 2. Hydrous speciation in Chaitén pyroclastic and effusive obsidians as deter-
ence) between these totals. On sample wafers that were too thin mined by synchrotron-FTIR. The amount of molecular H2O and OH  groups dissolved
(or dry) to produce good peaks at 5200 cm  1, we used the in the silicate glass are given on the y- and x-axes, respectively. The total water
1630 cm  1 peak to determine the amount of molecular H2O contents are indicated by dashed diagonal guidelines on the figure. The data fall into
(Stolper, 1982), and added that to the value from the two distinct groups, one comprising tuffisite bombs, vein-filling ash and obsidian
lava, and the other including obsidian pyroclasts from the Plinian and sub-Plinian
4512 cm  1 peak to determine the total water. We only used
activity (transitional cone). Because the proportions of OH  and molecular H2O vary
results in which there was o5% relative difference between the with cooling history (e.g., Stolper, 1989) whereby rapid quenching at higher
total water content inferred from the sum of 4512 and 1630 cm  1 temperatures preserves larger amounts of hydroxyl water for a given total water
values and that of 3570 cm  1. Spot spacing in linear profiles was content, the two fields of data reflect relatively fast quenching in the explosive phases
compared to obsidian erupted as tuffisite bombs and lava. The relatively lower
either 5 or 10 mm and 10–30 mm in 2D maps.
amounts of hydroxyl and correspondingly higher molecular water contents in bomb,
tuffisite, and lava flow obsidians is consistent with protracted cooling and accordingly
lower closure temperatures. This interpretation supports the notion that the bombs
3. Results came from what was probably dense, partly degassed magma forming along conduit
margins and then was periodically excavated during pulses of explosive activity. The
continuity between the bomb speciation data and that of the lava flows suggests that
H2O and OH  concentrations plot in two distinct clusters corre-
the formation of lava is the physical and chemical extension of the path followed by
spond to the pyroclastic obsidians, and the ensemble comprised of the bomb and tuffisite magma.
lava, tuffisite bombs, and tuffisite vein ash (Fig. 2). The Plinian and
sub-Plinian pyroclasts contain more OH  at fixed total water content Castro et al., 2005). The total H2O loss at vein and particle margins
than the effusive and tuffisite-bomb obsidians. This pattern reflects is relatively low ( 25% of far field) and appears not to have been
distinct thermal histories corresponding to different magma ascent sufficient to strip significant quantities of water from the melt.
(and cooling) histories (Stolper, 1989; Zhang et al., 1995). Relatively However, the ability of cracks to channel gases derived distally
rapid ascent of the Plinian eruptives resulted in higher quenching could have fostered outgassing, or the expulsion of previously
temperatures, explaining their elevated OH  . By contrast, slower exsolved volatiles to the surroundings (e.g., Stasiuk et al., 1996;
magma ascent enabled protracted cooling and conversion of more Chouet et al., 2005), especially if the cracks were numerous and
OH  to H2O in the tuffisite bombs and lava (Stolper, 1989; Zhang remained open for a sufficient amount of time. Since the H2O-
et al., 1995). These interpretations are verified by cooling rate concentration gradients developed over time, they can be used to
calculations (Zhang et al., 2000) made on the Plinian, tuffisite-bomb, estimate the timescales that the veins were active and also place
and effusive obsidians (Supplementary material; Fig. S1). These data bounds on the local diffusive degassing rates into cracks.
