Jessop Et Al., 2016, Linear Fissures and Caldera Rign Produce Pyroclastic Flows

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Earth and Planetary Science Letters 454 (2016) 142–153

Contents lists available at ScienceDirect

Earth and Planetary Science Letters


www.elsevier.com/locate/epsl

Are eruptions from linear fissures and caldera ring dykes more likely
to produce pyroclastic flows?
D.E. Jessop a,b,∗ , J. Gilchrist b , A.M. Jellinek b , O. Roche a
a
Laboratoire Magmas et Volcans, Université Blaise Pascal, CNRS, IRD, OPGC, Clermont-Ferrand, France
b
Department of Earth, Ocean and Atmospheric Sciences, University of British Columbia, Vancouver, Canada

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

Article history: Turbulent volcanic jets are produced by highly-energetic explosive eruptions and may form buoyant
Received 26 November 2015 plumes that rise many tens of kilometres into the atmosphere to form umbrella clouds or collapse to
Received in revised form 27 August 2016 generate ground-hugging pyroclastic flows. Ash injected into the atmosphere can be transported for
Accepted 1 September 2016
many hundreds of kilometres with the potential to affect climate, disrupt global air travel and cause
Available online 4 October 2016
Editor: T.A. Mather
respiratory health problems. Pyroclastic flows, by contrast, are potentially catastrophic to populations
and infrastructure close to the volcano. Key to which of these two behaviours will occur is the extent to
Keywords: which the mechanical entrainment and mixing of ambient air into the jet by large (entraining) eddies
turbulent entrainment forming the jet edge changes the density of the air–ash mixture: low entrainment rates lead to pyroclastic
volcanic jets flows and high entrainment rates give rise to buoyant plumes. Recent experiments on particle-laden
pyroclastic flows (multi-phase) volcanic jets from flared and straight-sided circular openings suggest that the likelihood
caldera formation for buoyant plumes will depend strongly on the shape and internal geometry of the vent region. This
column collapse
newly recognised sensitivity of the fate of volcanic jets to the structure of the vent is a consequence of a
complex dynamic coupling between the jet and entrained solid particles, an effect that has generally been
overlooked in previous studies. Building on this work, here we use an extensive series of experiments on
multi-phase turbulent jets from analogue linear fissures and annular ring fractures to explore whether
the restrictive vent geometry during cataclysmic caldera-forming (CCF) eruptions will ultimately lead a
relatively greater frequency of pyroclastic flows than eruptions from circular vents on stratovolcanoes.
Our results, understood through scaling analyses and a one-dimensional theoretical model, show that
entrainment is enhanced where particle motions contribute angular momentum to entraining eddies.
However, because the size of the entraining eddies scales approximately with vent width, the extent
of entrainment is reduced as the vent width becomes small in comparison to its length. Consequently,
our work shows that for specified mass eruption rates, the high length-to-width ratio vents typical of
CCF events are more likely to produce pyroclastic flows. We suggest that the enigmatic trend in the
geological record for the largest CCF eruptions to produce pyroclastic flows is an expected consequence
of their being erupted through continuous or piece-wise continuous caldera ring fractures.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction the atmosphere and, more critically, on the extent to which ambi-
ent air is entrained into the jet (e.g. Woods, 2010). Both of these
Volcanic jets are turbulent, particle-laden (multi-phase) flows phenomena are hazardous: ash dispersal and fallout can affect avi-
that may rise as a buoyant plume high into the atmosphere, ation, infrastructures, and cause health problems at distances up to
forming an umbrella cloud, or undergo catastrophic gravitational several hundreds of kilometres from the eruption source, whereas
collapse to produce devastating pyroclastic flows (PFs). Which of pyroclastic flows pose a particular threat to populations living on
these behaviours occurs depends on the source conditions govern- the slopes of volcanoes. Furthermore, very large eruptions that
ing the strength and shape of the jet (e.g. Wilson et al., 1980, loft large quantities of ash and aerosols high into the stratosphere
and references therein), the stratification (i.e. density profile) of have been linked to drastic climate change (e.g. Rampino and Self,
1992).
A remarkable feature in the deposits related to the largest
*Corresponding author at: Laboratoire Magmas et Volcans, Université Blaise Pas-
catastrophic caldera-forming (CCF) events is that ignimbrites
cal, CNRS, IRD, OPGC, Clermont-Ferrand, France.
E-mail address: d.jessop@opgc.fr (D.E. Jessop). formed by pyroclastic flows are commonly much more voluminous

http://dx.doi.org/10.1016/j.epsl.2016.09.005
0012-821X/© 2016 Elsevier B.V. All rights reserved.
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 143

Fig. 1. Ratio of ignimbrite volume, V i , to total volume erupted, V t , produced for 22 silicic eruptions as a function of the eruption magnitude, M. As a visual guide, we show
a suggested trend for the ignimbrite/total deposit volume ratio with magnitude as a dashed line. Low M corresponds to eruptions from stratovolcanoes and high M to CCF
eruptions, with a transition at around M ≈ 7.5. We note that few caldera-forming eruptions occur at M < 7.5. The deposits shown are, in order of decreasing M: Fish Canyon
Tuff (FC Tuff); Younger Toba Tuff (YT Tuff); Oruanui Ignimbrite; Cerro Galán Ignimbrite (CGI); Bishop Tuff; Bandelier Tuff (BT); Tambora 1815; Mount Mazama; Kurile Lake;
Santorini (Minoan eruption); Novarupta; Quilotoa; Krakatau 1883; Santa Maria 1902 (SM 1902); Ksudach V (K-V); Lascar (Soncor); Khangar; Pinatubo 1991; Ceboruco (Jala);
Mt. St. Helens 1980-05-18 (MSH 1980); Puyhue–Cordón Caulle 2012 (PCC 2012); Montagne Pelée (Pelée-P1). Three eruptions whose total volumes are likely to be greatly
underestimated are highlighted by the dashed ellipse. References and details for these deposits are given in §5.3 and Table 3.

than deposits resulting from ash fall out from sustained buoy- along continuous (e.g. Mt. Mazama, Bacon, 1983; Long Valley
ant plumes (e.g. Lipman, 1984; Cole et al., 2005; Cas et al., 2011; Caldera, Hildreth and Mahood, 1986; Wilson and Hildreth, 1997)
Ferguson et al., 2013). Furthermore, many caldera-related deposit or partly continuous (e.g. Poris formation, Brown and Branney,
sequences consist of basal ash-fall layers with upper thick ig- 2004; Permian Ora Ignimbrite, Willcock et al., 2013; Bad Step
nimbrites, suggesting a shift in eruptive dynamics (Druitt and Tuff, Branney et al., 1992) ring fractures may follow initial cham-
Sparks, 1984). The size of eruptions can be inferred from the vol- ber decompression (Druitt and Sparks, 1984; Roche and Druitt,
ume of their deposits using the magnitude scale (Mason et al., 2001). Ultimately this drives a protracted eruption that exploits
2004), the ring fractures to form a ring dyke with geometric properties
M = log10 (mt ) − 7 (1) that vary as the eruption proceeds (e.g. Druitt and Sparks, 1984;
Lipman, 1984; Hildreth and Mahood, 1986; Folch and Martí, 2004;
where mt = ρb V t is the total mass of the deposit (expressed in
Kennedy et al., 2008). In addition to ring dyke vent shapes, large
kg), ρb is the bulk density and V t is the total erupted (bulk) vol-
volume and mass eruption rate (MER) eruptions from purely lin-
ume. Some of the best-documented examples of CCF eruptions and
ear vents (e.g. a dyke extending to the surface) have been in-
smaller stratovolcano eruptions are shown in Fig. 1, where we have
ferred purely from field data (e.g. Soldier Meadow Tuff, Korringa,
assumed that all volumes are minimum values and that the ratio
1973) and from field data along with contemporary observations
of ignimbrite (V i ) to total volume, V i / V t , is approximately pre-
(Tarawera 1886, Walker et al., 1984). Until recently, it was not well
served. Our compilation demonstrates a general increase in V i / V t
understood how a linear vent could retain its shape and hence sus-
with M and, further, highlights that CCF eruptions have potentially
tain a high MER as, in general, stresses in the country rock tend
disproportionately high V i / V t even for their relatively high values
to cause the vent to evolve to a more stable, circular form with a
for M, as identified in the geological record. This class of eruptions
generally have M > 7.5 which corresponds roughly to the transi- more restricted MER. However, as Costa et al. (2011) showed, in re-
tion from the “hyperactive” to “super-eruption” regime as defined gions of crustal extension (e.g. Taupo, NZ, Cole et al., 2010; Sierra
by Jellinek and DePaolo (2003) and Jellinek (2014). In the latter, Madre Occidental, Mexico, Aguirre-Diaz and Labarthe-Hernandez,
magma is stored gradually in very large magma chambers, rather 2003), even moderate extensional stress (40 MPa) is sufficient to
than being erupted periodically from stratovolcanoes at about the sustain the linear vent shape and thus produce very large MER
rate it is produced in the mantle. CCF eruptions that ultimately (∼1010 kg/s) eruptions.
occur in this storage regime are driven by the collapse of large Whether the general trends shown in Fig. 1 are significant
calderas and are distinguished in the geological record by volumi- and are thus indicative of distinctive dynamics in caldera-forming
nous pyroclastic flow deposits (ignimbrites) (Mason et al., 2004). eruptions compared to eruptions from circular vents on strato-
Although straight and flared circular vent geometries are typ- volcanoes, or are simply a consequence of preservational bias in
ical of eruptions from stratovolcanoes, eruptions from linear fis- their deposits is unclear. Nevertheless, the data admit a provoca-
sures and ring dykes related to caldera subsidence, which may tive question: are eruption columns from caldera ring fractures
in part be structurally controlled, dominate the geological record inherently unstable because of their large size, or are they un-
of caldera-forming eruptions (e.g. Lipman, 1997; Aguirre-Diaz and stable because eruption column dynamics are strongly affected by
Labarthe-Hernandez, 2003; Mason et al., 2004; Kennedy et al., the linear or annular geometry through which the particle–gas
2008; Cole et al., 2010). During caldera-forming events, subsi- mixture is injected into the atmosphere. In this paper we use ana-
dence of relatively thin, laterally-extensive chamber roofs, and logue experiments and theory to test and confirm the hypothesis
144 D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153

