Вы находитесь на странице: 1из 51

Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

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

Review

Structure of volcano plumbing systems: A review of


multi-parametric effects
Alessandro Tibaldi
Department of Earth and Environmental Sciences, University of Milan-Bicocca, Piazza della Scienza 4, 20126 Milan, Italy

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

Article history: Magma is transported and stored in the crust mostly through networks of planar structures (intrusive sheets),
Received 31 December 2014 ranging from vertical dykes to inclined sheets and horizontal sills, and magma chambers, which make up the
Accepted 21 March 2015 plumbing system of volcanoes. This study presents an overview of plumbing systems imaged at different depths
Available online 11 April 2015
and geodynamic settings, in order to contribute to assessing the factors that control their geometry. Data were
derived from personal field surveys and through the analysis of publications; observations include local lithology
Keywords:
Plumbing system
and tectonics of the host rock with special reference to local fault kinematics and related stress tensor, regional
Dyke tectonics (general kinematics and far-field stress tensors), geology and shape of the volcano, topographic set-
Sill tings, and structural and petrochemical characteristics of the plumbing system. Information from active volca-
Volcano noes and eroded extinct volcanoes is discussed; the shallow plumbing system of active volcanoes has been
Stress reconstructed by combining available geophysical data with field information derived from outcropping sheets,
Magma chamber morphometric analyses of pyroclastic cones, and the orientation and location of eruptive fissures. The study of
eroded volcanoes enabled to assess the plumbing system geometry at deeper levels in the core of the edifice
or underneath the volcano-substratum interface. Key sites are presented in extensional, transcurrent and con-
tractional tectonic settings from North and South-America, Iceland, the Southern Tyrrhenian Sea and Africa.
The types of sheet arrangements illustrated include swarms of parallel dykes, diverging rift patterns, centrally-
inclined sheets, ring and radial dykes, circum-lateral collapse sheets, sills, and mixed members. This review
shows that intrusive sheet emplacement at a volcano depends upon the combination of several local and regional
factors, some of which are difficult to be constrained. While much progress has been made, it is still very challeng-
ing to forecast the likely paths and geometry of sheet propagation and emplacement during volcanic unrest
events.
© 2015 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
2. Definition of a magma plumbing system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3. Propagation and arrest of intrusive sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4. Deep plumbing system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.1. Felsic magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2. Mafic magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5. Shallow plumbing system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6. Plumbing system in the interior of volcanoes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
6.1. Tectonic vs. magmatic components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
6.2. Intrusive sheets and orientation of tectonic stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.2.1. Vertical σ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.2.2. Vertical σ3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.2.3. Vertical σ2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.3. Topographic influence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7. Plumbing systems and surface deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8. Plumbing systems at calderas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

http://dx.doi.org/10.1016/j.jvolgeores.2015.03.023
0377-0273/© 2015 Elsevier B.V. All rights reserved.
86 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

1. Introduction of sheet intrusions from the source magma chamber and their propaga-
tion to the surface, which again requires knowledge of the structure of
This paper addresses the fast-growing field of studies on the struc- the plumbing system and of the host rock.
ture of magma plumbing systems below and inside volcanoes, the Classically, studies of plumbing systems focused on igneous process-
sites where processes like magma transfer, storage and evolution es that are elucidated by sampling of exposed rocks and laboratory anal-
occur. A magma plumbing system can be defined as a network of con- yses, more recently integrated with experimental approaches. The new
duits along which magma moves, interconnected with chambers field of studies of “volcanotectonics” devoted to reconstructing the
where magma accumulates. A plumbing system may be connected structure of plumbing systems uses field data, analogue modelling and
to the surface, at least during transient times, to which magma is numerical modelling. The experimental approaches are rapidly growing
transported producing eruptions. During transfer from the deep fluid in number and quality, but it is fundamental to anchor and validate their
source and subsequent storage in the crust, magmas are subject to a se- results with field truth. In recent years, exposed plumbing systems have
ries of processes that lead to their differentiation. From the shallow been studied in sufficient detail with modern techniques of structural
chambers to the surface, magmas are subject to further processes that geology in order to get information aimed at a better understanding of
eventually dictate the type, intensity and duration of eruption. All the entire process from crustal magma storage to eruption at the
these processes are influenced by a series of parameters that include, surface.
for example, depths of differentiation, amount of wall-rock assimilation, The present paper contributes to the knowledge of magma plumb-
rates and timescales of magma generation, and times of storage (Annen ing systems by reviewing the most relevant literature and integrating
and Zellmer, 2008). Magma storage and ascent are above all tightly it with field data mostly collected by the author. The focus is on the anal-
linked to the structure and state of stress of the crust (Chaussard ysis of the structure of plumbing systems in the uppermost crust and in-
and Amelung, 2014), but in turn magma intrusions might control side volcanoes, where more data obtained from field evidence and
plate boundary evolution (Acocella, 2014). As a consequence, the recon- geophysics are available. This is integrated with a summary of the liter-
struction of the structure and geometry of the plumbing system is of ature on the deeper part of plumbing system above the melt generation
paramount importance for understanding how the subvolcanic engine zone. These data provide a backdrop for understanding the entire
works. magma plumbing system and the processes that take place within it.
The assessment of volcanic hazard is also dependant on the compre- Through a series of examples, from deeply eroded volcanoes to the sur-
hension of the structure of magma plumbing systems. Processes acting face of active volcanoes, I describe the various parameters that control
at open conduit volcanoes, and leading to paroxistic explosions, have the geometry of plumbing systems; these in fact are sensitive to multi-
been recently addressed taking into consideration the structure of the ple factors that frequently work together to dictate the final configura-
shallower conduits. Chouet et al. (1997), for example, studied the tion of the conduit array. Intact cones and volcanoes that experienced
wave fields of tremors and explosions at Stromboli Volcano, Italy, dem- lateral failure or caldera collapse are taken into consideration.
onstrating that the source of this phenomenon is localised beneath One of the most complex conditioning factors is represented by the
the summit crater in the shallower part of the plumbing system at tectonic settings; the tectonic influence on magma migration and volca-
depths b 200 m. The model for degassing and explosion occurrence is nism has received a lot of attention recently, but several issues are still
consistent with a vertical, NE–SW-striking crack-like conduit. This ge- open and controversial. As an example, for decades volcanism and re-
ometry fully coincides with the field observations carried out at gional extensional tectonics have been thought to be tightly linked, as
Stromboli's plumbing system, which crops out in the more dissected
parts of the volcano (Pasquarè et al., 1993; Tibaldi, 1996, 2001). Later,
more detailed geophysical studies on the active conduit (Chouet et al.,
2008) and field data on Holocene conduits (Tibaldi, 2003; Corazzato
et al., 2008) put further constraints on conduit geometry suggesting a
dip towards NE.
Also the evaluation of the areas most prone to the opening of new
vents and eruptive fissures is intimately linked to the understanding
of the structure of plumbing systems, especially at volcano-tectonic
rift zones on volcano slopes (e.g. Bonali et al., 2011). Mafic magma, in
fact, is normally supplied to the surface along planar and mostly
steeply-dipping intrusive sheets that may group to form dyke swarms
(Dieterich, 1988; Carracedo, 1994; Moore et al., 1994; Walter and
Schmincke, 2002), and eventually volcano-tectonic rift zones formed
by hundreds of such parallel dykes (Fiske and Jackson, 1972; Walker,
1999). These rift zones can be studied at the surface by analysing the
orientation and location of eruptive fissures, vents and the morphomet-
ric characteristics of pyroclastic cones (Tibaldi, 1995), as well as by
interferometric methods (Massonnet and Sigmundsson, 2000). At
the same time, much information can be obtained by studying in the
field the eroded parts of the rift zones where sheet intrusions are ex-
posed. Field studies of thousands of sheet intrusions also show that
most sheets become arrested on their way to the surface, and that un-
rest commonly does not lead to an eruption (Fig. 1) (Gudmundsson Fig. 1. A magma plumbing system is represented as a network of vertical, inclined and hor-
and Brenner, 2005). The comprehension of this phenomenon depends izontal conduits that channel magma towards the surface, and a series of chambers where
on a better understanding of the physical conditions for the injection magma can be stored. Main nomenclature is shown. Not to scale.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 87

this type of stress state favours magma upwelling along vertical frac- and crystallisation. At caldera structures also felsic magmas are normal-
tures perpendicular to the regional least principal stress (σ3) that is hor- ly emplaced as ring dykes, although this concept has been argued by
izontal (Anderson, 1951; Cas and Wright, 1987; Watanabe et al., 1999). Legros et al. (2000) who suggest that magma ascent at caldera ring-
For arc volcanism occurring at convergent margins, Nakamura (1977) fissures is less favoured than at more restricted conduits.
stated that the overall tectonics of the arcs should be strike-slip (with These data indicate that the most of the magma is generally
σ3 and greatest principal stress, σ1, both horizontal) instead of compres- transported towards the surface through conduits that can be repre-
sional reverse (σ3 vertical). Strike-slip tectonics would allow magma to sented by tabular intrusions (sheets) with planar, curved, en-échelon,
ascend through vertical dykes parallel to the direction of σ1 (Nakamura or more complex geometries, but that invariably have a very high length
and Uyeda, 1980). By contrast, a pure contractional tectonic environ- (and depth) to thickness ratio. A magmatic sheet is thus a crystalline
ment, with reverse or transpressional faulting, is usually regarded as a rock that formed in a crack, primarily caused by the magma itself, in a
highly unfavourable setting for volcanism (Glazner, 1991; Hamilton, pre-existing rock body. In the field, intrusive sheets are classically divid-
1995; Watanabe et al., 1999), where only intrusive emplacement is ex- ed into three end-members based on their attitude and the relations
pected (Cas and Wright, 1987), although in reality local magma stress with the country rocks: vertical to steep-dipping dykes, inclined sheets,
can give rise to eruption also in this setting. In the present paper I inves- and sills, formed when magma is injected between rock layers forming a
tigate also these issues through examples that encompass different horizontal or gently-dipping sheet. In origin, dykes have been described
geodynamic settings including compressional and extensional regimes as tabular bodies that intrude normally or obliquely to the bedding
with normal, strike-slip and reverse fault kinematics. Plumbing systems plane of the host rock, to be called also “discordant intrusion”, while
form a major component of diverging plate margins and orogenic belts the term sill indicated a sub-parallel or “concordant intrusion”
at converging margins, and can be highly varied both spatially and tem- (Billings, 1972; Best, 1982; Hall, 1987). Other authors, like Bates and
porally, which has led to a number of controversies about their architec- Jackson (1987) adopted a terminology consistent with the intrusion at-
ture and evolution. titude and referred to dykes as vertical tabular intrusions and to sills as
The paper is structured in order to present a review and analysis of horizontal ones. Hatayama et al. (1980) define a dyke as a vertical intru-
many existing data and models, by introducing magma plumbing sys- sion, a sheet as a horizontal intrusion, and a sill as a horizontal and con-
tems and their settings, and the possible influencing parameters at dif- cordant intrusion. The definition of these parts of a plumbing system as
ferent depths. First we will investigate the deeper level from the fluid based on their relations with the bedding of the host rock is misleading,
source zone, then the upper crust zone, and finally the interior of volca- also because it implies that dykes have to produce cracks in order to
noes and their surface. propagate across the bedding, whereas sill intrusion is easier along
pre-existent weakness zones like bedding discontinuities. Although
2. Definition of a magma plumbing system some authors consider that magma filling already existing fractures is
an important dyke intrusion mechanism (e.g. Delaney et al., 1986;
Volcanoes are underlain by magmatic plumbing systems that trans- Bear et al., 1994; Delaney and Gartner, 1997; Valentine and Krogh,
port magma from the Earth's mantle and crust towards the surface. A 2006), it is important to note that not always do sills follow pre-
large fraction of this magma never reaches the surface and becomes so- existing discontinuities and not always do dykes propagate along self-
lidified in the Earth's interior as magmatic intrusions, providing clues to generated fractures. I here encourage to use a terminology that does
the past geometry of magma pathways. As globally assumed in the not bear reference to the relations with fractures and is based on inher-
Earth Sciences, the behaviour of a geological process in the past can be ent geometric attributes: “dykes” are those tabular intrusions that have
a key to forecast the future development of the same process. Thus a dip ≥ 76°, “inclined sheets” have a dip N 10° and b 76°, and “sills” have
the reconstruction of the recent plumbing system below a volcano can a dip ≤ 10°. This terminology is essentially descriptive of the present
help to assess the present magma paths at an active volcano, but, as sheet attitude and does not take into account their possible tilting due
we will see later on, plumbing systems are extremely dynamic and to post-intrusion deformation. Post-sheet regional tilting must be
changes can occur at a fast rate. retro-deformed in order to get the original sheet architecture.
A plumbing system can be represented as a network of vertical or in- Magma chambers can be totally or partially molten, and in the latter
clined conduits that channel magma towards the surface, and sills and case a large part may consist of a crystal mush that behaves as poroelastic
chambers where magma can be stored (Fig. 1). At a certain time, one (Maaløe and Scheie, 1982; McKenzie, 1984; Gudmundsson, 1987; Marsh,
or more of these members can be partially or totally molten. Further 1989; Sinton and Detrick, 1992; Marsh, 2000). A totally or partially mol-
magma injection can be supplied from a deeper source in the mantle, ten magma chamber can be recognised by geophysical methods due to
or from one or more active magma chambers in the crust. Magma cham- its size that usually is much larger than sheet intrusions and to its phys-
bers located at different levels in the plumbing system can act as a sink ical properties that differ from those of the host rock. Also, a completely
for magma from the deeper reservoir, and as a source for magma injec- solidified, buried magma chamber usually has different properties than
tions into the surrounding crust (Gudmundsson, 2012). the host rocks, which enable to recognise it. Solidified magma chambers
Magma conduits have been represented for decades with a cylindri- can have dimension spanning from tens of km3 (plutons) to small intru-
cal shape and a circular section. However, direct observations of sions with size b 1 km3, which have been named differently in the past, in
eruptions along fractures, ground deformation measures, and the distri- relation to their shape. This can be from spherical to strongly ellipsoid or
bution of seismicity associated with magma intrusions provide evidence tabular, and this geometric feature is very important since the shape of an
for the role of dykes in mafic, intermediate and felsic magma transport active magma chamber dictates the orientation of the stress exerted on
(Pollard et al., 1983; Rubin and Pollard, 1987; Peltier et al., 2005; the host rock by magma overpressure (Gudmundsson, 2006, 2012).
Yamaoka et al., 2005; Aloisi et al., 2006; Mattia et al., 2007). Conduits The more complex plumbing systems can be composed of a plexus of in-
have also been studied by field evidence in eroded volcanoes where tab- terconnected sills, inclined sheets, dykes and multiple magma chambers
ular sheets compose the bulk of the plumbing system (Gudmundsson, (Hildreth, 1981; Lahr et al., 1994; Donoghue et al., 1995; La Delfa et al.,
1987, 1988, 1990; Tibaldi, 2001; Gudmundsson, 2002; Corazzato et al., 2001; Preston, 2001; Dawson et al., 2004; Marsh, 2004; Sanchez et al.,
2006; Pasquarè and Tibaldi, 2007; Tibaldi et al., 2008a,b,c, 2009). In 2004; Cartwright and Hansen, 2006).
the case of andesitic volcanism, some authors still suggest that erup-
tions are more focused and the conduits are more cylindrical (Zellmer 3. Propagation and arrest of intrusive sheets
and Annen, 2008), as explained through progressive melting of host
rocks (Quareni et al., 2001) and in terms of the sharp increase in Magma propagation occurs along planes of weakness that are al-
magma viscosity close to the surface due to decompression, degassing ready present in the host rock, or along hydrofractures generated by
88 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

internal fluid pressure. Hydrofractures are primarily extension fractures although in most cases the fracture and the intrusive sheet are arrested
whose propagation can be modelled as mode I cracks (extension normal inside the rock succession. The resulting magma pressure induces stress
to the fracture wall) (Bahat, 1980; Gautneb et al., 1989; Gudmundsson concentration at the tip that, in turn, creates the fracture. Numerical
et al., 2001). Many studies have been devoted to understanding models and field data suggest that the magma front should lag behind
hydrofractures in general and significant progress has been made on the hydrofracture tip during its propagation, and thus there should be
dyke propagation in particular. Earlier papers proposed that a sufficient- a dry part of the fracture (Warpinski, 1985; Bonafede and Olivieri,
ly large buoyant force can cause fractures to propagate upwards 1995; Garagash and Detournay, 2000). Sometimes in the field it is pos-
(Weertman, 1971; Secor and Pollard, 1975), whereas other studies did sible to observe the fracture at the sheet tip that still hosts magma with
not consider buoyancy (Spence and Sharp, 1985). From numerical a considerably minor thickness, in the order of a few mm–cm (e.g.
modelling, Spence et al. (1987) and Lister (1990) obtained a solution Fig. 2A). Ahead of the tip, a hydrofracture opens if it has a suitable orien-
of fracture surface displacement that results in the formation of a dyke tation with respect to the surrounding stress field in the host rock, and if
with a bulbous head and a constant width tail. Lister and Kerr (1991) the tensile stresses exceed the rock strength. If the rock succession is not
showed that magma moves along dykes under a local balance between homogeneous and contains discontinuities and layers with different
buoyancy forces and viscous pressure drop. Rubin (1993) and Chen mechanical properties, the tensile stress generated at the fracture tip
et al. (2007) modelled dyke propagation mainly based on a given source cannot be large enough to overcome the resistance to fracture propaga-
pressure (see also Rubin, 1995a and references therein, and Roper and tion and the sheet can become arrested (e.g. Fig. 2B). Intrusive sheets
Lister, 2005). Buoyancy and source pressure have been modelled by can stop when they meet: (1) discontinuities; (2) stress barriers; or
Meriaux and Jaupart (1998) and Bonafede and Rivalta (1999). Analogue (3) rock layers with strongly contrasting Young's moduli. The opening
modelling on magma-fracture propagation has been carried out for ex- of discontinuities (1) with a different orientation from the propagating
ample by Maaløe (1987), Takada (1990), Menand and Tait (2002), Ito sheet, can dissipate stresses; when a horizontal discontinuity, like a
and Martel (2002) and Tibaldi et al. (2014). bedding plane, is encountered at the contact ahead of a hydrofracture
These papers suggest that buoyancy forces or magma overpressure tip, the hydrofracture may be unable to propagate through the horizon-
can contribute to the propagation of hydrofractures to the surface, tal, open discontinuity (Gudmundsson and Brenner, 2001, 2005;

Fig. 2. A. Photo (left) and sketch (right) of a sheet showing a sudden decrease in thickness in correspondence of the hydrofracture propagated at its tip and initial intrusion along the frac-
ture; note that only a few centimetres of the fracture's length are dry (Skye Island, UK). B. Photo (left) and sketch (right) of a dyke arrested at the contact with a discontinuity (bedding) and
a layer with different mechanical properties (a thick lava flow) (Reunion Island).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 89

Gudmundsson, 2002). (2) Stress barriers are layers with local stresses the contact with the thick (~ 2 m) lava flow. Experiments also show
unfavourable for intrusion propagation, such as layers that contain that soft layers can be more effective in arresting hydraulic fractures
stresses in excess of up to 5–10 MPa with respect to the adjacent than stiff layers (Charlez, 1997; Yew, 1997), as can be seen in the exam-
rocks (Gretener, 1969; Gudmundsson, 1986, 1990; Parsons et al., ple of Fig. 3A where an inclined sheet gradually splits into minor fingers
1992). Regarding point (3), ascending dykes can stop at the contact of and finally becomes arrested within a soft layer media made of poorly
bedding layers with different mechanical properties, such as in the ex- consolidated breccia deposits.
ample of Fig. 2B where a dyke is intruded into a succession of thin The termination of an intrusive sheet can have different shapes that
(b20 cm) lava flows and dominant breccia layers and then stops at represent the diverse settings for the stoping of the fracture propagation

Fig. 3. Photos (left side) and sketches (right side) of: A. an inclined sheet that gradually splits into minor fingers and finally becomes arrested within poorly consolidated breccia deposits
(northwest coast, Sao Miguel, Azores Archipelago), since soft layers can be more effective at arresting hydraulic fractures than stiff layers (Charlez, 1997; Yew, 1997); B. a gently-dipping
sheet that terminates with a series of secondary intrusions (asymmetric fingers) that depart upward (Stromboli volcano, Italy); and C. an arrested steeply-dipping sheet with an upward
symmetric fanning arrangement (Ross Lake, Washington, USA).
90 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

(Gudmundsson, 2003). The sheet can maintain a unique main plane propagation through heterogeneous geological media with local devia-
with a bulbous tip shape, or it can show sets of fingers that depart up- tions of the stress field, and in this case the segments' step direction is
ward (Fig. 3B) or in a fanning arrangement (Fig. 3C). The propagation non-systematic (Daniels et al., 2012). In layered host rocks, overlapping
of fingers uses more mechanical energy on the processes of flow, dila- sheet segments may tend to link together by folding and thinning of the
tion, and deformation than the propagation by a continuous sheet ge- rock between the overlapping segments (bridge) and by thickening of
ometry (Pollard et al., 1975) and thus can explain the arrest. Finger the sheet, by a propagating crack that is usually straight. In isotropic
formation may occur when local inhomogeneities in the rock mass host rocks, the termination of the single sheet segment tends to bend to-
along the wall of the sheet tip are located ahead of the bulk of the intru- wards the adjacent segment, breaking the bridge (Fig. 4C); this occurs
sion, and the magma may accelerate at the inhomogeneity as it finds because there is a mechanical interaction between the adjacent crack
less resistance there. Alternatively, fingers can form where there are tips, with the stress field at one tip altering the stress of the adjacent
perturbations in the cooling distribution in the magma sheet and propagating tip by inducing curved cracks (Nicholson and Pollard,
lobes with higher temperatures advance as the others have a greater 1985). At overlapping sheet segments, a local temperature decrease at
viscosity (Wylie et al., 1999). A sheet can also change its geometry one sheet segment might hinder further magma propagation here and
into an en-échelon pattern expressed by individual fingers that coalescence at the adjoining segment, with the possible effect of disper-
are oblique, in section view, with respect to the main sheet plane sion of the magma overpressure and stoping of the intrusion propaga-
(Fig. 4A). Sheet en-échelon segmentation commonly occurs in response tion. However, if magma flux is high enough, the original bridge can
to stress field rotation at an angle to the principal stress directions, fre- be completely cut across, resulting in the two original sheet segments
quently found approaching the surface, as also experimentally obtained now linked (Fig. 4D) (‘bridge xenoliths’ of Rickwood, 1990). Inflation
by Tibaldi et al. (2014) (Fig. 4B). En-échelon sheet fingers can show sys- can proceed until a complete coalescence has been reached and the con-
tematic step-overs between segments owing to tangential stress on tinuous planar sheet configuration has been restored (Hutton, 2009;
the sheet walls (Pollard, 1987). Segmentation may also reflect dyke Schofield et al., 2012).

