Sei sulla pagina 1di 9

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Earth and Planetary Science Letters 288 (2009) 455–462

Contents lists available at ScienceDirect

Earth and Planetary Science Letters


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e p s l

Effects of wall-rock elasticity on magma flow in dykes during explosive eruptions


A. Costa a,b,⁎, R.S.J. Sparks b, G. Macedonio a, O. Melnik b,c
a
Istituto Nazionale di Geofisica e Vulcanologia, Naples, Italy
b
Department of Earth Sciences, University of Bristol, Bristol, BS8 1RJ, UK
c
Institute of Mechanics, Moscow State University, Moscow, Russia

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

Article history: Magma flow during explosive volcanic eruptions has been described assuming rigid conduits with simple
Received 13 January 2009 cylindrical or planar geometries. Here we study the dynamics of explosive volcanic flows to take account of
Received in revised form 1 October 2009 the role of elastic deformation of the conduit influenced by local magmatic pressure. Three cases are
Accepted 11 October 2009
investigated: a dyke with elliptical cross-section, a cylindrical conduit and a deep dyke connected to a
Available online 1 November 2009
shallow cylinder. The model CPIUC (Macedonio et al., 2005) was used for simulations and generalized to
Editor: R.D. van der Hilst account for elastic deformations of the conduit cross-section area due to magmatic overpressure.
Fragmentation level is typically deeper in a dyke than in a cylinder. For flows in wide dykes pressure at
Keywords: the fragmentation depth can be lower than the surrounding lithostatic pressure by several tens of MPa,
conduit geometry indicating that the wall-rocks of the dyke will be unstable, constraining the dyke width and eventually
explosive eruption blocking the eruption. On the other hand, when the fragmentation level is shallow the corresponding
elastic effect lithostatic pressure is not large enough to close the dyke and eruptions from wide dykes are possible. The
dyke deformation behaviour changes drastically when we assume the conduit is a dyke at depth that evolves to a cylinder near
the surface. In this case even very wide dykes can be stable because the fragmentation level moves into the
cylindrical region where deformation is negligible.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction the conduit. We show that in explosive eruptions, as a consequence of


variations in magmatic pressure, induced changes in discharge rate
Models of explosive volcanic eruptions have commonly assumed and dyke cross-section can be very large.
flows in undeformable cylindrical conduits (e.g., Wilson et al., 1980; In order to isolate the effects of wall-rock elasticity and conduit
Dobran, 1992; Melnik, 2000; Mastin, 2002). Such models utilize shape we will keep the model of conduit flow as simple as possible.
advanced multiphase flow dynamics for understanding of the Also for this reason we will only use two sets of governing parameters
principal controls of the eruption regimes and intensity. However, representative of a silicic and basaltic magma, as reported in Table 1.
the magma transport in the brittle upper crust occurs dominantly in
dykes (e.g., Lister and Kerr, 1991; Rubin, 1995). Cylindrical conduits 2. Conduit model
can develop at shallow levels by near-surface explosive disruption but
the deeper feeding system still likely consists of a dyke or of system of In this section we present a model for magma flows in a conduit of
dykes. Although there have been some models of explosive eruptions elliptical cross-section with major and minor semi-axes changing
along linear fissures (e.g., Walker et al., 1984; Gilberti and Wilson, with depth. Quasi-static elastic deformation of the dyke is accounted
1990; Folch and Martí, 2009) these models have not incorporated the by an analytical solution that couples cross-section area with the
pressure-dependent deformation of the dyke. In the case of extrusive magma pressure. Moreover we assume that vertical variations in
eruptions recent models indicate that marked fluctuations on cross-sectional area of the dyke occur at length-scales that are much
extrusion rate result from a strong feedback between magma larger than the dyke width; therefore cross-section averaged variables
pressure, dyke width and flow rate (Costa et al., 2007a,b). Evidence can be used to describe magma flow in the conduit and vertical
for the stress field changes in the wall-rock of dykes due to variations deformations are much smaller than horizontal ones. The set of cross-
in magma pressure have also been found from studies of fault plane section averaged equations is derived from a generalization of the
solutions (Roman et al., 2006). In this paper we develop a model for equations used in Macedonio et al. (2005). Moreover with respect to
explosive eruptions, which takes account of elastic deformations of Macedonio et al. (2005) we have introduced melt and crystal
compressibility, i.e. the magma behaves as a compressible liquid
even when the pressure is greater than the bubble nucleation
⁎ Corresponding author. Istituto Nazionale di Geofisica e Vulcanologia, Naples, Italy.
Tel.: +39 081 6108 446; fax: +39 081 6108 351. pressure. If the pressure is lower than the nucleation pressure, the
E-mail address: costa@ov.ingv.it (A. Costa). gas exsolves and a liquid–gas mixture becomes much more

0012-821X/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2009.10.006
Author's personal copy