Fig. 2. Schematics of the turbulent structure of a volcanic jet from (left) an annular vent and (right) from a linear vent. We show the time-averaged velocity profiles, u (r , z)
and u (x, z) (red solid line) and a region of recirculation just above the annular vent. (Inset) particle–eddy coupling regimes according to St. Particles with St ∼ 1 can stretch
the eddies to which they are coupled. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)


that the particular mechanics of entrainment into multi-phase vol- ⎨  1 Stokesian – one-way (fluid–particle) coupling
canic jets erupting from partial (piece-wise continuous fissures) St = ∼ 1 Inertial – two-way “critical” coupling

or complete (continuous) caldera ring dykes ultimately favour jet  1 Ballistic – one-way (particle–fluid) coupling
instability and, in turn, the production of pyroclastic flows. An es-
sential conclusion of our work is that the transition in V i / V t –M
Particle motions are affected by flow accelerations and the struc-
space at M ≈ 7.5 in Fig. 1 may be be real: CCF eruptions from ring
tures and momentum fluxes carried by eddies within the flow
dykes are more likely to form devastating pyroclastic flows and
are, in turn, altered by particle accelerations when St ∼ 1 (inset
thick ignimbrite sequences than jets from typical stratovolcanoes
with axisymmetric vents. Critically, we will show that this expec- of Fig. 2 and Burgisser et al., 2005). In this two-way coupling
tation is related to not simply the geometry of volcanic vents but regime, particles are “critically coupled” to the flow, and we will
also dynamically related to the coupling of these jets to the parti- refer to them as “inertial particles”. By contrast, particles with
cles they carry into the atmosphere. St  1 are assumed to be perfectly coupled to eddies and particles
with St  1 are decoupled from the flow entirely and follow bal-
1.1. Particle–fluid interactions listic trajectories. Whereas perfect coupling or sedimentation may
be true of the very smallest or largest particles, respectively, vol-
Generally, most previous works assume implicitly that the ma- canic jets contain pumices, ash and lithic clasts of a very wide
jor effect of particles is to contribute only to the bulk density range of grain sizes, a certain portion of which are critically cou-
of the flow (Veitch and Woods, 2000). Models for jet evolution pled to the flow and may hence influence the entraining prop-
based on the entrainment hypotheses consequently require either erties of the jet (Crowe et al., 1997; Raju and Meiburg, 1997;
that particles are perfectly coupled (i.e. the particles exactly follow
Burgisser et al., 2005). To put these regimes into the volcanological
fluid streamlines) and contribute to the mean physical properties
context, we note that for many Plinian-style eruptions the propor-
of the flow or else that the particles decouple from the flow and
tion of the grain size distribution (GSD) that corresponds to the
sediment quasi-instantaneously from the jet margins and play no
St ∼ 1 condition is significant and may be typically 20–30%. (Fur-
further role in driving flow of the jet (e.g. Woods, 1988). A useful
metric for particle–fluid stress coupling in a turbulent flow is the ther details are given in §3.1 and Appendix A.1, supplementary
Stokes number, material.) In addition to the effects of particle–fluid coupling on
entrainment, particle sedimentation from the jet margins can also
τp ρ p d2p u e affect the mean density and internal dynamics of these flows. Par-
St = = , (2)
τe 18 f μ le ticle sedimentation from the jet margins can be quantified through
which is a ratio of the response time of a particle τ p to flow accel- a settling number, s = v s (d p )/ū where v s (d p ) is the settling ve-
erations imparted on an eddy overturn time τe ∼ le /u e , where u e is locity of particles of size d p (Burgisser et al., 2005). Under most
proportional to the mean vertical flow speed. In (2), ρ p and d p are natural conditions, however, settling of St ∼ 1 particles is a much
the particle density and diameter, respectively, μ is the fluid dy- slower process compared to eddy overturning in the jet and the in-
namic viscosity, f is a drag factor of O (1) (Burgisser et al., 2005). ertial timescales of these particles. Hence sedimentation has a neg-
Three different particle–fluid coupling regimes can be defined ac- ligible effect on the dynamics of the jet (Woods and Bursik, 1991;
cording the value of St: Carazzo and Jellinek, 2012; Jessop and Jellinek, 2014).
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 145

1.2. The combined effect of vent geometry and inertial particles on and increase the size of entraining eddies, hence enhancing en-
entrainment trainment. Inertial particles ejected from straight-sided vents carry
momentum vertically to produce thin, stretched eddies to hence
Typically, the rate of turbulent entrainment is characterised by reduce entrainment. As the combination of vent shape and in-
relating the inflow rate of ambient air to the mean rise rate of ertial particles affects the mixing of the jet, the structure of the
a turbulent jet through an entrainment coefficient (Morton et al., ash cloud and hence the sedimentation regime are affected in turn
1956), (Carazzo and Jellinek, 2012). This result predicts substantially dif-
ferent ash-fall deposit architectures produced by jets from straight-
αe ≡ u in /ū , (3) sided and flared vents and an evolution in sedimentation regime
with vent erosion during an eruption (Carazzo and Jellinek, 2012,
where u in is the speed at which ambient air is entrained into the
JJ14). This picture is in marked contrast to previously proposed
jet and ū is the mean vertical jet velocity (Morton et al., 1956;
models for the dynamics of volcanic jets that generally assume a
Woods, 2010 and Fig. 2). It is commonly assumed that αe is a
point source (e.g. Woods, 1988, 2010).
constant with values in the range 0.1 for momentum dominated
jets to 0.15 for buoyant plumes (Morton et al., 1956; Fischer et al.,
2. A simple model for entrainment dynamics
1979), though more recent work has shown that αe may in fact
vary as the buoyancy, velocity and turbulent shear stress profiles
In order to entrain and mix ambient fluid (such as atmospheric
evolve with height (Kaminski et al., 2005; Paillat and Kaminski,
air) into a turbulent jet, the eddies at the edge of the jet must
2014a).
do work to penetrate, deform, and overturn the density interface
A mechanical clue to the potential for differing dynamics of en- defining the edge of the jet. The work to ultimately irreversibly
trainment into volcanic jets from circular, linear and annular vents mix the jet and ambient fluid is extracted from the kinetic en-
is illustrated in Fig. 2. We infer from experimental investigations ergy (KE) carried by the jet. The ratio of KE available to the
of incompressible jets from annular vents that, just downstream change in gravitational potential energy in overturning the inter-
of the caldera roof, a zone of steady recirculation forms and that face is commonly expressed in terms of a local Richardson number,
the jet remains annular for some distance (e.g. Ko and Chan, 1978; Ri = g  b/ū 2 where b and ū are the local values of jet radius and
Del Taglia et al., 2004). This annular flow causes the dynamics of mean axial velocity, respectively (e.g. Linden, 1973). The reduced
the jets with annular vents to differ from jets with circular vents, gravity is g  = g (ρa − ρ )/ρ0 with ρ , ρa and ρ0 the densities of the
notably in the strain rates close to the vent which, in turn, affect jet, local ambient fluid and a reference location (typically taken to
entrainment into the jet. be the density of the ambient fluid at the level of the vent), respec-
Vent geometry affects the dynamics of a multi-phase volcanic tively, and g the gravitational acceleration. Ri equally expresses the
jet in two ways: i) the cross-sectional area modulates the MER, local balance between stabilising buoyancy and destabilising iner-
and eruptions with high MER tend to produce jets that rise high tial forces (Linden, 1973). However, as a metric for jet strength in
into the atmosphere to form buoyant plumes, whereas low MER our experiments, it is convenient to define Ri in terms of condi-
eruptions tend to form collapsing fountains (Wilson, 1976); ii) vent tions controlled at the source,
shape and internal geometry influences the structure of entrain-
ing eddies as well as the trajectories of inertial particles (that is g 0 L 0
Ri0 = , (4)
with St ∼ 1) with knock-on effects for the ability of the relatively ū 20
dense jet to entrain ambient air and become buoyant (Jessop and
Jellinek, 2014). The effect of vent geometry has been explored in where, for all√variables, a subscript 0 represents a value at the
the context of particle-free, trans-sonic volcanic jets (e.g. Wilson et source, L 0 = A 0 /π s is a length scale for a given nozzle shape
al., 1980; Koyaguchi et al., 2010; Ogden, 2011). Under these condi- defined by the flux through the opening area, A 0 and s = 0 for lin-
tions, vent radius, crater structure and pressure at the vent govern ear vents and s = 1 for annular vents. Rectangular (linear) vents
the MER. are defined by a width, 2l w , and breadth, 2lb , (Fig. 3b) hence
Entrainment into jets from linear vents without particles has A 0 = 4lb2 / where L 0 = 2lb  −1/2 and  = lb /l w . Annular nozzles are
been well studied (e.g. Kotsovinos, 1977; Paillat and Kaminski, defined by a gap size, d, and overall diameter, d0 (Fig. 3c) so that
2014b). In experiments on jets from vents that are much longer A 0 = π d20 /4(2 −  2 ) where the nozzle aspect ratio is  = d/d0 .