Fig. 4. A. Photo of en-échelon fingers in section view departing from a single intrusive sheet (Skye Island, UK). B. Photo of an analogue model, physically scaled, of a sheet intrusion showing
the upward protrusion of en-échelon fingers approaching the free topographic surface; white and black boxes have 1 cm side. C. Photo (left side) and sketch (right side) of two partially
overlapping sheet segments whose tips bend towards the adjacent segment, typical of isotropic host rocks (Skye Island, UK). D. Photo (left side) and sketch (right side) of a bridge xenoliths
isolated by two partially overlapping sheet segments linked together (Skye Island, UK).
B. After Tibaldi et al. (2014).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 91

4. Deep plumbing system interconnected static microfractures (porosity networks). For the
lower crust, magma diapirism is thought to be a common mechanism
Information on the deeper part of magma plumbing systems mostly of magma transport (Whitehead and Luther, 1975; White and
comes from geophysical and petrological research. Outcrops of Chappell, 1977; Pitcher, 1979; Hildreth, 1981; Marsh, 1982). In the
the deeper portions of plumbing systems are extremely rare and deeper section of plumbing systems, in fact, it is considered that the
are not representative of the actual root zone, where melts start to as- thermal and pressure conditions are such as to allow for the ductile
cend. Much research has been carried out on both felsic and mafic flow of the host rock around the rising magma mass, whereas at
root zones, resulting in different models also depending on magma upper levels (middle to upper crust) this is regarded as thermally and
geochemistry. mechanically unrealistic. Moreover, the diapiric ascent model is also
controversial since numerical modelling indicates that magma solidifi-
4.1. Felsic magmas cation should occur due to the slow ascent (Clemens and Mawer,
1992). The model of flow through porosity networks is accepted, for ex-
In granitic systems, a classical view is that percentages within the ample, in sediments at low confining pressures, but the model is
range 30–50% of melt fraction may allow for the mobilisation of zones questioned for geopressured zones since it cannot explain all aspects
hundreds of metres in lateral extent (Wickham, 1987). For example, of the observed large-scale fluid flow (Nunn, 1996). The third model
in the Trois Seigneurs massif in the Pyrenees (France), metamorphic of propagation relies on fluid-filled fractures and is considered valid
rocks pass gradationally into migmatitic biotite–sillimanite gneisses throughout the lithosphere. If diapirs rise in the lower crust, their up-
and then into a biotite granite–quartz diorite body (Wickham, 1987). ward pressure can induce fracturing in the roof rocks, where fluid-
Within the metamorphic sequence there are ponds of granite with di- filled fractures can propagate much faster than diapirs, thus overcoming
ameters ranging from l m to l km across. Melting at 670–700 °C, 3.0– the problem of time-dependant solidification.
4.0 kbar under water-rich conditions was followed by segregation and Tectonics plays a major role not only in contributing to the creation
movement of melt into fractures. When large enough, these bodies in- of heat flux conditions, but also in governing the weakness zones that
truded the overlying metasediments, but were soon halted as they be- may favour magma upwelling. Several examples, mainly based on
came water-saturated. The zone of granite ponds corresponds to the field data, indicate that melt upwelling in the deeper parts of plumbing
zone of partial melting, segregation and collection (a) in the division systems was channelled along transcurrent zones (Mitjavila et al., 1997;
of plumbing systems of Atherton (1993), above which there are the Olivier et al., 1999; Rossetti et al., 2000; Rosenberg, 2004; Weinberg
transport zone (b) and the emplacement zone (c), where freezing oc- et al., 2004; Marcotte et al., 2005). This is mainly related to two facts:
curs (Fig. 5). In zone (a), with increasing melt proportion there is a (1) transcurrent faults are vertical shear zones that extend downward
change of the mechanical properties of the rock melt that changes to lower crustal levels, whereas reverse and thrust faults and normal
from linearly-elastic to elastic–plastic and finally to viscous; this is faults are less steep and tend to acquire a horizontal geometry with
reflected in the flow behaviour that therefore will vary with melt frac- depth (Fig. 6) (Davis et al., 1996) and (2) magma can enter ductile ex-
tion, and thus in the formation of the deeper part of plumbing systems tensional jogs within shear zones as envisaged for example by Aydin
(Shaw, 1965; Murase and McBirney, 1973; Shaw, 1980; Wickham, et al. (1990), D'Lemos et al. (1992) and Chiarabba et al. (2004a). Recent-
1987). In the case of low melt content, small batches of magma will be ly it has been suggested that even transpressional tectonics can be an ef-
able to rise and will soon stop. In the case of large melt fraction, such ficient mechanism for moving magma through the lithosphere (Saint
as for example in the case of high thermal gradients (80–100 °C/km), Blanquat et al., 1998), although magma upwelling under transpression
large magma batches can ascend giving rise to huge plutons, especially might result in the movement of only a small volume of magma to the
if associated with contemporaneous rifting or local extension related to surface (Marcotte et al., 2005), and finally an in-depth review by
pull-apart structures or releasing bends within transcurrent shear zones Tibaldi et al. (2010) shows several examples of volcanism along
and the consequent horizontal regional ductile extension (Atherton, transpressional, transcurrent and transtensional fault zones. For felsic
1993). Similarly, also Zellmer (2008, 2009) based on global correlations magmas it is also interesting to note a close temporal relation between
between eruptive style, surface heat flux and convergence rates at dif- peak anatexis and regional strike-slip displacements, suggesting that
ferent volcanic arcs, infers that the rate of melt production in the mantle strike-slip motions may initiate along zones that were thermally soft-
wedge ultimately controls the deep dynamics of magma transfer. ened by anatexis and that regional tectonics and granite formation
From zone (a) of melt production, fluids should find their pathway may be inextricably linked processes (D'Lemos et al., 1992). Based on
to move upwards through zone (b). Models of fluid transport in the lith- this, the shape of granitic intrusions should be represented by elongated
osphere propose diapiric ascent or fluid flow through networks of vertical bodies with a horizontal major axis from parallel to sub-parallel
to the strike of a wrench zone (e.g. Farris et al., 2006). Notwithstanding
this, some authors indicate that granitic plutons are mostly thin with
aspect ratios (thickness to width) up to 1:15 and a flat shape (e.g.
Atherton, 1993). Geophysical and structural studies showed that some
plutons are sill-like, low aspect-ratio, tabular bodies (Bridgwater et al.,
1974; Lefort, 1981; Cruden, 1998; Petford et al., 2000), whereas
magma chambers and plutons are often represented as spherical bodies
(e.g. Pitcher, 1979).
The importance of major, deep crustal fracturing in triggering felsic
magma melting and assisting magma upwelling has been put forward
by many authors (e.g. Rittmann, 1962). They suggest that the deep
plumbing system of granitic magmas is dominated by intrusions that
frequently use vertical and horizontal extensional features linked to
strike-slip faulting as in the British Tertiary Province and Irish
Caledonides (Leake, 1990), at the Coastal Batholith of Peru (Pitcher,
1978), at the Main Donegal granite (Hutton, 1982), at the Peruvian
Fig. 5. Subdivision of a granitic plumbing system modified after Atherton (1993) for Cor-
Cordillera Batholith (Petford and Atherton, 1992), in Scotland
dilleran setting, with: (a) zone of partial melting, segregation and collection; (b) transport (Watson, 1984), in southern Greenland (Harrison et al., 1990), and at
zone; and (c) emplacement zone where freezing occurs. the Boulder and Tobacco Root Batholiths (Montana, USA, Schmidt
92 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 6. Deep magma plumbing system is conditioned also by the geometry of tectonic weakness zones that can channel magma flow. Transcurrent faults are vertical shear zones that extend
downward at lower crustal levels and can capture ascending magmas. On the contrary, reverse faults with their typical ramp and flat geometry tend to stop dykes or divert dykes into sills.
Finally, normal faults are shallower-dipping than strike-slip faults and tend to acquire a horizontal geometry with depth where they exert a lower influence on ascending magmas.

et al., 1990). Most authors thus agree that zone (b) (Fig. 5) of plumbing beneath the Bolivia Altiplano, suggest that partially molten granitic bod-
system of granitic magmas should be characterised by sheet-like com- ies ascend by way of a diapiric mechanism through the hot ductile crust.
plexes of intrusions that give rise to the conduit system below plutons,
the latter also being controlled by shear zones. This corresponds to the 4.2. Mafic magmas
mechanism of transport and emplacement in the middle crust.
In zone “c” of Atherton (1993) storage processes occur and most In the case of mafic magma, the deep plumbing system is considered
magma is emplaced in the form of plutons. Storage includes a series of to be governed by a higher mobility than in the case of a felsic magma at
mechanisms that can be summarised by two end-members, although parity of other conditions. Mafic magma is transported through cracks
it is reasonable to consider that a spectrum of intermediate possibilities in the lithosphere that give rise to swarms of dykes. These dykes can
exists. One end-member is represented by a plutonic body that results be connected with deep to shallow magma chambers, or they may di-
from a single, connected and fairly closed batch of magma that freezes rectly convey magma from the mantle towards the surface without a
while rising through the crust. This mechanism can also be viewed as shallow reservoir, as derived for example from the very high magma
a diapir that has been disconnected from its source (Paterson et al., supply rate of 0.7 km3/yr during the 1959 Kilauea Iki eruption
2010). The other end-member is represented by a plutonic body that (Takada, 1999) respect to the values of 0.02 to 0.18 km3/yr during
is a frozen part of a former complex plumbing system that remains ac- 1952–1983 (Dvorak and Dzurisin, 1993). Nevertheless, as for felsic
tive for an extended duration. In this case the plumbing system is magmas, mafic magmas mostly also stop rising in the crust. Although
huge, evolves over time, is used during several magma upwelling direct observations of exhumed deep intrusions are rare, some outcrops
events, and remobilises material from older pulses or from host rock indicate that mafic magma bodies emplaced within or at the base of the
(Paterson et al., 2010). Pluton growth can be further complicated lower crust, result from the accretion of successive magma pulses. For
by other processes such as: magma pulses moving back down the example, a deep section of a mafic plumbing system has been exhumed
magma pathway during rise of other pulses; loss of magma from the after orogenic uplift and erosion at the Ivrea mafic body in the Italian
plutonic system by volcanic eruptions; reheating of crystal mush Alps, where the evidence of individual pulses of magma can be directly
zones and consequent continuous movements in a magma conduit; in- observed (Quick et al., 1994).
ternal differentiation processes. A long-lasting scientific discussion re- An explanation for the deep magma accumulation along a horizontal
volved around the question if magmas more commonly ascend by zone was initially proposed by Ryan (1987) and Lister and Kerr (1991)
continuous upwelling or by pulse-like flow. Abundant field evidence at the level of neutral buoyancy (LNB). Here it was retained that magma
suggests that the latter mechanism might be more common than previ- density equals the density of the host rock and buoyancy becomes null,
ously thought; several sheeted dyke and sill complexes indicate that favouring sill emplacement. This hypothesis was revised in more recent
plutonic bodies have been constructed through magma pulsing years based on the fact that several dykes cut through the LNB and even
(Pitcher and Berger, 1972; Hardee, 1982; Hutton, 1982; Lagarde et al., through rock successions with a lower density than magma. Sills in fact
1990; Paterson and Vernon, 1995; McNulty et al., 1996; Vigneresse can emplace due to other mechanical and physical conditions that will
and Bouchez, 1997; Paterson and Miller, 1998; Wiebe and Collins, be discussed later. If the magma supply rate is low, sills can freeze,
1998; Johnson et al., 1999; Miller and Paterson, 2001; Miller and whereas if the magma supply rate is high, they may combine and result
Miller, 2002; Memeti et al., 2010). in the formation of a magma chamber (Gudmundsson, 1990, 1995).
In the middle-upper crust, dyking is hence considered as the most Once formed, the magma chamber can further stop other ascending
consistent magma transfer mechanism (Clemens and Mawer, 1992; dykes, thus growing in size. Below continental crust, mafic magmas
Clemens et al., 1997; Petford et al., 2000). Granite intrusion along ductile can be trapped near the Moho or within the lower crust. Below oceanic
extensional shear zones is considered the most suitable process for up- island volcanoes, it has been suggested that basaltic to gabbroid rocks
welling and storage and this solves the room problem posed by pluton are denser than basaltic magmas and thus LNB conditions are not
emplacement by diapirism (Hutton et al., 1990). A model that instead reached within the lower crust. In this condition, a LNB can be present
extends diapirism also to the middle crust is the one proposed by a few kilometres below the volcano or even inside the volcanic edifice,
Miller and Paterson (1999); these authors suggested that felsic giving rise to very shallow magma chambers (Klügel et al., 2005).
magma can rise and be emplaced as ‘visco-elastic diapirs’ through a se- Deep tabular horizontal sills have also been individuated by means of
ries of multiple magma batches within regional tectonic deformation, seismic profiles: layering in the lower crust has been interpreted as
accompanied by downward movements of the host rocks through mul- due to the presence of sills (Fuchs et al., 1987; Wenzel and Brun,
tiple processes including brittle deformation and stoping (Cabello et al., 1991; Franke, 1992). Exposure of very deep lithospheric section at the
2006; Farris et al., 2006; Zak and Paterson, 2006). Recently, diapirism Oman Ophiolite shows that magma-induced fracturing and intrusion
has been extended up to the middle–upper crust by del Potro et al. propagation are common processes that occur even at the Moho level
(2013) who, based on the inversion of high-resolution gravity data (Nicolas et al., 1994).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 93

The occurrence of mafic intrusions in the lower crust has also been extending between 3 and 9 km of depth and 8–10 km in length beneath
modelled by numerical simulations by several authors (Petford and the central and eastern sectors of the volcano.
Gallagher, 2001; Annen and Sparks, 2002; Dufek and Bergantz, 2005; At Kilauea Volcano (Hawaii) several studies have been carried out on
Annen et al., 2006a), who showed that if the composition of the crust the plumbing system in the last 45 years. Wright (1971) suggested a
in contact with the mafic intrusions is amphibolitic, partial melting of vertical plumbing system located below the summit zone of Kilauea
the crust is very limited. This on one side implies that large quantities with a shallow magma chamber at a few km depth, a storage zone at
of silicic melts can be generated by incomplete crystallisation of the in- 20–25 km of depth, and a region of partial melting at 65 km of depth.
truded mafic magma (Dufek and Bergantz, 2005; Annen et al., 2006b), These features account for the high pressure crystal fractionation
and on the other side that a deeper plumbing system should preserve emphasised by Clague (1987) and the xenolith populations in the
its original architecture given by tabular intrusions, because the con- lavas that require a magma reservoir at the base of the oceanic crust.
tacts with the host rocks are subject to limited modifications. Dahm The presence of a magma chamber at a depth N 16 km has been sug-
(2000) modelled the possible deep plumbing system at convergent gested also by Frey et al. (1990) based on geochemical data. One of
margins. He found that fluid-filled fractures can propagate through the first and most complete models of the Kilauea plumbing system
the inhomogeneous stress field of the asthenospheric mantle wedge was presented by Ryan (1988), who outlined the presence of a deep
above a subduction zone. The magma paths depend on the dip angle main conduit zone located down to a depth of 60 km below the summit
of the subducting slab, the distance from the wedge corner, the subduc- zone of the volcano (Fig. 8). This conduit is elongated and probably is
tion velocity, the mantle viscosity, and the apparent buoyancy force of composed of a series of vertical and sub-vertical dykes which are linked
dykes. In particular, dykes propagate upwards parallel to the slab if to a long feeding system located in the uppermost kilometres below the
the angle of subduction is N 45°, whereas dykes are bended in sub- volcanic rift. Based on microgravimetry data, it appears that the dense
horizontal curved sills, and thus magma is trapped without reaching central core of the volcano extends south of the summit, implying a
the lithospheric layer, in the case of shallow subduction angles (Dahm, structural continuity between the Southwest Rift Zone and the East
2000). Rift Zone (Kauahikaua et al., 2000). The surface of the dense core
More information comes from recent geophysical studies that allow beneath the East Rift Zone is steeper to the south than to the north,
imaging of the entire crust down to the source region of the magma in suggesting southward migration of more recent eruptive activity
the uppermost mantle, as for example at the deep plumbing system of (Kauahikaua, 1993). Microgravimetry also suggests possible connec-
Mt Etna volcano; Chiarabba et al. (2004b) show the structure beneath tions of the plumbing system at depths of 10–14 km between different
Mt Etna down to a depth of 24 km through tomographic inversions of volcanoes of the Island of Hawaii. These connections reach a depth com-
P- and S-wave arrival times from local earthquakes (Fig. 7). Tomogra- patible with the oceanic crust and their existence would suggest that
phy shows the lower plumbing system made by a main narrow conduit magma may have exploited linear zones of weakness in the oceanic
zone, 4–6 km long, beneath the central part of the volcano between 9 plate before the conduits became separated, or that island-building vol-
and 18 km of depth, and a wide region of low Vp in the uppermost man- canoes can originate along the rift zones of older volcanoes (Kauahikaua
tle (below 34 km of depth) associated with the magma source region. et al., 2000). The main conduits of Kilauea are punctuated by a series of
The upper plumbing system is characterised by an intrusive complex flat bodies that represent various magma chambers highlighted by
below the southern craters and the Valle del Bove, and another complex aseismic zones. The maximum volume of the aseismic zone is 40 km3

Fig. 7. The deep plumbing system of Mt Etna (Italy) has been revealed by tomographic inversions of P- and S-wave arrival times from local earthquakes down to 24 km depth. Two to-
mographic sections (oriented SSE–NNW and SW–NE) show the presence of a high velocity body (HVB) interpreted as a main solidified intrusive body. Magma intrusions prevalently affect
the western and north-western border of the HVB, following NNW–SSE and NE–SW structures.
After Chiarabba et al. (2004b), reproduced under license number 3490690796656 of Oct 16, 2014.
94 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 8. One of the first and most complete models of the Kilauea plumbing system was presented by Ryan (1988), who outlined the presence of a deep main conduit zone located down to a
depth of 60 km below the summit zone of the volcano. This elongated conduit is composed of a series of vertical and sub-vertical dykes which are linked to a long feeding system located in
the uppermost kilometres below the volcanic rift.
Reproduced under license number 3490680301921 of Oct 16, 2014.

(Klein et al., 1987); however, it must be noted that an aseismic zone this volcano at a depth N 10 km. Evidence from the 2011 eruption sug-
may include a crystal-mush zone and a hot, ductile region surrounding gests that magma migrated towards the volcano along inclined path-
the magma reservoir. More recent studies showed that rapid variations ways. These pathways coincide with a conductor zone located through
of 206Pb/204Pb isotopes on a time scale of years to decades require short- resistivity analyses. This type of deep to medium-depth magmatic sys-
term changes in the pathway that melt takes from the source zone to tem implies a component of lateral transport of magma and contributes
surface (Greene et al., 2013, and references therein). This suggests to the recent discussion on magma chambers that are in an offset posi-
that the plumbing system is more complicated than expected and that tion from the volcanoes they are connected with. Another recent exam-
it might be made of interconnected conduits with various geometries. ple comes from Ishizuka et al. (2008) who found geophysical and
Moreover, the integration of previous geophysical and geochemical petrological evidence of long-distance lateral horizontal magma trans-
data with new findings suggests the presence of a deeper magma cham- port below the submarine/subaerial volcanic chain of the northern Izu
ber at about 10–12 km depth (Lin et al., 2014), another at 3–5 km depth arc (Fig. 9B). Lava outpoured in different parts of the chain resulted to
(Fiske and Kinoshita, 1969; Klein et al., 1987), and a shallow chamber at come from the same basaltic primary magma, and this suggests that
about 1 km depth (Cervelli and Miklius, 2003; Johnson et al., 2010). magma was laterally transported for at least 20 km in the middle to
As opposed to the geometry of the plumbing systems of Mt Etna lower crust (10–20 km deep here). Later on, the same magmas experi-
and Kilauea volcanoes, geophysical studies at the Shinmoe–Dake volca- enced crystal fractionation and accumulation at a shallow magma
no, Japan, indicate the presence of a magma feeding system with a gen- chamber with further transport at distance b 5 km. Ishizuka et al.
eral oblique dip (Fig. 9A). Magnetotelluric data from Aizawa et al. (2008) also suggested that the long-distance magma transport has
(2014) combined with petrographic data show the presence of a deep been controlled by a regional extensional stress regime, while the
basaltic–andesitic magma chamber located in an offset position from shorter (b5 km) distance transport has been controlled by a local stress
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 95

Fig. 9. Models of oblique and horizontal deep magma plumbing systems. A. Geophysical studies at the Shinmoe–Dake volcano, Japan indicate the presence of a magma feeding system with
a general oblique dip. Magnetotelluric data from Aizawa et al. (2014) combined with petrographic data show the presence of a deep basaltic–andesitic magma chamber located in an offset
position from this volcano at a depth N 10 km. Evidence from the 2011 eruption suggests that the magma migrated towards the volcano along inclined pathways. B. Ishizuka et al. (2008)
found evidence of long-distance lateral horizontal magma transport below the submarine/subaerial volcanic chain of the northern Izu arc: lava emitted in different parts of the chain re-
sulted to derive from the same basaltic primary magma; this enabled to conclude that magma was laterally transported for at least 20 km in the middle to lower crust (10–20 km deep).
Later on, the same magmas experienced crystal fractionation and accumulation at shallow magma chamber with further transport at distance b5 km.

regime linked to the load of the volcanic chain. Similarly, significant supply rates for oceanic island volcanoes (~ 0.0002 km/yr, Amelung
lateral magma migration, in the order of ~ 15 km, has been found at and Day, 2002). Hildner et al. (2012) proposed that the age of the un-
El Hierro Island (Canary Archipelago) during the July–October 2011 derlying lithosphere could control the presence of one or more cham-
seismic–volcanic crisis (González et al., 2013). On October 2011, a sub- bers within the upper part of the plumbing system, because the older
marine fissure eruption occurred at a distance of 15 km from the initial oceanic lithosphere is much cooler and denser and tends to sink due
earthquake loci. Geodetic analysis reveals two distinct shallow magma to advanced thermal contraction together with a deepening of the
reservoirs, at 9.5 ± 4.0 km and 4.5 ± 2.0 km, vertically offset from magma chambers. González et al. (2013) similarly suggested that
one another and linked by vertical conduits. the main controlling factor is the thermomechanical structure of the
crust, which in turn depends on the deep mantle magma supply rate
5. Shallow plumbing system and on the age of the oceanic lithosphere. Fialko and Rubin (1998)
and Grandin et al. (2012) claimed that the presence of double shallow
The shallow part of plumbing systems, that is to say the first magma chambers is linked to the presence of two horizons of neutral
kilometres below volcanoes, is better constrained than the deeper part buoyancy, where magma accumulation occurs preferentially and its lat-
due to the larger accessibility of outcrops at eroded volcanoes and to eral migration is facilitated, giving rise to eruptions along rift zones. The
more feasible geophysical studies. The plumbing system can go from deeper horizon associated with olivine-rich cumulates exists, for exam-
very simple dyke complex, made of tens to thousands parallel vertical ple, at a depth of 6–10 km for Hawaii and at ~9.5 km at El Hierro, and a
dykes that link the magma chamber to the uppermost volcano conduit, shallower one at 2–4 km at Hawaii and ~ 4.5 km at El Hierro (Ryan,
to a very complex intrusive plexus made of dykes, sills and inclined 1988; González et al., 2013).
sheets, which interconnect multiple shallow magma chambers. For a Shallow magma chambers have been detected also at several eroded
long time it has been believed that the simpler kind of plumbing system volcanoes for example in the British Tertiary Igneous Province (United
without shallow chambers can be found at oceanic island volcanoes, but Kingdom and Ireland) and in Iceland (Walker, 1958, 1960; Bald et al.,
more recent observations indicate the opposite (Amelung and Day, 1971; Walker, 1974, 1975; Fridleifsson, 1977; Gudmundsson, 1990,
2002; Klügel et al., 2005; González et al., 2013). The possible absence 1995, 2002; Klausen, 2004, 2006; Tibaldi and Pasquarè, 2008; Tibaldi
of shallow magma chambers has been interpreted as due to local pro- et al., 2011). These chambers may host magma with the same pressure
cesses that can modify the structure of a volcano such as catastrophic as the host rock and in this case their presence does not perturb sensibly
flank collapses (Amelung and Day, 2002; Tibaldi, 2004; Tibaldi et al., the surrounding stress field. However, the presence of magma is linked
2008a, 2008b, 2008c; Manconi et al., 2009) or due to poor knowledge to high temperatures and a different rheology than the host rock, condi-
of the subvolcanic system. tions that can perturb the propagation of dykes from below (Karlstrom
At El Hierro Island (Canary archipelago), geodetic data indicate the et al., 2010). Chambers that instead are significantly overpressured with
presence of a shallow magma reservoir at a depth of 4.5 ± 2.0 km be- respect to the stress in the host rock can perturb the stress field in the
neath the volcano flank of the southern rift (Gonzalez et al., 2013), al- surrounding region; below the chamber there may be a rock volume in-
though a deeper chamber has been suggested by Becerril et al. (2013). side which the stress produced by magma forces is large enough to per-
At the Island of Hawaii, beneath Kilauea Volcano there are two main turb the trajectories of rising dykes, focusing them towards the magma
magma storage zones at shallow level, at ~ 1 km and ~ 3 km depth chamber (Fig. 10A) (Karlstrom et al., 2009), although, as far as I know,
(Cervelli and Miklius, 2003; Baker and Amelung, 2012). At Piton de la there are no field examples of such focusing below eroded magma
Fournaise, two active magma chambers have been detected at ~ 2.5 chambers.
and ~7 km of depth (Peltier et al., 2009). The Fernandina Island volcano On the contrary, there are plenty of examples of the structure of
(Galápagos Islands) has two levels of magma storage at depths of ~ 1 plumbing systems immediately above shallow magma chambers, espe-
and ~ 5 km (Chadwick et al., 2011; Bagnardi and Amelung, 2012). cially in the case of mafic magmas. These plumbing systems are made of
Other volcanoes of the Galápagos Islands have a single shallow crustal swarms of inclined sheets that have a typical arrangement consisting in
magma reservoir, detected geodetically, such as at Alcedo, Cerro Azul, variable dips that tend to converge on a central area. The definition of
Darwin, Sierra Negra, and Wolf volcanoes. The existence of shallow inclined sheet was first used by Harker (1904) for intrusions at Skye,
magma chambers has been attributed to the existence of a high melt while the term “cone sheet” was later introduced by Bailey et al.
supply rate from the deep mantle by Clague and Dixon (2000). (1924) to describe the occurrence of sheets arranged in a conical
González et al. (2013) suggest that at the Canary Islands the high melt swarm geometry around a main intrusion. Diverse models have been
supply rate is not enough for the existence of shallow magma reservoir, introduced to explain the formation of cone sheets, which have been
because the volcanoes of this archipelago have the lowest magma widely studied due to their importance in magma transport, also
96 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 10. Chambers that are significantly overpressured with respect to the stresses in the host rock can perturb the stress field in the surrounding region. A. Below the chamber there may be
a rock volume inside which stresses produced by magma force are large enough to perturb the trajectories of rising dykes, focusing them towards the magma chamber. B. Boundary el-
ement results showing a pressurised spherical magma chamber with an internal magmatic excess pressure of 5 MPa as the only loading; the trajectories of the greatest principal stress
depart at the right angle from the chamber walls, creating a radial pattern.
A. Redrawn after Karlstrom et al. (2009). B. Redrawn after Gudmundsson (1998).

laterally from the central zone of a volcano; cone sheets are crucial also Within a centrally-inclined sheet swarm, the sheet dip may vary
in terms of the possibility of locating the magma chamber based on their from one sheet to another and also at the same sheet. Anyway, it is
geometry and of the fact that the piecemeal intrusion of hundreds of in- worth noting that outcrops of centrally-inclined sheet swarms are usu-
clined sheets induces a major upward deformation of the host rock (Le ally limited in lateral/vertical extent and different geometries of inclined
Bas, 1971). Bailey et al. (1924) suggested that any increase in the pres- sheets can be found; accordingly, different models are proposed to
sure of a magma reservoir at depth would superimpose a tensional re- explain them. These models put forward: (1) concave-downward
gime on the host rock and favour the formation of inward-dipping (trumpet-shaped) sheets with increasing dip closer to the magmatic
fractures, the intrusion of which produces cone sheets (Fig. 11A). A de- source (Fig. 12A) (Phillips, 1974); in this model, sheets are missing in
crease in the magma chamber pressure instead may promote instability the central part; (2) concave-upward (bowl-shaped) sheets with de-
in the roof rock resulting in subsidence along outward-dipping or sub- creasing sheet dip with depth from a pressurised magma chamber
vertical fractures, the intrusion of which produces “ring dykes” (Fig. 12B) (Phillips, 1974); also here, sheets are missing in the central
(Fig. 11B) (Bailey et al., 1924). Anderson (1936) considered that ring part; (3) radial planar sheets from a spherical magma chamber
dykes should be emplaced along shear fractures that dip outward in (Fig. 12C) (Chadwick and Dieterich, 1995; Gudmundsson, 1998);
the range of 60–70° becoming shallower upward. Robson and Barr and (4) planar parallel to sub-parallel sheets originated from a lobate
(1964) and Durrance (1967) put forward the importance of shear (sill-like) magma chamber (Fig. 12D) (Gudmundsson, 1998; Tibaldi
planes in guiding the emplacement of cone sheets, although now it is et al., 2011; Bistacchi et al., 2012).
recognised that most sheets occupy extensional fractures. The model In westernmost Iceland, along the southern coast of the Snaefellsnes
of Phillips (1974) suggests that cone sheets are restricted to a concentric Peninsula that is essentially made up of Tertiary–Quaternary basalts,
swarm departing from the shoulders of a magma chamber and follow- there are two major intrusions, each surrounded by a centrally-
ing shear fractures, and thus they are missing above the centre of the inclined sheet swarm. The intrusions correspond to the Midhyrna gab-
magma chamber, something usually different in the field (Fig. 11C). bro and Lysuskard granophyre (Upton and Wright, 1961) and both in-
Shear fractures form as rotational strain develops along inward- truded Tertiary basaltic lava flows. The Midhyrna and Lysuskard
dipping strips that delimit roof uplifts relative to the immobile side intrusions are surrounded and intruded by two centrally-dipping
regions. Phillips (1974) also suggested that cone sheets could extend sheet swarms (Fig. 13A) (Tibaldi et al., 2013); these sheets show no
upward in sub-vertical shear fractures with a bowl-like geometry gradual variation in dip with distance from the focus area, are rectilinear
(Fig. 12) or they may curve into sub-horizontal sills upwards with a in section view, and intrude with the same geometry the main intrusive
trumpet-shape (Fig. 12). In the present paper I prefer to use the term bodies as well as the layered Tertiary lavas (Fig. 13B). The diameter of
“centrally-inclined sheet swarm” because it is more descriptive and both sheet swarms is about 12 km, the average sheet thickness is
does not bear reference to the previous cone sheet models. 0.63 m, and the average dip is 28° (Tibaldi et al., 2013), lower than
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 97