456 A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462

Table 1 where A = π a b is the cross-section area, V is the vertical mixture


Parameters used in the simulations. velocity, g is the gravity acceleration and fft is the friction term
expressed as fft = 4 μρ a
2 2
Notation Description Value + b
V (e.g., Costa et al., 2007a,b) below
a2 b 2
Basaltic Silicic the fragmentation level and fft = 0 above the fragmentation level
xtot Concentration of dissolved gas (wt.%) 2 5 (μ denotes the magma viscosity, here assumed constant). Assuming a
xc Crystal fraction (wt.%) 5 5 homogeneous mixture, magma density is (e.g., Macedonio et al.,
L Conduit length (m) 6000 6000 2005):
LT Cylinder-dyke transition depth (m) 2000 2000
wT Cylinder-dyke transition width (m) 200 200
1 x 1−xe −xc x
ρl0 Density of the melt phase (kg m− 3) 2500 2500 = e + + c ð3Þ
ρc0 Density of the crystal (kg m− 3) 3000 3000 ρ ρg ρl ρc
ρr Host rock density (kg m− 3) 2600 2600
T Magma temperature (K) 1273 1023 where ρg is the gas density, xe is the exsolved gas mass fraction, and xc
μ Effective Magma viscosity (Pa s) 5 · 103 5 · 105 is the crystal mass fraction. The exsolved and the dissolved gas mass
Pch Magma chamber pressure (MPa) 123–183 123–183
fraction can be expressed as:
s Solubility coefficient (Pa− 1/2) 6.8 · 10− 8 4.1 · 10− 6
n Solubility exponent 0.7 0.5
xtot −xd n
G Static host rock rigidity (GPa) 6.0 6.0 xe = ð1−xc Þ; xd = s P ð4Þ
ν Poisson ratio 0.3 0.3 1−xd
β Bulk modulus of melt/crystal (GPa) 10 10
where xtot is the initial total gas mass fraction, xd is the dissolved gas
mass fraction; the exponent n and the constant s in the solubility law
compressible. Here we assume that no supersaturation is required for are assumed to be independent of pressure, but dependent on the
bubble nucleation. magma composition only (see Table 1).
In order to keep the assumptions simple, we assume a fixed Here we assume that the gas phase behaves as a perfect gas and
pressure as the boundary condition at the base of the conduit. This the condensed phases are compressible:
assumption is applicable when a dyke starts from a large magma
ρg = P = ðRg TÞ; ρl = ρl0 ð1 + P = βÞ ð5Þ
source for which the time of pressure change is much longer than
magma ascent time through the conduit. We explore different values
where Rg is the gas constant and T is the temperature; β denotes the
of the magma pressure. Thus we can also implicitly explore temporal
bulk modulus of melt (and/or crystals) and it is assumed to be equal to
variations on the premise that magma chamber pressure slowly
10 GPa, i.e. similar to typical bulk modulus characterizing host rocks
declines with time in explosive eruptions. At the top of the conduit
(e.g., Huppert and Woods, 2002).
choked flow conditions are assumed.
In order to integrate the above equations avoiding the singularity
at the outlet (exit condition) in Eqs. (1) and (2), we introduce the two
2.1. Governing equations
characteristic velocities, namely the “isothermal” sound velocity in
the erupting magma, a, (Macedonio et al., 2005): and the wave
The flow in the conduit is governed by the cross-section averaged
velocity in a deformable tube, c, (e.g., Young, 1808; Shapiro, 1977;
mass and momentum balance equations under quasi-steady state
Paıdoussis, 2006):
conditions. The conduit cross-section is assumed to be an ellipse with
axes a and b variable with depth z. The magma can enter the conduit in    
1 dρ 1 ρ dA
either the homogeneous or the bubbly flow regime, and exits in the ≡ and ≡ ð6Þ
a2 dP T c2 A dP T
particulate flow regime, after fragmentation. For simplicity we assume
that fragmentation occurs when the gas volumetric fraction α reaches and rewrite Eqs. (1) and (2) using the Mach number M = V/a as
a critical value αc corresponding to the bubble maximum packing independent variable. After analytical manipulations we obtain:
(typically αc ≈ 0.75, e.g., Sparks, 1978). However, the choice of a
different fragmentation criterion, such as critical strain rate threshold dP ρa2 ½g + fft −M2 fA 
(e.g., Papale, 1999) or bubble overpressure (e.g., Melnik, 1999) will not =−    ð7Þ
dM M 1 + ρa dP + γ2 ðg + fft Þ−ð1 + ρa dP
da da
M2 ÞfA
produce qualitatively different results. We fix the pressure at the
conduit bottom, defined as Pch (magma chamber top), and assume
choked flow condition at the conduit exit. Since the time-scale of the dz a2 ½1−M 2 ð1 + γ2 Þ
=    ð8Þ
pressure variations at the conduit base (hours) is much longer than the dM M 1 + ρa dP + γ2 ðg + fft Þ−ð1 + ρa dP
da da
M 2 ÞfA
travel time of the magma in the conduit (a few minutes), we can
approximate the flow in the conduit as steady state. Both the gas and 2 2 2