than they are wide, velocity profiles across the widths of these Therefore, L 0 = d0 /2 2 −  2 . Ri in the jet evolves with height
flows are similar in form to those characteristic of axisymmet- and, generally, we may write Ri ∝ |Ri0 |. A coefficient of propor-
ric jets (Morton et al., 1956). Consequently, a similar entrainment tionality, β , relates |Ri0 | to Ri so we may therefore write, for an
hypothesis (cf. eq. (3)) can be applied along with similar condi- annular vent,
tions for the likelihood that a volcanic jet will rise as a buoyant  2
plume or collapse to form a pyroclastic flows. Because of the high Ri ū 0 1
β= ∝ √ , (5)
vent aspect ratio of jets erupted from fissures, Glaze et al. (2011) |Ri0 | ū 2 −  2
proposed that such eruptions will inevitably involve more exten-
whereas the definition of L 0 for a rectangular nozzle gives
sive entrainment relative to jets from circular sources for specified
 2
mass eruption rates. However, whether this picture is true and ū 0 √
how it is affected by the particulate component of eruptions is β∝ . (6)

unclear.
In most studies, the explicit and coupled mechanical effects of These two expressions for β indicate the extent to which KE avail-
the three-dimensional geometry of a vent and the presence of in- able at the source (i.e. KE ∼ ū 2 ) is extracted for mixing, depending
ertial particles in the jet on entrainment are neglected. Jessop and on vent geometry,  . However, following JJ14, to simplify discus-
Jellinek (2014), henceforth referred to as JJ14, showed that inertial sion of our results we introduce a more intuitive metric λ = 1/β :
particles influence entrainment because vent geometry influences when λ is large, more energy is supplied to do the work of over-
the trajectories of inertial particles and these, in turn, stretch the turning the stratification and cause mixing.
entraining eddies (see inset of Fig. 2). In particular, inertial par- Steady jets can be described through coupled equations for
ticles ejected from flared vents contribute angular momentum to the conservation of flux of specific volume, q = b2 ū, momentum,
146 D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153

Fig. 3. a) Schematic of experimental set up and the measurements taken. Definitions of dimensions for b) rectangular and c) annular vents. An image of a collapsing fountain
from an experiment using the linear vent.

m = b2 ū 2 , and buoyancy, f = g  b2 ū. Equivalently, a closed system and jets were formed through various straight-sided nozzles which
can also be obtained by replacing f by the conservation of par- had openings that were either annular or rectangular in cross sec-
ticle mass flux, p = φ b2 ū, and the effect of dilution on g  (see tion (Fig. 3b, c). The resulting turbulent jets had a source Reynolds

appendix of Carazzo and Jellinek, 2012). Here, φ is the particle vol- number, Re0 = ū 0 D 0 /ν , of O 103 or greater (Table 1) indicating
ume fraction (i.e. monodisperse GSD) and thus the jet bulk density fully turbulent conditions in these unbounded flows. Here, ν is the
is ρ = (1 − φ)ρa + φ ρs where ρs is the particle density. In this case, kinematic mixture viscosity and the length scale D 0 is the nar-
the effects of environmental stratification are encompassed in the rowest nozzle gap (d or lb ). We recorded the morphology of these
g  parameter. Particle mass flux is a balance between the upward jets and measured the jet opening angle, 2θ , as a function of φ0 ,
transport of particles by the jet and particle loss through sedimen-
Ri0 and St0 as shown in Fig. 3. We explicitly control these source
tation, though the latter is comparatively small for the jets con-
parameters so as to reproduce conditions inferred for natural erup-
sidered here. In the “top-hat” formalism (cf. Morton et al., 1956;
tions (Carazzo and Jellinek, 2012, JJ14, and see also below). In
Woods, 2010) these conservation principles can be expressed as
some experiments, a density stratification was achieved by adding
(for axisymmetric jets)
a lower layer of salty water and carefully filling the rest of the
dq √ dm dp tank with fresh water. This “two-layer” setup has been shown to
= 2αe m, m = g  q2 , = 0. (7)
dz dz dz reproduce the effects of more realistic quasi-linear density gradi-
ents if the density difference and layer height are carefully chosen
It is apparent from these relationships that the jet radius can be
expressed as b = q/m1/2 . The jet angle, θ , is then given by to match the value of the linear gradient (Carazzo and Jellinek,
2012). Although injected mixtures are always denser than the am-
db |Ri0 | bient fluid at the source, entrainment, dilution and particle loss
= 2αe − = tan θ. (8)
dz 2λ can cause the jets to become buoyant plumes (e.g. Woods, 2010;
Carazzo and Jellinek, 2012). Some of our experiments consequently
For a linear jet, an equivalent expression can be obtained by writ-
ing the two-dimensional conservation equations as per Paillat and produce buoyant plumes while others lead to collapsing fountains.
Kaminski (2014a) and differentiating to obtain In other experiments, only fresh water was used to fill the tank
√ (Table 1).
db 4 2 We used particles with a mean diameter of 250 μm and stan-
= √ αe − |Ri0 | = tan θ. (9)
dard deviation of 50 μm, determined by passing the materials
dz π λ
through a graded range of sieves. All the particles were made of
The system of equations (7) is sufficient to model the rise and silica and had a density of 2.5 g/cm3 . The particles were chosen so
spread of turbulent jets as all the details of how fluid is entrained
that, for the largest eddies formed at the source, St0 = O (0.1 − 1).
into the flow are contained within αe (Morton et al., 1956). We
The particle size and typical experimental conditions give s0 =
note that, for multi-phase flows, αe must therefore implicitly de-
O (0.01 − 0.1) (Table 1).
scribe the influence of the source geometry and inertial particles
We measured the angle from the vertical of the jet, θ , from
on entrainment dynamics whereas λ ∝ (ū /ū 0 )2 is a measure of
a stack of 3–5 images taken from a digital camera each taken at
how much KE has been extracted from the flow relative to how
much was present at the source. We will make experimental mea- intervals of 1 s and infer the steady-state entrainment from this
surements of θ as a function of the source conditions and use (8) information via (8). We carefully manufactured our annular vents
to infer how αe and λ vary according to these conditions. to ensure that they were as symmetric as possible so no varia-
tion in entrainment rate linked to varying gap size around the vent
3. Experiments was observed in the experiments. Finally, we noted whether the jet
formed in each experiment formed a buoyant plume or a collaps-
We performed isothermal experiments similar to JJ14 where a ing fountain and these findings are reported in Table 1. Because
particle–water mixture was injected into a tank of water to pro- buoyancy reversals cannot occur in unstratified environments, we
duce a particle-laden jet (Fig. 3a). The injection rate was constant only report collapsing fountains for experiments with stratification.
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 147

Table 1
Experimental conditions used in this study. A dash in the St0 and s0 fields indicates that no particles were present for that experiment. When the tank was stratified, we
report whether the experiment formed a buoyant plume (BP) or a collapsing fountain (CF).