Fig. 11. Models of cone sheet and ring dyke emplacement in a vertical section view: A. an expanding magma chamber induces a radial σ1 (continuous lines) and concentric σ3 (segmented
lines) parallel to the vertical section view, cone sheets are emplaced along tension fractures (diverging arrows). B. Magma chamber deflation induces shear zones at the periphery along
which ring dykes may intrude. C. An expanding magma chamber produces tensional fractures (diverging arrow) along which cone sheets may intrude with a downward-concave shape,
and/or shear zones with rotational strain along which cone sheets may intrude with an upward-concave shape. Note that emplacement of cone sheets is not foreseen above the central part
of the magma chamber.
A. and B. Modified after Bailey et al. (1924) and Anderson (1936). C. Modified after Phillips (1974).

other sheet swarms in Iceland whose average dip is 34°. Farther south overlying succession that forms the original volcano, up to a thickness
there is the Thverfell centrally-inclined sheet swarm that was emplaced of 2 km, has been eroded. Bell et al. (1994) stated that centrally-
mostly within almost isotropic hyaloclastite deposits (Pasquarè and inclined sheet magmas were generated by large degrees of partial melt-
Tibaldi, 2007; Tibaldi et al., 2008c) (Fig. 14). Also in this case the sheets ing of a depleted mantle source, combined with assimilation of Lewisian
do not show a systematic dip decrease outwards from the focus zone. gneisses, and fractional crystallisation that occurred in a magma cham-
The Cuillin Igneous Complex in the Island of Skye (UK) represents ber subjacent to the final site of emplacement and undergoing periodic
the deep core of a Tertiary, large basaltic volcano. It is composed of replenishment. Also in this case, the about one thousand sheets mea-
coarse-grained layered and unlayered basic and ultrabasic huge intru- sured do not show a systematic variation of dip angle with distance
sive bodies that are in turn intruded by centrally-inclined sheets. The from the focus zone (Tibaldi et al., 2011) as can be seen in Fig. 15. It is
98 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 12. Possible geometries of centrally-inclined sheet swarms resulting from internal excess magma pressure: A. radial planar sheets from a spherical magma chamber; B. concave-up-
ward (bowl-shaped) sheets from a spherical magma chamber; C. concave-downward (trumpet-shaped) sheets from a sill-shaped magma chamber; and D. planar parallel sheets from a
laccolith-like chamber.
A. After Chadwick and Dieterich (1995) and Gudmundsson (1998). B. After Chadwick and Dieterich (1995) and Gudmundsson (1998). C. After Phillips (1974) and Chadwick and Dieterich
(1995). D. After Bistacchi et al. (2012).

particularly worth noting the similarity of the average dip of the in- interpreted this architecture of the centrally-inclined sheets as the effect
clined sheets near the centre of the swarm with respect to those of magma forces exerted by an ellipsoidic magma chamber with a flat
cropping out at the periphery. This is also quantified by the distribution upper surface. Also seismic observations (e.g., Barker and Malone,
of sheet dip measured along four transects across the whole swarm 1991; White et al., 2008) and geodetic inversions from active volcanic
(Fig. 16). Also at the Tejeda complex, in the island of Gran Canaria, areas (e.g., Newman et al., 2006) suggest an ellipsoidal shape for
there is a centrally-inclined sheet swarm with homogeneous dip angles magma chambers located at a few kilometres depth. At Ardnamurchan
measured across the whole swarm (Fig. 17). Schirnick et al. (1999) (NW Scotland) there is a centrally-inclined sheet swarm with a

Fig. 13. A. Photo of the western centrally-dipping sheet swarm of the Lysuskard–Midhyrna intrusions. B. Graph of sheet dip angle vs. distance from the centre of each centrally-dipping
sheet swarm: box and circles represent the data of the two swarms; note that these sheets show no gradual variation in dip attitude with distance from the focus area, are rectilinear
in section view, and intrude with the same geometry the main intrusive bodies as well as the layered Tertiary lavas. Box shows the location.
A. Modified after Tibaldi et al. (2013).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 99

Fig. 14. A. Photo of the western part of Thverfell (Iceland) centrally-inclined sheet swarm: note the constant dip angle of the sheets. B. Photo and C. sketch of part of the sheet swarm where
it is possible to observe that some of the intrusions cross the boundary between lava flows and the underlying hyaloclastite breccias. Box shows the location.

diameter of 18 km (Richey and Thomas, 1930; Emeleus, 2009). This ex- vary in thickness from a few centimetres to about 60 m, with a mean
ample has been recently used by Magee et al. (2012) to question the for- thickness of 2–6 m (Gudmundsson et al., in press). The data of
mation of this sheet pattern by proposing an alternative model. These Fig. 18B indicate that the average dip angle is 41° and the average thick-
authors claimed that the Ardnamurchan sheet swarm may be linked ness is 0.93 m. The average inclined sheet dip angle in the British Isles
to laterally propagating regional dykes, sourced from laterally adjacent has been found at ~45° (Billings, 1972; Hills, 1972), and at Gran Canaria
magmatic systems, which are deflected around a central complex by at ~41° (Schirnick et al., 1999), and at some individual sheet swarms in
stress field interference. Iceland at 34° (Gudmundsson et al. 1996). The average sheet thickness
Since the geometry and location of centrally-inclined sheet of 0.93 m found from the data of Fig. 18B is larger than the one found
swarms is considered as mainly dependant from the geometry of the in some individual swarm in Iceland, which is around 0.1 m
overpressured magma chamber, it is important to quantify the maxi- (Gudmundsson et al., 1996) but identical to the average thickness of
mum distance the sheets can reach. In fact, if the stress exerted by the the 1128 sheets in the Thverartindur Central Volcano (Klausen, 2004).
magma chamber overpressure on the surrounding rock guides the de- These field examples indicate the effects of overpressured magma
velopment of these inclined sheets, we can estimate the distance by chambers on the overlying host rock. These data can be compared
which this stress dominates. I have collected most of the available with numerical modelling that takes into account the variation in the
data on well exposed centrally-inclined sheet swarms in order to evalu- stress field around an overpressured magma chamber. This overpres-
ate their width (Fig. 18A). Here it is possible to see that most swarms sure can develop through a series of different causes that include injec-
have diameters ranging from a minimum of 5 km to a maximum of tion of new magma from below into the chamber, volatile exsolution,
20 km, with an average value of 12.3 km. In order to quantify also melting or solidification of wall rocks, and magmatic differentiation
the typical thickness of centrally-inclined sheets, I collected data (Tait et al., 1989; Folch and Martì, 1998; Annen and Sparks, 2002). The
from different locations in Iceland and in the British Isles and overpressure considered large enough to produce fracture in the cham-
summarised them in the graph of Fig. 18B. In this graph it is possible ber walls and propagation of intrusive sheets depends on the criterion
to observe the distribution of dip angles and thickness for a population of rock rupture; Sartoris et al. (1990) considered a value of 1–10 MPa
of 2340 sheets, with the larger number of intrusions with dip angles in large enough to exceed the tensile strength of the surrounding rocks
the range of 20–65° and thickness of 0.2–1.8 m (largest thickness up under a tensile mechanism of rupture, a value adopted also by other au-
to 10–12 m). It has been shown that in some centrally-inclined sheet thors (e.g. Atkinson and Meredith, 1987; Gudmundsson, 1988). Other
swarms, shallower-dipping sheets are thinner than steeper-dipping authors suggest that this is likely an underestimation (e.g. Karlstrom
sheets (Gudmundsson, 2003), which is not consistent with the data pre- et al., 2009) and proposed greater values (10–100 MPa) based on ther-
sented above. This might be due to the fact that it has been carefully mal considerations of long-distance dyke propagation and on the com-
avoided to include in the sheet population of Fig. 18B any steeply- position and tectonic settings of the chamber (Rubin, 1995b; Jellinek
inclined sheets that might in reality represent regional dykes, which is and DePaolo, 2003).
to say to intrusions that are not related to the central magma chamber. A pressurised magma chamber changes the overall stress pattern of
Regional dykes, in fact, are much thicker in average than inclined sheets the host rock significantly, as shown by numerical and analogue model-
linked to local magma chambers: In Iceland, for example, regional dykes ling (e.g. McLeod and Tait, 1999). In the case of a pressurised spherical
100 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 15. Photos of the centrally-inclined sheets at the Cuillin Igneous Complex at the Isle of Skye (UK). Note that the sheets have the average dip angle both in the core of the swarm
(A) where the highest frequency of intrusion is observed, and in the peripheral part (C). B. Location of the photos.

magma chamber, boundary element results show – in the case of inter- concluded that neglecting the dyke effect gives in any case good quali-
nal magmatic excess pressure of 5 MPa as the only loading – that the tative agreement in the pattern of principal-stress trajectories, but sub-
trajectories of the greatest principal stress (σ1, ticks) depart at right stantially different estimates of the stresses. Based on the above, it can
angle from the chamber walls, creating a radial pattern in section view be assumed that magma escaping from the chamber walls should follow
(Fig. 10B) (Gudmundsson, 1998). Based on the work of Anderson the direction of less resistance giving rise to intrusive sheets that prop-
(1951), it is commonly assumed that magma paths select the orienta- agate along planes that contain σ1 and σ2 (i.e. normal to σ3). Fig. 10C
tion of “least resistance” by opening along planes across which the shows the boundary element results of the stress field around a vertical
normal stress is the lowest. The consequence is that magma paths fol- section of a spherical magma chamber subject to remote horizontal ten-
low and trace trajectories perpendicular to σ3 within the stress field oc- sile stress of 5 MPa as the only loading. The trajectories of the σ1 are
curring before sheet propagation (Stevens, 1911; Anderson, 1936, 1938; mostly vertical to sub-vertical suggesting that dykes departing from
Odé, 1957). This conclusion has been questioned by McKenzie et al. the magma chamber have a vertical geometry at a short distance.
(1992) and Meriaux and Lister (2002) who pointed out that several Fig. 10D presents the same setting as in Fig. 10A but with a flat
theoretical works ignored the fact that dykes radically alter the sur- magma chamber (laccolith-like shape). Although we are aware of the
rounding stress field as they propagate. Meriaux and Lister (2002) limitation of considering only magma overpressure as the acting stress,
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 101

Fig. 16. Distribution of the sheet dip angles measured along four transects across the whole centrally-inclined sheet swarm of the Cuillin Complex (Island of Skye, UK). Note the similarity
between the average dip angle of the inclined sheets near the centre of the swarm and those cropping out at the periphery.
Redrawn after Tibaldi et al. (2011).

this example is useful to discern the contribution of the magma cham- with respect to the magma chamber. Assuming the case of intrusive
ber force in dictating the intrusive sheets pattern. In this case in fact, sheets that follow paths corresponding to surfaces perpendicular to
the trajectories of σ1 above the magma chamber are mostly parallel to σ3 (thus containing σ1 and σ2), this case corresponds to the forma-
each other, due to the flat upper surface of the chamber, resulting in a tion of a ring dyke complex. When σ3 is low, the dip of ring dykes
series of inclined sheets whose dip angle does not vary significantly. at 2–3 km depth is 65–80°. Within a distance from the central axis
As a matter of fact, the host rock is subject to further stresses of dif- of about 1–1.2 diameters of the magma chamber, radial dykes are
ferent origins, including the lithostatic component and the tectonic favoured for intermediate overpressures, and centrally-inclined sheets
component of the regional far field; Russo et al. (1996, 1997), Russo dominate in the axial domain for higher (≥10 MPa) overpressures.
and Giberti (2000), van Wyk de Vries and Matela (1998), Grosfils The dip angles of centrally-inclined sheets are predicted in the
(2007) and Martì and Geyer (2009) present numerical models where range 45–65° for overpressure ≥ 20 MPa. In all cases, radial dykes
the gravitational body force and tectonic boundary stresses are consid- dominate beyond the 1–1.2 diameter limit. By simulating different
ered together with the stresses transmitted by the magma chamber. depths of magma chamber in the range 2–3.5 km, the only difference
Bistacchi et al. (2012), in particular, focused on the possible trajectories is that inclined stress trajectories consistent with centrally-inclined
of intrusive sheets modelled under a total stress field resulting from sheets appear for lower overpressures in the case of a shallower
three components of boundary forces (Ranalli, 1995): (1) the superpo- magma chamber. By using different possible shapes of topography,
sition of diverse magma overpressures in the magma chamber; (2) the results of stress trajectories at 2–3 km depth do not vary sensibly,
presence of different free topographic surfaces ranging from a flat to- apart when there is a caldera depression that produces variation in
pography to a positive topography, simulating the load of a volcanic ed- the trajectories. At shallower depths, the topography influence in-
ifice, or to a negative surface corresponding to a caldera; (3) regional creases. The effect of an increasing gravitational force or of an increas-
tectonics with diverse remote stresses, plus a body force (4) corre- ing regional extensional tectonics is to steepen the sheets above the
sponding to gravitational acceleration. The analysis used a set of magma chamber, whereas the effect of increasing magma overpres-
quasi-static Finite Element Method mechanical models that showed sure (in the case of a chamber with a lobate shape) or increasing re-
the configuration of the resulting stress field (Fig. 19) under diverse gional horizontal compression is to decrease the sheet dip angle.
pressures in an oblate magma chamber. With negative values, which In the field, it has been frequently observed that centrally-inclined
correspond to a deflating chamber, the trajectories of σ1 dip outwards sheets are associated with radial dykes; this can be clearly seen, for

Fig. 17. Geological–structural section of the Tejeda complex, on the island of Gran Canaria, with a centrally-inclined sheet swarm that shows homogeneous dip angles across the whole
swarm. The authors interpreted the geometry of the centrally-inclined sheets as the effect of the magma force exerted by an ellipsoidic magma chamber with a flat upper surface.
Redrawn after Schirnic et al. (1999).
102 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 18. A. Graph of the diameter of well exposed centrally-inclined sheet swarms; it is possible to note that most swarms have a diameter that ranges from a minimum of 5 km to a max-
imum of 20 km, with an average value of 12.3 km. B. Graph of the distribution of dip angles (X axis) and thickness (Y axis) of 2340 sheets surveyed at the centrally-inclined sheet systems of
Iceland and British Islands (Ardnamurchan). Note the larger number of intrusions with dip angles in the range of 20–65° and thickness of 0.2–1.8 m.
B. Data after Magee et al. (2012).

example, at the Cuillin Complex centrally-inclined sheet swarm at Skye values from 10− 4 to 10−2 m/s have been classically considered
Island (UK), where Tibaldi et al. (2011) recognised three zones with dif- (Huppert and Sparks, 1981; Sparks et al., 1984), although recent find-
ferent characteristics: (1) a central zone in the core of the centrally- ings show that values up to 11–17 m/s for basaltic magmas are possible
inclined sheet swarm where inclined sheets dominate; (2) a more (Lloyd et al., 2014), as well as for andesitic magmas from 10− 3 to
external zone where inclined sheets coexist with vertical dykes; and 10−2 m/s (Ruprecht and Plank, 2013) and up to 1 m/s even for rhyolitic
(3) an outer zone where dykes dominate (Fig. 20). In zone 2) several magmas (Castro and Dingwell, 2009).
generations of dykes are interleaved with generations of inclined sheets, All this description assumes that intrusive sheets follow planes of
suggesting quick changes in the state of stress and consequent geome- fracturing that contain σ1 and σ2. Some field evidence shows another
try of intrusion. These field data are consistent with the above described possibility: at the Isle of Skye it has been observed that some inclined
numerical models in which the geometry of intrusions above a magma sheet intruded along shear planes, although the great majority followed
chamber changes at a critical distance that depends on the state of stress fractures with an opening mode I (Fig. 21). Lower angle shear planes
resulting from the combination of the various sources. In zone 1) the show normal motions, whereas steeper planes have reverse motions
paths of the intrusive sheets are controlled by the magma chamber (Bistacchi et al., 2012). These planes form a conjugate set bisected by
overpressure. In zone 2) the intrusive paths can rapidly change follow- the dominant dilational sheets, and are thus consistent with each
ing pulsations of the magma chamber: with large inflation the zone of other. The frequently found scatter in centrally-inclined sheet dip angles
centrally-inclined sheets expand outwards, whereas with a decrease can thus be explained as the result of intrusions along both mode I frac-
of the overpressure, the magma chamber influence decreases and ture planes and conjugate shear planes with a higher- or lower-than-
dykes develop. In zone 3) the regional background stress dominates average dip angle. The presence of reverse shear planes has been
and intrusive sheets departing form the same magma chamber assume found also by Schirnick et al. (1999) in Gran Canaria and by Mathieu
a dyke-like geometry. Galland et al. (2014), based on analogue models, and van Wyk de Vries (2009) at the Mull volcano in Scotland.
suggested that the shift from cone sheets to dykes is controlled by the The centrally-inclined sheets contribute to magma transport up-
magma dynamics expressed by the ratio between the viscous stresses ward: the sheets located above the centre of the magma chamber may
in the flowing magma to the host rock strength; dykes form preferen- be steeper and tend to transport magma vertically into the volcano,
tially when their magma source is deep compared to its size, or when whereas those located more externally tend to move magma laterally
magma influx rate (or viscosity) is low, and vice-versa for cone sheets. with respect to the volcano, possibly contributing to lateral (eccentric)
Some analogue models show that these cone sheets can depart from eruptions at the volcano's foot. Anyway, only a small fraction of these
dykes, although field data are necessary to confirm this hypothesis. sheets is capable of bringing magma to the surface and most remain
The importance of considering the possible effects of the velocity of trapped into the shallower crust. This might occur through bending or
magma upwelling is reinforced by the fact that it is becoming clear stoping. As regards bending, inclined sheets and dykes can rotate
that very diverse values are possible for both mafic and felsic magmas; to a shallower dip angle and reach a sill-like configuration. An old
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 103

Fig. 19. Results of quasi-static Finite Element Method mechanical models showing the configuration of the stress field around an oblate magma chamber that incorporates: (1) diverse
magma overpressures; (2) the presence of different free topographic surfaces ranging from a flat topography to a positive topography, simulating the load of a volcanic edifice, or to a
negative surface corresponding to a caldera; (3) regional tectonics with different remote stress values, plus a body force (4) corresponding to gravitational acceleration. Results are
shown on a vertical cross-sectional symmetry plane of the 2D axial-symmetric model. Each figure corresponds to an increasing magma chamber pressure. Vector symbols show orienta-
tion of σ1 and σ2; σ3 is always perpendicular to the σ1–σ2 plane, σ2 axes plot as small dots in areas where they are perpendicular to the plane of the cross-section. Colour shows the values
of least compressive to tensional σ3 stress axis; red indicates the negative values, related to the emplacement of magma in tensional fissures.
Modified after Bistacchi et al. (2012).
104 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 20. At the Cuillin Complex cone sheet system (Skye Island, UK), there are three zones with different characteristics: (1) a central zone in the core of the cone sheet where inclined
sheets dominate; (2) a more external zone where inclined sheets coexist with vertical dykes; and (3) an outer zone where dykes dominate.
Modified after Tibaldi et al. (2011).

explanation for the dyke-sill transition was the effect of neutral buoyan- (Menand et al., 2010). Among all these possibilities, field data suggest
cy forces (Corry, 1988). Nevertheless, much field evidence shows that that the further propagation of sheets is strongly influenced by litholog-
sheets were injected through host rock successions dominantly com- ic boundaries in the host rock. As an example, at the Thverfell eroded
posed of fragmented deposits, like hyaloclastites and breccias, the den- volcano (SW Iceland), the centrally-inclined sheets bend and reach a
sity of which is lower than magma density (Pasquarè and Tibaldi, 2007; horizontal geometry as they approach the contact between a lava suc-
Tibaldi and Pasquarè, 2008; Tibaldi et al., 2008c), without any geomet- cession and the underlying hyaloclastites (Fig. 22A–B). As another ex-
rical change, so neutral buoyancy cannot be the main explanation. ample, a dyke bends into a sill at a similar contact at the Stardalur
Similarly, Thomson (2007) suggested that the presence of lithological eroded volcano (SW Iceland) (Fig. 22C), and a similar configuration
contrasts, particularly ductile horizons such as overpressured shales, can be seen also at Nysiros volcano (Greece), where a dyke bends into
may guide sill formation at any depth below the neutrally buoyancy a sill near a contact between hyaloclastites and lavas (Fig. 22D). This
level. Another explanation may be the intersection of the intrusive field evidence suggests that dykes and inclined sheets can be diverted
sheet with an already existing horizontal, freely slipping joint or stopped at contacts between layers with lithologies having very dif-
(Weertman, 1980) or in a similar way, the intersection between a ferent stiffness (Gudmundsson, 1986; Kavanagh et al., 2006).
weak bedding plane and a steep normal fault at shallow depths Changes in intrusion geometry can also derive from dyke–dyke in-
(Gaffney et al., 2007). The upward propagation of dykes and inclined teractions. Numerical and analogue simulations have been developed
sheets can be diverted also due to the generation of stress barriers, i.e. to analyse how ascending dykes might interact, showing that their
layers with local stresses unfavourable for the intrusion propagation self-induced stress field may locally be more significant than the region-
(Gretener, 1969; Gudmundsson, 1986, 1990; Parsons et al., 1992; al stress field (Kühn and Dahm, 2008). Coeval ascending dykes may
Gudmundsson, 2011). The deviation from a dyke to a sill has also been converge towards regions of increased dyke densities and may lead to
reproduced experimentally by Kavanagh et al. (2006) at the interface the formation of sills; alternatively, they can change their dip and
between upper, rigid layers overlaying lower, weaker layers. Valentine magma paths by getting closer to each other, or they can merge crating
and Krogh (2006) explained dyke bending in terms of local stress rota- dykes with increasing magma flux (Ito and Martel, 2002; Canon-Tapia
tion along 3-D variations on a normal fault plane. Finally, dykes can and Merle, 2006), or they can have a higher probability of becoming
bend into sills in the case of remote tectonic compression, as they adjust arrested (Jin and Johnson, 2008). The dyke–dyke interaction is greater
their trajectory reaching a geometry parallel to the σ1–σ2 plane when the intrusion is more than some dyke lengths away from the
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 105

Fig. 21. Photos taken at the Cuillin cone sheet in the Skye Island (UK), showing examples of inclined sheets intruded along shear planes with reverse motion (R) and normal motion (N),
although most followed fractures with an opening mode I (extension normal to the fracture wall) (T). Lower angle shear planes show normal motions, whereas steeper planes have re-
verse motions, and both form a conjugate set bisected by the dominant dilational sheets (T), and are thus consistent with each other. The frequently found scatter in cone sheet dip angles
can thus be explained as the result of intrusions along both mode I fracture planes and conjugate shear planes with a higher- or lower-than-average dip angle.

feeding reservoir or the free surface, whereas the interaction is general- were obtained by Peltier et al. (2005) who, using seismic data, analysed
ly small when the horizontal tensional stress is large compared to the the direction and velocity of dyke propagation from a shallow reservoir
compressional stresses induced by the emplacement of a single dyke at a depth of 3 km at Piton de la Fournaise (Reunion Island) between
(Kühn and Dahm, 2008). 1998 and 2004. They found a change from vertical ascent to slower lat-
Numerical and analogue models indicate that also the load exerted eral propagation when the dyke reached the Dolomieu cone base. This
by a volcano can influence the trajectory of the magma path below. change was attributed to the presence of a fractured zone as well as to
Pinel and Jaupart (2004) found, by numerical modelling, that if a volca- the effect of volcano load. The final geometry of a sheet is also a function
no is present, at small distances from the axis, confining stresses due to of the distribution of magma overpressure within the intrusion; dykes
the volcano load hinder vertical magma propagation. This produces hor- with either uniform internal overpressure or with a simple viscous pres-
izontal dyke propagation and, in case the dyke reaches the surface, the sure drop propagate across much greater distances than the corre-
formation of a distal eruptive centre. They found also that, everything sponding unpressurised cracks, and uniformly loaded dykes tend to
else being equal, a decreasing magma supply rate or a decreasing propagate with less curvature than those with a viscous pressure drop
magma viscosity generate eruptive centres at increasing distances (Meriaux and Lister, 2002).
from the focal area. Gaffney and Damjanac (2006) found a similar ten- Several field examples show that crosscutting relationships between
dency of magma flow to be diverted away from a highland zone (like generations of intrusive sheets may give clues to the parameters
a volcano) towards the lowland. By analogue modelling, Kervyn et al. that control the geometry of magma conduits. For example, as already
(2009) showed that the local loading stress field, due to the presence introduced before, at the eroded Thverfell volcano (Iceland) the
of a volcano, favours rising magma away from the volcano centre if a centrally-inclined sheets bend until they become horizontal at the con-
central conduit is not established or is blocked. Experiments show that tact between the lava succession and the underlying hyaloclastites
the load compressive stress stops rising dykes as they approach the (Fig. 22A–B). The resulting sills stacked on top of each other and created
cone stress field especially in the case of dykes with limited overpres- a laccolith-like body, showing the difficulty of magma uprising to the
sure and steep volcanic cones (Fig. 23). Dyke overpressure builds up surface (Tibaldi et al., 2008c). However, the stacked sills are crosscut
as intrusion continues and dykes extend laterally until their tips rise by a vertical dyke that affected all the exposed rock succession
vertically again, leading to eruptions. This might explain the presence (Fig. 24A–B). A similar situation can be seen at another site of the
of volcanic centres located at the base of a volcano, whose distance same eroded Thverfell volcano, where inclined sheets intruded the
from the volcano's axis depends on volcano slope angle, magma over- lava succession and are in turn cut by a series of parallel, sub-vertical
pressure and substratum thickness, whereas extensional processes fa- dykes (Fig. 24C–D). Both cases reflect a change in the geometry of
vour magma propagation in the axial volcanic zone. Similar results the plumbing system, that first favoured lateral magma transport
106 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 22. Examples suggesting that dykes and inclined sheets can be diverted at contacts between layers with lithologies characterised by very different stiffness. Photo (A) and sketch (B) at
the Thverfell eroded volcano (SW Iceland), showing centrally-inclined sheets that bend attaining a horizontal geometry as they approach the contact between a lava succession and the
underlying hyaloclastites. At the Stardalur eroded volcano (SW Iceland) (C), a dyke bends into a sill at a similar contact. At the Nysiros volcano (Greece) (D), a similar configuration can be
seen where a dyke bends into a sill nearby a contact hyaloclastite/lava.