indicate significantly higher cooling rates in Plinian pyroclastic
obsidians (500 K s  1) compared to tuffisite-bomb ( 1 K s  1) and
effusive obsidians ( 0.01 K s  1). Most importantly, the data suggest 4. Discussion
that magma parcels corresponding to the lava and tuffisite bombs
followed similar Pressure–Temperature–time (P–T–t) paths, as their H2O-concentration gradients develop because H2O-solubility
hydrous speciations form a single trend (Fig. 2). Note that because the at the crack margin is lower due to a pressure drop during
total water content retained in obsidian (H2Ot; diagonal lines, Fig. 2) fragmentation. Melt fracturing due to combined shear and tensile
is a strong function of confining pressure (P) through the H2O- stress will create such a low-pressure condition (Tuffen and
solubility relation (e.g., Newman and Lowenstern, 2002), and hydrous Dingwell, 2005). If shortly after cracking, the pressure within
speciation is strongly temperature (T) and cooling rate dependent the cavities equilibrates to near-hydrostatic (Heiken et al., 1988),
(Stolper, 1989), then a trend formed by a time-series of samples, such then the resultant H2O-solubility drop driving diffusion at the
as the tuffisite bombs and later-erupted lavas, could record the time crack wall can be assessed from the solubility of H2O in rhyolite
variation of those two variables, i.e., a P–T–t path. Thus, the glassy melt (Newman and Lowenstern, 2002). For example,  0.8 wt%
tuffisite bombs, by virtue of their forming a hydrous-speciation trend H2O content has a solubility pressure of  5.3 MPa (at 825 1C),
with the lava samples (Fig. 2), likely record an intermediate step in and this corresponds to a depth of about 230 m, assuming magma
the degassing-induced transition to effusive behavior. density of 2300 kg m  3. If the pressure at this depth were to
Fig. 3 shows the H2O distributions in and around tuffisite decrease to hydrostatic during a cracking event, then the new
veins. Here, the H2O-content reflects the sum of molecular-H2O pressure would be 2.3 MPa, assuming a pressure gradient of
and OH  -groups. In all maps, H2O decreases within 50 to 300 mm  0.01 MPa m  1. The associated H2O-solubility shift would favor
of the crack or particle margins, consistent with diffusive degas- an equilibrium melt-H2O content of  0.53 wt%. These values
sing of magmatic water into cracks (e.g., Cabrera et al., 2011; (0.8–0.53 wt%) are comparable to the values measured along the
66 J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69

Fig. 3. H2Ot-concentration maps across (a) tuffisite-bearing obsidian, (b) vesicular rhyolite clast in a tuffisite vein, (c) a dense obsidian fragment with a vein and,
(d) tuffisite-bearing vein. Each profile plots the total water concentration (H2Ot), which was determined either from the broad asymmetrical peak positioned at 3570 cm  1,
or from summing the individual molecular H2O and OH  groups derived from 5200 cm  1 and 4500 cm  1, respectively. Clear white areas in photomicrographs (left and
lower right frames) are dense glass whereas the dark domains are fine tuffisite ash. Colorful spots on photomicrographs indicate analysis points. Vertical dashed lines on
the profiles in (a) and (b) indicate the glass–tuffisite matrix boundaries. All maps show depletion of H2O at the margins of veins or fragments, indicating that H2O-diffusion
occurred from the melt into a low-pressure tuffisite zone.

Fig. 4. 1D-diffusion-model fits to natural H2O concentration profiles. These models correspond to 625 1C. 825 1C results are shown in Supplementary Table 1. Best-fit
diffusion times are shown by the central curve and the bounding curves differ by an order of magnitude in time. (a) CHRIMB_2LHS 7.5 hrs, (b) Tuffvein_6 1.6 hrs,
(c) CHRIMB_3RHS 4.7 hrs and (d) TVDZ_45_2 1 hrs.
J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69 67

H2O profiles, reinforcing the interpretation that tuffisite veins uniformly close to 0.6 wt%. The precise degassing criterion for this
acted as low-pressure sinks for diffusive degassing. and other model runs was a 97% H2O decrease of the interval (i.e.,
We fitted H2O-concentration profiles with solutions to the 1D 0.97  0.2 wt%) marking the end of the simulation. This criterion
chemical diffusion equation in Cartesian coordinates (Castro et al., was chosen so because the sensitivity of the model decreases as
2005) in order to estimate diffusive degassing times. Diffusion is the concentration asymptotically approaches the boundary value
described by concentration at very long times. In principle, it would take
infinite time for the initial 0.8 wt% H2O to decay to the exact
@c
¼ rðDrcÞ ð1Þ 0.6 wt% limit across the full melt parcel. Degassing to lava-like
@t
levels (e.g., from  0.4 to about 0.2 wt% H2O) would take roughly
where c is the H2O concentration (wt%), t is time (s), and D twice as long as the interval from 0.8 to  0.6 wt% because of the
(m2 s  1) is the diffusion coefficient, a function of T, P, and c H2O-compositional dependence of D.