Expt. Nozzle shape  φ0 Re0 St0 s −Ri0 Class


1 Annular 0.09 0.000 10 347 – – 3.46 × 10−7 –
2 Annular 0.09 0.000 14 225 – – 8.42 × 10−6 –
3 Annular 0.09 0.000 13 929 – – 3.50 × 10−4 –
4 Annular 0.09 0.009 18 182 0.92 0.02 4.20 × 10−4 –
5 Annular 0.09 0.009 13 475 0.67 0.03 1.08 × 10−3 –
6 Annular 0.09 0.009 14 014 0.59 0.03 9.19 × 10−3 –
7 Annular 0.09 0.019 12 958 0.56 0.03 1.11 × 10−2 –

8 Annular 0.26 0.000 9 896 – – 1.16 × 10−5 –


9 Annular 0.26 0.000 9 912 – – 1.02 × 10−4 –
10 Annular 0.26 0.003 4 378 0.28 0.06 1.27 × 10−5 BP
11 Annular 0.26 0.009 4 409 0.29 0.06 3.34 × 10−4 BP
12 Annular 0.26 0.019 3 832 0.26 0.07 7.10 × 10−4 CF
13 Annular 0.26 0.019 3 268 0.22 0.08 1.59 × 10−2 CF

14 Annular 0.42 0.000 3 387 – – 6.49 × 10−5 BP


15 Annular 0.42 0.000 6 422 – – 3.66 × 10−4 –
16 Annular 0.42 0.000 3 617 – – 2.64 × 10−2 CF
17 Annular 0.42 0.000 3 398 – – 3.28 × 10−2 CF
18 Annular 0.42 0.009 7 768 0.09 0.12 2.24 × 10−2 –
19 Annular 0.42 0.020 6 260 0.07 0.14 5.40 × 10−3 –
20 Annular 0.42 0.020 7 061 0.08 0.13 5.73 × 10−3 CF
21 Annular 0.42 0.000 4 345 – – 6.03 × 10−3 BP

22 Linear 0.41 0.000 6 495 – – 9.53 × 10−5 –


23 Linear 0.41 0.000 6 571 – – 1.89 × 10−2 –
24 Linear 0.41 0.005 6 373 0.28 0.04 3.23 × 10−5 BP
25 Linear 0.41 0.009 6 524 0.29 0.04 7.77 × 10−4 BP
26 Linear 0.41 0.019 18 435 1.48 0.01 2.33 × 10−4 CF
27 Linear 0.41 0.019 15 970 1.26 0.02 5.80 × 10−4 CF
28 Linear 0.41 0.019 5 061 0.23 0.05 8.56 × 10−3 CF
29 Linear 0.41 0.022 5 800 0.26 0.04 8.16 × 10−5 BP
30 Linear 0.41 0.032 6 002 0.28 0.04 3.53 × 10−5 BP

3.1. Source particle volume fraction and the proportion of St ∼ 1 for narrow and broad/flared vents for particle-free and particle-
particles in natural eruptions laden jets. Also, to aid interpretation, we have included “isoen-
trainment” curves which are solutions to (8) with constant αe and
For our experiments to be fully scaled with natural erup- λ: the vertical differences in the isoentrainment curves represent
tions the values of φ0 need to comparable with those of Plinian- the different values of αe (curves are higher for larger αe ) whereas
style eruptions. However, φ0 is difficult to determine from ob- horizontal differences represent different values of λ, so that the
servations or field studies, so we employed the conduit flow “roll-off” is quicker for smaller λ. The source particle concentra-
model of La Spina et al. (2015) modified for dacitic magmas tion, φ0 is indicated by a colour scale.
(M. de’ Michieli Vitturi, pers. comm.). This model takes the mass In general, θ increases with  of annular nozzles. In addition
fraction of dissolved volatiles, n0 , magma temperature, magma particle-laden jets correspond to larger values of αe and higher
chamber depth and conduit radius (assuming constant width), r0 , values of λ. Particle-free jets from the  = 0.42 annular vent cor-
as inputs and makes predictions of MER and φ0 . We present our respond approximately to iso-entrainment curves with αe ≈ 0.125
full results in Appendix A.1 (supplementary material) and deter- and λ ≈ 1 (Fig. 4a). Approximately particle-free jets from the vent
mine φ0 by matching predicted and observed values of MER, u 0
with  = 0.26 correspond to αe ≈ 0.075, λ ≈ 0.01 whereas higher
and r0 . The model predictions agree well with those of Degruyter
concentration jets from a vent with the same  have αe ≈ 0.125,
et al. (2012) for the 1980-05-18 eruption of Mount Saint Helens
λ ≈ 0.05. The jet angles formed by jets with φ ≈ 0 from the vent
(MSH), USA and suggest that for a range of r0 and n0 typical of
with  = 0.09 suggest that αe ≈ 0.075 and λ ≈ 0.01 whereas more
Plinian eruptions, φ0 should be in the range 1–10%.
concentrated jets have αe ≈ 0.1 and λ ≈ 0.05. We summarise these
Furthermore, we calculated the range of effective St based
results in Table 2.
on the GSD and source/jet conditions in the 1979 eruption of
For rectangular vents, a distinctive feature is that the jet an-
Soufrière, St. Vincent (SSV), and MSH. We note that MSH was about
gles parallel and perpendicular to the long axes of the vent are
100 times greater in eruption magnitude than SSV. Again, details
of our calculations are given in the supplementary material. We not equal (Figs. 4b, 5a, b), and λ that is almost 5 times larger.
find that 32% of the GSD has St = O (1) for the SSV eruption and Furthermore, we find that θ is systematically smaller parallel to
18% for the MSH eruption. Here we infer by St = O (1) the portion the long axis of the jet than perpendicular to the long axis. How-
of the distribution with 0.3 ≤ St ≤ 3. ever, over a vertical distance of about 10–15L 0 the jet width
evolves to be approximately equal in both directions, as indicated
4. Results in Figs. 5a, b where the widths at 15L 0 cm are identical. This
suggests that the jets from a linear vent with  = 0.41 even-
We present our results for θ as a function of Ri0 in Fig. 4. Fig. 4a tually become axisymmetric, which is further evidenced by the
shows results for the annular nozzles of different  , whereas approximately circular deposits (Fig. 5c). We acknowledge, how-
Fig. 4b shows the same parameter space for a rectangular nozzle ever, that smaller  may produce only elliptical deposits. Particle-
with  = 0.41. We show solutions to (8) fitted to the data of JJ14 laden jets from this linear nozzle (φ ≈ 0.02) had αe ≈ 0.1 and
148 D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153

Fig. 4. Measured jet angles as a function of −Ri0 for (a) annular nozzles and (b) a linear nozzle ( = 0.41), where filled and empty symbols represent measurements
taken along the broad and narrow sides of the jet. The different shaped symbols in (top) represent different values of  . The colour of the symbol represents the particle
concentration, φ . We show solutions to (8) (isoentrainment curves) fitted to the data from JJ14 in a) ( = 1) for narrow (straight-sided) nozzles, broad and/or flared nozzles
and particle-free jets. We also show solutions to (8) as dotted/dashed curves to aid interpretation of the new data. (For interpretation of the references to colour in this
figure, the reader is referred to the web version of this article.)

Table 2 5. Discussion and volcanological context


Values of αe and λ for the different vent and particle concentrations tested.
Nozzle  Condition/φ αe λ 5.1. The distinct contributions of vent geometry and inertial particles to
8/9 0.42 Particle free/φ ≈ 0 0.125 0.25
jet dynamics
8/9 0.42 Particle laden – –
10 0.26 Particle free/φ  0.01 0.100 0.01 Our results show that the entrainment coefficient, αe , is af-
10 0.26 Particle laden/φ = 0.019 0.125 0.05 fected by both the particle volume fraction, φ , and the gap size,  .
12 0.09 Particle free 0.075 0.01 Entrainment is affected by gap size in two competing ways: first,
12 0.09 Particle laden 0.100 0.05
small gaps generate an initial maximum eddy size that is small in
11 0.41 φ  0.005 0.075 1 comparison to the jet diameter, for example d0 for a jet from an
11 0.41 φ = 0.019 0.100 1
annular vent. Small eddies carry less angular momentum and con-
11 0.41 φ = 0.032 0.125 1
sequently penetrate the ambient fluid to a relatively lesser extent
leading, in turn, to reduced entrainment. As the largest eddies in a
jet govern entrainment and their size is always constrained by the
λ ≈ 1. For nearly equivalent Ri0 , larger φ increased θ , indicating initial gap size, small gaps act to decrease αe . This is particularly
a larger αe , whereas jets with φ ≈ 0 had a decreased θ , indicating critical for high aspect ratio (i.e. elongated) vents. A second influ-
lower αe . ence of the gap size on entrainment is through the shear strain
Using our estimations for αe and λ for the different vent rate at the source γ̇ = u 0 /l, which is dimensionally the inverse of
shapes,  and φ (resumed in Table 2) we derived the proportion the eddy turnover time or the eddy rotation frequency. For nar-
of source to local KE, noting that (ū /ū 0 )2 ∝ λ f ( ) with f ( ) given row gaps, l is small. Thus, for a given average flow velocity, u 0 ,
by (5) and (6). The effect of  on (ū /ū 0 )2 is shown in Fig. 6. We shear strain rates and eddy rotation rates are high, which enhance
find that, for φ = 0 from an annular vent, (ū /ū 0 )2 decreases with entrainment.
decreasing  . For an annular vent with φ > 0, (ū /ū 0 )2 increases Our results also suggest that the relative balance of the com-
by nearly an order of magnitude as  increases from 0.09 to 0.26. peting effects of gap size on the angular momentum flux carried
Whereas we do not have enough data to establish trends for φ > 0 by eddies and on the eddy turnover time vary with vent geome-
in rectangular vents of any  , for an approximately equivalent  try. For jets from annular vents, a small gap relative to the overall
and φ = 0, (ū /ū 0 )2 is about an order of magnitude greater for a vent diameter (small  ) have small eddies and a reduced αe in
linear vent than an annular vent. turn. For linear vents, by contrast, the eddy size and jet width
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 149