the condensed phase (liquid and crystals) are assumed to have the where γ = ac2 and fA = aA ddAz . Eqs. (7) and (8) are integrated between
same velocity. As a first order approach, the isothermal assumption is the initial Mach number at the conduit inlet, Mi, and the Mach number
justified by the large heat capacity of the magmatic mixture and the at the conduit exit, Mo, which is set equal to Mo = 1/(1 + γ2)1/2,
very effective heat transfer between the condensed phase and rapidly defining the choked flow condition. Since, for convenience, we fix the
expanding gas (Wilson et al., 1980; Buresti and Casarosa, 1989). conduit length, L, instead of the Mach number at the conduit entrance,
Under these assumptions the governing mass and momentum the value of Mi is searched, using standard shooting methods, until the
equations are: integration of Eq. (8) gives z = L at the conduit exit, where M = Mo.
Lengths of the dyke semi-axes a and b depend on the magmatic
∂ overpressure ΔP as follows (Mériaux and Jaupart, 1995; Costa et al.,
ðρAVÞ = 0 ð1Þ 2007a,b):
∂z
ΔP
and aðzÞ = a0 ðzÞ + ½2ð1−νÞb0 ðzÞ−ð1−2νÞa0 ðzÞ ð9aÞ
2G
∂V 1 dP ΔP
V =− −g−fft ð2Þ bðzÞ = b0 ðzÞ + ½2ð1−νÞa0 ðzÞ−ð1−2νÞb0 ðzÞ: ð9bÞ
∂z ρ dz 2G
Author's personal copy

A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462 457

Here ΔP = P − ρrgz is the difference between the local magmatic For estimating the value of rigidity we use the empirical
pressure, and the lithostatic pressure is defined as the overpressure relationship between static and dynamic values for typical ranges of
for positive values and underpressure for negative values, G is the the Young modulus E of rocks (Wang, 2000):
rigidity of wall-rock, ν is Poisson's ratio, a0 and b0 are the initial
unpressurized values of the semi-axes. If we assume a transition
from a dyke at depth to a cylindrical Estatic ðGPaÞ = 0:415 Edynamic ðGPaÞ−1:056: ð10Þ
 conduit the value of a0 is para-
z−LT
meterized as: a0 ðzÞ = A1 arctan + A2 , where LT and wT are the
wT
position and the vertical extent of the transition zone and constants
A1 and A2 are calculated to satisfy conditions a0(L) = R, where R is the Assuming Edynamic ≈ 40 GPa and a Poisson ratio ν = 0.3 we obtain
E
þνÞ ≈6 GPa:
a rigidity Gstatic = 2ð1static
constant radius of the conduit in the cylindrical region, and a0(0) = a0
dyke width at the bottom (see sketch in Fig. 1). The value of b0 can be
either calculated assuming conservation of the cross-section area of
the unpressurized dyke, or can be specified independently. In this way 2.2. Limitations
we can prescribe a transition from a dyke to a cylinder at LT,
approximate the behaviour of dyke in the limit a ≫ b and recover the With the model presented above we will study in a simple way the
case of a cylinder when a = b = R. influence of conduit walls deformation on the eruption dynamics. The

Fig. 1. Sketch of the conduit geometry investigated: i) a conduit having a purely cylindrical cross-section is obtained assuming a0 = b0 = R; ii) a dyke having a purely elliptical cross-
section is obtained assuming a0 ≫ b0 and LT/L = 0; iii) a dyke-to-cylinder conduit is obtained when a0 ≫ b0 and LT/L ≠ 0. Modified after Costa et al. (2007b).
Author's personal copy