(Zhang et al., 1991; Zhang and Behrens, 2000). Eq. (1) was solved Diffusive degassing will become less effective with time,
numerically by an explicit finite-difference method. The model unless the fracture density is high. When fractures are numerous
assumes that H2O-molecules are the diffusing species, a pre-crack and closely spaced, the degassing rate is enhanced because there
melt-H2O concentration equal to the plateau value away from the are more surfaces to degas from. This point is illustrated in Fig. 5,
crack, and a fixed-H2O concentration at the stationary crack
margin equal to the low value on the profile. The H2O, once
diffused into the crack, is assumed to instantaneously flow away
from the crack so that the crack-boundary H2O concentration
remains fixed. We assume, furthermore, that the preserved
natural concentration profiles did not evolve after the eruption,
that is, we assume diffusion stopped after relatively rapid
quenching of tuffisite bombs.
We ran models at 825 and 625 1C, values which bracket the
eruption temperature (Castro and Dingwell, 2009) and the
approximate calorimetric glass transition (Giordano et al., 2008),
respectively. H2O-concentration profile model-fits (Fig. 4) range
from minutes (at 825 1C) to several tens of hours (at 625 1C; Table
S3, Supplementary material). These times ( 102–105 s) represent
possible crack ‘‘lifetimes’’ in that diffusion would have started at
the moment of fracture and continued until the bombs were
quenched upon eruption. Tuffisite-veined magma that was not
ejected may have had longer degassing intervals, but other
processes could have acted to render the veins impermeable,
such as welding or clogging with ash (Tuffen et al., 2003; Tuffen
and Dingwell, 2005).
The tuffisite vein lifetimes represent possible timescales over
which local diffusive degassing and distal gas channeling could
have taken place. Are the timescales sufficient for fracture degas-
sing to defuse explosive rhyolite eruptions? Clearly, during the
time that they were active, the veins could have allowed passage
of pressurized vapor and pyroclastic material from deeper and
distal parts of the magma, thereby relieving pressure. That the
calculated fracture lifetimes (h) are much shorter than the
duration of sustained pyroclastic venting (several days), suggests
that, if these timescales represent most fracture lifetimes, then
many cycles of fracturing, gas venting, and healing would have to
occur to effectively alleviate magma vapor overpressure in the
conduit (Stasiuk et al., 1996). It is important to note, however,
that despite the likeliness of repeated fracturing, degassing, and
healing (Tuffen et al., 2003), the mere pyroclastic origin of the
bombs suggests that the ability of tuffisite veins to act as high- Fig. 5. (A) Plot of melt-fracture degassing time versus fracture spacing for two
permeability pathways for previously exsolved vapor to escape hypothetical fracture geometries, plane-parallel (pln) and spherical (sph) and two
was at times limited and may have lead to overpressure in the eruption temperatures (see also Fig. S2). The broad vertical gray stripe is the time
between the beginning of the explosive eruption and the first appearance of lava
conduit and explosive fragmentation. at the vent based on direct observations at volcán Chaitén in May 2008. The
The effectiveness of local diffusive degassing from fractures dashed and solid lines show times to degas magma from 0.8 and 0.6 wt% H2O
depends on the pervasiveness of fracturing and the rate of across a range of fracture spacing. The time required to degas magma is indicated
degassing from those fractures. H2O-concentration profiles and by selecting a particular fracture spacing, e.g., 1 mm, and then extrapolating over
to the dashed or solid lines, and finally reading the corresponding time value on
associated model-fits indicate that diffusive degassing is both
the abscissa. Fracture degassing across the stated H2O concentration interval is
spatially limited and time consuming. For example, near-com- clearly faster when the magma is hotter and has a spherical distribution of
plete diffusive degassing of a 200 mm-wide melt parcel from fractures compared to plane-parallel fractures. In (B), a step-wise degassing
0.8 to about 0.6 wt% would require about 5  104 s ( 14 h) at history involving three intervals (from 0.8 to 0.2 wt%) and two fracture geometries
825 1C, and about 5  105 s ( 140 h) at 625 1C. These estimates is investigated. Bold and horizontal arrows show the cumulative time of degassing
for the spherical fracture networks (solid arrows) and the planar fracture networks
were determined by running the diffusion model for a sufficient (dashed). In both cases highly fractured magma (o1 mm spacing) is required to
amount of time so as to reduce the initial composition of 0.8 wt% degas melt within the timeframe observed for the explosive-to-effusive transition
H2O by approximately 0.2 wt%, so that the final composition is at volcán Chaitén.