(Fig. 5) consistent with Glaze et al. (2011). This mixing property


of the flow can explain the approximately circular deposits for
 = 0.41 (Fig. 5c). However, the vertical extent over which this
anisotropy in entrainment will persist is expected to increase as 
becomes small (and hence the magnitude of anisotropy becomes
large). Thus, we expect deposits to be increasingly elliptical in
shape as  → 0.
Glaze et al. (2011) argued that, compared to jets from circular
vents, jets from linear vents have a large perimeter compared to
their cross-sectional area and so entrain a greater volume of fluid
for a given mean plume speed. However, our experiments suggest
that for similar conditions of φ0 , Ri0 and approximately equal L 0 to
the experiments of JJ14 (see curves for  = 1 in Fig. 4a), αe is lower
for linear vents. Furthermore, recent work on purely linear jets and
plumes (i.e.  → 0) has shown, assuming Gaussian profiles for ve-
locity and buoyancy, that (Paillat and Kaminski, 2014a, 2014b)
 −1/2 √
1 + ζ2 2 2 6I 4
αe,linear = Ri 1 − (1 + ζ ) + , (10)
2 3 2

∂ f (x∗ , z) ∗
I4 = j (x , z)dx∗ , (11)
∂ x∗
−∞

where ζ is the ratio of the widths of the buoyancy and velocity


profiles, I 4 is a dimensionless integral shape function involving the
dimensionless velocity profile, f (x∗ , z), and turbulent shear stress
profile, j (x∗ , z). A similar expression has been derived for round
jets (Kaminski et al., 2005) where
Fig. 5. Experimental jet formed by a linear vent (expt. 23) looking a) perpendicular
  −1
and b) parallel to the long axis of the vent, the latter having a larger jet angle 2 C
(given in degrees on both images). Scale bars, given in terms of L 0 , are shown in αe,round = Ri 1 − (1 + ζ 2 ) + (12)
both images and the horizontal black bars at z = 15L 0 are of equal length. The pixel 3 2
to physical length values are equal in both images. c) Example of a deposit formed
by a linear vent. We highlight the approximately circular shape of the deposit (solid where C is a combination of shape factors (similar to I 4 ) whose
line) and suggest that smaller  will give more elliptical deposits (dashed lines). value has been shown to be about 0.135 (Kaminski et al., 2005;

Carazzo and Jellinek, 2012). Assuming that 6I 4 = C , comparing
(10) and (12) shows that αe,linear ≤ αe,round for equal Ri, inde-
pendent of ζ . Hence we conclude that linear vents have a lower
entrainment efficiency than circular vents. When φ ≈ 0, our results
for a linear vent with  = 0.41 (αe,linear ≈ 0.075) and an annular
vent with  = 0.42 (αe,round ≈ 0.125) agree with these calculations.
The presence of particles resulted in αe increasing by 25–35%
for all vent geometries we tested. An increase of αe with φ was
also observed by JJ14 for  = 1 vents with of a large diame-
ter (see also Fig. 4a) giving αe = 0.16 compared to αe = 0.15
for particle-free jets. JJ14 deduced that the increased αe is re-
lated to the additional angular momentum imparted by particles
bound to eddies, which cause them to overshoot more deeply into,
deform and engulf the ambient fluid. Additionally, the vent exit
geometry (straight/flared, narrow/broad) focuses the particle mo-
mentum flux, causing eddy stretching: when the vent is narrow
Fig. 6. Prediction of KE in the jet relative to that present at the source as a function and straight, the particle momentum flux acts to vertically stretch
of  . The functions, f ( ) depend on the geometry and  of the vent and are given eddies, thus reducing the degree to which they intrude into the
by (5)–(6). ambient fluid. Adding particles to jets from  = 1 vents with nar-
row diameter therefore decreased αe from 0.15 to 0.09, despite
scale with each other. Consequently the effects of vent width on the increased angular momentum from the particles. Decreasing
the shear strain rate exert the strongest control on entrainment  leads to small eddies so we interpret the increased αe in the
and, accordingly, αe increases with  (Figs. 4 and 6). For example, relatively narrow  = 0.09 vents (where the gap size was approx-
in our experiments where the width of the linear vent was many imately equal to the narrow vent diameter of the experiments
times smaller than its length, the shear strain rate and αe were reported in JJ14) as the gain in angular momentum due to pres-
larger in the direction perpendicular to the long axis (i.e. l = l w ), ence of particles being of lesser effect than the vertical stretching
hence the larger jet spreading angles seen in Fig. 5. of eddies.
We observe that geometric anisotropy in entrainment into jets The work required to overturn and mix the density interface
imparted near the source of linear vents does not persist very high between jet and ambient fluid is extracted from the KE of the jet
up the jet: the turbulent spreading of internal velocity gradients (Linden, 1973). Our results show that more KE is expended to do
causes the jet to become axisymmetric by the time it rises ≈ 15L 0 the work of overturning and mixing when  is small (Fig. 6), due
150 D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153

to the lack of penetration of small eddies into the ambient fluid. shown to cause the block to tilt and shift laterally leading to az-
Hence, small gap sizes produce eddies that are small in comparison imuthal variation in the gap width (Folch et al., 2001; Kennedy and
to the overall jet diameter which thus require more KE to overturn, Stix, 2007; Kennedy et al., 2008). Assuming variation in gap width
mix and entrain fluid into the jet. of the order of what we considered in experiments on linear vents
that showed anisotropic entrainment, it is possible that natural
5.2. Collapse of jets during caldera-forming eruptions caldera eruptions may involve anisotropic entrainment because of
the variation in gap size during uneven sinking. Consequently, col-
Our results can be applied to identify and understand the con- umn collapse would occur soonest and even preferentially where
ditions in which catastrophic caldera-forming eruptions from spa- the vent is narrowest (hence having the locally lowest αe ), which
tially continuous or discontinuous ring fractures are more likely to may show up in the deposit stratigraphy. We note also that during
lead to jet instability, which causes collapse and the production of a piston-type caldera collapse, the width of an annular vent may
pyroclastic flows. The potential for a jet of given source conditions vary over time if collapse of the central piston is driven by steeply
to collapse can be determined from the conservation equations (7). inclined reverse faults, which opens the gap, while marginal blocks
We numerically integrated this system of equations using a 4th- slide along associated normal faults, which closes the gap (Roche
order Runge–Kutta solver with suitable source conditions for q0 , et al., 2000). Such a pattern would lead to complex eruptive feed-
m0 and p 0 (Fig. 7). We consider particle-laden jets that are nega- back effects in the eruptive dynamics, leading to alternating phases
tively buoyant at the source and that will collapse if they do not of low or high entrainment rates and hence possible successive
go through a buoyancy reversal. Hence if, for given initial condi- buoyant jet phases and collapse events.
tions, g  remains negative at all altitudes then we determine that Jets from linear vents collapse at lower Ri0 and φ0 than for
the jet will collapse. round or annular jets with equivalent conditions (cf. data points
Fig. 7 shows a Ri0 –φ0 regime diagram for the formation of for  = 0.42, annular vent and  = 0.41, linear vent in Fig. 7). This
buoyant plumes/collapsing fountain from circular, annular and lin- reflects the comparatively low values of αe estimated from the jet
ear vents. We include our experiments that formed a buoyant angle data. Very large MER eruptions from linear vents have rel-
plume (BP) or a collapsing fountain (CF) (Table 1) as well as those atively large width compared to their length (Costa et al., 2011),
reported in Carazzo and Jellinek (2012) which are for a round so are likely to have  similar to our experiments and so will also
jet with  = 1. We calculate the theoretical transition from buoy- have similar, low αe . Large MER jets from linear vents are there-
ant plumes to collapsing fountains for the present model as per fore prone to collapse as, not only are they very dense as φ is
Carazzo and Jellinek (2012) and show these as curves, though we large for large MER (see Table A.1 of the supplementary material),
use a constant values of αe as determined from our experimen- but they also have relatively small entrainment coefficients and are
tal data (Table 2). We also give the theoretical transition using a hence less stable. This supports field observations of large volumes
variable entrainment coefficient as was reported in Carazzo and of ignimbrite deposits linked to linear vent areas (Korringa, 1973;
Jellinek (2012). It is immediately apparent from the experimen- Aguirre-Diaz and Labarthe-Hernandez, 2003).
tal data that collapse is more likely when φ0 and/or Ri0 are large
but our experiments also indicate that collapses occur in annu- 5.3. Super eruptions produce a greater proportion of pyroclastic flows
lar and linear vents at lower values of φ0 and Ri0 than in circular
vents. The theoretical transition is strongly dependent on the value The relationship between V i / V t and M suggested in Fig. 1 is
of αe , and the transition curves shift towards lower φ0 and Ri0 not likely to be straightforward and, particularly, the transition to
as αe decreases, a trend which is also seen in the experimental CCF regimes marked by the emplacement of massive ignimbrites
data as  decreases (equivalent in itself to decreasing αe ). This at M ≈ 7.5 is not well defined. Accordingly, we test this through
finding supports our observation that αe decreases with decreas- a compilation of 22 silicic eruptions that represent a range of M
ing  for annular vents (Fig. 4a). The theoretical transitions come from 4.6–9.1 for which at least two components out of the outer-
close to capturing the actual transition when the particle fraction caldera ignimbrite, intra-caldera fill and tephra fall volumes are
is greater than about 0.02, but perform more poorly when the known. Our collated data, shown in Fig. 1 and given in Table 3,
particle fraction approaches zero. As Table 2 shows, αe is lower represent the best-documented examples of 13 caldera-forming
for particle-free flows so that we expect a transition from larger eruptions and 9 smaller stratovolcano eruptions. We note that few
αe at large φ0 to lower αe at low φ0 . Hence theoretical models eruptions with M < 7.5 form calderas: Pinatubo 1991, Krakotoa
using a constant entrainment coefficient are poorly suited to cap- 1883 and the Minoan eruption of Santorini are the most famous
turing the rise and spread of particle-laden jets, particularly from examples. As per Mason et al. (2004) we take V t to be the sum of
non-round vents. A variable entrainment model (e.g. Carazzo and the tephra fall, intra-caldera filling and outer caldera ignimbrites.
Jellinek, 2012, shown as dotted curve in Fig. 7) may be more suc- The intra-caldera fill volume by ignimbrite is hard to constrain as
cessful but there is currently no data with which the integral shape i) caldera floor depth of often hard to determine unless there is
factors in (10) and (12) can be constrained when jets contain par- full-depth faulting and ii) intra-caldera fill also contains caldera
ticles or are produced from non-round vents. collapse material (syn- or post-eruptive, e.g. Lipman, 1984) and
In this context, we highlight that calderas have an overall di- hence is often poorly constrained in terms of the actual volume
ameter on the order of tens to hundreds of kilometres for a gap erupted. Assuming that the ignimbrite and total erupted volumes
width (i.e. the ring fault width) on the order of a hundred metres are minimum values and that, in general, the V i / V t ratio is ap-
(Lipman, 1984) so that caldera is of the order of 0.1 or less. The proximately preserved, the volume of ignimbrite produced during
shift in transition curves as  decreases shows that collapse oc- silicic eruptions as a proportion of the total volume generally in-
curs for lower φ0 and Ri. Hence we conclude that, for equal source creases with the eruption magnitude (Fig. 1). Furthermore, the data
conditions, volcanic jets produced by annular ring fractures during suggest that CCF eruptions have potentially disproportionately high
caldera forming eruptions are more prone to collapse and the em- V i / V t even for their high values of M as identified in the geologi-
placement of very-large scale ignimbrite deposits than jets from cal record. These eruptions generally have M > 7.5.
circular vents. Calderas that form in eruptions where M < 7.5 (e.g. Pinatubo
Our experiments involved symmetric annular vents. In nature, 1991) have  ≈ 1 and we expect these eruptions to proceed as
however, during subsidence of piston-style calderas (e.g. Lipman, per stratovolcano eruptions where  = 1. However, along with the
1997), feedback between the sinking block and magma has been Krakatau and Minoan eruptions, there is a distinct preservational
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 151