458 A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462

model has several important assumptions that must be overcome in and effective viscosity. As will be demonstrated, the depth of the
future research. fragmentation level plays an important role in determining the
dynamics of magma rise in an elastic dyke.
• Conduit flow is assumed to be i) steady state, ii) isothermal, and In the following two sub-sections we compare three different
iii) and quasi-1D. The fragmentation criterion is based on the critical conduit geometries of a conduit having: i) a purely cylindrical cross-
volume concentration of bubbles and there is no gas escape either section, ii) an elongated elliptical cross-section, and iii) an elongated
through the magma or to surrounding wall-rocks. Discussion about elliptical cross-section at depth with a transition to a cylindrical cross-
the validity of some of these assumptions can be found in Melnik section at shallower depth (see Fig. 1).
(2000) and Woods and Koyaguchi (1994).
• Eqs. (9a) and (9b) are based on the assumption that elastic 3.1. Pure elliptical dyke versus pure cylinder
deformation can be locally described in the plain strain approxi-
mation (Muskhelishvili, 1963; Mériaux and Jaupart, 1995). This In this section we compare the main differences between a
implies that relationships (9a) and (9b) can be applied only if the cylindrical conduit and an elliptical dyke.
local length-scales characterizing vertical changes in the conduit Fig. 2 compares typical profiles of three conduits: i) a cylinder with
geometry and in the fluid pressure are much larger than the a radius R = 20 m; ii) a dyke with a0 = 1000 m and b0 = 20 m; and
transversal dimensions of the conduit, i.e. the diameter conduit iii) a dyke with a0 = 1000 m and b0 = 20 m at depth that evolves to a
D = 2R or the dyke thickness HD = 2b. Moreover, local relationships cylinder with a radius R = 20 m at depth LT = 2000 m. The conduit
(Eqs. (9a) and (9b)) are not valid near the free surface where both length, L, is assumed to be 6000 m long in all three cases. Three
boundary and topographic effects can be important. Purely elastic magma chamber pressures of 123 MPa, 153 MPa and 183 MPa are
rock properties are assumed and no rock failure is allowed. This reported (153 MPa corresponds to the lithostatic pressure at the
latter assumption can be assessed a posteriori if solutions are found magma chamber depth L = 6000 m). As expected, deformation is
that are likely to exceed the wall-rock strength. negligible in the case of the pure cylinder, while deformation in the
• The shape of the cross-section area of the conduit is assumed to be dyke, as reflected in variations of cross-sectional area, is pronounced
elliptic. In reality, the dyke cross-section is not necessarily elliptical (Fig. 2b). Pressure variations with depth are similar in both cases with
and can instead have cuspate ends (e.g., Rubin, 1995). A dyke with a pronounced tendency to large underpressures reaching maximum
cuspate ends can only remain open for positive fluid pressures. values at the fragmentation depth (Fig. 2a). In the case of a dyke, the
However, solidification of magma in the region near the tips can minor semi-axis b (i.e., the cross-section) reaches the minimum at the
make dyke geometry more similar to the elliptic cross-section that fragmentation depth forming a constriction, where the underpressure is
we assumed in this study. Moreover, for the case a ≫ b, the flow can at a maximum. The underpressure minimum is much more pronounced
localise in several vents along the dyke and both the cross-section in the dyke. There is no sonic transition at this constriction because, due
averaged approximation and Eqs. (9a) and (9b) can become to large magma viscosities below the fragmentation level, the velocity is
violated. In these cases, for example, magma fragmentation can typically much lower than sonic. The depth of this constriction and
occur at different depths along the dyke. In order to overcome these maximum in underpressure, moves deeper with decreasing pressure in
limitations a fully 3D approach should be adopted, although the magma chamber (Fig. 2c and d). The variations of gas volumetric
modelling is likely to be very complicated and possibly outside of fraction and velocity are similar and are shown in Fig. 2c and d. The
the current computational capabilities. region of maximum underpressure at the fragmentation level is where
• Many near-surface volcanic conduits are approximately cylindrical but conduits wall failure is most likely if the underpressure exceeds some
narrow with depth. Our models keep the conduit diameter constant strength threshold (Macedonio et al., 1994; Barnett, 2008). The
and assume choked flow at the surface. Strictly for a diverging conduit calculated underpressures are very large and exceed the typical tensile
that gets wider with decreasing depth the choke point may be at a strength of rocks of 10 to 20 MPa. The implications of these very large
constriction in the conduit rather than at the vent with supersonic flow underpressures are discussed further below.
in the diverging part. For the dyke the width increases above the These features are qualitatively similar in all the other cases
pressure minimum and the divergence gives the possibility of analyzed below although there are some quantitative differences
supersonic flows. Thus more complex multiphase models will be depending on width, aspect ratio (a0/b0) and shape of the dyke. For
needed to address the effects of flared shapes on flow dynamics. the sake of comparison, we consider the case when the unperturbed
• For sake of simplicity, since we consider cross-section averaged cross-section area is conserved, so that the minor semi-axis b0
variables only, magma properties are treated in an approximate decreases as a0 increases. As expected, deformation is stronger with
way. This includes equilibrium water exsolution, absence of gas increasing a0 (Fig. 3a). Since increasing a0 increases the perimeter,
overpressure with respect to magma pressure and constant the fragmentation level becomes deeper, because magma fragmen-
viscosity assumption. A more realistic description of the effective tation occurs at a fixed pressure in the current model, and the
viscosity should account for the coupling with dissolved water, pressure decreases more rapidly due to the larger friction. After
energy loss, viscous dissipation and two-dimensional effects (e.g., fragmentation the pressure decreases much more slowly mainly due
Costa et al. 2007c; Vedeneeva et al. 2005). to the low density of the particulate dispersion.
Fig. 3b shows the case of a set of conduits having a fixed aspect
3. Results ratio a0/b0 = 50. In this case, since magma pressure decreases as a0
and b0 decrease, the fragmentation level is deeper for smaller dykes in
In case of explosive conduit flows there is a transition from a comparison with larger dykes (see normalized cross-section profile in
laminar flow of bubble-rich melt, in the lower part of the conduit, to a Fig. 3b). Moreover, dykes having the same aspect ratio show a similar
turbulent flow of particle-rich hot gas, in the upper part of the conduit. variation of the minor-axis deformation with depth.
As magma rises from a depth of a few kilometres, its pressure Fig. 3c shows the effect of increasing the values of a0 for a fixed b0
decreases and gas bubbles nucleate. With further decompression the (i.e. b0 = 20 m), from a pure cylinder, a0 = b0, to an elongated ellipse,
mixture accelerates and expands breaking the magma into fragments. a0 ≫ b0. As shown in Fig. 3c, for larger a0 there is a stronger variation
The mixture leaves the vent as a gas-particle jet with a velocity equal of the cross-section b0.
to the local sonic velocity of the mixture (e.g., Melnik, 2000; Mastin, Finally Fig. 3d shows the cross-section variation of two identical
2002). The two zones, above and below the fragmentation level, have unperturbed dykes in the case of two different magmas, i.e. silicic and
different properties such as pressure drop rate, velocity, gas fraction, basaltic (see Table 1). In this case the fragmentation depth becomes
Author's personal copy