(horizontal or inclined) and then vertical magma upwelling. This can be the dyke swarm has a fractionated character with wide compositional
explained in terms of a change in the stress state within the host rock, variations. Geshi (2005) interpreted this compositional change by frac-
induced by a rotation of the σ3 axis from vertical (sill case) or gentle- tional crystallisation due to the decline in the supply of less fractionated
dipping (inclined sheet) to horizontal. This change is the effect of a de- hot magma into the reservoir. The change in the geometry of the plumb-
crease of the horizontal far field stress, as supported by the fact that the ing system from centrally-inclined sheets to dyke has thus been
dykes that intruded the whole rock succession are perpendicular to the interpreted as reflecting a change from conditions with large overpres-
orientation of the Icelandic Rift extension axis. If we consider that the sure and high magma supply, to conditions with pressure decrease in
stacked sills and the centrally-inclined sheets of the Thverfell example the magma chamber and lower magma supply. During the decline of
were located below the volcano, the successive dyke intrusions indicate magma replenishment into the reservoir, dykes reach an orientation
that extensional tectonics may favour upward magma propagation in perpendicular to the tectonic σ3 (Geshi, 2005).
the axial volcanic zone, consistent with the modelling of Kervyn et al. Most of the above described papers assume that sheets intrude along
(2009). A similar situation was encountered at the Otoge Complex planes containing σ1 and σ2. However, a sheet could propagate oblique
(Japan), where field evidence shows a centrally-inclined sheet system to this plane if it invades a preexisting fracture that is misaligned
crosscut by a swarm of more recent parallel dykes (Geshi, 2005). Petro- with respect to the principal stress directions; sheets that occupy
graphic analyses indicate that the centrally-inclined sheets have a less faults or preexisting joints are described in several studies (Johnson,
fractionated character with little compositional variations, whereas 1961; Curtie and Ferguson, 1970; Gudmundsson, 1983; Delaney and
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 107

Fig. 23. Analogue models showed that the local loading stress field, due to the presence of a volcano, favours magma ascent away from the volcano centre. Dyke overpressure builds up as
intrusion continues and dykes extend laterally until their tips rise vertically again, leading to eruption.
Modified after Kervyn et al. (2009).

Gartner, 1997). Ziv et al. (2000) presented a numerical model for sheet along very diverse paths during the seismo-volcanic crisis of 2011–
intrusion into preexisting fractures that are oblique to the principal 2013 at El Hierro (Canary Islands). Magma intruded southward and
stresses. They conclude that it appears to be very difficult for a sheet then northward along a N–S-striking pre-existing fracture zone, and
to follow a preexisting fault unless one or more of the following condi- successively migrated along an E–W-striking fracture zone. The differ-
tions are met: (1) the fracture is nearly perpendicular to σ3; (2) the re- ent intrusions were guided by a combination of two magma pressure
solved shear stress on the fracture is small compared to the excess sources located at diverse position and depth, and the presence of the
magma pressure; and (3) the effective ambient sheet-normal stress is discontinuities.
small compared to the rock tensile strength. The same authors suggest
that it may be quite difficult for sheets emerging from midcrustal to 6. Plumbing system in the interior of volcanoes
lower crustal depths to follow faults that are oblique to σ3. One of the
main observations is related to the fact that rocks are usually fractured As we look at the structural characteristics of plumbing systems
medium with possible pervasive discontinuities. A preexisting fault within volcanoes, we have to consider further parameters that play a
with a different orientation from the σ1–σ2 plane acting at the time of role in dictating the final configuration. As we have seen in the previous
sheet intrusion is usually surrounded and intersected by other faults chapters, the structure of magma conduits is essentially governed by the
and joints of various lengths. Field data demonstrate in fact that combination of preexisting structures and the total state of stress of the
magma paths can change orientation at short distance when the rock host rock. The latter results from several possible inputs, including far-
is densely fractured (e.g. Fig. 25). Thus the presence of these secondary field tectonics, lithostatic loading and magma force, but as Earth's sur-
discontinuities makes it even more difficult for the magma to follow a face is approached, topography contributes more sensibly to the total
misaligned pre-existing fault (Ziv et al., 2000). In spite of this sugges- stress field. In order to distinguish the various components of the stress
tion, a recent work by García et al. (2014) demonstrated, by seismolog- sources that contribute to the total stress field inside a volcanic edifice, I
ical data and numerical modelling, that magma migrated several times will start analysing the tectonic component vs. the magma component.
108 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 24. Examples of changes in the geometry of magma plumbing system at shallow level at the eroded Thverfell volcano (Iceland). A. Sketch and B. photo of stacked sills crosscut by a
vertical dyke that affected all the exposed rock succession. C. Sketch and D. photo of centrally-inclined sheets intruded into the lava succession and, in turn, cut by a series of parallel, sub-
vertical dykes. Both settings reflect a change in the geometry of the plumbing system that first favoured lateral magma transport and then vertical magma upwelling. This can be explained
in terms of a change in the stress state in the host rock promoted by a rotation of the σ3 axis from vertical (sill case) or gentle-dipping (inclined sheet) to horizontal.

6.1. Tectonic vs. magmatic components stress dominates in the volcano, whereas Stromboli reflects the control
of far-field tectonic stress. More in detail, Alicudi developed a radial pat-
Field data show that the geometry of magma plumbing system in tern of dykes that has been caused by a dominating magma force along a
volcanoes changes in different tectonic settings and with the variation central main conduit zone. Magma exerts an isotropic force on the sur-
of the tectonic stress magnitude. To introduce these concepts, I will rounding rocks that produces hydraulic fracturing followed by radial
use the examples of Alicudi and Stromboli volcanoes (Aeolian Arc, dyking (Fig. 27A). In this case the magma flow direction should be hor-
southern Tyrrhenian Sea, Italy, Fig. 26): both volcanoes have an izontal or sub-horizontal and outward-directed. Another explanation
age b 100 ka BP, lie above the same substratum, have a similar size for the radial pattern refers to the distribution of the gravitational
(rising about 2.5 km from sea bottom), and are mainly of basaltic and stresses due to the load of the edifice. The load controls the trajectories
basaltic–andesitic composition; however, they have different patterns of the σ1 that become subparallel to the volcano slopes while σ3 is tan-
of dykes (or eruptive fissures that represent their homologous at the gential (Dieterich, 1988). A third explanation is that radial dykes may be
surface). The evolution of the summit, emergent part of Alicudi has fed from a deeper level of the magma plumbing system at the basement
been characterised by several main episodes of cone growth alternating of the volcano (Geshi, 2008) (Fig. 27). In this third case, the source may
with lateral volcano-tectonic collapses directed southward (Tibaldi be located beneath the centre of the radial dyke structure where a
et al., 2005). During the growth stages, dyking occurred with a radial pressurised magma chamber with a vertically elongated spindle shape
pattern, which can be defined as a group of dykes that have a similar will induce a zone of compression in which the radial dykes can be
frequency in every direction with regard to the axis of the volcanic edi- emplaced (Chadwick and Dieterich, 1995; Gudmundsson, 1998);
fice (Fig. 26A). Several dyke generations, suggested also by some magma flow should be oriented vertically or obliquely and upward-
crosscutting relationships, show the persistence of radial dyking that re- directed. This process will be encouraged when the central conduit is
flects the stability over time of the plumbing system geometry. Oceano- solidified (Porreca et al., 2006). Other examples of radial dykes include
graphic studies by Calanchi et al. (1995) show that no large faults are Fernandina volcano at the Galápagos islands (Chadwick and Dieterich,
present beneath Alicudi volcano. Also the growth of Stromboli volcano 1995), Summer Coon volcano in Colorado (Poland et al., 2004, 2008),
was repeatedly interrupted by caldera collapses and lateral collapses Kliuchevskoi volcano in Kamchatka (Takada, 1997) and Komochi volca-
(Pasquarè et al., 1993; Tibaldi, 1996, 2001). The plumbing system ac- no in Japan (Geshi, 2008).
companying the various growth phases was strongly dominated by On the contrary, Stromboli shows a main swarm of rectilinear paral-
dyking along a NE-trending weakness axial zone crossing the summit lel dykes that crosscuts the summit zone of the volcano and has been ac-
of the volcano (Fig. 26B). Inside this zone, single dykes strike between tive since 100 ka BP. Another two minor zones of dyking have been
NNE and ENE. Relationships between the unconformity surfaces of the active at 85–20 ka BP (Corazzato et al., 2008) and b 13 ka BP (Tibaldi,
various volcano growth phases and these dykes indicate that the NE- 2001) and will be discussed later. Rectilinear dyke swarms and their
trending weakness zone persisted over the entire history of the volcano surface expression in the form of aligned volcanic centres and eruptive
(Tibaldi, 1996, 2003). fissures are defined as “volcano-tectonic rift zones” (VTRZ). VTRZ are
These two volcanoes can thus be considered examples of end mem- usually aligned with the regional tectonic stress field, i.e. perpendicular
bers of the possible patterns of plumbing systems; all other conditions to the σ3 and parallel to the horizontal greatest principal stress (σHmax),
being similar, Alicudi represents the end member where local magma which indicates a control on intrusion geometry exerted by far-field
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 109

Fig. 25. A. Photo and B. sketch of a dyke intruded in a rock affected by several preexisting
fractures with diverse orientations (Asja Peninsula, west Iceland). Note the sharp change
in dyke geometry along the intrusion path.

stresses. Other examples are Askja and Krafla in Iceland (Gudmundsson


and Nilsen, 2006), Erta Ale, Ale Bagu, Dabbahu and Gabho in Afar
(Barberi and Varet, 1970; Wright et al., 2006), Tongariro in New
Zealand (Nairn et al., 1998), Nyiragongo and Nyamuragira in Congo
(Komorowski, 2002). The interpretation of the radial pattern at Alicudi
and of the VTRZ pattern at Stromboli in terms of the control exerted by
magma forces vs. tectonic forces, respectively, is consistent with the
geodynamic setting of the Aeolian Volcanic Arc (Fig. 26C). The western
part of the arc has poorly developed faults and in the late Quaternary
has been subject to low tectonic loading or to compression (Neri et al.,
2003; Goes et al., 2004; Argnani et al., 2007; Di Roberto et al., 2008).
This situation is suitable for magma stress dominating over tectonic
stress. The eastern part of the arc, instead, has been subject to regional
extension with a NW–SE trend of σ3, creating suitable conditions for Fig. 26. Alicudi (A) and Stromboli (B) volcanoes (Aeolian Arc, southern Tyrrhenian Sea,
the development of the NE–SW rift at Stromboli. Italy) have an age b 100 ka BP, lie above the same substratum, have similar size and geo-
Between these two end-members there may be intermediate set- chemical characteristics, but have different dyke patterns. Alicudi shows a radial dyking
tings where both magma pressure and tectonic stresses can interact, pattern, whereas Stromboli has a main NE-striking rift zone active since 100 ka ago and
two minor zones of dyking (active at 64–36 ka BP and b13 ka BP). The two end-member
giving rise to more complex magma path patterns within volcanoes. A
dyke patterns reflect the effect of a hydraulic force exerted by magma in the case of
classical example is the dyke pattern at Spanish Peaks (Colorado, Alicudi, and a control exerted by regional tectonics at Stromboli. C. Shows the tectonic set-
USA), a site which has been studied for decades (Fig. 28A). Dykes are tings of the two volcanoes where it is possible to observe that Stromboli lies in an area af-
arranged radially around the Spanish Peaks stock and have been fected by extension normal to its rift zone.
interpreted to result from radial stresses caused by the pressurised Data of Alicudi and regional tectonics from Tibaldi et al. (2005); data of Stromboli from
Pasquarè et al. (1993) and Tibaldi, 1996, 2001).
stock (Odé, 1957). At larger distances from the stock, dykes gradually
bend until they reach an average N80° strike. All more recent works
agree that the overall dyke orientations result from the combination
110 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

edifice as at Askja, Iceland (Gudmundsson and Nilsen, 2006) and Erta


Ale, Tat Ali and several other volcanoes in the Afar region (Barberi and
Varet, 1970; Acocella and Neri, 2009). Sometimes the rift zone may ex-
tend beyond the base of the volcano with dykes reaching as far as tens of
km from the volcano perimeter, such as at Krafla in Iceland and at sev-
eral volcanoes in the Afro-Arabian Rift (Gudmundsson, 1995; Ebinger
and Casey, 2001; Sigmundsson, 2006). This orientation of the regional
stress tensor with a vertical σ1 gives way to swarms of vertical, parallel
dykes inside the volcano, a percentage of which are capable of crossing
the summit part of the cone or its flanks, giving rise to aligned pyroclas-
tic cones and fissure eruptions. This situation is the simplest from a
point of view of the structure of the shallower magma feeding system,
although it may be complicated by the interaction with topography, as
we will see in the next chapter. Another complication is represented
by the possible development of sills also within volcanoes under an ex-
tensional stress field. In fact, sills have been observed, for example, in
extensional settings associated with active/recent volcanoes such as
Stromboli and Nysiros, and at eroded centres like in the Paiute Ridge
area (Nevada, USA) (Valentine and Krogh, 2006) and in the North Atlan-
tic Igneous Province (Hansen, 2006). It is worth noting that in general
these sills have been observed at lithological boundaries (Fig. 22D), sug-
gesting that their geometry has been controlled by differences in stiff-
ness within the rock succession (Gudmundsson, 1986), as already
mentioned.

6.2.2. Vertical σ3
Volcanoes can also overlie zones of active compression with hori-
zontal σ1 and σ2 (for a review see Tibaldi et al., 2010) and there are
some field and experimental data that indicate which planes are the
prime candidates to host magma paths in such contractional settings.
Magma migration along pre-existing faults is consistent with several
cases found in different geodynamic settings and spanning from evi-
dence at deep crustal level (e.g. Rosenberg, 2004), to near surface
level (e.g. Tibaldi et al., 2010). Below volcanoes, examples from different
contractional settings indicate that the control exerted by ramp and flat
Fig. 27. Sketch of possible models of propagation of radial dykes in volcanoes. A. Magma structures can occur after the main collision event or during active
exerts an isotropic force along the central conduit walls that produces hydraulic fracturing thrust movements (syn-kinematic). Examples of the first case are the
followed by radial dyking; this is favoured in the case of an open conduit. B. Radial dykes early Proterozoic Chilimanzi granites (Zimbabwe) that were emplaced
may be fed from a deeper level of the magma plumbing system at the basement of the vol-
50 Ma after a continental collision event (Fig. 29A) (Dirks and Jelsma,
cano and propagate upward; this is more favoured in the case of a closed conduit.
B. Modified after Geshi (2008). 1998). These granites intruded the crust along sills coinciding with
sub-horizontal thrust and along inclined sheets corresponding to
ramp structures. In Nigeria, the Neoproterozoic Rahama monzogranite
pluton was emplaced 60 Ma after a collision event (Ferré et al., 1997,
of radially distributed stresses and remotely applied regional tectonic 2002). Magma was emplaced along reverse inclined shear zones
stresses (Muller and Pollard, 1977; Smith, 1987), although Muller interpreted as thrust ramps with structures of the intrusive rocks sug-
(1986) observed that there are four diverse dyke populations indicating gesting some late syn-kinematic deformations (Fig. 29B) (Ferré et al.,
a rotation of the remote tectonic stress field. The remote stresses caused 2012). The Miocene tonalites of the Hidaka Belt (Japan) were emplaced
the reorientation of the distal parts of dykes that reach a strike parallel along both ramp and flat structures of a duplex system (Fig. 29C)
to σHmax. A similar pattern has also been obtained by Meriaux and (Shimura, 1992; Toyoshima et al., 1994). Regarding examples of syn-
Lister (2002) by calculation of the trajectories of dyke paths following kinematic intrusions, the Palaeozoic Wyangala granites (Australia)
mode I opening of fractures. When dykes intercept the volcano slopes, were emplaced shortly after a continental collision event (Paterson
they produce radial fissure eruptions and fractures in the central upper- et al., 1990; Tobisch and Paterson, 1990). In particular for the Yarra plu-
most part of the edifice, and parallel fractures, eruptive fissures and ton, the data of Tobisch and Paterson (1990) are consistent with magma
aligned pyroclastic cones along two opposite volcano flanks (Fig. 28B) ascent along a ductile, reverse shear zone that was still moving during
(Nakamura, 1977; Nakamura et al., 1977; Meriaux and Lister, 2002, magma cooling, and this attests to syn-kinematic emplacement
and references therein). (Fig. 29D). Halliday et al. (1987) showed that alkaline magmas in NW
Scotland intruded along the Moine thrust zone and documented the
6.2. Intrusive sheets and orientation of tectonic stress tensor persistence of magmatism, over 30 Ma, in a belt shortened by thrusting.
Martí et al. (1992) found that the Cenozoic magmatism of the Valencia
6.2.1. Vertical σ1 trough (Spain) is characterised by a first cycle of Early to Middle Mio-
Regional tectonic stresses can propagate and dominate across a vol- cene age with calc-alkaline rocks emplaced during thrust development
cano with the result that they can guide the geometry of magma paths. under a compressional regime. Mkweli et al. (1995) studied the Umlali
Where the active regional tectonics is characterised by a vertical σ1, vol- Thrust Zone between the granite–greenstone rocks of the Zimbabwe
canoes are usually elongated perpendicularly to σ3 and show a volcano- craton and the granulite-facies rocks of the Limpopo Belt, demonstrat-
tectonic rift zone crossing the summit part of the cone or the entire ing the syn- to post-kinematic intrusion of porphyritic granites.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 111

Fig. 28. A. At Spanish Peaks (Colorado, USA) dykes (red lines) are arranged radially around the Spanish Peaks stock and have been interpreted as resulting from radial stresses caused by the
pressurised stock (Ode, 1957); at greater distances from the stock, dykes gradually bend and, in a distal position, assume a parallel strike that results from remotely applied regional tec-
tonic stresses (Muller and Pollard, 1977; Smith, 1987). Black lines here represent the calculated trajectories of dyke paths following mode I opening of fractures. B. When dykes intercept
the volcano slopes, they produce radial fissure eruptions and fractures in the central uppermost part of the edifice, and parallel fractures, eruptive fissures and aligned pyroclastic cones
along two opposite volcano flanks (Nakamura, 1977; Nakamura et al., 1977).
A. Modified after Meriaux and Lister (2002).

The analogue experiments carried out by Ferré et al. (2012) to sim- structures, similarly to the previously described field examples. They
ulate basic and felsic magma upwelling in a contractional setting show conclude that thrust ramp structures provide an inclined pathway for
that fluids of different viscosities can move along ramp and flat magma to ascend along pre-existing mechanical discontinuities. Ferrè

Fig. 29. Examples of control exerted by ramp and flat structures on magma intrusions. A. Early Proterozoic Chilimanzi granites (Zimbabwe) emplaced 50 Ma after a continental collision
event along ramp and flat structures (Dirks and Jelsma, 1998). B. Neoproterozoic Rahama monzogranite pluton emplaced 60 Ma after a collision event along ramp structures (Ferré et al.,
1997, 2002, 2012). C. Miocene tonalites of the Hidaka Belt (Japan) emplaced along both ramp and flat structures of a duplex system (Shimura, 1992; Toyoshima et al., 1994). D. Yarra pluton
of the Palaeozoic Wyangala granites (Australia) emplaced along a ductile, reverse shear zone that was still moving during magma cooling (Paterson et al., 1990; Tobisch and Paterson,
1990).
112 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

et al. (2012) also suggest that additional room for magma can be provid- thrust fault, most of the eruptions have occurred in the upper part of
ed by local dilational jogs along the thrust plane. Similar experiments the edifice. In the central part of the volcano, the magmatic conduits
carried out by Galland et al. (2007a) (Fig. 30A), show that fluids can are in the form of subvertical andesitic dykes, striking almost E–W. On
move upwards along thrust planes at the leading edges of a plateau. In the eastern and southern flanks, E–W trending alignments of volcanic
the experiment, vertical tension fractures formed on the surface of the domes probably follow underlying fractures (Galland et al., 2007a,b).
plateau, in a direction almost parallel to the imposed shortening. This Also in this case, dykes intruding the volcano are parallel to the direction
situation has been attributed by Galland et al. (2007a) to superposition of σ1. At Guagua Pichincha volcano, Ecuador (Legrand et al., 2002),
of a local load, because of the uplifted area, on the regional stress field. and Miyakejima volcano, Japan (Fujita et al., 2004), where the tectonic
At a shallow depth, the stress field was mainly due to the load of the settings are also compressional, seismic data suggest subhorizontal
uplifted area, so that the greatest stress was vertical, whereas at deeper magma transport at depth, followed by subvertical transport nearer
levels the stress was mainly due to regional compression, with a hori- the surface. Several geophysical data have been collected also at
zontal greatest stress. In any case, the vertical tension fractures did not Mt Redoubt in Alaska, during the 1989–90 and 2009 eruptions; Lahr
act as conduits in these experiments. et al. (1994) showed that the location of an earthquake swarm of the
As regards an example of volcanism and contractional tectonics at 1989–90 eruption is consistent with a narrow conduit steeply dipping
shallow level, at Iwate volcano (NE-Japan) the focal mechanism analysis to the NE (i.e. a NW–SE-striking dyke). The eruption of 2009 was imme-
of the M = 6.1, 1998 earthquake indicates reverse faulting with the hor- diately preceded by a period marked by volcano-tectonic events indi-
izontal principal compressive stress axis in an E–W direction (http:// cating a NE–SW-trending P-axis orientation corresponding to the
hakone.eri.u-tokyo.ac.jp/vrc/erup/iwate.html). N–S-striking reverse ac- horizontal displacement of the conduit walls in the same direction
tive faults are present in this area: The activity of one of these, the (Roman and Gardine, 2013), a finding that can be interpret as the possi-
Nishine-fault, was estimated to be about 0.7 m/1000 years by the ble effect of a NW–SE dyke. These NW–SE dykes are parallel to the direc-
Japan Active-Fault Research Group. GPS surveys showed a steady and tion of the horizontal, regional σ1 (Nakamura, 1977; Nakamura et al.,
continuous extension between southern and northern sites of the volca- 1977, 1980; Koehler et al., 2012). A plexus of dykes and sills has also
no in 1998 (Miura et al., 2000). Based on these results, it has been sug- been recognised beneath the northern flank of Mt Redoubt at deeper
gested that intrusion of an E–W-striking dyke occurred at around 10 km level (Benz et al., 1996). At the active Trident volcano (Alaska),
of depth below the summit of the volcano, thus resulting in a magma Wallmann et al. (1990), based on the presence of a set of NW–SE fis-
path parallel to the direction of σ1. At Tromen volcano (Chile, Galland sures crossing the margins of the 1912 Novarupta crater, local plate-
et al., 2007a), lying above the hanging wall of a coeval west-dipping motion vectors and regional stress orientation, suggest that a NW–SE

Fig. 30. Examples of experiments where volcanoes in contractional settings have been modelled. In A, intrusion of magma in the shallower crust along thrust planes. In B, C and D, photos in
plan view (to the left) and fault traces (to the right) of experiments of volcanic cones lying above substrate reverse faults, dipping to the left, with different positions with respect to the
cone: B. cone summit located on the footwall block, C. cone summit located above the surface trace of the substrate fault, and D. cone summit located on the hanging-wall block; E. section
view of the deformation zone inside the cone of case C.
A. Modified after Galland et al. (2007a). E. Modified after Tibaldi (2008).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 113

feeder dyke propagated from a reservoir beneath Trident volcano to the


Novarupta vent. Anyway, the contemporaneous formation of the Mt
Katmai caldera represents a possible magmatic connection between
the reservoir beneath Katmai and Trident; this connection has been
interpreted in terms of a propagating sill or the failure of septa between
NE-aligned batches of magma (Hildreth, 1987; Wallmann et al., 1990;
Lowenstern et al., 1991).
The experiments of Tibaldi (2008) were mostly aimed at simulating
deformation within a cone lying above a reverse fault, by varying the lo-
cation of the fault with respect to the cone (Fig. 30B–D). Fractures par-
allel to the contraction direction did not form, probably because the
reverse/thrust faults affecting the substrate are not arcuate in plan
view, as opposed to those simulated in the experiments by Galland
et al. (2007a). In Tibaldi's (2008) experiments, the propagation of the
rectilinear (in plan view) reverse faults from the substrate across the
volcanic cone results in the formation of a steep normal fault zone (F1
in Fig. 30C, D and E) and a reverse low angle fault (F2). This implies
the developing of a local extensional field with a horizontal σ3 parallel
to the general contraction direction in the summit part of the cone.
Summarising the above, all these data suggest that in contractional
tectonic settings, magma ascending can be partitioned between a
deeper transport by inclined sheets and a shallow transport by vertical
dykes. In the crust below a volcano, magma may migrate horizontally
along flat structures and upward along ramp structures. Not necessarily
this magma migration may lead to eruptions, but in any case it emerges
that pre-existing contractional discontinuities in the crust my act as pre-
ferred magma pathways upwards, even during active compressional
deformation. Inside a volcano, magma upwelling may occur along the
steeper-dipping fault zone F1 in case the reverse fault propagates
through the cone and splits upwards. This magma migration along the
steeper fault is facilitated by the very local extension and is corroborated
by field evidence of eruptive vents close to or along F1 zones, such as
at Trohunco caldera and Los Cardos–Centinela volcanic complex
(Argentina). In several other cases, dykes within the volcanic cone are
parallel to the regional σ1, and also sill emplacement seems to be a fre-
quent mechanism of intrusion.

6.2.3. Vertical σ2
As an example of magma paths in strike-slip tectonic settings
(vertical σ2), Pasquarè and Tibaldi (2003) studied by field and analogue
data several volcanoes of the Bicol Peninsula (Luzon, Philippines),
where transcurrent tectonics dominate; they concluded that at depth
magma most likely used the main NW-striking strike-slip regional
faults, whereas at shallower level, the dykes followed fractures sub-
parallel to σ1 and sub-perpendicular to σ3. A similar change in dyke ori-
entation with depth was also discovered in another transcurrent zone at
Galeras volcano (Colombia, Tibaldi and Romero, 2000). This volcano is
also characterised by the presence of a sector collapse towards the
east within which is the active crater zone (Fig. 31A). The orientation
of this collapse matches exactly the results obtained by means of
analogue modelling by Lagmay et al. (2000), who showed that sigmoi-
dal deformation zones form on the volcano flanks close to the underly-
ing fault, along the right side of a right-lateral strike-slip fault, or along
the left side of a left-lateral strike-slip fault. If we compare their ana-
logue model deformed above a right-lateral strike-slip fault, we observe
that the collapse of Galeras volcano occurred in the same position
and orientation (Fig. 31B–C). This deformation zone above a strike-slip
fault should be the most prone to develop failures, intrusions and Fig. 31. A. DEM of Galeras volcano (Colombia) with location of sector collapse and main
eruptions. right-lateral strike-slip faults affecting the cone; the sketch has been rotated for easier
comparison with the analogue model of B–C. B. Photo and C. interpretative sketch of ana-
Holohan et al. (2007), who analysed by analogue modelling the in- logue modelling of a volcano located above a right-lateral strike-slip fault. Note the
teractions between structures associated with regional tectonic strike- matching between the Galeras example and the experiment.
slip deformation and volcano-tectonic caldera subsidence, suggested a A. Modified after Tibaldi and Romero (2000). B–C. Modified after Lagmay et al. (2000).
similarity between the roof-dissecting Riedel shears and Y-shears
appearing in their models and the regional strike-slip faults that dissect inside pull-apart basins (Fig. 32A). According to Holohan et al. (2007),
the central floors of the Negra Muerta (Riller et al., 2001; Ramelow et al., these fault systems might be regarded as preferential pathways for
2006) and Hopong calderas, with indication of volcanoes developed the ascent of magma and other fluids before, during, or after caldera
114 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

that most dykes conformed to an en-échelon pattern and were affected


by a right-lateral shear component along the intrusion planes during
propagation.
Summing up the above, data suggest that at volcanoes lying in
transcurrent settings, magma ascending can be partitioned at plumbing
systems that have a different structure depending on depth. The deeper
magma transport below a volcano can be dominated by dyking along
the main, rectilinear segment of a strike-slip fault zone, or at associated
releasing bends. The shallow magma transport within a volcanic cone
can occur along dykes oriented parallel to σ1 (T structures) or oblique
to σ1 (Riedel shears).