Fig. 7. Regime diagram describing the theoretical transition between (stable) buoyant plumes and (unstable) collapsing fountains, given by the collapse model (7), as a
function of φ at the source, φ0 , and the source Ri, −Ri0 . The data points are the experiments from this study (cf. Table 1) and from Carazzo and Jellinek (2012) (CJ12) for
a cylindrical vent (i.e.  = 1). The various lines give the theoretical transition from buoyant plumes to collapsing fountains using a top-hat model with different (constant)
values of αe as suggested from our experiments, and the variable entrainment model of CJ12. The arrow shows how they shift to the left as  decreases. Coloured dashed
lines show the approximate transitions indicated by the data. In the legend text, BP = buoyant plume, CF = collapsing fountain. (For interpretation of the references to
colour in this figure, the reader is referred to the web version of this article.)

Table 3
Age and erupted volume of some historic and pre-historic silicic eruption deposits found in the literature. This record represents some of the best-documented examples
of caldera-forming and non-caldera-forming eruptions. We give both the total (bulk) volume, V t , as well as the (bulk) volume that formed ignimbrite deposits, V i , and the
eruption magnitude, M. We also indicate which erupted components went into estimating V t : O – outflow ash sheets/ignimbrite deposit; I – intra-caldera fill (ignimbrite +
lithics); A – ash/tephra deposit.

Volcano/Caldera Eruption/Formation Age Total vol. Ignimbrite V t calc Magnitude References


(Ma) (V t /km3 ) (V i /km3 ) (M/[–])
La Garita (USA) Fish Canyon Tuff 27.8 5000 2500 O+I 9.1 Lipman (1997)
Toba (IDN) Younger Toba Tuff 0.074 2800 2000 O+I+A 8.8 Rose and Chesner (1987)
Taupo (NZL) Oruanui Ignimbrite 0.027 1170 740 O+I+A 8.5 Wilson (2001)
Cerro Galán (ARG) Cerro Galán Ignimbrite 2.2 801 486 O+I 8.3 Folkes et al. (2011)
Long Valley (USA) Bishop Tuff 0.73 750 200 O+I+A 8.3 Hildreth and Mahood (1986)
Valles Caldera (USA) Bandelier Tuff 1.6 298 89 O+I+A 7.9 Cook et al. (2016)
Tambora (IDN) Tambora 1815 0 175 25 O+A 7.6 Self et al. (1984)
Crater Lake (USA) Mt. Mazama 0.007 155.8 39.8 O+A 7.6 Druitt and Bacon (1986)
Kurile Lake (RUS) Kurile Lake 0.008 155 76 O+I+A 7.6 Ponomareva et al. (2007)a
Santorini (GRC) Minoan Er. 0.004 105.7 79.1 O+I+A 7.4 Johnston et al. (2014)
Mount Katmai (USA) Novarupta 0 28 11 O+A 6.8 Hildreth and Fierstein (2000)
Quilotoa (ECU) Quilotoa 0.001 21.3 2.5 O+A 6.7 Mothes and Hall (2008)
Krakatau (IDN) Krakatau 1883 0 20 13.6 O+A 6.7 Sigurdsson et al. (1991)
Santa Maria (GTM) Santa Maria 1902-10-25 0 20 0 O+A 6.7 Williams and Self (1983)
Ksudach (RUS) Ksudach V 0.002 18 3 O+I+A 6.6 Braitseva et al. (1996)
Lascar (CHL) Soncor 0.026 15 5 O+A 6.6 Gardeweg et al. (1998)
Khangar (RUS) Khangar 0.007 12 2 O+A 6.5 Ponomareva et al. (2007)a
Pinatubo (PHL) Pinatubo 1991 0 8.4 5.5 O+A 6.3 Scott et al. (1996)
Ceboruco (MEX) Jala Pumice 0.001 4.4 0.1 O+A 6.0 Browne and Gardner (2005)
Mount Saint Helens (USA) MSH 1980-05-18 0 1.5 0.2 O+A 5.6 Tilling et al. (1993)
Puyhue–Cordón Caulle (CHL) Puyhue–Cordón Caulle 2012 0 1 0 O+A 5.4 Pistolesi et al. (2015)
Mt. Pelée (FRA) P1 0.001 0.2 0 O+A 4.6 Carazzo et al. (2012)
a
References therein.

bias in the deposits of these three events due to their proximity therefore have   1. Owing to the influence of  on jet stability
to open water. For example, strong winds carried much of the ash as discussed above, these eruptions are prone to forming collapsing
from the 1991 Pinatubo eruption away to the southwest where it fountains rather than buoyant plumes, and this tendency increases
deposited in the South China Sea (Paladio-Melosantos et al., 1996). with M as  decreases further. Hence large CCF events deposit rel-
Similar processes may have affected the Krakatau and Minoan atively larger volumes of ignimbrite compared to the total volume
eruptions making volume estimations for the ash-fall component erupted. Thus, even given the assumptions that we make, we pre-
unreliable and potentially underestimating, resulting in higher es- dict a regime change in terms of V i / V t as the eruption style goes
timations of V i / V t . For this reason, we excluded these three events from stratovolcano to CCF events according to the geometry of the
(highlighted by the dashed ellipse) when determining the trends vent.
for V i / V t as a function of M.
These trends in Fig. 1 can be understood in terms of our exper- 6. Conclusions
imental results as follows. Increasing M increases the size of the
volcanic vent (up to M ≈ 7.5) and the caldera outer diameter (for Motivated by the curious trend of cataclysmic, caldera-forming
M > 7.5). However, the width of the ring dyke does not scale with eruptions to emplace very large amounts of ignimbrite relative to
the outer diameter, hence increasing M when M > 7.5 is equiv- the total erupted volume, we undertook a study into the stability
alent to decreasing  in our experiments. The largest eruptions of particle-laden jets under conditions that are commonly found
152 D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153