A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462 459

Fig. 2. Variation along the conduit of pressure difference ΔP (a), normalized cross-section (b), gas volumetric fraction (c), magma velocity (d), for three different magma chamber
pressures (Pch = 123, 153, and 183 MPa respectively; see labelled values) in the case of the three conduit geometries investigated: cylinder (dashed line), dyke (grey line) and dyke-
to-cylinder (black line). Panel a: profile of overpressure (i.e. pressure inside the conduit minus the lithostatic pressure) for the value of a0 and b0 reported in the inset, dashed line
indicates the rock failure threshold. Panel b: profile of normalized cross-section (i.e., ratio between the pressurized cross-section, πab, and the unperturbed cross-section, πa0b0).
Panel c: Profile of magma gas volumetric fraction. Panel d: Profile of magma velocity inside the conduit. Corresponding discharge rates range from 1 × 106 kg/s to 6 × 106 kg/s
depending on geometry and pressures (see Fig. 4).

shallower in the basaltic case because of the smaller friction and the whereas for the case of the dyke evolving to a cylinder the mass flux is
smaller water content (see Table 1). This implies that in the latter case 4 to 6 times larger.
the dyke can remain open more easily because the lithostatic pressure
at the fragmentation level is smaller. 4. Discussion and implications

3.2. The case of conduit composed of a dyke at the bottom that evolves to Within the limitations of the model, our results show that for a
a cylinder at shallower depths dyke embedded in an elastic medium, magma flow dynamics during
explosive eruptions can differ dramatically from the case of an
The transition from an elongated dyke to a cylinder introduces undeformable conduit typically assumed in most conduit models in
significant differences in terms of deformation, fragmentation depth the literature. This is because the shape of the dyke itself is strongly
and flow rate. In the case of a dyke that evolves to a cylinder a influenced by local magma pressure. These effects are more
constriction in the dyke does not develop because the fragmentation pronounced for long dykes having a large a0/b0 aspect ratio. For the
level becomes shallower, remaining in the cylindrical region where same cross-section area, an elliptical cross-section with large a0/b0
the deformation is negligible (Fig. 2). The deformation in the dyke aspect ratios has a deeper fragmentation level with respect to a
region becomes smoother and the cross-section profile does not show cylindrical cross-section, because the friction term is smaller in the
a constriction as in the case of the purely elliptical cross-section. There latter case. When the fragmentation level is deep (i.e., at large
is of course another kind of constriction at the transition between the lithostatic pressures), the cross-section area decreases markedly near
dyke and cylindrical conduit. The pressure variations with depth are the fragmentation level and forms a constriction where large under-
also very different with much higher pressures being maintained, and pressures are calculated. These results indicate that dyke systems in
the pressure increasing with depth until shallow levels. explosive eruptions are very unstable. Two types of instability are
Besides moving the fragmentation level to shallower depths, the indicated. First the width of the underpressured constriction can tend
transition from a dyke to a cylinder is identified as a very efficient to zero. For the typical values reported in Table 1, a dyke system
mechanism that allows larger mass fluxes for the same magma having a0 of the order of a km tends to collapse even for aspect ratio
chamber pressure (Fig. 4). For example, in the case reported in Fig. 4, a0/b0 ≲ 100. In order to have longer dykes stable, mechanisms that
the mass fluxes of the pure cylinder and pure dyke are similar, produce shallower fragmentation levels need to be considered. Given
Author's personal copy