68 J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69

which shows the effect of fracture spacing on the degassing time Despite the high degree of brecciation extant in this narrow zone,
of a 1 m cube of magma (Supplementary material). In this fracture-degassing would not have likely lead to wholesale
simulation, two or more plane-parallel or spherical shaped cracks degassing of the bulk magma due to the relatively localized
are introduced in a body of melt and then the 1-D diffusion nature of the brecciated zone. In other words, magma in the
equation is solved at the new melt–crack interfaces to track marginal shear zone would have contributed very little (o1 vol%)
amount of diffusive degassing from the cracks with time. The to the large flux of dense, coherent, and degassed lava emerging
diffusion equation is solved either in Cartesian or spherical from the vent (Pallister et al., 2010; Lara, 2008).
coordinates depending on whether plane-parallel or spherical In order for melt-fracturing to effectively degas magma, fractures
shaped crack geometries are modeled, respectively must access pockets of magma that already contain a significant
(Supplementary material). We assume that once H2O diffuses amount of exsolved volatiles, i.e., vesicular magma, as this would tap
across the melt–crack interface, it instantaneously flows away a source of readily expulsed volatiles and create new surfaces
from the crack boundary. In this way, the cracks are assumed to through which diffusive degassing could operate. These conditions
be infinitely permeable, which allows the H2O content in the melt could be met if cracks propagated from the dense conduit margins
at the crack margin to stay constant, thereby upholding the inward towards vesicular magma, or if the melt fractures nucleated
constant composition boundary condition. These conditions com- and grew in the bubbly magma itself (Okumura et al., 2010). The
prise the most favorable (i.e., the fastest) scenario for the melt- presence of rare, but well-preserved tuffisites in vesicular rhyolite at
fracture degassing mechanism to work because the crack-bound- Chaitén (Fig. 1c and e) suggests that melt fracturing of bubbly
ary H2O-concentration remains constant, fostering the largest magma may indeed render degassed magma; the sample shown in
possible composition gradient across the parcel of melt. Calcu- Fig. 1 is among the most degassed ( 0.1 wt%) samples collected at
lated degassing timescales are thus minimum times needed to Chaitén (Supplementary Table S2). We would expect these veins to
alleviate melt of a pre-defined amount of H2O by way of diffusive form by the same shear-induced process, only much more easily
fracture degassing. As the number of cracks increases, the crack than in dense non-vesicular magma due the presence of bubbles,
spacing decreases (see Supplementary Fig. S2) and so does the which promote non-Newtonian rheology (Llewellin et al., 2002),
degassing timescale (Fig. 5A). As seen in Fig. 5A, the time to strain localization (e.g., Okumura et al., 2008, 2009, 2010), reduced
diffusively degas melt from 0.8 to  0.6 wt% decreases by several magma strength (e.g., Martel et al., 2000; Okumura et al., 2010) and a
orders of magnitude ( 4107– o104 s) as the plane-parallel frac- lower viscous-brittle threshold (e.g., Zhang, 1999; Okumura et al.,
ture spacing decreases from about 1 cm to 100 mm. Cracks having 2010; Laumonier et al., 2011). Whether or not this ‘‘foam-fracturing’’
a spherical distribution, analogous to fully brecciated magma, mechanism could prompt effusive behavior remains an open ques-
degas nearly an order of magnitude faster than an arrangement of tion. As shown by Okumura et al. (2010), moderately vesicular
plane-parallel cracks. rhyolite melts subjected to simple shear will fracture and degas
These simulations show that dense, non-vesicular magma can locally, but bulk degassing of the hydrous bubbly magma outboard of
effectively degas if it is thoroughly fragmented. Note, however, the fracture would still require repeated, rapid and efficient healing
that this simulation (Fig. 5A) represents only a single degassing of early formed fractures (Yoshimura and Nakamura, 2010) followed
interval (i.e., from 0.8 to  0.6 wt%) and degassing to the level by fracturing in other hydrous domains.