during the formation of calderas. The fissures and ring dykes that Appendix A. Supplementary material
form during caldera collapse, and transport the gas and ash mix-
ture to the surface, can be characterised as being either linear or Supplementary material related to this article can be found on-
annular vents. Furthermore, catastrophic caldera-forming eruptions line at http://dx.doi.org/10.1016/j.epsl.2016.09.005.
involve a significant proportion of inertial particles, that is particles
with a Stokes number of the order of unity, that are critically cou- References
pled to the flow. Hence both these vent geometries and inertial
particles are critical to the stability of eruption columns and ul- Aguirre-Diaz, G., Labarthe-Hernandez, G., 2003. Fissure ignimbrites: fissure-source
origin for voluminous ignimbrites of the Sierra Madre Occidental and its rela-
timately determine whether an eruption predominantly produces
tionship with basin and range faulting. Geology 31 (9), 773–776.
buoyant plumes or pyroclastic flows. Bacon, C.R., 1983. Eruptive history of Mount Mazama and Crater Lake Caldera, Cas-
We reproduced the dynamics of jets formed in explosive cade Range, USA. J. Volcanol. Geotherm. Res. 18 (1–4), 57–115.
caldera-forming eruptions using scaled laboratory experiments Braitseva, O.A., Melekestsev, I.V., Ponomareva, V.V., Kirianov, V.Y., 1996. The caldera-
forming eruption of Ksudach volcano about cal. AD 240: the greatest explosive
where a particle–water mixture was injected into a tank of wa-
event of our era in Kamchatka, Russia. J. Volcanol. Geotherm. Res. 70 (1–2),
ter. In particular, we varied the shape and aspect ratio of the vents 49–65.
through which these jets were formed in order to reproduce a Branney, M.J., Kokelaar, B.P., McConnell, B.J., 1992. The Bad Step Tuff: a lava-like
range of vent types seen in nature. Our results were combined rheomorphic ignimbrite in a calc-alkaline piecemeal caldera, English Lake Dis-
trict. Bull. Volcanol. 54 (3), 187–199.
with a one-dimensional entrainment model to show that the pres-
Brown, R.J., Branney, M.J., 2004. Event-stratigraphy of a caldera-forming ignimbrite
ence of inertial particles and the aspect ratio of the vent play an eruption on Tenerife: the 273 ka Poris Formation. Bull. Volcanol. 66 (5),
important role in determining the dynamics of volcanic jets and 392–416.
the related entrainment of the ambient fluid or the collapse of Browne, B.L., Gardner, J.E., 2005. Transport and deposition of pyroclastic. Bull. Vol-
the particle-laden mixture. Using this model, we measured en- canol. 67 (5), 469–489.
Burgisser, A., Bergantz, G.W., Breidenthal, R.E., 2005. Addressing complexity in labo-
trainment through an entrainment coefficient along with another ratory experiments: the scaling of dilute multiphase flows in magmatic systems.
parameter that compared the amount of work required to overturn J. Volcanol. Geotherm. Res. 141 (3–4), 245–265.
and mix the density interface forming the edge of the jet to the Carazzo, G., Jellinek, A.M., 2012. A new view of the dynamics, stability and longevity
amount of kinetic energy available to do this mixing. More specif- of volcanic clouds. Earth Planet. Sci. Lett. 325, 39–51.
Carazzo, G., Tait, S., Kaminski, E., Gardner, J.E., 2012. The recent Plinian explosive
ically, our experimental results and analysis favour three specific
activity of Mt. Pelee volcano (Lesser Antilles): the P1 AD 1300 eruption. Bull.
conclusions: Volcanol. 74 (9), 2187–2203.
Cas, R.A.F., Wright, H.M.N., Folkes, C.B., Lesti, C., Porreca, M., Giordano, G., Viramonte,
i) The kinetic energy (KE) extracted from turbulent jets to over- J.G., 2011. The flow dynamics of an extremely large volume pyroclastic flow, the
2.08-Ma Cerro Galan Ignimbrite, NW Argentina, and comparison with other flow
turn the density interface compared to the amount available
types. Bull. Volcanol. 73 (10), 1583–1609.
at the source is large when the vent aspect ratio is small. The Cole, J.W., Milner, D.M., Spinks, K.D., 2005. Calderas and caldera structures: a review.
ratio of KE extracted to KE available decreases as the aspect Earth-Sci. Rev. 69 (1–2), 1–26.
ratio increases. Cole, J.W., Spinks, K.D., Deering, C.D., Nairn, I.A., Leonard, G.S., 2010. Volcanic and
ii) The addition of inertial particles to jets produced by straight- structural evolution of the Okataina Volcanic Centre: dominantly silicic vol-
canism associated with the Taupo Rift, New Zealand. J. Volcanol. Geotherm.
sided annular and linear vents increased the entrainment co- Res. 190 (1–2), 123–135. SI.
efficient because of the additional angular momentum trans- Cook, G.W., Wolff, J.A., Self, S., 2016. Estimating the eruptive volume of a large py-
ported by turbulent particle-laden eddies. roclastic body: the Otowi Member of the Bandelier Tuff, Valles caldera, New
iii) The entrainment coefficient decreases when the aspect ratio Mexico. Bull. Volcanol. 78 (2).
Costa, A., Gottsmann, J., Melnik, O., Sparks, R.S.J., 2011. A stress-controlled mecha-
of the vent decreases because eddies produced in jets from
nism for the intensity of very large magnitude explosive eruptions. Earth Planet.
these vents are small compared to the diameter of the jet and Sci. Lett. 310 (1–2), 161–166.
therefore they penetrate less far into the ambient fluid and Crowe, C., Schwarzkopf, J., Sommerfeld, M., Tsuji, Y., 1997. Multiphase Flows with
hence entrain less fluid. Droplets and Particles. Taylor & Francis.
Degruyter, W., Bachmann, O., Burgisser, A., Manga, M., 2012. The effects of out-
gassing on the transition between effusive and explosive silicic eruptions. Earth
Reduced entrainment stems from mechanical effects imparted Planet. Sci. Lett. 349, 161–170.
as the jet exits low aspect ratio vents and these jets are unable to Del Taglia, C., Blum, L., Gass, J., Ventikos, Y., Poulikakos, D., 2004. Numerical and
undergo a buoyancy reversal. When applied to volcanic jets, these experimental investigation of an annular jet flow with large blockage. J. Fluids
findings support the conclusion that jets from caldera-forming Eng. 126 (3), 375–384.
Druitt, T.H., Bacon, C.R., 1986. Lithic breccia and ignimbrite erupted during the col-
eruptions are more likely to collapse and so produce pyroclastic lapse of Crater Lake Caldera, Oregon. J. Volcanol. Geotherm. Res. 29 (1–4), 1–32.
flows. Therefore we should indeed expect to find that, by volume, Druitt, T.H., Sparks, R.S.J., 1984. On the formation of calderas during ignimbrite
ignimbrite forms a larger portion of the deposits from caldera- eruptions. Nature 310 (5979), 679–681.
forming eruptions but that this relationship is not a straightfor- Ferguson, C.A., McIntosh, W.C., Miller, C.F., 2013. Silver Creek caldera: the tectoni-
cally dismembered source of the Peach Spring Tuff. Geology 41 (1), 3–6.
ward function of eruption magnitude and, particularly, caldera size.
Fischer, H., List, E., Koh, R., Imberger, J., Brooks, N., 1979. Mixing in Inland and
Coastal Waters. Academic Press.
Acknowledgements Folch, A., Codina, R., Marti, J., 2001. Numerical modeling of magma withdrawal dur-
ing explosive caldera-forming eruptions. J. Geophys. Res., Solid Earth 106 (B8),
16163–16175.
The authors thank editor Tamsin Mather and three anony- Folch, A., Martí, J., 2004. Geometrical and mechanical constraints on the formation
mous reviewers for their constructive comments. We also thank of ring-fault calderas. Earth Planet. Sci. Lett. 221 (1–4), 215–225.
Folkes, C.B., Wright, H.M., Cas, R.A.F., de Silva, S.L., Lesti, C., Viramonte, J.G., 2011.
J. Unger, for technical support and help with the experimental set-
A re-appraisal of the stratigraphy and volcanology of the Cerro Galan volcanic
up, M. de’ Michieli Vitturi for providing the conduit flow model system, NW Argentina. Bull. Volcanol. 73 (10), 1427–1454.
and R. Gallo for help in compiling the geological data. DEJ was Gardeweg, M.C., Sparks, R.S.J., Matthews, S.J., 1998. Evolution of Lascar Volcano,
funded by a Région Auvergne-Vancouver (Canada) post-doctoral northern Chile. J. Geol. Soc. 155 (1), 89–104.
fellowship, and experiments reported here were financed by a Glaze, L.S., Baloga, S.M., Wimert, J., 2011. Explosive volcanic eruptions from linear
vents on Earth, Venus, and Mars: comparisons with circular vent eruptions.
grant from the INSU, CNRS 2015 ALEAS programme. JG and AMJ J. Geophys. Res., Planets 116.
were supported by NSERC. This work represents LMV/ClerVolc con- Hildreth, W., Fierstein, J., 2000. Katmai volcanic cluster and the great eruption of
tribution 180. 1912. Geol. Soc. Am. Bull. 112 (10), 1594–1620.
D.E. Jessop et al. / Earth and Planetary Science Letters 454 (2016) 142–153 153