6.3. Topographic influence

As introduced before when discussing the Stromboli example, this


volcano shows other two zones of dyking (Fig. 26B). Between 85 and
20 ka BP, a N–S-trending magma feeding zone developed in the SSW
side of the island, interpreted as the effect of a possibly southward ex-
tension of the magma chamber (Corazzato et al., 2008). After 13 ka
BP, sheets injected in the NW side of the volcano and along its summit
Fig. 32. In strike-slip fault zones, volcanism can occur at pull-apart basins (A); at releasing part, depicting a horse-shoe shape in plan view, open northwestward
bend structures (B); directly along rectilinear strike-slip faults (C); and at the tips of the (i.e. towards the sea) (Tibaldi, 2001) (Fig. 26B). These sheets follow
main strike-slip faults (horsetail structures (D). The most common dyke orientation inside
the escarpments created by a series of nested sector collapses
the volcanoes located within these diverse fault settings is represented by red lines. Stra-
tovolcanoes, shield volcanoes, pyroclastic cones and domes may occur at all the above (“circum-lateral collapse sheets”) that affected the northwest volcano
types of strike-slip fault structures, whereas calderas are preferentially located within side (Fig. 33), and the most recent failure produced the present shape
pull-apart basins. of the Sciara del Fuoco depression. The sheets' strike and dip tend to ro-
tate along the margins of the sector collapse zone and their geometry is
rendered graphically in Fig. 34A, showing dyke attitude compared with
the local geometry of the collapse scarp as the angle between the two
dip directions. This indicates that all these dykes, with only two excep-
formation. Busby and Bassett (2007) document that the intrabasinal tions, intruded along planes striking parallel to the collapse scarp. More-
lithofacies of the Santa Rita Glance Conglomerate record repeated intru- over, most dykes dip into the same direction as the collapse scarps, i.e.
sion and emission of small volumes of magma emplaced along a main towards the interior of the collapse depression. Along the uppermost
strike-slip fault, as frequently found (Bellier and Sebrier, 1994; Bellier part of the collapse zone, the local orientation and location of the col-
et al., 1999), and along releasing fault bends (sketch in Fig. 32B). lapse scarp coincides with the NE-trending volcano-tectonic rift zone
Marra (2001), studying the Mid-Pleistocene volcanic activity in the (red zone in Fig. 33A), and here the dykes strike NE–SW. It is also
Alban Hills (Central Italy), documented that a local, clockwise block ro- worth noting that in a section perpendicular to the sector collapse
tation between parallel N–S strike-slip faults might have generated local scarp, dyke frequency is maximum at the scarp and decreases outward,
crustal decompression, enabling volatile-free magma to rise from deep with dykes finally disappearing at 500 m of distance (Fig. 33B). All
reservoirs beneath the Alban Hills and feeding fissure lava flows. sheets have been sampled (Fig. 33A); the geochemical characteristics
Chiarabba et al. (2004a), on the basis of a shallow seismic tomography of the sheets cropping out along the southern part of the Sciara del
of Vulcano Island (Aeolian Arc, Italy) observe that at a depth N 0.5 km, Fuoco depression are represented in Fig. 34B, and those located along
the rise of magma is controlled by NW–SE fractures associated with the northern and eastern part of the Sciara del Fuoco in Fig. 34C
the activity of the NW–SE-striking, right-lateral strike-slip to oblique- (Corazzato et al., 2008). A total of 33 sheets (triangles) have the same
slip, Tindari–Letojanni fault system (Fig. 32C) (Mazzuoli et al., 1995). characteristics of the lava deposits of the Neostromboli volcano growth
At shallower depth (i.e. b0.5 km), the plumbing system of the volcano phase and thus can be regarded as the shallow feeding system of that
is mainly controlled by N–S-striking secondary faults, which are per- volcanic phase. These sheets were emplaced following the first huge
pendicular to the regional σ3 and thus are not transcurrent faults but sector collapse towards NW that occurred about 13 ka BP. This sector
rather extensional fractures. Also Aydin et al. (1990) observed that at collapse clearly reflects a major change in the shape of the volcano;
strike-slip fault zones, magma preferentially rises at the surface along huge morphological changes influence the orientation of the stress
the associated secondary extensional structures rather than along the due to the reorganisation of the gravity force following a new distribu-
main strike-slip fault segments (e.g. Fig. 32D). Scarrow et al. (1997) tion of rock masses (Voight and Elsworth, 1997; Tibaldi, 2001). Rock
found in Antarctica sets of shoshonitic dykes that were emplaced mass removal due to lateral failure produces: (1) debuttressing of the
along strike-slip faults in late Cretaceous times. Rossetti et al. (2000) il- cone flank; (2) new local directions of extension; (3) the formation of
lustrate how the effusive and intrusive rocks belonging to the McMurdo a fracture pattern around the depression amphitheatre; and (4) the for-
Volcanic Group (Antarctica) were fed by dykes along a crustal-scale, mation of lithological boundaries. These aspects have deserved an
non-coaxial transtensional shear zone where the strike-slip component increasing attention in recent times, in terms of new field data, numer-
increased over time. By analogue modelling, Corti et al. (2001) showed ical modelling and analogue experiments. Aspects 1) and 2) are related
that magma is emplaced at depth along faults parallel to the main shear to the influence of topography on dyke paths, whereas items 3) and
zone but uprises to the surface along cracks that are orthogonal to the 4) are more related to rock mechanics.
orientation of the extension. Finally, Eriksson et al. (2014) studying Regarding issues 1) and 2), results of finite difference numerical
the Álftafjörður volcano (Iceland) by AMS techniques found that dykes modelling of the state of stress within the Stromboli volcano before
mostly propagated horizontally from a shallow magma chamber within the first huge sector collapse towards NW and after the failure, show
a transtensive shear zone. Although in other cases in the same Iceland, important changes in terms of both orientation of stress tensor and
AMS results show different directions of magma flow in dykes (Kissel stress magnitude (Fig. 35) (Tibaldi et al., 2008b). In particular, a closer
et al., 2010; Eriksson et al., 2011), the example of Álftafjörður shows view of the area of the sector collapse shows the marked reorientation
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 115

Fig. 33. A. Aerial oblique view of the Sciara del Fuoco depression at Stromboli (Italy) generated by a series of lateral collapses, with all the intrusive sheets (yellow, with relative number of
sample), emplaced along the scarps. The main NE–SW volcano tectonic rift is represented by the red zone. B. Photo taken along the coast of Stromboli showing a section view of the lateral
collapse escarpment and intrusive sheets; the diagram quantifies the frequency of dykes moving from the collapse depression outward.

of the stress tensors close to the collapse scarps, with the minimum the western steep scarp of the Valle del Bove, at a distance of few tens
compressive stress reaching a trend normal to the collapse surface. of meteres up to 500 m from the rock cliff (McGuire and Pullen, 1989;
This local configuration favours the intrusion of sheets parallel to the McGuire et al., 1990, 1991). Also during the July–August 2001 magmatic
collapse scarps and dipping steeply towards the depression, consistent event, field data indicated a deviation of the strike of a shallow dyke that
also with field data summarised in Fig. 34A. Other examples of the influ- followed the local topography (Tibaldi and Groppelli, 2002). A configu-
ence of local cone slope on dyke geometry have been observed at ration similar to Stromboli has also been observed for the dykes intrud-
Mt. Etna, Italy, during historical dyke-fed eruptions (McGuire and ed along lateral collapse scarps at Tenerife, Canary Islands (Gottsmann
Pullen, 1989; McGuire et al., 1990, 1991; Tibaldi and Groppelli, 2002). et al., 2008; Delcamp et al., 2012), Tahiti–Nui (French Polynesia)
During the 1978–1979 and 1983 magmatic events, feeder dykes propa- (Hildenbrand et al., 2004), and Piton de la Fournaise (Reunion Island)
gated from the summit zone of Mt Etna towards the western wall of the (Peltier et al., 2009). At Fogo Island (Cape Verde archipelago) a recent
Valle del Bove. Geophysical data and structural data provided insights N–S swarm of dykes, emplaced below an older collapse structure,
into the changes in azimuth of feeder dykes at shallow depths, prior to show an en-échelon geometry associated with a seaward displacement
their intersection with the surface, as they deviated from an originally of the eastern flank of the volcano (Day et al., 1999). In the presence of a
southeasterly propagation direction, into a southerly one, parallel to high scarp, such as the one between the top of Vesuvius and Mt Somma,
116 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

the difference in topography can create a zone of buttressing that hin-


ders lateral dyke propagation across the scarp under ordinary excess
magmatic pressures (Acocella et al., 2006).
Also several analogue models indicate the differences in sheet con-
figuration based on the volcano morphology, with the results shown
in Fig. 36. In the case of an intact cone, dykes propagate parallel to the
σHmax and perpendicular to the σHmin, which are radial and concentric,
respectively (Fig. 36, upper part) (McGuire and Pullen, 1989). If a lateral
collapse depression is cut into the cone, later intrusions follow
the scarps of the collapse depression if the initial intrusive source is lo-
cated near the scarps, interpreted as the effect of unbuttressing and the
consequent reorganisation of the local stress field (Fig. 36, middle)
(Acocella and Tibaldi, 2005). Similar results have been obtained for a
volcano with an unstable flank that has not collapsed yet (Walter and
Troll, 2003). If the initial intrusion is located below the collapse depres-
sion, the resulting dyke will propagate along the collapse axis, i.e. along
the mean points of the collapse in map view (Acocella and Tibaldi,
2005).
Regarding the other effects of the rock mass removal due to a lateral
failure, which are 3) the formation of a fracture pattern around the de-
pression amphitheatre and 4) the formation of lithological boundaries,
these favour the capture of propagating sheets due to higher permeabil-
ity and stoping effect, respectively. For example, a field analysis con-
ducted along the margins of the sector collapse that affected Ollague
volcano (Chile–Bolivia border) showed the very high density of frac-
tures formed along the collapse scarps (Figs. 37A–B) with respect to
the same lava deposits observed far away from the collapse zone
(Fig. 37C). These fractures may result from brittle deformation during
the sliding of the collapse blocks, as well as from debuttressing effect
due to rock mass removal. Preferential dyke intrusions along fractures
parallel to collapse scarps have been observed also at Piton de la
Fournaise (Reunion Island) by Letourneur et al. (2008). Once the
amphitheatre depression created by a sector collapse is infilled by suc-
cessive volcano growth, a major lithological boundary may form at the
interface between the pre- and post-collapse deposits. In the case of a
strong difference in rock stiffness at this interface, an ascending dyke
can be diverted and become parallel to the lithological/mechanical
boundary.
In the case of elongated volcanoes, topography can also play a funda-
mental role. In very large volcanoes like at the Hawaiian islands, it has
been proposed that gravity forces can control the orientation of rectilin-
ear dyke swarms that tend to be parallel to the volcano major axis (Fiske
and Jackson, 1972), once space for dyking is provided by slip of the vol-
cano flank along deep faults (Dieterich, 1988). At some elongated volca-
noes, anyway, the dyke pattern is represented by a central rectilinear
volcanic rift zone passing outwards into fan-shaped dyke systems at
the two opposite volcano flanks (Fig. 36, bottom part), defined as “di-
verging volcano-tectonic rifts” (Tibaldi et al., 2014). In their pioneering
analogue experiments, Fiske and Jackson (1972) suggested that this ge-
ometry might be linked to the elongated shape of a volcano. Better-
scaled models and field data recently suggested that the formation of
these diverging rifts is not specifically linked to substrate lithology and
mechanical behaviour. Volcanoes with diverging rifts have typical elon-
gation (major/minor axis) b 0.88 and V N 10 km3 (mostly N300 km3),
Fig. 34. A. Geometry of intrusive sheets surveyed along the lateral collapse escarpments of
and the central rift zone is normal to the regional σHmin (Tibaldi et al.,
Stromboli volcano (Italy) around the Sciara del Fuoco depression, expressed as number of 2014). If the regional σHmin is oblique to the volcano elongation axis,
dykes (Y axis) versus the angle between the dip direction of dyke and the dip direction of dyke geometry in the edifice axial zone is controlled by elongation
the collapse scarp (X axis). B. Geochemical characteristics of the intrusive sheets cropping and thus by local gravity σ3, but dyke strike becomes perpendicular to
out along the southern part of the Sciara del Fuoco depression, and C. along the northern
σHmin when dykes intrude the more external areas of the volcano. If a
and eastern part of the Sciara del Fuoco depression. A total of 33 sheets (triangles) belong
to the Neostromboli volcano growth phase (label NEOSTR), 2 sheets belong to the previ- dyke is injected under volcano flanks with slope inclination b 50°, at
ous Vancori growth phase and crop out far away from the Sciara del Fuoco depression the edifice terminations magma paths diverge outwards and crosscut
(see Vancori label in Fig. 32A for the location of these two sheets) and the other 3 sheets slopes at high angle (Fig. 36). These authors point out the diverging vol-
were emplaced during post-Neostromboli growth phases. canic rift system is an underestimated structural pattern that can be
A. Modified after Tibaldi (2003). C. Modified after Corazzato et al. (2008).
more common than previously recognised. For example at Mt Etna the
diverging rift system is superimposed on other dyke patterns, like
the West rift, lying above hidden faults (Mattia et al., 2007) and the
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 117

Fig. 35. Results of finite difference numerical modelling by FLAC of the state of stress within the Stromboli volcano before the first huge sector collapse towards NW and after the failure,
computed along a NW-striking section. On the left, the sections show the complete pre- and post-collapse stress configurations (upper and lower sections respectively), in terms of both
vertical stress contours and principal stress tensors (negative values indicate compression). A strong reorganisation in the stress distribution occurs, in terms of both magnitude and ori-
entation. The right sections at a closer view show the marked reorientation of the stress tensors close to the collapse scarps, with the minimum compressive stress normal to the collapse
surface. This local configuration favours the intrusion of dykes parallel to the collapse scarps and dipping at high angles towards the depression, consistent also with field evidence.
Modified after Tibaldi et al. (2008b).

dykes surrounding the Valle del Bove depression, whose geometry is Mt Cameroon (Favalli et al., 2012), Miyakejima (Nakada et al., 2005)
interpreted to have been guided by the depression debuttressing and Fuji (Takada, 1997) in Japan, Kliuchevskoi in Russia (Ozerov et al.,
(McGuire and Pullen, 1989). More clear field examples of diverging 1997), Nyamuragira in Congo (Hayashi et al., 1992), and Newberry in
rifts can be seen at El Hierro (Canary Islands) (Gee et al., 2001), the USA (MacLeod et al., 1995).

Fig. 36. Sketch of the main configurations of intrusive sheets at volcanoes under the influence of topography and indication of the stress field guiding the geometry of magma paths. Upper
part of the figure: circular intact cone; middle part: circular cone with a sector collapse depression; lower part: elongated volcano.
118 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 37. Photos taken at the Ollague volcano (Chile–Bolivia border). A. Field view of the depression left by the Holocene sector collapse; B. photo taken close to the collapse scarp showing
the pervasive fracturing of the lava flows; C. photo of the same lava flows taken far away from the collapse depression.

The experiments of Tibaldi et al. (2014) also suggest that in the case coalescence of adjacent cones, by reproducing the spreading of an edi-
of steep slopes, the topography of a volcano exerts a dominant control fice composed of overlapping volcanoes. Their results suggest that
on magma path orientation, determining sheet rotation until they spreading edifices of similar age that partially overlap tend to develop
become parallel to the local edifice slope. The steep slope actually repre- a rift zone approximately perpendicular to the boundary of both volca-
sents a debuttressing that produces a local σ3 oriented perpendicularly noes, and this causes the two edifices to grow together and develop an
to the topography. Regarding the threshold value of slope steepness, elongated topographic ridge. However, in the case of partially overlap-
these experiments show that when sheets approach a slope dipping at ping volcanoes of different ages that spread at different rates, the
35°, they propagate perpendicularly to the slope. This has been observed resulting rift will be parallel to their boundary, causing the two edifices
also with slopes as steep as 45° by Walter and Troll (2003). Hence, the to structurally separate from each other.
transition to slope-parallel sheets may occur with a slope angle N 45– Regarding general topographic effects, Gaffney and Damjanac
50°, but more research is needed to better constrain this value. (2006) suggested that magma uprising along a fracture that runs from
Walter et al. (2006) investigated the geometry of shallow magma a highland to an adjacent lowland, can be diverted away from the high-
paths related to the deformation of large volcanic edifices built by the land towards the lowland. This is in line with the results of Kervyn et al.

Fig. 38. Models of deformation of the surface topography above a shallow propagating dyke tip: A. uplift zones on each side of the dyke tip and no uplift above the dyke tip with propa-
gation of normal faults and fissures at the surface. B. Uplift of the zone above the dyke tip with propagation of reverse faults and fissures at the surface (Gudmundsson et al., 2008). C. Photo
and D. sketch of the dramatic effects caused by the very shallow intrusion of a dyke in northern Iceland during the Krafla “fires” of 1975–1984.
A. Modified after Mastin and Pollard (1988).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 119

(2009) who investigated, by analogue modelling, the control of volcano to imprecise results on the geometry and depth of the dyke
load on magma ascent and vent location. Their results show that if a (Gudmundsson, 2003). Model 2) is valid only in the case of pre-
central conduit is blocked, the volcano loading stress favours eruption existing graben structures above a propagating dyke, and is more effec-
of rising magma away from the volcano summit. Folch et al. (2000) tive for layered host rocks. The effects of the intrusion are to open pre-
studied the numerical solutions of ground displacement caused by an existing normal faults and, later on, to close the lower sectors of the
overpressurised magma chamber, finding that correct calculation and faults and encourage reverse slip along them. If the magma front has
interpretation of displacements can be reached only taking into consid- propagated up into the graben, dyke emplacement uplifts the crustal
eration also topography. A recent paper by Maccaferri et al. (2014) de- block between the boundary faults, forming a horst (Gudmundsson
veloped numerical modelling of dyke propagation near the shoulders et al., 2008).
of rift valleys. They found that the location of the shallow magma path Sill emplacement also requires important deformation of the host
is governed by the competition between gravitational unloading pres- rock that depends on the stiffness of the rock succession and the thick-
sure and tectonic stretching. When gravitational unloading dominates, ness, velocity and structure of the intrusion. Frequently, sills are saucer-
ascending dykes are diverted towards the rift shoulders, whereas shaped and consist of an axis-symmetrical flat inner horizontal intru-
when tectonic stretching dominates, shallow dykes propagate within sion connected outward and upward to transgressive inclined sheets,
rift. This means that rifts with mild topographic expression may concen- ending in a flat outer sill (Planke et al., 2005; Polteau et al., 2008;
trate dyke paths within the graben valleys. Galland et al., 2009; Galerne et al., 2011; Gudmundsson and Løtveit,
2014). Some authors proposed that saucer-shaped sills form along the
7. Plumbing systems and surface deformation level of magma neutral buoyancy and the feeder dyke is located at one
side of the saucer structure (Fig. 39A) (Bradley, 1965; Francis, 1982;
Although it has been widely recognised that magma intrusion may
create regional uplift (McKenzie, 1984; Saunders et al., 2007; Ernst,
2014, and references therein), only recently the relations between shal-
low magma emplacement and local surface deformation field have be-
come clearer. Observations during episodes of dyke emplacement at
subaerial volcanic rift zones have shown the relation between dyking
and slip increment along existing faults and fissures and the formation
of new fissures (Abdallah et al., 1979; Bjornsson et al., 1979; Pollard
et al., 1983; Wright et al., 2006). More in detail, direct field observations
(Larsen et al., 1979) and continuously recording measurements
(Hauksson, 1983), confirmed by more recent studies (Belachew et al.,
2011, and references therein), showed that fault and fissure motions
at the surface are correlated in time with the arrival of the earthquake
swarm front, several hours after the initiation of dyke propagation.
This suggested that faulting and fissuring were triggered by dyke intru-
sion (Rubin and Pollard, 1988). Above the propagating dyke tip, stress
concentration induces compression in the surrounding host rock and
consequent faulting up to, or beyond, the dyke plane at depth; this in-
duces a typical deformation pattern at the surface given by: (1) uplift
of a region several kilometres across with a maximum value at the
edges of a central graben located above the dyke tip; (2) an abrupt tran-
sition from graben flank to graben floor; and (3) subsidence of a
concave-upward graben floor by one to several times the amount of
flank uplift (Rubin and Pollard, 1988). At the surface, near the dyke,
this process is expressed by two possible models: (1) uplift zones on
each side of the dyke tip and no uplift above the dyke tip (Fig. 38A)
(Pollard et al., 1983; Mastin and Pollard, 1988; Rubin and Pollard,
1988) and (2) uplift of the zone above the dyke tip (Fig. 38B)
(Gudmundsson et al., 2008; Abdelmalak et al., 2012). The effects can
be dramatic, causing heavy deformation as shown in a field example
(Figs. 38C–D) from northern Iceland, regarding one of the dyking and
rifting events that took place during the Krafla “fires” of 1975–1984. In
model 1) the stress is not transferred directly above the dyke tip but it
is concentrated at its two sides, creating two areas with vertical fissures
and normal faults dipping towards the dyke. Two maxima of surface
ground breaking may occur in the case of a vertical dyke, one on each
side of the dyke plane, and are separated by a distance roughly equal
to twice the depth of the dyke top (Mastin and Pollard, 1988). An in-
clined sheet would produce asymmetrical uplift of the two side zones. Fig. 39. Existing models of saucer-shaped sill and laccolith emplacement mechanisms:
A. sill emplacement controlled at the level of neutral buoyancy (LNB): Sills are fed laterally
Subsidence above the dyke tip is not explained by the dyke alone, but
from one part of the outer sill; B. sill emplacement along horizontal discontinuity: Sills are
it requires the development of a graben bounded by normal faults that fed radially from the central inner sill. C. Gilbert's (1877) model for laccolith intrusion with
extent down to the dyke tip level or even deeper. Once the propagating rigid strata and faulted periphery. D. Laccolith with continuous flexure of overburden stra-
crack in front of a dyke tip cuts the surface, the relative downward dis- ta. E. Model of symmetric laccolith-induced deformation about intrusive plane. F. Mixed
placement over the dyke changes to upward displacement (Pollard flexure-fault deformation at a laccolith periphery, inclined sheet climbing may initiate at
forced fold hinges or at a fault.
et al., 1983). For a layered crust with weak contacts, the stress magni- A. Modified after Bradley (1965) and Francis, 1982. B. Modified after Malthe-Sørenssen
tude and configuration above a propagating dyke may be more compli- et al. (2004). D. Modified after Koch et al. (1981). E. Modified after Koch et al. (1981).
cate and the straightforward inversion of geodetic surface data may lead F. Modified after Thomson (2007).
120 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Chevallier and Woodford, 1999; Goulty, 2005), whereas others claim 20 km intruded at a depth of 14–18 km and produced about 6 cm of up-
that the feeder dyke is located below the central part of the saucer struc- lift of an elliptical area in plan view with an axis of more than 20 km
ture (Fig. 39B) (Pollard and Johnson, 1973; Hansen et al., 2004; Galerne (Fig. 40C) (Wicks et al., 2006). The modelling of interferometry and
et al., 2011 and references therein). In some large sill complexes it has GPS data suggests the presence of a sill gently dipping towards NNW
been observed that individual saucer-shaped sills are interconnected that explains the asymmetric shape in plan view (Fig. 40D). Anyway,
to other overlying or underlying sills. These connections may represent also in this case prudence should be used since, based on field examples
a model of plumbing system where magma does not migrate along ver- elsewhere, the large dimension of the sill in the horizontal plane does
tical feeder dykes but moves along horizontal layers until it breaches the not fit with the very small uplift.
rock succession along inclined sheets, and then rotates again to reach
horizontal attitude (Muirhead et al., 2012): This pattern resembles the 8. Plumbing systems at calderas
“ramp and flat” structure of compressional tectonics. These magma
feeding structures have been named “interconnected sill-feeding-sill The plumbing systems at calderas are here reviewed separately due
networks” (Cartwright and Hansen, 2006). The first models of host to the complex structure of calderas and their marked morphological,
rock deformation were proposed for the gradual growth of a single sill structural and dynamic differences in comparison to intact volcanic
into a laccolith. In case of rigid strata below the intrusion plane, two cones. Several studies have focused on calderas all over the world, and
main models suggest asymmetric deformation of the host rock where a good number of review papers have been recently published dealing
the intrusion is compensated by uplift of the overburden; in one case with: the structure of the caldera (Cole et al., 2005; Acocella, 2007),
peripheral continuous faults accommodate the intrusion (Fig. 39C) the tectonic settings (Hughes and Mahood, 2011), the petrographic
(Gilbert, 1877), and in the other case the sill/laccolith emplacement is characteristics of the caldera magma reservoir (Cashman and
accommodated by peripheral bending of strata (Fig. 39D) (Koch et al., Giordano, 2014), the characteristics of the ring faults (Geyer and
1981). Another model shows a laccolith-induced deformation that is Martí, 2014), and caldera subsidence in the specific extensional tectonic
symmetric about the intrusive plane (modified after Koch et al., 1981). setting (Carlino et al., 2014). Here I focus on the geometry of intrusions
More recent studies consider a possible mixed mode of deformation at calderas.
by flexural folding and faulting (Fig. 39E); the viscous dissipation of Calderas can be essentially associated with basaltic, peralkaline, an-
magma along the length of sills may enable them to propagate faster desitic–dacitic and rhyolitic magmas. Basaltic calderas are linked to
and induce non-elastic deformations in the surrounding rocks. This cre- shield volcanoes such as those at hotspots (e.g.: Hawaii and Galápagos
ates the conditions for faulting of the rock succession at the peripheral islands) (Simkin and Howard, 1970; Decker, 1987), or in other settings
zones of the sill and, eventually, inclined sheet climbing may initiate such as convergent margins (e.g. Masaya Volcano in Nicaragua,
at forced fold hinges or at a fault (Thomson, 2007). Since, in many in- Williams and Stoiber, 1983). Peralkaline calderas are typically associat-
stances, field data indicate that laccoliths lack internal discontinuities, ed with zones of rifting (Cole et al., 2005) like at Pantelleria (Mahood,
suggesting that magma was injected as a single pulse or a series of and Hildreth, 1986) and Mayor Island, New Zealand (Cole, 1990). An-
quickly coalescing pulses (Roni et al., 2014 and references therein), a desitic–dacitic calderas are mostly related to subduction zones like in
large amount of recent studies focused on the growth of a single sill the case of Krakatau, Indonesia (Self and Rampino, 1981), and Santorini,
both by numerical computation (e.g.: Zenzri and Keer, 2001; Zhao Greece (Druitt et al., 1999). Rhyolitic calderas are characterised by huge
et al., 2008; Michaut, 2011 and references therein) and analogue exper- volumes of erupted pyroclastic products and usually large collapse de-
iments (e.g.: Dixon and Simpson, 1987; Roman-Berdiel et al., 1995). An pressions. Examples comprise the Valles caldera, New Mexico, USA
interesting case is represented by the laccoliths of Elba Island (Italy) that (Smith and Bailey, 1968), the Cerro Galan, Argentina (Francis et al.,
also grew without internal discontinuities, suggesting continuous feed- 1978), and the Campi Flegrei, Italy (Barberi et al., 1991).
ing of the magma (Westerman et al., 2015). These laccoliths are one A caldera forms mainly after a sufficient amount of magma has been
above the others but emplaced as separate sheets without coalescing. withdrawn during an eruption, and the consequent underpressure in
They failed to form a composite pluton and instead gave rise to a the magma chamber causes its roof to collapse (Lipman, 1997). Another
“Christmas tree geometry”, with each magma batch that created its possibility is that overpressure within the magma chamber induces first
own room by lifting the rock overburden (Westerman et al., 2015). doming and fracturing of the overburden, followed by magma migra-
Other recent field studies indicate that, instead of growing in thick- tion and caldera collapse (Gudmundsson et al., 1997). Although the
ness or ramping upward, sills can give rise to stacked piles of horizontal first model has been applied to several cases, it has also been criticised
to sub-horizontal sheeted intrusions (Boudier et al., 1996; Menand, in view of the fact that the underpressure in the magma chamber should
2008; Tibaldi and Pasquarè, 2008; Annen, 2009). The first sill frequently hinder the continuation of the eruption until the critical value to trigger
intrudes at a discontinuity given by the contact between two rock layers the collapse has been reached; another factor is the inconsistency, at
with different stiffness, followed by a second sill that is emplaced at the some calderas, between the volume of erupted material and the col-
contact with the previous one. The multiple intrusion process may end, lapse volume, and the contradiction between the dip direction of the
as for example was the case at Maiden Creek (Horsman et al., 2005), theoretical caldera faults in comparison to field data (Gudmundsson
where a small intrusion was formed. However, if enough magma is and Nilsen, 2006).
fed from below, the process may go on through the emplacement of From the point of view of the possible structure of a caldera, this may
several stacked sills generating a sheeted laccolith (Menand, 2008; result from different evolutionary processes that involve single or mul-
Tibaldi and Pasquarè, 2008; Tibaldi et al., 2008c). In the case of a single tiple collapses, each with possible different geometries, and also possi-
sill, some InSAR-based modelling indicates that the emplacement is ac- ble phases of resurgence. A categorisation was attempted by Lipman
companied at the surface by uplift of circular to elliptical areas. For ex- (1997) who suggested five end-member structural styles: Plate or pis-
ample, modelling of the 1994 unrest episode at Eyjafjallajokull ton, piecemeal, trapdoor, downsag and funnel (Fig. 41). Plate/piston col-
volcano (Iceland) suggests that a sill intrusion induced deformation lapse involves the subsidence of a coherent block of rock along a ring
at the surface in the line of sight direction of the Radar satellite fault (Fig. 41A), like at Creede caldera (Steven and Lipman, 1976). The
(Pedersen and Sigmundsson, 2004). The interferometric fringes high- piecemeal case refers to a caldera where the collapsing block is
light a circular deformation pattern (Fig. 40A) that is consistent with subdivided into a series of discrete secondary blocks (Fig. 41B) such as
the modelling of an about 0.36-m-thick sill source, intruded at a depth the Scafell caldera, U.K. (Branney and Kokelaar, 1994). The trapdoor
of 4.5–6.0 km (Fig. 40B); however, the authors acknowledge that fur- style corresponds to an asymmetric collapse with block subsidence on
ther investigations are necessary to better constrain the data. At Yellow- one side where a ring fault develops, and tilting on the hinged other
stone caldera (USA), since 1996 an expanding sill with a length of about side (Fig. 41 C). Examples of the above are the Valles caldera, U.S.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 121