observed in the May 2008 lava (  0.2 wt%) would require more
time and finer brecciation (Fig. 5B). At Chaitén, the  10-day-long
period (  106 s) preceding the onset of effusion establishes a time 5. Concluding remarks
limit for fracture degassing to operate. In Fig. 5B, we assess the
amount of time required to degas the magma from 0.8 wt% to To conclude, the short length scales over which diffusive-
0.2 wt% using a series of three equal-sized degassing ‘‘steps’’: degassing operates, and the relatively short timescales that the
0.8 to 0.6 wt%, 0.6 to  0.4 wt% and 0.4 to  0.2 wt%. These steps fragment-laden tuffisite veins remain open means that fracturing
were chosen to resemble the approximate H2O-concentration in rhyolite melt cannot serve as a time-effective magma degassing
decreases occurring in response to pressure drops within cracks mechanism. This finding belies recent suggestions that melt-
that form episodically and assuming their internal pressure fracturing effectively promotes the oft-observed explosive-to-
adjusts from lithostatic (or magmastatic) to hydrostatic after a effusive transition at silicic volcanoes (Cabrera et al., 2011). While
short time. We performed the calculation in this way to assess melt fracturing may facilitate local degassing of pulverized
order-of-magnitude estimates of degassing timescale to which, magma strictly at the conduit margins (Gonnermann and
we can compare to the explosive–effusive timeline at Chaitén. Manga, 2003), it could also augment lateral degassing of vesicular
The even-valued pressure drops and correspondingly equal- magma in the conduit interior by virtue of providing long path-
spaced degassing intervals are not meant to imply that this ways through which gas can escape. The effectiveness of shear
specific degassing pattern was followed by the magma. As seen fractures to alleviate volatile overpressure in conduits and stave off
in Fig. 5B, degassing from 0.8 to  0.2 wt% H2O could occur within explosions ultimately requires high fracture densities, textures and
the permitted 10-day-long timeframe if magma were fragmented geometries that foster high permeability, and long fracture lifetimes
to grains smaller than  1 mm and erupting at 825 1C. This prior to their welding shut. Future work should examine the role that
fracture density is probably unrealistically high for an effusive tuffisites in permeable foams play (Eichelberger et al., 1986; Fig. 1c
eruption anywhere within the conduit except for at its margins, and f), as these may provide a faster mechanism for degassing large
where the conditions for shear fracturing are met (Gonnermann volumes of melt in shallow volcanic conduits by virtue of the higher
and Manga, 2003). The width of the fractured zone (d) in a surface area of bubbly magma, which could favor higher rates of
coherent-magma-filled cylindrical conduit can be estimated from diffusive and advective gas loss.
a formula (d ECRm0.1) provided by Gonnermann and Manga
(2003). Here, C is a constant, 0.01 Pa s  0.1, R is the conduit radius
(m) and m is the relaxed melt viscosity (Pa s). For reasonable Acknowledgments
values of melt viscosity (106–108 Pa s) and a conduit radius of
50 m, which we estimated from aerial views of the dome in the The authors thank the Australian Synchrotron and Advanced
first week of its eruption (Smithsonian Institution, 2008), the Light source for the beamtime required to perform this study.
width of the brecciated margins would be between 1 and 3 m. Valuable field support was provided by the USGS Volcano Hazards
J.M. Castro et al. / Earth and Planetary Science Letters 333–334 (2012) 63–69 69

Program and SERNAGEOMIN. The authors appreciate the con- Newman, S., Epstein, S., Stolper, E., 1988. Water, carbon dioxide and hydrogen
structive comments of Caroline Martel and an anonymous isotopes in glasses from the ca. 1340 A.D. eruption of the Mono craters,
California: constraints on degassing phenomena and intitial volatile content. J.
reviewer, which greatly improved the manuscript. HT was sup- Volcanol. Geotherm. Res. 35, 75–96.
ported by a Royal Society University Research Fellowship and BC Newman, S., Lowenstern, J.B., 2002. VolatileCalc: a silicate melt–H2O–CO2 solution
the Marie Curie Action RHEA-254407. model written in Visual Basic for Excel. Comput. Geosci. 28, 597–604.