460 A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462

Fig. 3. Profile of the normalized cross-section (i.e., ratio between the pressurized cross-section, πab, and the unperturbed cross-section, πa0b0) for a set of dykes having the same
unperturbed cross-section (a), the same aspect ratio (c), the same dyke thickness (b), and the same geometry but different magma composition (d). The values of a0 and b0 are
reported in the insets.

that real geological systems are likely heterogeneous a long dyke may calculations commonly exceed the tensile strength of rocks. Thus
have a strong tendency to close locally and thus localise the explosive wall-rock disintegration seems very likely to develop in explosive
flow elsewhere. Second the underpressures are very large and our dykes flows. Again taking account of geological heterogeneity
suggests that local implosion of the dyke walls in the region of the
constriction can also block the conduit and localise flow. For an
eruption from a simple dyke, when deformation at the fragmentation
depth is large, wall-rock erosion is expected to be stronger for magma
flows in dykes in comparison to cylinders.
A very efficient mechanism to stabilise explosive flows is to form a
localised cylindrical flow region in the shallow part of the conduit. If
we assume a conduit that is an elliptical dyke at depth evolving to a
cylinder near the surface, the behaviour of the system changes
dramatically. In fact the presence of a cylindrical region reduces the
friction in the upper part moving the fragmentation level to shallower
depths, where the lithostatic pressure is smaller. The cylindrical
geometry is much more stable and pressures in the flowing magma
tend to be larger; the magma is typically overpressured rather than
underpressured.
As implications, we expect that if an eruption starts from a long
dyke, the dyke tends to be blocked quickly from collapse and the
eruption eventually restarts, after a re-pressurization of the system
from one or few focussed vents along the dyke, or along a portion of
the dyke, or only from an active portion of the dyke that may migrate
Fig. 4. Discharge rate vs magma chamber pressure Pch for the three different conduit
along the fissure during the eruption. For basaltic magma, since the
geometries investigated: cylinder (dashed line), dyke (grey line) and dyke-to-cylinder fragmentation level can be very shallow, the dyke can remain open
(black line). The values of a0 and b0 are reported in the insets. during the eruption, allowing for much more continuous activity.
Author's personal copy