Fig. 40. A. Interferometric fringes of the deformation field from 06/09/1992 to 15/08/1995 that accompanied the 1994 unrest episode at Eyjafjallajokull volcano (Iceland) in the line of sight
direction of the Radar satellite, and B. artificial fringes obtained by modelling a sill source, about 0.36-m-thick intruded at a depth of 4.5–6.0 km. C. Interferometric fringes of the deforma-
tion field at Yellowstone caldera (USA) from the summer of 1996 to the summer of 2000 characterised by an expanding sill, and D. result of modelling of the source as a sill gently dipping
towards NNW.
B. Modified after Pedersen and Sigmundsson (2004). D. Modified after Wicks et al. (2006).

(Heiken et al., 1990), and Kumano caldera, Japan (Miura, 1999). the caldera floor dips gently towards the collapse centre. Bolsena
Downsagging takes place where ring faults do not form or do not (Italy) is an example of a downsag caldera (Walker, 1984), as well as
reach the surface (Fig. 41D), and the overburden deforms by bending the Gross Brukkaros system in Namibia (Stachel et al., 1994). Funnel cal-
without fracturing. In this case there are no distinct caldera walls and deras occur through the complete destruction of the caldera floor that is

Fig. 41. Different styles of caldera collapse. (A) Plate or piston, (B) piecemeal, (C) trapdoor, (D) downsag, (E) funnel with complete destruction, and (F) funnel with concentric ring faults.
See text for details.
Redrawn after Lipman (1997).
122 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Fig. 42. Relations between the size of magma chamber and of caldera. (A) In the case of caldera forming by overpressure, an expanding chamber with a sill-like shape (lower high/width
ratio) may induce steeply-dipping ring faults at the chamber edges that will guide the caldera collapse. The caldera and magma chamber size should correspond. (B) A laccolith-like shape
(higher H/W) can favour the nucleation of shear stress in correspondence of specific points of maximum curvature, inducing the development of ring faults. (C) An overpressurised magma
chamber with a laccolith-like shape may also produce arching of the overburden with development of local extension at the extrados and caldera occurrence. In cases (B) and (C) the
caldera width is much lower than the magma chamber size.
Redrawn after Aizawa et al. (2006).

Fig. 43. (A) Crater Lake caldera and (B) related model of magma propagation from the growing chamber through a series of steeply-dipping dykes diverging upward from the sides of the
magma chamber. (C) Toba caldera (northern Sumatra) and (D) related model of stacked sills linked by vertical dykes, superimposed on the distribution of the horizontally polarised shear
wave speed (shown above 20 km in depth), dashed line = the low-velocity area below the caldera that might have been affected by the super-eruption of 74 ka ago.
A. and B. After Karlstrom et al. (2015). C. and D. After Jaxybulatov et al. (2014).
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 123

broken into series of blocks that chaotically subside deeper towards the U.K. (Moore and Kokelaar, 1998; Troll et al., 2002), and at younger cal-
collapse centre (Fig. 41E), or by blocks that are displaced downward deras such as Rallier-du-Baty in Kerguelen Archipelago (Bonin et al.,
along concentric ring faults that become deeper towards the zone of 2004). At modern calderas, data are mostly of the geophysical and sub-
deepest collapse (Fig. 41F). An example of the latter type is the Guayabo ordinately geological type, like at the Campi Flegrei (Piochi et al., 2014),
Caldera, Costa Rica (Halinan, 1993). Miyakejima (Geshi et al., 2002; Geshi, 2009), Piton de la Fournaise
Although several publications have been dedicated to the study of (Michon et al., 2007), Rabaul (Mori and McKee, 1987), Sierra Negra
the structure and dynamics of calderas, a lot of work is still required to (Jónsson, 2009), and Nyamulagira (Wauthier et al., 2013). Based on sev-
clarify the geometry of their plumbing system. The necessity of further eral field and geophysical data, the classical model of caldera plumbing
investigations on the structure of sub-caldera magmatic systems is system comprises a magma chamber and a system of transient conduits
even more important if we take into consideration the possible connec- that develop during caldera forming. This “Standard Model” as termed
tion between location, dimension and shape of a magma chamber and by Gualda and Ghiorso (2013), is represented by a single, long-lived,
the related caldera; it has been shown that in the case of caldera forming usually horizontally-elongated magma chamber, characterised by dom-
by underpressure of the magma chamber, the width of the caldera de- inating melt. This model has been recently applied, for example, to Mt
pression is similar to the diameter of the underlying chamber (Aizawa Mazama (Oregon Cascades, U.S.), a long-lived (400 ka) volcanic centre
et al., 2006 and references therein). In the case of caldera forming by whose activity culminated with the 50 km3 climactic eruption of
overpressure, an expanding chamber with a sill-like shape, i.e. with a 7.7 ka BP that produced the Crater Lake caldera (Fig. 43A) (Bacon and
lower high/width (H/W) ratio, may induce the development of Lanphere, 2006; Wright et al., 2012). Based on field data and numerical
steeply-dipping ring faults above the chamber edge that will guide the work by Karlstrom et al. (2015), a centralised oblate spheroidal magma
caldera collapse (Fig. 42A). As a consequence, also in this case the calde- chamber was responsible for the climactic eruption. The chamber
ra size and magma chamber size should correspond. In the case, instead, growth, fed by a deep magma influx, culminated in the caldera forma-
of an overpressurised magma chamber with a laccolith-like shape tion. Based on the distribution in space and time of the eruptive vents
(higher H/W), the shape of the roof of the chamber may favour the nu- preceding the Crater Lake caldera failure, Karlstrom et al. (2015) propose
cleation of shear stress in correspondence of specific points of maxi- that magma propagated from the growing chamber through a series of
mum curvature, inducing the development of ring faults (Fig. 42B) steeply-dipping dykes diverging upward from the sides of the magma
(Mandl, 1988). A laccolithic shape of the magma chamber may also re- chamber (Fig. 43B). This model, more related to silicic magmas, suggests
sult in the arching of the overburden with the development of local ex- the focusing of dykes all around the perimetry of a growing magma
tension at the extrados; the combination of this extension with the chamber, providing a plausible mechanism for clustering of eruptions
upward-directed magma push may lead to the development of a caldera around the Mazama centre. The eruptions at the central main volcano
(Fig. 42C) (Aizawa et al., 2006). In both cases, the caldera width is much are fed by a vertical dyke that drains the chamber and releases accumu-
smaller than the magma chamber size. lated overpressure; once this central event has taken place, eruptions
Field data are scarce in comparison to other plumbing systems and could return to proximal areas as the effective chamber volume and re-
are mostly collected at eroded ancient calderas, like for example Scafell lated deviatoric stresses are reduced. This model differs from the geo-
(Branney and Kokelaar, 1994; Kokelaar et al., 2007) and Glencoe in the metrical distribution of inclined sheets that propagate above a flat,

Fig. 44. A. Model proposed for the plumbing system of resurgent calderas where several inclined sheets propagated upward from the magma chamber, and most became arrested at layer
contacts with contrasting mechanical properties. A number of inclined sheets intersected the ring faults that have been previously developed during the caldera collapse, and became
deflected up along the fault to form multiple ring dykes. (B) Model of ring dykes directly linked with the underlying magma chamber.
A. After Browning and Gudmundsson (2015). B. After Saunders (2005).
124 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

mafic magma chamber, as introduced in the previous sections (compare


with Figs. 10 and 17).
In recent years, there has been a growing evidence that sub-caldera
plumbing systems may be made of a series of superimposed sills inter-
connected by inclined sheets and vertical dykes. In particular, the exis-
tence of a complex magma storage region is postulated, where the
magma reservoir is composed of multiple melt lenses within a crystal
mush, or by multiple sills stacked into a completely crystalline rigid
framework (Cashman and Giordano, 2014). For example, in the deeply
eroded calderas of the southern Rocky Mountains, apart from the main
plutons, relatively minor intrusions near the calderas are characterised
by megadykes, up to some hundreds of metres wide, and tabular sill-
laccolith bodies (Lipman, 2007). In general, the presence of several
stacked tabular interconnected magma bodies can reconcile several ob-
servations of crystals, carried by the transporting melt, which have been
stored at a range of pressures and temperatures (Cashman and
Giordano, 2014). These crystals have been erupted when the tabular
bodies at different depths have been tapped. A modern example
comes from the felsic Toba caldera (northern Sumatra), which has
long been regarded as a supervolcano (Fig. 43C). Using ambient-noise
seismic tomography below the caldera, Jaxybulatov et al. (2014) ob-
serve an anisotropy that may reflect a fine-scale layering linked to the
presence of many partially molten sills in the crust below a depth of
7 km (Fig. 43D).
Another model has been proposed for the plumbing system of resur-
gent calderas (Browning and Gudmundsson, 2015). It has been derived
from field data, integrated with numerical modelling, at the eroded
Hafnarfjall volcano (western Iceland), an extinct stratovolcano with a
caldera of predominantly basaltic and subordinately andesitic composi-
tion. Here several inclined sheets propagated upward from the magma
chamber, and most became arrested at layer contacts with contrasting
mechanical properties (Fig. 44A). A number of inclined sheets
intersected the ring faults that have previously developed during calde-
ra collapse, and became deflected up along the fault to form multiple
ring dykes. This model differs from previous models of ring dykes,
which were directly linked with the underlying magma chamber (e.g.
Saunders, 2005) (Fig. 44B). Saunders (2005) also proposes that arching
of the caldera floor induces extension in the extrados, whereas in the in- Fig. 45. Sketch in section view of possible intrusive histories found in the western Penin-
trados compression is caused by both arching and the force exerted lat- sular Ranges Batholith (Mexico and U.S.) by Johnson et al. (2002). In the first stage (A), a
erally by the intruding ring dykes, preventing intrusions in the caldera set of conical fractures (black lines) above a rising magma chamber develops, with the
block. Several eroded calderas show that substantial amounts of possible emplacement of cone sheets (red lines). In the successive stage (B), the collapse
of a caldera along ring-faults is followed by intrusion of ring dykes and volcanism. In
magma are intruded along ring faults that channel the magma flow up-
(C) a nested intrusion develops with possible further cone sheets.
ward (Clough et al., 1909; Bailey and Maufe, 1960; McCall and Bristow,
1965; Smith and Bailey, 1968; Almond, 1977; Sparks, 1988; Miura,
1999; Johnson et al., 2002; Sewell et al., 2012). In some case, ring
dykes have been channelling magma for a long time, so as to favour
the mingling of felsic and basic magmas, as observed at the Ossipee
ring complex (New Hampshire, U.S.) (Kennedy and Stix, 2007). Not- complexes emplaced at depths of up to 18 km, suggesting that caldera
withstanding the frequent identification of this intrusion type, recent subsidence can extend to mid-crustal levels or that different processes
studies have attempted to suggest alternative geometries at some can produce ring dykes.
ring dyke complexes, although these cases remain controversial (e.g. As we have seen, several models have been introduced, reflecting
O'Driscoll et al., 2006; Stevenson et al., 2008). the possible variability of the structure of plumbing systems at calderas.
In a study of several intrusive bodies and calderas in the western This variability has been linked to the diverse size and geometry of
Peninsular Ranges Batholith (Mexico and U.S.), Johnson et al. (2002) in- magma chambers, and also to the possible structural style of the caldera
dividuated the widespread presence of ring dykes. They suggested block. Their interaction gives rise to different configurations of the
that development of subvolcanic ring complexes in the Peninsular plumbing system in the overburden that may be composed of inclined
Ranges Batholith involved a three-stage sequence: (1) a buoyant or sheets, stacked sills and dykes. A further complication may be due to
overpressured magma chamber induces fracturing of the overlying the interaction with regional tectonics. As an example, the Los Azufres
rocks, with the formation of inward-dipping conical fractures and caldera complex, in the central part of the Mexican Volcanic Belt, is a
cone sheets (Fig. 45A); (2) the successive loss of magma from the cham- sub-circular depression (27 × 26 km) that resulted from several nested
ber causes collapse of the roof by near-vertical ring faults and ring dyke collapses of latest Miocene–Pliocene age (Ferrari et al., 1991, 1993). The
intrusions (Fig. 45B); and (3) resurgence takes place due to chamber in- caldera depression is filled by a fluvio-lacustrine sequence and several
flation and/or intrusion of a nested pluton (Fig. 45C). A similar sequence volcanic domes of Pliocene–Pleistocene age, as well as cinder cones
was found also at the deeply eroded, felsic complex of the Vallehermoso and associated lava flows of Pleistocene age (Fig. 46A). Domes have a
Caldera (La Gomera, Canary Islands) by Rodríguez-Losada and dacitic to rhyolitic composition whereas cinder cones and lavas are ba-
Martinez-Frias (2004). Johnson et al. (2002) also found some ring saltic. Some of the younger deposits are offset by normal faults mainly
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 125

Fig. 46. A. Distribution of silicic volcanic domes, basaltic cinder cones and normal faults at the Los Azufres caldera complex (central Mexican Volcanic Belt), suggesting the control of re-
gional tectonics across the caldera depression. (B) The different cases of distribution of volcanoes with respect to a caldera give clues to the possible patterns of shallow magma plumbing
systems. See main text for details.
A. Redrawn after Ferrari et al. (1991). B. Taken from Walker (1984) and Geyer and Martí (2008).

striking E–W to WSW–ENE, whereas the oldest deposits are also affect- with centres located both along the caldera rim and along regional faults
ed by a series of slip planes of different orientations that should repre- crossing the caldera depression (Fig. 46A). This interaction favoured the
sent segments of caldera ring faults. Some of the E–W to WSW–ENE presence of silicic centres in the middle of the caldera depression and
faults show a very subordinate left-lateral strike-slip component, a ge- along the ring faults, whereas the new basaltic magmas within the cal-
ometry and kinematics consistent with the regional tectonics (Ferrari dera were guided by regional faults. As a consequence, the plumbing
et al., 1991). The distribution of vents and other data indicates that the system in this example should be represented by silicic dykes emplaced
Los Azufres caldera complex was affected by post-caldera volcanism along ring faults and regional faults, with the possibility that these dykes
126 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

are vertically connected with the felsic magma chamber, or might rep- propagation can be modelled as mode I cracks, although field observa-
resent captured inclined sheets. tions also indicate that, in a much lower number of cases, magma over-
Regional tectonics can also have an influence on the development of pressure may induce shear faults along which magma intrusion can take
ring dykes. As shown by numerical simulations of Walter (2008), an in- place. These findings open questions about the kinematics, geometry
truding ring-dyke may follow pre-existing faults and thus abruptly stop and propagation modes of intrusive fractures; we have to admit that un-
or change its direction. This is the case of the Erongo complex (Namibia), certainty still exists about the factors that control the different modes of
where a ring dyke formed only to the northwest of a SW–NE-striking re- growth of small magma chambers vs. large igneous bodies. Further
gional fault passing through the caldera (Wigand et al., 2004). studies, which may couple field data and modelling, are required, espe-
The Los Azufres example is also consistent with other cases suggest- cially integrated with a more in depth analysis of the influence of the
ing that calderas are long-lasting volcanic complexes whose activity in rates of magma transport and accumulation in relation to diverse tec-
most cases does not end after caldera collapse: in fact, resurgence can tonic settings and fracture propagation at different depths. Moreover,
take place and/or volcanism can resume, guided by the development diapiric ascent is still recognised especially for granitic bodies, and
of a plumbing system that can be different from the one that led to thus it is debatable how the deep distributed source region of melt-
the caldera-forming event. Geyer and Martí (2008), also integrating filled pores is linked to the shallow sheeted plumbing system.
data from Walker (1984), reviewed the distribution of post-caldera The presence of a shallow magma chamber is a prime candidate to
vents at more than 160 Quaternary calderas in different geodynamic control the geometry of the sheets that typically display a circumferen-
settings, providing clues to the possible patterns of magma uprising. tial pattern – in plan view – above the chamber, giving rise to centrally-
They proposed nine categories of distribution of post-caldera volcanic inclined sheet swarms. Their arrangement is dictated by the presence of
activity (surface distribution of vents in Fig. 46B). Assuming that the an oblique stress tensor (Fig. 47, upper apex of the triangle). The local
shallow geometry of the plumbing system was mirrored by the different oblique orientation of the σ1 is a function of the shape and overpressure
distributions of volcanic centres, I suggest different magma paths in the of the chamber. The consequent σ1 trajectories condition the arrange-
various cases (magma feeding systems in Fig. 46B): a single large or ment of the intrusions that can range from radially-inclined sheets
small volcano can be located in different positions in the caldera, corre- (in vertical section view) diverging from a spherical chamber, to
sponding to single main conduits (cases 1–4). Vents along the ring faults centrally-inclined sheets with similar dip angle in the case of lobate
may be linked to ring dykes (case 5), as took place for example at the magma chambers. Above the central part of the chamber there is the
20 × 30 km wide Tondano caldera (Indonesia) in 1952 and 1971 greatest chance to have subvertical to vertical conduits that drain
where ring-dyke intrusions and eruptive activity occurred at two or magma towards the axis of the volcano. This chance is increased by
more ring vents (Lecuyer et al., 1997). A circular or elliptical distribution the presence of a horizontal σ3 below the volcano axis (Fig. 47, right
of vents outside the caldera suggests the existence of inclined cone apex of the triangle) that can be induced by regional extensional tecton-
sheets (case 6). A defined straight line of vents across the caldera sug- ics or local deformation, for example due to volcano spreading (Borgia,
gests a rectilinear dyke zone (case 7) or the control of regional faults 1994; van Wyk de Vries and Francis, 1997). With increasing distance
(case 8), as at the above-mentioned Los Azufres caldera, and at the from the magma chamber, the effects of magma overpressure decrease
Rabaul caldera (Papua New Guinea) where eruptive activity simulta- and the intrusive sheets can change their attitude from inclined sheets
neously occurred at opposite sides in 1878 and 1937 (Mori and to vertical dykes laterally with respect to the chamber. These dykes
McKee, 1987; Nairn et al., 1995). Finally, distributed vents inside the cal- can be parallel to each other, hence forming a swarm perpendicular to
dera may suggest conduits and dykes of different geometries linked to a the regional σ3, or can assume a radial pattern if the influence of region-
highly-fractured caldera block (case 9), as for example at the resurgent al tectonics is poor. Above the magma chamber, inclined sheets and
Ischia Island caldera (Italy) (Tibaldi and Vezzoli, 1998). dykes can bend into a sill-like attitude following a re-orientation of
the stress tensor (Fig. 47, left apex of the triangle). This can be caused
9. Conclusions by several situations that involve the presence of stress barriers and
major mechanical contrasts between different lithotypes (Menand,
The data presented in this work indicate that several parameters can 2011). Centrally-inclined sheet swarms are more widespread at mafic
influence the structure of a magma plumbing system. The classical magma chambers, but have been also found at felsic bodies. Magma mi-
models of plumbing systems characterised by upward propagation of gration along sills and inclined sheets may occur at different scales,
magma by buoyancy forces have been questioned by more recent find- going from tens of metres to tens of kilometres, with the potential of
ings that indicate magma overpressure as the main engine. The stress producing eruptions in unexpected locations. Analysis of the literature
field induced by local magma pressure sources, in turn, interacts with shows a great amount of works that use numerical or analogue model-
the rheological boundaries represented by rock layers with large differ- ling approaches to unravel these processes, but there is a comparative
ences in stiffness, and with physical discontinuities such as bedding, lack of field data. For example, only seventeen field-based studies of
joints and faults. This interaction may occur at any depth, as suggested centrally-inclined sheet swarms are sufficiently detailed to obtain quan-
for example by magma sources in the substratum that are in an offset titative data. Anchoring the results of modelling to quantitative ground
position from the volcano. This is reflected also in the available most truth is fundamental to establish the thresholds of zones of influence of
complete images of plumbing systems, as for example at Hawaii and tectonic stresses versus other stress sources, and this might definitively
Mt Etna, which show complex arrays of vertical dykes, inclined sheets benefit from investigations at eroded volcanic centres and exhumed
and sills. Local isotropic stresses or time-frequent rotations of stress ten- plumbing systems. Besides, large field data sets should be integrated
sors might lead to intricate assemblages of intrusions of various geome- in a more interdisciplinary way with geophysical and petro-chemical
tries, as for example below Mt Redoubt volcano. data.
Most field and geophysical recent research indicates not only that At shallower crustal levels, there may be a strong interference be-
magma moves along planar sheets and that circular/elliptical conduits tween tectonic stresses, the stresses related to volcano topography,
(in section view) are extremely rare, but there is increasing evidence and the stresses exerted by magma overpressure. The tectonic stress
that also laccoliths and magma chambers are composed by stacked field may affect both plumbing systems within volcanoes and below
planar sill intrusions. Similarly, vertical sheeted intrusions may charac- them; if extensional stresses dominate, a rectilinear volcano-tectonic
terise larger plutonic bodies. Sheets propagate by self-induced rift zone can guide magma upwelling below the volcano as well as
hydrofracturing or by using pre-existent weakness planes, if these are across the cone with the injection of a swarm of parallel dykes (lower
suitably oriented with respect to the stress field. Most field evidences left part of the graph of Fig. 48). In the case of a transcurrent regime,
indicate that hydrofractures are primarily extension fractures whose dykes can follow the main strike-slip fault zone and/or secondary
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 127

Fig. 47. Graph that summarises some of the main parameters that influence the upper part of a magma plumbing system. The apexes of the triangle are located in correspondence of stress
tensors with different orientations: the upper apex has oblique σ1, σ2 and σ3; under this stress field centrally-inclined sheets are favoured. The local oblique orientation of the σ1 is a func-
tion of the shape and overpressure of the chamber. The consequent σ1 trajectories condition the arrangement of the intrusions that can range from centrally-inclined sheets diverging from
a spherical chamber to centrally-inclined sheets with similar dip angle in the case of lobate magma chambers. With increasing distance form the magma chamber, or under an increasing
effect of σ3, dykes dominate as in the right apex of the triangle where σ2 and σ3 are horizontal. Above the magma chamber, inclined sheets and dykes can bend to reach a sill attitude
following a re-orientation of the stress tensor that locally has horizontal σ1 and σ2, as in the left apex of the triangle.