Hildreth, W., Mahood, G.A., 1986. Ring-fracture eruption of the Bishop Tuff. Geol. Pistolesi, M., Cioni, R., Bonadonna, C., Elissondo, M., Baumann, V., Bertagnini, A.,
Soc. Am. Bull. 97 (4), 396–403. Chiari, L., Gonzales, R., Rosi, M., Francalanci, L., 2015. Complex dynamics of
Jellinek, M., 2014. Volcanic bipolar disorder explained. Nat. Geosci. 7 (2), 84–85. small-moderate volcanic events: the example of the 2011 rhyolitic Cordón
Jellinek, A.M., DePaolo, D.J., 2003. A model for the origin of large silicic magma Caulle eruption, Chile. Bull. Volcanol. 77 (1).
chambers: precursors of caldera-forming eruptions. Bull. Volcanol. 65 (5), Ponomareva, V., Melekestsev, I., Braitseva, O., Churikova, T., Pevzner, M., Sulerzhit-
363–381. sky, L., 2007. Late Pleistocene–Holocene volcanism on the Kamchatka peninsula,
Jessop, D.E., Jellinek, A.M., 2014. Effects of particle mixtures and nozzle geometry Northwest Pacific region. In: Eichelberger, J., Gordeev, E., Izbekov, P., Kasahara,
on entrainment into volcanic jets. Geophys. Res. Lett. 41, 1–6. M., Lees, J. (Eds.), Volcanism and Subduction: The Kamchatka Region, 1st edi-
Johnston, E.N., Sparks, R.S.J., Phillips, J.C., Carey, S., 2014. Revised estimates for tion. Wiley.
the volume of the late Bronze Age Minoan eruption, Santorini, Greece. J. Geol. Raju, N., Meiburg, E., 1997. Dynamics of small, spherical particles in vortical and
Soc. 171 (4), 583–590. stagnation point flow fields. Phys. Fluids 9 (2), 299–314.
Kaminski, E., Tait, S., Carazzo, G., 2005. Turbulent entrainment in jets with arbitrary Rampino, M.R., Self, S., 1992. Volcanic winter and accelerated glaciation following
buoyancy. J. Fluid Mech. 526, 361–376. the Toba super eruption. Nature 359 (6390), 50–52.
Kennedy, B., Stix, J., 2007. Magmatic processes associated with caldera collapse at Roche, O., Druitt, T.H., 2001. Onset of caldera collapse during ignimbrite eruptions.
Ossipee ring dyke, New Hampshire. Geol. Soc. Am. Bull. 119 (1–2), 3–17. Earth Planet. Sci. Lett. 191 (3–4), 191–202.
Kennedy, B.M., Jellinek, A.M., Stix, J., 2008. Coupled caldera subsidence and stirring Roche, O., Druitt, T., Merle, O., 2000. Experimental study of caldera formation. J. Geo-
inferred from analogue models. Nat. Geosci. 1 (6), 385–389. phys. Res., Solid Earth 105 (B1), 395–416.
Ko, N.W.M., Chan, W.T., 1978. Similarity in initial region of annular jets – 3 configu- Rose, W.I., Chesner, C.A., 1987. Dispersal of ash in the great Toba eruption, 75 ka.
rations. J. Fluid Mech. 84 (4), 641–656. Geology 15 (10), 913–917.
Korringa, M.K., 1973. Linear vent area of Soldier Meadow Tuff, an ash-flow sheet in Scott, W.E., Hoblitt, R.P., Torres, R.C., Self, S., Martinez, M.M.L., Nillos, T., 1996. Py-
northwestern Nevada. Geol. Soc. Am. Bull. 84 (12), 3849–3865. roclastic flows of the June 15, 1991, climactic eruption of Mount Pinatubo. In:
Kotsovinos, N.E., 1977. Plane turbulent buoyant jets, part 2: turbulence structure. Fire and Mud: Eruptions and Lahars of Mount Pinatubo, Philippines. University
J. Fluid Mech. 81 (JUN9), 45–62. of Washington Press, Seattle, pp. 545–570.
Koyaguchi, T., Suzuki, Y.J., Kozono, T., 2010. Effects of the crater on eruption column Self, S., Rampino, M.R., Newton, M.S., Wolff, J.A., 1984. Volcanological study of the
dynamics. J. Geophys. Res., Solid Earth 115. Great Tambora eruption of 1815. Geology 12 (11), 659–663.
La Spina, G., Burton, M., de’ Michieli Vitturi, M., 2015. Temperature evolution during Sigurdsson, H., Carey, S., Mandeville, C., Bronto, S., 1991. Pyroclastic flows of the
magma ascent in basaltic effusive eruptions: a numerical application to Strom- 1883 Krakatau eruption. Eos 72 (36), 377–381.
boli volcano. Earth Planet. Sci. Lett. 426, 89–100. Tilling, R.I., Topinka, L.J., Swanson, D.A. (Eds.), 1993. Eruptions of Mount St. Helens:
Linden, P.F., 1973. Interaction of a vortex ring with a sharp density interface – Past, Present, and Future. U.S. Geological Survey.
a model for turbulent entrainment. J. Fluid Mech. 60 (3), 467–480. Veitch, G., Woods, A.W., 2000. Particle recycling and oscillations of volcanic eruption
Lipman, P.W., 1984. The roots of ash flow calderas in western North America: columns. J. Geophys. Res., Solid Earth 105 (B2), 2829–2842.
windows into the tops of granitic batholiths. J. Geophys. Res., Solid Earth 89, Walker, G.P.L., Self, S., Wilson, L., 1984. Tarawera 1886, New Zealand – a basaltic
8801–8841. Plinian fissure eruption. J. Volcanol. Geotherm. Res. 21 (1–2), 61–78.
Lipman, P.W., 1997. Subsidence of ash-flow calderas: relation to caldera size and Willcock, M.A.W., Cas, R.A.F., Giordano, G., Morelli, C., 2013. The eruption, pyro-
magma-chamber geometry. Bull. Volcanol. 59 (3), 198–218. clastic flow behaviour, and caldera in-filling processes of the extremely large
Mason, B.G., Pyle, D.M., Oppenheimer, C., 2004. The size and frequency of the largest volume (>1290 km3 ), intra- to extra-caldera, Permian Ora (Ignimbrite) Forma-
explosive eruptions on Earth. Bull. Volcanol. 66 (8), 735–748. tion, Southern Alps, Italy. J. Volcanol. Geotherm. Res. 265, 102–126.
Morton, B.R., Taylor, G.I., Turner, J.S., 1956. Turbulent gravitational convection from Williams, S.N., Self, S., 1983. The October 1902 Plinian eruption of Santa-Maria vol-
maintained and instantaneous sources. Proc. R. Soc. A, Math. Phys. Eng. Sci. 234 cano, Guatemala. J. Volcanol. Geotherm. Res. 16 (1–2), 33–56.
(1196). Wilson, L., 1976. Explosive volcanic eruptions, III: Plinian eruption columns. Geo-
Mothes, P.A., Hall, M.L., 2008. The Plinian fallout associated with Quilotoa’s 800 yr phys. J. R. Astron. Soc. 45 (3), 543–556.
BP eruption, Ecuadorian Andes. J. Volcanol. Geotherm. Res. 176 (1), 56–69. Wilson, C.J.N., 2001. The 26.5 ka Oruanui eruption, New Zealand: an introduction
Ogden, D., 2011. Fluid dynamics in explosive volcanic vents and craters. Earth and overview. J. Volcanol. Geotherm. Res. 112 (1–4), 133–174.
Planet. Sci. Lett. 312 (3–4), 401–410. Wilson, C.J.N., Hildreth, W., 1997. The Bishop Tuff: new insights from eruptive
Paillat, S., Kaminski, E., 2014a. Entrainment in plane turbulent pure plumes. J. Fluid stratigraphy. J. Geol. 105 (4), 407–439.
Mech. 755. Wilson, L., Sparks, R.S.J., Walker, G.P.L., 1980. Explosive volcanic eruptions, IV: the
Paillat, S., Kaminski, E., 2014b. Second-order model of entrainment in planar turbu- control of magma properties and conduit geometry on eruption column be-
lent jets at low Reynolds number. Phys. Fluids 26 (4). haviour. Geophys. J. R. Astron. Soc. 63 (1), 117–148.
Paladio-Melosantos, M.L.O., Solidum, R.U., Scott, W.E., Quiambao, R.B., Umbal, J.V., Woods, A.W., 1988. The fluid dynamics and thermodynamics of eruption columns.
Rodolfo, K.S., Tubianosa, B.S., Reyes, P.J.D., Alonso, R.A., Ruelo, H.B., 1996. Tephra Bull. Volcanol. 50 (3), 169–193.
falls of the 1991 eruptions of Mount Pinatubo. In: Fire and Mud: Eruptions and Woods, A.W., 2010. Turbulent plumes in nature. Annu. Rev. Fluid Mech. 42, 391–412.
Lahars of Mount Pinatubo, Philippines. University of Washington Press, Seattle, Woods, A.W., Bursik, M.I., 1991. Particle fallout, thermal disequilibrium and volcanic
pp. 513–535. plumes. Bull. Volcanol. 53 (7), 559–570.

You might also like