A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462 461

These results show that the evolution of volcanic conduit geometry pressure gradient between surface and the magma chamber and
has a strong impact on the flow dynamics. A number of permutations decreases the friction in the shallow part of the system. All these
of conduit evolution can be envisaged. First, a simple dyke is not a effects can contribute to eruption intensification. In stage 2 of the
stable configuration for explosive flows. We calculate that, as the dyke eruption the pyroclastic fall deposits become finer grained, lithic-rich
increases in length, the width of the constriction tends to decrease and and stratified, and some pyroclastic density currents were generated.
for large aspect ratios the solutions predicted by the model became This stage is consistent with the increasing instability of the conduit
unstable as the constriction width tends to zero. The underpressures walls and repeated choking of the vent, as well as a decrease in
around the constriction also increase as chamber pressure decreases pressure gradient from the chamber as the chamber pressure declines.
and reach values that can exceed the tensile strength of most rocks. In stage 3 there was extrusion of a lava dome, some Vulcanian
Thus in some eruptions the initial constriction could become the focus explosions, and subsidence of the vent fill along ring fractures.
of wall-rock disruption (brecciation), erosion and the supply of lithics
which are then transported by the explosive flow out of the vent. This
5. Conclusions
process could also cause an eruption to stop if the disruption blocks the
vent with lithics. If disruption is localised, as seems likely due to
We have shown that conduit geometry and the coupling between
geological heterogeneity, flow should be thereafter focussed through
magma pressure, flow, and conduit wall deformation have major
this disrupted region. A low pressure cavity structure could form
controls on the dynamics of explosive eruptions. Comparison of
which then acts as region of weakness. One possibility is that overlying
steady-state, one-dimensional models of explosive flows in a
rocks could subside into this cavity and thereby eventually form a
cylindrical conduit and in dykes of comparable cross-sectional area
tapered volcanic pipe to the surface.
shows that much larger underpressures are experienced in the dyke
Evidence for the formation of cavities linked to dykes below and
flow. The dyke models show a pronounced difference between the
volcanic pipes above is observed in kimberlite pipes. Here a deep root
lithostatic pressure and the flow pressure at the fragmentation level
zone is observed at the base of kimberlite pipes (Hawthorne, 1975),
where the dyke width is at a minimum. In some models the dyke
which is typically recognised by the cross-sectional area increasing
tends to zero thickness and when such conditions arise the dyke may
with depth, marked irregularity in cross-sectional shape, and by
close and the eruption could stop or the flow could become localised.
evidence of intense disruption such as breccias and blind re-entrants.
The underpressures are calculated as typically tens of MPa at depths of
Such root zones typically occur at depths of 1 to 3 km, comparable to
a few kilometres and wall-rock failure is anticipated.
the depths calculated in the models for the low pressure constriction.
Models of explosive flows through a dyke connected to a
The study recognises that the formation of a shallow cylindrical
cylindrical near-surface conduit exhibit stabilised explosive flows
conduit fed from a deeper dyke will have a marked influence on
and mostly result in overpressures and fragmentation in the
eruption dynamics. Phreatic eruptions occur early in many explosive
cylindrical conduit above the transition between the deeper dyke
eruptions forming shallow craters that should connect at depth with
and shallower conduit. Flow rates increase by factors of several-fold
the dyke system. We have also identified a mechanism of starting the
for a system with the same cross-sectional area as a simple dyke or
formation for a cylindrical conduit due to severe underpressures in
simple cylinder. Such a configuration could develop early in an
explosive flows in dyke systems and the intrinsic instability that
eruption if a significant cylindrical conduit is formed by phreatic or
should lead to flow focussing. In both cases the changing conduit
phreatomagmatic explosions prior to the dyke reaching the surface.
geometry coupled with strong coupling of magma flow, pressure and
Alternatively this configuration could develop later by wall-rock
cross-section in the dyke play a key role in controlling the eruption.
implosion due to the large underpressures developed in the initial
Some qualitative remarks can be made on the time evolution of
dyke system. These ideas provide a new framework for the
eruptions. In eruptions where previous phreatic or magmatic
interpretation of pyroclastic sequences from major explosive erup-
eruptions have already created a significant and quite deep near-
tions, and may also explain some previously enigmatic features in
surface cylindrical conduit, the initial activity will immediately
kimberlites where a wide root zone is commonly observed beneath a
involve the case of high discharge with overpressured flow. However,
tapering volcanic pipe.
if the dyke reaches the surface at the start of the eruption, the large
underpressures may initiate the formation of a cylindrical conduit by
implosion of the conduit walls at depth, as described here. Acknowledgements
The Huayaputina eruption in Peru in about 1600 AD provides an
example of a Plinian eruption with stratigraphic features and vent AC, OEM and RSJS acknowledge NERC Grants (NE/C509958/1) and
morphology consistent with the processes indicated from the model support from the Royal Society International collaboration fund. RSJS
results. This eruption involved the discharge of about 11 km3 of acknowledges a Royal Society-Wolfson Merit Award and a European
magma in a major Plinian eruption (Adams et al., 2001; Lavallée et al., Research Council Advanced Grant. OEM acknowledges support from
2006). The sequence of events inferred from the stratigraphy by Russian Foundation for the Basic Research (07-01-12041). We are
Lavallée et al. (2006) indicates that the eruption occurred along a grateful to Prof. L. Wilson for useful comments that improved the
fissure from a deep magma chamber, consistent with early dyke flow. clarity of the paper.
The deposits from the first stage of the eruption show several layers of
lithic clasts interpreted by Lavallée et al. (2006) as episodic conduit References
erosion. In terms of our model we suggest that the outbursts of lithics
represent time when the very large underpressures led to wall-rock Adams, N.K., de Silva, S.L., Self, S., Salas, G., Schubring, S., Permenter, J.L., Arbesman, K.,
2001. The physical volcanology of the 1600 eruption of Huaynaputina, southern
implosion and entrainment. The eruption also increased in intensity
Peru. Bull. Volcanol. 62, 493–518.
during stage 1, as indicated by reverse grading of pumice with time, Barnett, W.P., 2008. The rock mechanics of kimberlite volcanic pipe excavation.
consistent with progressive widening and deepening of the a shallow J. Volcanol. Geotherm. Res. 174, 29–39.
Buresti, G., Casarosa, C., 1989. One-dimensional adiabatic flow of equilibrium gas-
cylindrical conduit with time. Increase in intensity with time may
particles mixtures in long vertical ducts with friction. J. Fluid Mech. 203, 251–272.
reflect several processes, including the effect documented in this Costa, A., Melnik, O., Sparks, R.S.J., Voight, B., 2007a. The control of magma flow in dykes on
paper of a geometry where a cylindrical conduit is linked to a dyke at cyclic lava dome extrusion. Geophys. Res. Lett. 34, L02303. doi:10.1029/2006GL027466.
depth. The increasing intensity tendency could also be related to the Costa, A., Melnik, O., Sparks, R.S.J., 2007b. Controls of conduit geometry and wallrock elasticity
on lava dome. Earth Planet. Sci. Lett. 260 (1–2), 137–151. doi:10.1016/j.epsl.2007.05.024.
effects of the shallow cylindrical conduit increasing in depth and Costa, A., Melnik, O., Vedeneeva, E., 2007c. Thermal effects during magma ascent in
width with time, which increases fragmentation depth, increases the conduits. J. Geophys. 112 (B12). doi:10.1029/2007JB004985.
Author's personal copy