fractures. In the case of contractional tectonics, sheets may assume Fig. 48), as can be inferred from the Mt Etna and Stromboli data. For
more complex and variable geometries. Within a volcanic edifice, if slopes steeper than 40–50°, sheets may assume an attitude parallel to
the stresses induced by magma overpressure along the main central the slope.
conduit zone are sufficiently high, a radial dyke pattern might develop. Based on the available literature, it appears that much more studies
The influence of the stresses related to topography increases ap- are required to elucidate the characteristics of magma feeding systems
proaching the volcano slopes with more probable effects at distances in contractional tectonic settings, especially with an interdisciplinary
b0.5 km (increasing influence towards the right side of the graph of approach. Similarly, the effects of slopes on the trajectories of magma
paths need further studies based on field data integrated with numeri-
cal and analogue modelling. Since eroded calderas are not very frequent
on Earth, the scarcity of field data on the structure of their plumbing sys-
tems is even more noticeable.
In conclusion, the data illustrated in the present work indicate that
several parameters can influence the structure and location of a
magma plumbing system, and that it is extremely difficult, if not impos-
sible, to assign a given model to a volcano without a detailed complete
knowledge of its geological, structural, petro-geochemical and geomor-
phological characteristics. The ultimate magma path is governed by the
inner volcano structure, and by the stress field in the host rock existing
prior to sheet emplacement and by changes in the stress field induced
by the emplacement itself. As the assessment of the possible magma
paths leading to a new eruptive centre is one of the main goals of volca-
nic hazard studies, it is very important to better understand how far a
stress field can be perturbed by the emplacement of previous sheets,
and why and when sheet propagation might be deviated or halted.
These questions should be better addressed by an integrated approach
that may combine natural examples with experimental works, and by
increasing the data sets on the distribution of mechanic properties in-
side volcanoes and of the active stress field in the substrate and within
Fig. 48. Scheme of the most frequent sheet geometries as resulting from the interference a volcanic edifice. This knowledge can be achieved especially by desir-
between different tectonic stress fields (stress tensors at Y axis) and the distribution of able further geological and geophysical exploration and in-situ stress
the stresses related to topography (increasing influence towards the right side of the
graph). The debuttressing effect of topography is larger at b0.5 km from the volcano
measurements and observations that, coupled with numerical model-
slope and for slopes steeper than 40–50°; in these cases sheets might reach an attitude ling and instrument monitoring of a volcano, can lead to a better under-
parallel to the slope. standing of these processes.
128 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Acknowledgements Bacon, C.R., Lanphere, M.A., 2006. Eruptive history and geochronology of Mount Mazama
and the Crater Lake region, Oregon. Geol. Soc. Am. Bull. 118 (11), 1331.
Bagnardi, M., Amelung, F., 2012. Space-geodetic evidence for multiple magma reservoirs
I thank the Editors Lionel Wilson and Joan Martì and the Managing and subvolcanic lateral intrusions at Fernandina volcano, Galápagos Islands.
Editor Timothy Horscroft for requesting me to write this review, after J. Geophys. Res. 117, B10406. http://dx.doi.org/10.1029/2012JB009465.
Bahat, D., 1980. Hertzian fracture, a principal mechanism in the emplacement of the
an invited talk I gave at the 2014 EGU Meeting in Vienna. My research British Tertiary intrusive centres. Geol. Mag. 117, 463–470.
contribution to this review benefitted from extensive field work with Bailey, E.B., Maufe, H.B., 1960. The geology of Ben Nevis and Glen Coe and the surrounding
several colleagues, to mention Giorgio Pasquaré, Alberto Renzulli, country. Mem Geol Surv Scotlandp. 307.
Bailey, E.B., Clough, C.T., Wright, W.B., Richey, J.E., Wilson, G.V., 1924. The Tertiary and
Derek Rust, Federico Pasquaré Mariotto, Claudia Corazzato, Fabio Bonali,
post-tertiary geology of Mull, Loch Aline, and Oban. Geological Survey of Scotland
and many others. The research was carried out in the framework Memoir, Sheet. 44 ((Scotland): 445 pp.).
of MIUR, FIRB, INGV and NATO projects, and under the aegis of the Baker, S., Amelung, F., 2012. Top-down inflation and deflation at the summit of Kilauea
volcano, Hawaii observed with InSAR. J. Geophys. Res. 117, B10406. http://dx.doi.
International Lithosphere Program, Task Force II. F. Jim Cole and Agust
org/10.1029/2012JB009123.
Gudmundsson are acknowledged for their useful comments on an ear- Bald, N., Noe-Nygaard, A., Pedersen, K., 1971. The Kroksfjordur central volcano in North-
lier version of this paper. Pasquaré Mariotto is also acknowledged for his West Iceland. Acta Naturalia Islandica II (10) (29 pages).
review of the English grammar. Barberi, F., Varet, J., 1970. The Erta Ale volcanic range (Danakil depression, northern Afar,
Ethiopia). Bull. Volcanol. 34, 848–917.
Barberi, F., Cassano, E., La Torre, P., Sbrana, A., 1991. Structural evolution of Campi Flegrei
caldera in light of volcanological and geophysical data. J. Volcanol. Geotherm. Res. 48,
References 33–49.
Barker, S.E., Malone, S.D., 1991. Magmatic system geometry at Mount St. Helens modeled
Abdallah, A., Courtillot, V., Kasser, M., LeDain, A.Y., Lepine, J.C., Robineau, B., Ruegg, J.C., from the stress field associated with post-eruptive earthquakes. J. Geophys. Res. 96
Tapponnier, P., Tarantola, A., 1979. Relevance of Afar seismicity and volcanism to (B7), 11,883–11,894. http://dx.doi.org/10.1029/91JB00430.
the mechanics of accreting plate boundaries. Nature 282, 17–23. Bates, R., Jackson, J.A., 1987. Glossary of Geology. 3rd edition. McGraw-Hill Book Company
Abdelmalak, M.M., Mourgues, R., Galland, O., Bureau, D., 2012. Fracture mode analysis and (788 pp.).
related surface deformation during dyke intrusion: results from 2D experimental Bear, G., Beyth, M., Reches, Z., 1994. Dikes emplaced into fractured basement, Timna
modelling. Earth Planet. Sci. Lett. 359–360, 93–105. Igneous Complex, Israel. J. Geophys. Res. 99, 24039–24050.
Acocella, V., 2007. Understanding caldera structure and development: an over- Becerril, L., Galindo, I., Gudmundsson, A., Morales, J.M., 2013. Depth of origin of magma in
view of analogue models compared to natural calderas. Earth Sci. Rev. 85 (3), eruptions. Sci. Rep. 3.
125–160. Belachew, M., Ebinger, C., Coté, D., Keir, D., Rowland, J.V., Hammond, J.O.S., Ayele, A., 2011.
Acocella, V., 2014. Structural control on magmatism along divergent and convergent plate Comparison of dike intrusions in an incipient seafloor-spreading segment in
boundaries: overview, model, problems. Earth Sci. Rev. 136, 226–288. Afar, Ethiopia: seismicity perspectives. J. Geophys. Res. Solid Earth (1978–2012)
Acocella, V., Neri, M., 2009. Dike propagation in volcanic edifices: overview and possible 116 (B6).
developments. Understanding Deformation and Stress in Active Volcanoes. Bell, B.R., Claydon, R.V., Rogers, G., 1994. The petrology and geochemistry of cone-
Tectonophysics, Special issue 471, pp. 67–77. sheets from the Cuillins Igneous Complex, Isle of Skye: evidence for combined
Acocella, V., Tibaldi, A., 2005. Dike propagation driven by volcano collapse: a general assimilation and fractional crystallization during lithospheric extension. J. Petrol.
model tested at Stromboli, Italy. Geophys. Res. Lett. 32 (8). http://dx.doi.org/10. 35, 1055–1094.
1029/2004GL022248. Bellier, O., Sebrier, M., 1994. Relationship between tectonism and volcanism along the
Acocella, V., Porreca, M., Neri, M., Massimi, E., Mattei, M., 2006. Propagation of dikes at Great Sumatran Fault zone deduced by SPOT image analyses. Tectonophysics 233,
Vesuvio (Italy) and the effect of Mt. Somma. Geophys. Res. Lett. 33 (8). 215–231.
Aizawa, K., Acocella, V., Yoshida, T., 2006. How the development of magma chambers af- Bellier, O., Bellon, H., Sebrier, M., Sutanto, M.R.C., 1999. K–Ar age of the Ranau tuffs;
fects collapse calderas: insights from an overview. Geol. Soc. Lond., Spec. Publ. 269 implications for the Ranau Caldera emplacement and slip-partitioning in Sumatra
(1), 65–81. (Indonesia). Tectonophysics 312, 347–359.
Aizawa, K., Koyama, T., Hase, H., Uyeshima, M., Kanda, W., Utsugi, M., Ryokei, Y., Yusuke, Benz, H.M., Chouet, B.A., Dawson, P.B., Lahr, J.C., Page, R.A., J.A.H, 1996. Three-dimensional
Y., Takeshi, H., Kenichi, Y., Shintaro, K., Watanabe, A., Koji, M., Yasuo, O., 2014. Three- P and S wave velocity structure of Redoubt Volcano, Alaska. J. Geophys. Res. 101 (B4),
dimensional resistivity structure and magma plumbing system of the Kirishima 8111–8128.
Volcanoes as inferred from broadband magnetotelluric data. J. Geophys. Res. Solid Best, M.G., 1982. Igneous and Metamorphic Petrology. W.H. Freeman and Company, New
Earth 119, 198–215. York (630 pp.).
Almond, D.C., 1977. The Sabaloka igneous complex, Sudan. Philos. Trans. R. Soc. Lond. 287, Billings, M.P., 1972. Structural Geology. Prentice-Hall Inc., Englewood Cliffs, New Jersey,
595–633. USA (606 pp.).
Aloisi, M., Bonaccorso, A., Gambino, S., 2006. Imaging composite dyke propagation Bistacchi, A., Tibaldi, A., Pasquarè, F.A., Rust, D., 2012. The association of cone-sheets and
(Etna, 2002 case). J. Geophys. Res. Solid Earth 111. http://dx.doi.org/10.1029/ radial dykes: data from the Isle of Skye (UK), numerical modelling, and implications
2005JB003908. for shallow magma chambers. Earth Planet. Sci. Lett. 339–340, 46–56.
Amelung, F., Day, S., 2002. InSAR observations of the 1995 Fogo, Cape Verde, eruption: Bjornsson, A., Johnsen, G., Sigurdsson, S., Thorbergsson, G., Tryggvason, E., 1979. Rifting of
implications for the effects of collapse events upon island volcanoes. Geophys. Res. the plate boundary in north Iceland 1975–1978. J. Geophys. Res. 84, 3029–3038.
Lett. 29 (12), 1606. http://dx.doi.org/10.1029/2001GL013760. Bonafede, M., Olivieri, M., 1995. Displacement and gravity–anomaly produced by a shal-
Anderson, E.M., 1936. Dynamics of formation of cone-sheets, ring-dikes, and cauldron low vertical dyke in a cohesionless medium. Geophys. J. Int. 123, 639–652.
subsidence. Proc. Roy. Soc. Edinb. 56, 128–157. Bonafede, M., Rivalta, E., 1999. On tensile cracks close to and across the interface between
Anderson, E.M., 1938. The dynamics of sheet intrusion. Proc. Roy. Soc. Edinb. 58, 242–251. two welded elastic half-space. Geophys. J. Int. 138, 410–434.
Anderson, E.M., 1951. The Dynamics of Faulting and Dyke Formation With Applications to Bonali, F., Corazzato, C., Tibaldi, A., 2011. Identifying rift zones on volcanoes: an example
Britain. Oliver and Boyd, White Plains, New York. from La Réunion Island, Indian Ocean. Bull. Volcanol. 73, 347–366. http://dx.doi.org/
Annen, C., 2009. From plutons to magma chambers: thermal constraints on the accumu- 10.1007/s00445-010-0416-1.
lation of eruptible silicic magma in the upper crust. Earth Planet. Sci. Lett. 284 (3), Bonin, B., Ethien, R., Gerbe, M.C., Cottin, J.Y., Féraud, G., Gagnevin, D., Moine, B., 2004. The
409–416. Neogene to Recent Rallier-du-Baty nested ring complex, Kerguelen Archipelago
Annen, C., Sparks, R.S.J., 2002. Effects of repetitive emplacement of basaltic intrusions on (TAAF, Indian Ocean): stratigraphy revisited, implications for cauldron subsidence
thermal evolution and melt generation in the crust. Earth Planet. Sci. Lett. 203, mechanisms. Geol. Soc. Lond., Spec. Publ. 234 (1), 125–149.
937–955. http://dx.doi.org/10.1016/S0012-821X(02)00929-9. Borgia, A., 1994. Dynamic basis of volcanic spreading. J. Geophys. Res. Solid Earth
Annen, C., Zellmer, G.F. (Eds.), 2008. Dynamics of Crustal Magma Transfer, Storage and (1978–2012) 99 (B9), 17791–17804.
Differentiation. Geological Society, London, Special Publications 304, pp. 1–13. Boudier, F., Nicolas, A., Ildefonse, B., 1996. Magma chambers in the Oman ophiolite: fed
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006a. The genesis of intermediate and silicic from the top and the bottom. Earth Planet. Sci. Lett. 144 (1), 239–250.
magmas in deep crustal hot zones. J. Petrol. 47 (3), 505–539. Bradley, J., 1965. Intrusion of major dolerite sills. Trans. R. Soc. N. Z. 3, 2755.
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006b. The sources of granitic melt in deep hot zones. Branney, M.J., Kokelaar, P., 1994. Volcanotectonic faulting, soft-state deformation,
Trans. R. Soc. Edinb. 97 (4), 297–309. and rheomorphism of tuffs during development of a piecemeal caldera, English
Argnani, A., Serpelloni, E., Bonazzi, C., 2007. Pattern of deformation around the central Lake District. Geol. Soc. Am. Bull. 106, 507–530.
Aeolian Islands: evidence from multichannel seismics and GPS data. Terra Nova 19, Bridgwater, D., Sutton, J., Watterson, J., 1974. Crustal downfolding associated with igneous
317–323. http://dx.doi.org/10.1111/j.1365-3121.2007.00753.x. activity. Tectonophysics 21, 57–77.
Atherton, M.P., 1993. Granite magmatism. J. Geol. Soc. 150, 1009–1023. Browning, J., Gudmundsson, A., 2015. Caldera faults capture and deflect inclined sheets:
Atkinson, B.K., Meredith, P.G., 1987. The theory of subcritical crack growth with application an alternative mechanism of ring dike formation. Bull. Volcanol. 77 (1), 1–13.
to minerals and rocks. In: Atkinson, B.K. (Ed.), Fracture Mechanics of Rock. Academic, Busby, C.J., Bassett, K.N., 2007. Volcanic facies architecture of an intra-arc strike-slip basin,
New York, pp. 111–166. Santa Rita Mountains, Southern Arizona. Bull. Volcanol. 70, 85–103.
Aydin, A., Schultz, R.A., Campagna, D., 1990. Fault-normal dilatation in pull-apart basins: Cabello, G.C., Garza, R.M.L.D., Argote, R.C., Flores, E., Ramírez, A.O., Rivera, H., Böhnel, J.
implications for relationship between strike-slip fault and volcanic activity. In: Leed, 2006. Geology and paleomagnetism of El Potrero pluton, Baja California:
Boccaletti, M., Nur, A. (Eds.), Active and Recent Strike-slip Tectonics. Annales understanding criteria for timing of deformation and evidence of pluton tilt during
Tectonicae Special Issue, pp. 45–52. batholith growth. Tectonophysics 424, 1–17.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 129