Okumura, S., Nakamura, M., Tsuchiyama, A., Nakano, T., Uesugi, K., 2008. Evolution
of bubble microstructure in sheared rhyolite: formation of a channel-like
bubble network. J. Geophys. Res. 113, B07208, http://dx.doi.org/10.1029/
Appendix A. Supporting information 2007JB005362.
Okumura, S., Nakamura, M., Takeuchi, S., Tsuchiyama, A., Nakano, T., Uesugi, K.,
2009. Magma deformation may induce non-explosive volcanism via degassing
Supplementary data associated with this article can be found through bubble networks. Earth Planet. Sci. Lett. 281, 267–274.
in the online version at http://dx.doi.org/10.1016/j.epsl.2012.04. Okumura, S., Nakamura, M., Nakano, T., Kentaro, U., Tsuchiyama, A., 2010. Shear
024. deformation experiments on vesicular rhyolite: implications for brittle frac-
turing, degassing, and compaction of magmas in volcanic conduits. J. Geophys.
Res. 115, B06201, http://dx.doi.org/10.1029/2009JB006904.
Pallister, J.S., Diefenbach, A.K., Griswold, J., Munoz, J., Lara, L.E., Valenzuela, C.,
References Burton, W.C., Keeler, R., 2010. Volumes and Eruptions for the 2008–2009
Chaitén Rhyolite Lava Dome. Abstract no. V21D-2350. American Geophysical
Union.
Cabrera, A., Weinberg, R.F., Wright, H.M.N., Zlotnik, S., Cas, R.A.F., 2011. Melt Rust, A.C., Cashman, K.V., Wallace, P.J., 2004. Magma degassing buffered by vapor
fracturing and healing: a mechanism for degassing and origin of silicic flow through brecciated conduit margins. Geology 32, 349–352.
obsidian. Geology 39, 67–70. Rust, A.C., Cashman, K.V., 2007. Multiple origins of obsidian pyroclasts and
Castro, J.M., Dingwell, D.B., 2009. Rapid ascent of rhyolite magma at Chaitén implications for changes in the dynamics of the 1300 B.P. eruption of New-
volcano, Chile. Nature, 461. (doi:10.1038). berry Volcano, USA. Bull. Volcanol. 69, 825–845.
Castro, J.M., Manga, M., Martin, M.C., 2005. Vesiculation rates of obsidian domes Smithsonian Institution, 2008. Chaitén. Bull. Global Volcanism Network 33 (5).
inferred from H2O concentration profiles. Geophys. Res. Lett. 32, L21307, http: Stasiuk, M.V., Barclay, J., Carroll, M.R., Jaupart, C., Ratté, J.C., Sparks, R.S.J., Tait, S.R.,
//dx.doi.org/10.1029/2005GL024029. 1996. Degassing during magma ascent in the Mule Creek vent (USA). Bull.
Chouet, B., Dawson, P., Arciniega-Ceballos, A., 2005. Source mechanism of Volcanol. 58, 117–130.
Vulcanian degassing at Popcatepetl Volcano, Mexico, determined from wave- Stolper, E., 1982. Water in silicate glasses: an infrared spectroscopic study.
form inversions of very long period signals. J. Geophys. Res. 110, B07301 Contrib. Mineral. Petrol. 81, 1–17.
http://dx.doi.org/10.1029/2004JB003524. Stolper, E., 1989. Temperature dependence of the speciation of water in rhyolitic
Eichelberger, J.C., Carrigan, C.R., Westrich, H.R., Price, R.H., 1986. Non-explosive melts and glasses. Am. Mineral. 74, 1247–1257.
silicic volcanism. Nature 323, 598–602. Taylor, B.E., Eichelberger, J.C., Westrich, H.R., 1983. Hydrogen isotopic evidence of
Eichelberger, J.C., 1995. Silicic volcanism: ascent of viscous magmas from crustal rhyolitic magma degassing during shallow intrusion and eruption. Nature 306,
reservoirs. Annu. Rev. Earth Planet. Sci. 23, 41–63. 541–545.