462 A. Costa et al. / Earth and Planetary Science Letters 288 (2009) 455–462

Dobran, F., 1992. Nonequilibrium flow in volcanic conduits and application to the Muskhelishvili, N., 1963. Some Basic Problems in the Mathematical Theory of Elasticity.
eruptions of Mt. St. Helens on May 18, 1980 and Vesuvius in AD 79. J. Volcanol. Noordhof, Leiden, The Netherlands.
Geotherm. Res. 49, 285–311. Paıdoussis, M.P., 2006. Wave propagation in physiological collapsible tubes and a
Folch, A., Martí, J., 2009. Time-dependent chamber and vent conditions during proposal for a Shapiro number. J. Fluids Struct. 22, 721–725.
explosive caldera-forming eruptions. Earth Planet. Sci. Lett. 280 (1–4), 246–253. Papale, P., 1999. Strain-induced magma fragmentation in explosive eruptions. Nature
doi:10.1016/j.epsl.2009.01.035. 397, 425–428.
Gilberti, G., Wilson, L., 1990. The influence of geometry on the ascent of magma in open Roman, D.C., Neuberg, J., Luckett, R.R., 2006. Assessing the likelihood of volcanic
fissures. Bull. Volcanol. 52, 515-512. eruption through analysis of volcanotectonic earthquake fault-plane solutions.
Hawthorne, J.B., 1975. Model of a kimberlite pipe. Phys. Chem. Earth 9, 1–15. Earth Planet. Sci. Lett. 248, 244–252. doi:10.1016/j.epsl.2006.05.029.
Huppert, H.E., Woods, A.W., 2002. The role of volatiles in magma chamber dynamics. Rubin, A.M., 1995. Propagation of magma-filled cracks. Annu. Rev. Planet. Sci. 23,
Nature 420, 493–495. 287–336.
Lavallée, Y., de Silva, S., Salas, G., Byrnes, J.M., 2006. Explosive volcanism (VEI 6) without Shapiro, A.H., 1977. Steady flow in collapsible tubes. J. Biomech. Eng. 99, 126–147.
caldera formation: insight from Huaynaputina volcano, southern Peru. Bull. Sparks, R.S.J., 1978. The dynamics of bubble formation and growth in magmas: a review
Volcanol. 68, 333–348. and analysis. J. Volcanol. Geotherm. Res. 3, 1–37.
Lister, J.R., Kerr, R.C., 1991. Fluid mechanical models of crack propagation and their Vedeneeva, E., Melnik, O., Barmin, A., Sparks, R.S.J., 2005. Viscous dissipation in
application to magma transport in dykes. J. Geophys. Res. 96, 10049–10077. explosive volcanic flows. Geophys. Res. Lett. 32 (5). doi:10.1029/2004GL020954.
Macedonio, G., Dobran, F., Neri, A., 1994. Erosion processes in volcanic conduits and an Walker, G.P.L., Self, S., Wilson, L., 1984. Tarawera 1886, New Zealand, a basaltic plinian
application to the AD 79 eruption of Vesuvius, Earth Planet. Sci. Lett. 121, 137–152. fissure eruption. J. Volcanol. Geotherm. Res. 21, 61–78.
Macedonio, G., Neri, A., Martí, J., Folch, A., 2005. Temporal evolution of flow conditions Wang, Z., 2000. Dynamic versus static elastic properties of reservoir rocks: SEG Books,
in sustained magmatic explosive eruptions. J. Volcanol. Geotherm. Res. 143 (1–3), vol. 19, pp. 531–539.
153–172. doi:10.1016/j.jvolgeores.2004.09.015, 2005. Wilson, L., Sparks, R.S.J., Walker, G.P.L., 1980. Explosive volcanic eruptions — IV. The
Mastin, L.G., 2002. Insights into volcanic conduit flow from an open-source numerical control of magma properties and conduit geometry on eruption column behaviour.
model. Geochem., Geophys., Geosyst. 3 (7). doi:10.1029/2001GC000192. Geophys. J. R. Astron. Soc. 63, 117–148.
Melnik, O., 1999. Fragmenting magma. Nature 397, 394–395. Woods, A.W., Koyaguchi, T., 1994. Transitions between explosive and effusive eruptions
Melnik, O., 2000. Dynamics of two-phase conduit flow of high-viscosity gas-saturated of silicic magma. Nature 370, 641–644.
magma: large variations of sustained explosive eruption intensity. Bull. Volcanol. Young, T., 1808. Hydraulic investigations, subservient to an intended Croonian Lecture
62, 153–170. on the motion of the blood. Philos. Trans. R. Soc. Lond. 98, 164–186.
Mériaux, C., Jaupart, C., 1995. Simple fluid dynamics models of volcanic rift zones. Earth
Planet. Sci. Lett. 136, 223–240.

Potrebbero piacerti anche