Giordano, D., Russell, J.K., Dingwell, D.B., 2008. Viscosity of magmatic liquids: a Tuffen, H., Dingwell, D.B., Pinkerton, H., 2003. Repeated fracture and healing of
model. Earth Planet. Sci. Lett 271, 123–134. silicic magma generate flow banding and earthquakes? Geology 31,
Gonnermann, H.M., Manga, M., 2003. Explosive volcanism may not be an 1089–1092.
inevitable consequence of magma fragmentation. Nature 426, 432–435. Tuffen, H., Dingwell, D.B., 2005. Fault textures in volcanic conduits: evidence for
Heiken, G., Wohletz, K., Eichelberger, J., 1988. Fracture fillings and intrusive seismic trigger mechanisms during silicic eruptions. Bull. Volcanol. 67,
pyroclasts, Inyo Domes, California. J. Geophys. Res. 93, 4335–4350. 370–387.
Ihinger, P.D., Hervig, R.I., McMillan, P.F., 1994. Analytical methods for volatiles in Westrich, H.R., Stockman, H.W., Eichelberger, J.C., 1988. Degassing of rhyolitic
glasses. In: Carroll, M.R., Holloway, J.R. (Eds.), Volatiles in Magmas. Reviews in magma during ascent and emplacement. J. Geophys. Res. 93, 6503–6511.
Mineralogy, 30; 1994, pp. 67–90. Woods, A.W., Koyaguchi, T., 1994. Transitions between explosive and effusive
Jaupart, C., Alle gre, C.J., 1991. Gas content, eruption rate and instabilities of eruptions of silicic magmas. Nature 370, 641–644.
eruption regime in silicic volcanoes. Earth Planet. Sci. Lett. 102, 413–429. Westrich, H.R., Eichelberger, J.C., 1994. Gas transport and bubble collapse in
Klug, C., Cashman, K.V., 1996. Permeability development in vesiculating magmas: rhyolitic magma: an experimental approach. Bull. Volcanol. 56, 447–458.
implications for fragmentation. Bull. Volcanol. 64, 87–100. Yoshimura, S., Nakamura, M., 2010. Fracture healing in a magma: an experimental
Lara, L., 2008. The 2008 eruption of the Chaitén Volcano, Chile: a preliminary approach and implications for volcanic seismicity and degassing. J. Geophys.
report. Andean Geol. 36, 125–129. Res. 115, B09209, http://dx.doi.org/10.1029/2009JB000834.
Laumonier, M., Arbaret, L., Burgisser, A., Champallier, R., 2011. Porosity redistribu- Zhang, Y., Stolper, E.M., Wasserburg, G.J., 1991. Diffusion of water in rhyolitic
tion enhanced by strain localization in crystal-rich magmas. Geology 39, glasses. Geochim. Cosmochim. Acta 55, 441–456.
715–718. Zhang, Y., Stolper, E.M., Ihinger, P.D., 1995. Kinetics of the reaction H2OþO ¼2OH
Llewellin, E.W., Mader, H.M., Wilson, S.D.R., 2002. The rheology of a bubbly liquid. in rhyolitic and albitic glasses: preliminary results. Am. Mineral. 80, 593–612.
Proc. R. Soc. London A 458, 987–1016. Zhang, Y., 1999. A criterion for the fragmentation of bubbly magma based on
Martel, C., Dingwell, D.B., Spieler, O., Pichavant, M., Wilke, M., 2000. Fragmentation brittle failure theory. Nature 402, 648–650.
of foamed silicic melts: an experimental study. Earth. Planet. Sci. Lett. 178, Zhang, Y., Xu, Z., Behrens, H., 2000. Hydrous species geospeedometer in rhyolite:
47–58. improved calibration and application. Geochim. Cosmochim. Acta 64,
Newman, S., Stolper, E.M., Epstein, S., 1986. Measurement of water in rhyolitic 3347–3355.
glasses: calibration of an infrared spectroscopic technique. Am. Mineral. 71, Zhang, Y., Behrens, H., 2000. H2O diffusion in rhyolite melts and glasses. Chem.
1527–1541. Geol. 169, 243–262.

You might also like