Calanchi, N., Romagnoli, C., Rossi, P.L., 1995. Morphostructural and some petrochemical Day, S.J., Heleno da Silva, S.I.N., Fonseca, J.F.B.D., 1999. A past giant lateral collapse and
data from the submerged area around Alicudi and Filicudi volcanic islands (Aeolian present-day flank instability of Fogo, Cape Verde Islands. J. Volcanol. Geotherm. Res.
Arc, southern Tyrrhenian Sea). Mar. Geol. 123, 215–238. 94, 191–218.
Canon-Tapia, E., Merle, O., 2006. Dyke nucleation and early growth from pressurized Decker, R.W., 1987. Dynamics of Hawaiian volcanoes: an overview. Volcanism in Hawaii.
magma chambers: insights from analogue models. J. Volcanol. Geotherm. Res. 158 US Geol. Surv. Prof. Pap. 1350, 997–1018.
(3), 207–220. del Potro, R., Díez, M., Blundy, J., Camacho, A.G., Gottsmann, J., 2013. Diapiric ascent of
Carlino, S., Tramelli, A., Somma, R., 2014. Caldera subsidence in extensional tectonics. Bull. silicic magma beneath the Bolivian Altiplano. Geophys. Res. Lett. 40 (10), 2044–2048.
Volcanol. 76, 870. http://dx.doi.org/10.l007/s00445-014-0870-2. Delaney, P.T., Gartner, A.E., 1997. Physical processes of shallow mafic dike emplacement
Carracedo, J.C., 1994. The Canary-Islands—an example of structural control on the growth near the San Rafael Swell, Utah. Geol. Soc. Am. Bull. 109, 1117–1192.
of large oceanic-island volcanoes. J. Volcanol. Geotherm. Res. 60, 225–241. Delaney, P.T., Pollard, D.D., Ziony, J.I., Mckee, E.H., 1986. Field relations between dikes and
Cartwright, J., Hansen, D.M., 2006. Magma transport through the crust via interconnected joints: emplacement processes and paleostress analyses. J. Geophys. Res. 91, 4920–4983.
sill complexes. Geology 34 (11), 929–932. Delcamp, A., Troll, V.R., van Wyk de Vries, B., Carracedo, J.C., Petronis, M.S., Pérez-Torrado,
Cas, R.A.F., Wright, J.V., 1987. Volcanic Successions. Allen & Unwin, London (528 pp.). F.J., Deegan, F.M., 2012. Dykes and structures of the NE rift of Tenerife, Canary Islands:
Cashman, K.V., Giordano, G., 2014. Calderas and magma reservoirs. J. Volcanol. Geotherm. a record of stabilisation and destabilisation of ocean island rift zones. Bull. Volcanol.
Res. 288, 28–45. 74, 963–980. http://dx.doi.org/10.1007/s00445-012-0577-1.
Castro, J.M., Dingwell, D.B., 2009. Rapid ascent of rhyolitic magma at Chaiten volcano, Di Roberto, A., Rosi, M., Bertagnini, A., Marani, M.P., Gamberi, F., Del Principe, A., 2008.
Chile. Nature 461, 780–784. http://dx.doi.org/10.1038/nature08458. Deep water gravity core from the Marsili Basin (Tyrrhenian Sea) records
Cervelli, P., Miklius, A., 2003. The shallow magmatic system of Kilauea Volcano. U. S. Geol. Pleistocenic–Holocenic explosive events and instability of the Aeolian Archipelago,
Surv. Prof. Pap. 1676, 149–163. (Italy). J. Volcanol. Geotherm. Res. 177, 133–144. http://dx.doi.org/10.1016/j.
Chadwick Jr., W.W., Dieterich, J.H., 1995. Mechanical modelling of circumferential and jvolgeores.2008.01.009.
radial dike intrusion on Galapagos volcanoes. J. Volcanol. Geotherm. Res. 66, 37–52. Dieterich, J.H., 1988. Growth and persistence of Hawaiian volcanic rift zones. J. Geophys.
Chadwick, W.W., Jónsson, S., Geist, D.J., Poland, M., Johnson, D.J., Batt, S., Harpp, K.S., Ruíz, Res. Solid Earth 93, 4258–4270.
A., 2011. The May 2005 eruption of Fernandina volcano, Galápagos: the first circum- Dirks, P.H.G.M., Jelsma, H.A., 1998. Horizontal accretion and stabilization of the Archean
ferential dike intrusion observed by GPS and InSAR. Bull. Volcanol. 73, 679–697. Zimbabwe Craton. Geology 26 (1), 11–14.
http://dx.doi.org/10.1007/s00445-010-0433-0. Dixon, J.M., Simpson, D.G., 1987. Centrifuge modelling of laccolith intrusion. J. Struct. Geol.
Charlez, P.A., 1997. Rock Mechanics, 2: Petroleum Applications. Editions Technip, Paris. 9 (1), 87–103.
Chaussard, E., Amelung, F., 2014. Regional controls on magma ascent and storage in vol- D'Lemos, R.S., Brown, M., Strachan, R.A., 1992. Granite magma generation, ascent and em-
canic arcs. Geochem. Geophys. Geosyst. 15, 1407–1418. http://dx.doi.org/10.1002/ placement within a transpressional orogen. J. Geol. Soc. Lond. 149, 487–490.
2013GC005216. Donoghue, S.L., Gamble, J.A., Palmer, A.S., Stewart, R.B., 1995. Magma mingling in an an-
Chen, Z., Jin, Z.H., Johnson, S.E., 2007. A perturbation solution for dyke propagation in an desite pyroclastic flow of the Pourahu member, Ruapehu volcano, New Zealand.
elastic medium with graded density. Geophys. J. Int. 169 (1), 348–356. J. Volcanol. Geotherm. Res. 68, 177–191.
Chevallier, L., Woodford, A., 1999. Morpho-tectonics and mechanism of emplacement of Druitt, T.H., Edwards, L., Mellors, R.M., Pyle, D.M., Sparks, R.S.J., Lanphere, M., Davies, M.,
the dolerite rings and sills of the western Karoo, South Africa. S. Afr. J. Geol. 102 Barriero, B., 1999. Santorini volcano. Geol. Soc. Lond. Mem. 19.
(1), 43–54. Dufek, J., Bergantz, G.W., 2005. Lower crustal magma genesis and preservation: a stochastic
Chiarabba, C., Pino, N.A., Ventura, G., Vilardo, G., 2004a. Structural features of the shallow framework for the evaluation of basalt–crust interaction. J. Petrol. 46 (11), 2167–2195.
plumbing system of Vulcano Island Italy. Bull. Volcanol. 66, 477–484. Durrance, E.M., 1967. Photoelastic stress studies and their application to a mechanical
Chiarabba, C., De Gori1, P., Patanè, D., 2004b. The Mt. Etna plumbing system: the contri- analysis of the tertiary ring-complex of Ardnamurchan, Argyllshire. Proc. Geol.
bution of seismic tomography. Mt. Etna: Volcano Laboratory, Geophysical Mono- Assoc. 78, 289–318.
graph Series 143. AGU, pp. 191–204. Dvorak, J.J., Dzurisin, D., 1993. Variations in magma-supply rate at Kilauea volcano,
Chouet, B., Saccorotti, G., Martini, M., Dawson, P., De Luca, G., Milana, G., Scarpa, R., 1997. Hawaii. J. Geophys. Res. 98, 22255–22268.
Source and path effects in the wave fields of tremor and explosions at Stromboli Ebinger, C.J., Casey, M., 2001. Continental breakup in magmatic provinces: an Ethiopian
Volcano, Italy. J. Geophys. Res. Solid Earth (1978–2012) 102 (B7), 15129–15150. example. Geology 29, 527–530.
Chouet, B., Dawson, P., Martini, M., 2008. Shallow-conduit dynamics at Stromboli Emeleus, C.H., 2009. Ardnamurchan central complex, bedrock and superficial deposits. In:
Volcano, Italy, imaged from waveform inversions. Geol. Soc. Lond., Spec. Publ. 307 1:25,000 Geology Series. British Geological Survey. Scale 1:25 000, 1 sheet.
(1), 57–84. Eriksson, P.I., Riishuus, M.S., Sigmundsson, F., Elming, S.A., 2011. Magma flow directions
Clague, D.A., 1987. Hawaiian xenolith populations, magma supply rates, and development inferred from field evidence and magnetic fabric studies of the Streitishvarf compos-
of magma chambers. Bull. Volcanol. 49, 577–587. ite dike in east Iceland. J. Volcanol. Geotherm. Res. 206, 30–45.
Clague, D.A., Dixon, J.E., 2000. Extrinsic controls on the evolution of Hawaiian ocean island Eriksson, P.I., Riishuus, M.S., Elming, S.Å., 2014. Magma flow and palaeo-stress deduced
volcanoes. Geochem. Geophys. Geosyst. 1 (4), 1010. http://dx.doi.org/10.1029/ from magnetic fabric analysis of the Álftafjörður dyke swarm: implications for shal-
1999GC000023. low crustal magma transport in Icelandic volcanic systems. Geol. Soc. Lond., Spec.
Clemens, J.D., Mawer, C.K., 1992. Granitic magma transport by fracture propagation. Publ. 396, SP396–6.
Tectonophysics 204, 339–360. Ernst, R.E., 2014. Large Igneous Provinces. Cambridge University Press.
Clemens, J.D., Petford, N., Mawer, C.K., 1997. Ascent mechanisms of granitic magmas: Farris, D.W., Haeussler, P., Friedman, R., Paterson, S.R., Saltus, R.W., Ayuso, R., 2006.
causes and consequences. In: Holness, M.B. (Ed.), Deformation Enhanced Fluid Trans- Emplacement of the Kodiak batholith and slab-window migration. Geol. Soc. Am.
port in the Earth's Crust and Mantle. Chapman & Hall, London, pp. 145–172. Bull. 118, 1360–1376.
Clough, C.T.H., Maufe, H.B., Bailey, E.B., 1909. The cauldron subsidence of Glencoe and the Favalli, M., Tarquini, S., Papale, P., Fornaciai, A., Boschi, E., 2012. Lava flow hazard and risk
associated igneous phenomena. Q. J. Geol. Soc. Lond. 65, 611–678. at Mt. Cameroon volcano. Bull. Volcanol. 74, 423–439. http://dx.doi.org/10.1007/
Cole, J.W., 1990. Structural control and origin of volcanism in the Taupo volcanic zone, s00445-011-0540-6.
New Zealand. Bull. Volcanol. 52, 445–459. Ferrari, L., Garduno, V.H., Pasquarè, G., Tibaldi, A., 1991. Geological evolution of Los
Cole, J.W., Milner, D.M., Spinks, K.D., 2005. Calderas and caldera structures: a review. Azufres Caldera as a response to the regional tectonics. J. Volcanol. Geotherm. Res.
Earth Sci. Rev. 69 (1), 1–26. 47, 129–148.
Corazzato, C., Menna, M., Tibaldi, A., Renzulli, A., Francalanci, L., Petrone, C.M., Vezzoli, L., Ferrari, L., Garduno, V.H., Pasquarè, G., Tibaldi, A., 1993. The Los Azufres Caldera, Mexico:
Acocella, V., 2006. An integrated structural and petrochemical approach to unravel the result of multiple nested collapses. J. Volcanol. Geotherm. Res. 56, 345–349.
dike injection conditions and volcano flank instability at Stromboli (Italy). In: Ferré, E.C., Gleizes, G., Djouadi, M.T., Bouchez, J.L., Ugodulunwa, F.X.O., 1997. Drainage and
Thomson, K. (Ed.), Physical Geology of Subvolcanic Systems: Laccoliths, Sills, and emplacement of magmas along an inclined transcurrent shear zone: petrophysical
Dykes. Visual Geosciences, pp. 11–14 http://dx.doi.org/10.1007/s10069-006-0002-z. evidence from a granite–charnockite pluton (Rahama, Nigeria). In: Bouchez, J.L.,
Corazzato, C., Francalanci, L., Menna, M., Petrone, C., Renzulli, A., Tibaldi, A., Vezzoli, L., Hutton, D.H.W., Stephens, W.E. (Eds.), Granite: From Segregation of Melt to Emplace-
2008. What does it guide sheet intrusion in volcanoes? Petrological and structural ment Fabrics vol 8. Kluwer, Dordrecht, pp. 253–273.
characters of the Stromboli sheet complex, Italy. J. Volcanol. Geotherm. Res. 173, Ferré, E.C., Gleizes, G., Caby, R., 2002. Tectonics and post-collisional granite emplacement
26–54. in an obliquely convergent orogen: the Trans-Saharan belt, Eastern Nigeria. Precambrian
Corry, C.E., 1988. Laccoliths—mechanics of emplacement and growth. Geol. Soc. Am. Spec. Res. 114, 199–219.
Pap. 220 (110 pp.). Ferré, E.C., Galland, O., Montanari, D., Kalakay, T.J., 2012. Granite magma migration and
Corti, G., Bonini, M., Innocenti, F., Manetti, P., Mulugeta, G., 2001. Centrifuge models sim- emplacement along thrusts. Int. J. Earth Sci. 101 (7), 1673–1688.
ulating magma emplacement during oblique rifting. J. Geodyn. 31, 557–576. Fialko, Y.A., Rubin, A.M., 1998. Thermodynamics of lateral dike propagation, Implications
Cruden, A.R., 1998. On the emplacement of tabular granites. J. Geol. Soc. 155, 853–862. for crustal accretion at slow spreading mid-ocean ridges. J. Geophys. Res. 103 (B2),
Curtie, K.L., Ferguson, J., 1970. The mechanism of intrusion of lamprophyre dykes indicat- 2501–2514.
ed by “offsetting” of dykes. Tectonophysics 9 (525–535), 1970. Fiske, R.S., Jackson, E.D., 1972. Orientation and growth of Hawaiian volcanic rifts: the ef-
Dahm, T., 2000. Numerical simulations of the propagation path and the arrest of fluid- fect of regional structure and gravitational stresses. Proc. R. Soc. Lond. 329, 299–326.
filled fractures in the Earth. Geophys. J. Int. 141, 623–638. Fiske, R.S., Kinoshita, W.T., 1969. Inflation of Kilauea Volcano prior to its 1967–1968 erup-
Daniels, K.A., Kavanagh, J.L., Menand, T., Stephen, J.S.R., 2012. The shapes of dikes: tion. Science 165, 341–349.
evidence for the influence of cooling and inelastic deformation. Geol. Soc. Am. Bull. Folch, A., Martì, J., 1998. The generation of overpressure in felsic magma chambers by re-
124 (7–8), 1102–1112. plenishment. Earth Planet. Sci. Lett. 163, 301–314. http://dx.doi.org/10.1016/S0012-
Davis, G.H., Reynolds, S.J., Kluth, C., 1996. Structural Geology of Rocks and Regions. Second 821X(98)00196-4.
edition. Wiley, New York (776 pp.). Folch, A., Fernández, J., Rundle, J.B., Martí, J., 2000. Ground deformation in a viscoelastic
Dawson, P., Whilldin, D., Chouet, B., 2004. Application of near real-time radial semblance to medium composed of a layer overlying a half-space: a comparison between point
locate the shallow magmatic conduit at Kilauea Volcano, Hawaii. Geophys. Res. Lett. 31. and extended sources. Geophys. J. Int. 140 (1), 37–50.
130 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Francis, E.H., 1982. Magma and sediment — I. Emplacement mechanism of late Carbonif- Greene, A.R., Garcia, M.O., Pietruszka, A.J., Weis, D., Marske, J.P., Vollinger, M.J., Eiler, J.,
erous tholeiite sills in northern Britain. J. Geol. Soc. Lond. 139 (1), 1–20. 2013. Temporal geochemical variations in lavas from Kīlauea's Pu'u 'Ō'ō eruption
Francis, P.W., Hammill, M., Kretzschmar, G., Thorpe, R.S., 1978. The Cerro Galan Caldera, (1983‐2010): cyclic variations from melting of source heterogeneities. Geochem.
north-west Argentina and its tectonic setting. Nature 274, 749–751. Geophys. Geosyst. 14 (11), 4849–4873.
Franke, W., 1992. Phanerozoic structures and events in Central Europe. In: Bundell, D., Gretener, P.E., 1969. On the mechanics of the intrusion of sills. Can. J. Earth Sci. 6,
Freeman, R., Mueller, S. (Eds.), A Continent Revealed. Cambridge University Press, 1415–1419.
The European Geotraverse, pp. 164–180. Grosfils, E.B., 2007. Magma reservoir failure on the terrestrial planets: assessing the im-
Frey, F.A., Wise, W.S., Garcia, M.O., West, H., Kwon, S.T., Kennedy, A., 1990. Evolution of portance of gravitational loading in simple elastic models. J. Volcanol. Geotherm.
Mauna Kea volcano, Hawaii: petrologic and geochemical constraints on post-shield Res. 166 (2), 47–75.
volcanism. J. Geophys. Res. Solid Earth 95 (B2), 1271–1300. Gualda, G.A.R., Ghiorso, M.S., 2013. The Bishop Tuff giant magma body: an alternative to
Fridleifsson, I.B., 1977. Distribution of large basaltic intrusions in the Icelandic crust and the Standard Model. Contrib. Mineral. Petrol. 166, 755–775.
the nature of the layer 2–layer 3 boundary. Geol. Socam Bull. 11, 1689–1693. Gudmundsson, A., 1983. Form and dimensions of dykes in eastern Iceland. Tectonophysics
Fuchs, K., Bonjer, K.P., Gajewski, D., Lüschen, E., Prodehl, C., Sandmeier, K.J., Wilhelm, H., 95, 295–307.
1987. Crustal evolution of the Rhine-graben area. 1. Exploring the lower crust in Gudmundsson, A., 1986. Formation of crustal magma chambers in Iceland. Geology 14,
the Rhine graben rift by unified geophysical experiments. Tectonophysics 141 164–166.
(1–3), 261–275. Gudmundsson, A., 1987. Formation and mechanics of magma reservoirs in Iceland.
Fujita, E., Ukawa, M., Yamamoto, E., 2004. Subsurface cyclic magma sill expansions in the Geophys. J. R. Astron. Soc. Lond. 91, 27–41.
2000 Miyakejima volcano eruption: possibility of two-phase flow oscillation. Gudmundsson, A., 1988. Effect of tensile stress concentration around magma chambers
J. Geophys. Res. http://dx.doi.org/10.1029/2003JB002556. on intrusion and extrusion frequencies. J. Volcanol. Geotherm. Res. 35, 179–194.
Gaffney, E.S., Damjanac, B., 2006. Localization of volcanic activity: topographic effects on Gudmundsson, A., 1990. Emplacement of dykes, sills and crustal magma chambers at di-
dike propagation, eruption and conduit formation. Geophys. Res. Lett. 33, L14313. vergent plate boundaries. Tectonophysics 176, 257–275.
http://dx.doi.org/10.1029/2006GL026852. Gudmundsson, A., 1995. Infrastructure and mechanics of volcanic systems in Iceland.
Gaffney, E.S., Damjanac, B., Valentine, G.A., 2007. Localization of volcanic activity: 2. Effects J. Volcanol. Geotherm. Res. 64 (1), 1–22.
of pre-existing structure. Earth Planet. Sci. Lett. 263, 323–338. Gudmundsson, A., 1998. Magma chambers modeled as cavities explain the formation of
Galerne, C.Y., Galland, O., Neumann, E.-R., Planke, S., 2011. 3D relationships between sills rift zone central volcanoes and their eruption and intrusion statistics. J. Geophys.
and their feeders: evidence from the Golden Valley Sill Complex (Karoo Basin) and Res. 103 (B4), 7401–7412.
experimental modelling. J. Volcanol. Geotherm. Res. 202, 189–199. Gudmundsson, A., 2002. Emplacement and arrest of sheets and dykes in central volca-
Galland, O., Cobbold, P.R., de Bremond d'Ars, J., Hallot, E., 2007a. Rise and emplacement of noes. J. Volcanol. Geotherm. Res. 116, 279–298.
magma during horizontal shortening of the brittle crust: insights from experimental Gudmundsson, A., 2003. Surface stresses associated with arrested dykes in rift zones. Bull.
modelling. J. Geophys. Res. http://dx.doi.org/10.1029/2006JB004604. Volcanol. 65, 606–619. http://dx.doi.org/10.1007/s00445-003-0289-7.
Galland, O., Hallot, E., Cobbold, P.R., Ruffet, G., de Bremod d'Ars, J., 2007b. Volcanism in a Gudmundsson, A., 2006. How local stresses control magma-chamber ruptures, dyke in-
compressional Andean setting: a structural and geochronological study of Tromen jections, and eruptions in composite volcanoes. Earth Sci. Rev. 79, 1–31.
volcano (Neuquen province, Argentina). Tectonics http://dx.doi.org/10.1029/ Gudmundsson, A., 2011. Deflection of dykes into sills at discontinuities and magma-
2006TC002011. chamber formation. Tectonophysics 500 (1), 50–64.
Galland, O., Planke, S., Neumann, E.R., Malthe-Sørenssen, A., 2009. Experimental model- Gudmundsson, A., 2012. Magma chambers: formation, local stresses, excess pressures,
ling of shallow magma emplacement: application to saucer-shaped intrusions. and compartments. J. Volcanol. Geotherm. Res. 237–238, 19–41.
Earth Planet. Sci. Lett. 277 (3–4), 373–383. Gudmundsson, A., Brenner, S.L., 2001. How hydrofractures become arrested. Terra Nova
Galland, O., Burchardt, S., Hallot, E., Mourgues, R., Bulois, C., 2014. Dynamics of dikes 13, 456–462.
versus cone sheets in volcanic systems. J. Geophys. Res. Solid Earth 119 (8), Gudmundsson, A., Brenner, S.L., 2005. On the conditions of sheet injections and eruptions
6178–6192. in stratovolcanoes. Bull. Volcanol. 67, 768–782.
Garagash, D., Detournay, E., 2000. The tip region of a fluid-driven fracture in an elastic me- Gudmundsson, A., Løtveit, I.F., 2014. Sills as fractured hydrocarbon reservoirs: examples
dium. ASME J. Appl. Mech. 67, 183–192. and models. Geol. Soc. Lond., Spec. Publ. 374 (1), 251–271.
García, A., Fernández-Ros, A., Berrocoso, M., Marrero, J.M., Prates, G., De la Cruz-Reyna, S., Gudmundsson, A., Nilsen, K., 2006. Ring faults in composite volcanoes: structures,
Ortiz, R., 2014. Magma displacements under insular volcanic fields, applications to models, and stress fields associated with their formation. J. Geol. Soc. Lond.
eruption forecasting: El Hierro, Canary Islands, 2011–2013. Geophys. J. Int. http:// 269, 83–108.
dx.doi.org/10.1093/gji/ggt505. Gudmundsson, A., Marti, J., Turon, E., 1997. Stress fields generating ring faults in volca-
Gautneb, H., Gudmundsson, A., Oskarsson, N., 1989. Structure, petrochemistry and evolu- noes. Geophys. Res. Lett. 24, 1559–1562.
tion of a sheet swarm in an Icelandic central volcano. Geol. Mag. 126, 659–673. Gudmundsson, A., Berg, S.S., Lyslo, K.B., Skurtveit, E., 2001. Fracture networks and fluid
Gee, M.J.R., Masson, D.G., Watts, A.B., Mitchell, N.C., 2001. Offshore continuation of volca- transport in active fault zones. J. Struct. Geol. 23, 343–353.
nic rift zones, El Hierro, Canary Islands. J. Volcanol. Geotherm. Res. 105, 107–119. Gudmundsson, A., Friese, N., Galindo, I., Philipp, S.L., 2008. Dike-induced reverse faulting
Geshi, N., 2005. Structural development of dike swarms controlled by the change of in a graben. Geology 36, 123–126. http://dx.doi.org/10.1130/G24185A.1.
magma supply rate: the cone sheets and parallel dike swarms of the Miocene Gudmundsson, A., Pasquarè Mariotto, F.A., Tibaldi, A., 2014. Dykes, sills, laccoliths, and in-
Otoge igneous complex, Central Japan. J. Volcanol. Geotherm. Res. 141 (3), 267–281. clined sheets in Iceland. In: Breitkreuz, C., Rocchi, S. (Eds.), Laccoliths, Sills and Dykes,
Geshi, N., 2008. Vertical and lateral propagation of radial dikes inferred from the flow- Physical Geology of Shallow Level Magmatic Systems. Advances in Volcanology series
direction analysis of the radial dike swarm in Komochi Volcano, Central Japan. (in press).
J. Volcanol. Geotherm. Res. 173 (1), 122–134. Halinan, S., 1993. Non-chaotic collapse at funnel calderas: gravity study of the ring frac-
Geshi, N., 2009. Asymmetric growth of collapsed caldera by oblique subsidence during tures at Guayabo Caldera, Costa Rica. Geology 21, 367–370.
the 2000 eruption of Miyakejima, Japan. Earth Planet. Sci. Lett. 280, 148–158. Hall, A., 1987. Igneous Petrology. John Wiley & Sons Inc., New York (573 pp.).
http://dx.doi.org/10.1016/j.epsl.2009.01.027. Halliday, A.N., Aftalion, M., Parsons, I., Dickin, A.P., Johnson, M.R.W., 1987. Syn-orogenic
Geshi, N., Shimano, T., Chiba, T., Nakada, S., 2002. Caldera collapse during the eruption of alkaline magmatism and its relationship to the Moine Thrust Zone and the thermal
Miyakejima Volcano, Japan. Bull. Volcanol. 64, 55–68. state of the Lithosphere in NW Scotland. J. Geol. Soc. 144 (4), 611–617.
Geyer, A., Martí, J., 2008. The new worldwide collapse caldera database (CCDB): a tool for Hamilton, W.B., 1995. Subduction systems and magmatism. In: Smellie, J.R. (Ed.), Volca-
studying and understanding caldera processes. J. Volcanol. Geotherm. Res. 175 (3), nism Associated With Extension to Consuming Plate Margins. Geol. Soc. London
334–354. Spec. Publ. vol. 81, pp. 3–28.
Geyer, A., Martí, J., 2014. A short review of our current understanding of the development Hansen, D.M., 2006. The morphology of intrusion-related vent structures and their impli-
of ring faults during collapse caldera formation. Front. Earth Sci. Volcanol. 2, 22. cations for constraining the timing of intrusive events along the NE Atlantic margin.
Gilbert, G.K., 1877. Report on the Geology of the Henry Mountains. U.S. Government J. Geol. Soc. 163, 789–800.
Printing Office, Washington D.C. (160 pp.). Hansen, D.M., Cartwright, J.A., Thomas, D., 2004. 3D seismic analysis of the geometry
Glazner, A.F., 1991. Plutonism, oblique subduction, and continental growth: an example of igneous sills and sill junctions relationships. In: Davies, R.J., Cartwright, J.A.,
from the Mesozoic of California. Geology 19, 784–786. Stewart, S.A., Lappin, M., Underhill, J.R. (Eds.), 3D Seismic Technology: Application
Goes, S., Giardini, D., Jenny, S., Hollenstein, C., Kahle, H.-G., Geiger, A., 2004. A recent tec- to the Exploration of Sedimentary Basins. Memoirs. Geological Society, London,
tonic reorganization in the south-central Mediterranean. Earth Planet. Sci. Lett. 226, pp. 199–208.
335–345. http://dx.doi.org/10.1016/j.epsl.2004.07.038. Hardee, H.C., 1982. Incipient magma chamber formation as a result of repetitive intru-
González, P.J., Samsonov, S.V., Pepe, S., Tiampo, K.F., Tizzani, P., Casu, F., Fernández, J., sions. Bull. Volcanol. 45 (1), 1–49.
Camacho, A.G., Sansosti, E., 2013. Magma storage and migration associated with the Harker, A., 1904. The Tertiary igneous rocks of Skye. Mem. Geol. Surv. 1–481.
2011–2012 El Hierro eruption: implications for crustal magmatic systems at oceanic Harrison, T.N., Brown, P.E., Dempster, T.J., Hurron, D.H.W., Becker, S.M., 1990. Granite
island volcanoes. J. Geophys. Res. Solid Earth 118 (8), 4361–4377. magmatism and extensional tectonics in southern Greenland. Geol. J. 25, 287–293.
Gottsmann, J., Camacho, A.G., Martí, J., Wooller, L., Fernández, J., Garcia, A., Rymer, H., Hatayama, Y., et al., 1980. Chigaku Jiten (Geological Dictionary). The Association for Geo-
2008. Shallow structure beneath the Central Volcanic Complex of Tenerife from logical Collaboration, Heibonsha K.K., Tokyo (1612 pp., (in Japanese)).
new gravity data: implications for its evolution and recent reactivation. Phys. Earth Hauksson, E., 1983. Episodic rifting and volcanism at Krafla in north Iceland: growth of
Planet. Inter. 168 (3), 212–230. large ground fissures along the plate boundary. J. Geophys. Res. 88, 626–636.
Goulty, N.R., 2005. Emplacement mechanism of the Great Whin and Midland Valley dol- Hayashi, S., Kasahara, M., Tanaka, K., Hamaguchi, H., Zana, N., 1992. Major chemistry of
erite sills. J. Geol. Soc. 162, 1047–1056. recent eruptive products from Nyamuragire volcano, Africa (1976–1989).
Grandin, R., Socquet, A., Doubre, C., Jacques, E., King, G.C.P., 2012. Elastic thickness control Tectonophysics 209, 273–276.
of lateral dyke intrusion at mid-ocean ridges. Earth Planet. Sci. Lett. 319–320, 83–95. Heiken, G., Goff, F., Gardner, J.N., Baldridge, W.S., 1990. The Valles/Toledo Caldera Com-
http://dx.doi.org/10.1016/j.epsl.2011.12.011. plex, Jemez Volcanic Field, New Mexico. Annu. Rev. Earth Planet. Sci. 18, 27–53.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 131

Hildenbrand, A., Gillot, P., Le Roy, I., 2004. Volcano-tectonic and geochemical evolution of Klein, F.W., Koyanagi, R.Y., Nakata, J.S., Tanigawa, W.R., 1987. The seismicity of Kilauea's
an oceanic intra-plate volcano: Tahiti–Nui (French Polynesia). Earth Planet. Sci. Lett. magma system. U. S. Geol. Surv. Prof. Pap. 350, 1019–1186.
217, 349–365. Klügel, A., Hansteen, T.H., Galipp, K., 2005. Magma storage and underplating beneath
Hildner, E., Kluegel, A., Hansteen, T.H., 2012. Barometry of lavas from the 1951 eruption of Cumbre Vieja volcano, La Palma (Canary Islands). Earth Planet. Sci. Lett. 236 (1),
Fogo, Cape Verde Islands: implications for historic and prehistoric magma plumbing 211–226.
systems. J. Volcanol. Geotherm. Res. 217–218, 73–90. http://dx.doi.org/10.1016/j. Koch, F.G., Johnson, A.M., Pollard, D.D., 1981. Monoclinal bending of strata over laccolithic
volgeores.2011.12.014. intrusions. Tectonophysics 74 (3), T21–T31.
Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithospheric Koehler, R.D., Farrell, R.-E., Burns, P.A.C., Combellick, R.A., 2012. Quaternary faults and
magmatism. J. Geophys. Res. 86, 10153–10192. folds in Alaska: a digital database. Alaska Division of Geological & Geophysical Sur-
Hildreth, W., 1987. New perspectives on the eruption of 1912 in the Valley of Ten Thousand veys, Miscellaneous Publication 141 (31 pp.).
Smokes, Katmai National Park, Alaska. Bull. Volcanol. 49, 680–693. Kokelaar, P., Raine, P., Branney, M., 2007. Incursion of a large-volume, spatter-bearing
Hills, E.S., 1972. Elements of Structural Geology. second edition. John Wiley & Sons, New pyroclastic density current into a caldera lake: Pavey Ark ignimbrite, Scafell caldera,
York. England. Bull. Volcanol. 70, 23–54. http://dx.doi.org/10.1007/s00445-007-0118-5.
Holohan, E.P., van Wyk de Vries, B., Troll, V.R., 2007. Analogue models of caldera collapse Komorowski, J.C., 2002. The January 2002 flank eruption of Nyiragongo volcano (DRC):
in strike-slip tectonic regimes. Bull. Volcanol. http://dx.doi.org/10.1007/s00445-007- chronology, evidence for a tectonic rift trigger and impact of lava flows on the city
0166-x. of Goma. Acta Vulcanol. 14, 25–57.
Horsman, E., Tikoff, B., Morgan, S., 2005. Emplacement-related fabric and multiple sheets Kühn, D., Dahm, T., 2008. Numerical modelling of dyke interaction and its influence on
in the Maiden Creek sill, Henry Mountains, Utah, USA. J. Struct. Geol. 27 (8), oceanic crust formation. Tectonophysics 447 (1), 53–65.
1426–1444. La Delfa, S., Patane, G., Clocchiatti, R., Joron, J.L., Tanguy, J.C., 2001. Activity of Mount Etna
Hughes, G.R., Mahood, G.A., 2011. Silicic calderas in arc settings: characteristics, distribu- preceding the February 1999 fissure eruption: inferred mechanism from seismologi-
tion, and tectonic controls. Geol. Soc. Am. Bull. 123 (7–8), 1577–1595. cal and geochemical data. J. Volcanol. Geotherm. Res. 105, 121–139.
Huppert, H., Sparks, R.S., 1981. The fluid dynamics of a basaltic magma chamber Lagarde, J.L., Brun, J.P., Gapais, D., 1990. Formation of epizonal granitic plutons by in situ
replenished by influx of hot, dense ultrabasic magma. Contrib. Mineral. Petrol. 75 assemblage of laterally expanding magma. C. R. Acad. Sci. Paris 310 (II), 1109–1114.
(3), 279–289. http://dx.doi.org/10.1007/bf01166768. Lagmay, A.M.F., van Wyk de Vries, B., Kerle, N., Pyle, D.M., 2000. Volcano instability in-
Hutton, D.H.W., 1982. A tectonic model for the emplacement of the Main Donegal granite, duced by strike-slip faulting. Bull. Volcanol. 62, 331–346.
NW Ireland. J. Geol. Soc. Lond. 139, 615–631. Lahr, J.C., Chouet, B.A., Stephens, C.D., Power, J.A., Page, R.A., 1994. Earthquake classifica-
Hutton, D.H.W., 2009. Insights into magmatism in volcanic margins: bridge structures and tion, location and error analysis in a volcanic environment — implications for the
a new mechanism of basic sill emplacement—Theron Mountains, Antarctica. Pet. magmatic system of the 1989–1990 eruptions at Redoubt volcano, Alaska.
Geosci. 15 (3), 269–278. J. Volcanol. Geotherm. Res. 62, 137–151.
Hutton, D.H.W., Dempster, T.J., Brown, P.E., Becker, S.D., 1990. A new mechanism of gran- Larsen, G., Gronvold, K., Thorarinsson, S., 1979. Volcanic eruption through a geothermal
ite emplacement — intrusion in active extensional shear zones. Nature 343, 452–455. borehole at Namafjall, Iceland. Nature 278, 707–710.
Ishizuka, O., Geshi, N., Itoh, J., Kawanabe, Y., Tuzino, T., 2008. The magmatic plumbing of Le Bas, M.J., 1971. Cone-sheets as a mechanism of uplift. Geol. Mag. 108 (05), 373–376.
the submarine Hachijo NW volcanic chain, Hachijojima, Japan: long-distance magma Leake, B.E., 1990. Granite magmas: their sources, initiation and consequences of emplace-
transport? J. Geophys. Res. 113, B08S08. http://dx.doi.org/10.1029/2007JB005325. ment. J. Geol. Soc. Lond. 147, 579–589.
Ito, G., Martel, S.J., 2002. Focusing of magma in the upper mantle through dike interaction. Lecuyer, F., Bellier, O., Gourgaud, A., Vincent, P.M., 1997. Tectonique active du Nord-Est de
J. Geophys. Res. 107 (B10), 2223. http://dx.doi.org/10.1029/2001JB000251. Sulawesi (Indonesie) et controle structural de la caldeira de Tondano. C. R. Acad. Sci.
Jaxybulatov, K., Shapiro, N.M., Koulakov, I., Mordret, A., Landès, M., Sens-Schönfelder, C., Paris 325 (8), 607–613.
2014. A large magmatic sill complex beneath the Toba caldera. Science 346, 617–619. Lefort, P., 1981. Manaslu leucoganite – a collision signature of the Himalaya – a model for
Jellinek, A.M., DePaolo, D.J., 2003. A model for the origin of large silicic magma chambers: its genesis and emplacement. J. Geophys. Res. 86, 545–568.
precursors of caldera-forming eruptions. Bull. Volcanol. 65, 363–381. http://dx.doi. Legrand, D., Calahorrano, A., Guillier, B., Rivera, L., Ruiz, M., Villagomez, D., Yepes, H., 2002.
org/10.1007/s00445-003-0277-y. Stress tensor analysis of the 1998–1999 tectonic swarm of northern Quito related to
Jin, Z.H., Johnson, S.E., 2008. Magma‐driven multiple dike propagation and fracture tough- the volcanic swarm of Guagua Pichincha volcano, Ecuador. Tectonophysics 344,
ness of crustal rocks. J. Geophys. Res. Solid Earth (1978–2012) 113 (B3). http://dx.doi. 15–36.
org/10.1029/2006JB004761. Legros, F., Kelfoun, K., Martı́, J., 2000. The influence of conduit geometry on the dynamics
Johnson, R.B., 1961. Patterns and origin of radial dike swarms associated with West of caldera-forming eruptions. Earth Planet. Sci. Lett. 179 (1), 53–61.
Spanish peak and Dike Mountain, South-Central Colorado. Geol. Soc. Am. Bull. 72, Letourneur, L., Peltier, A., Staudacher, T., Gudmundsson, A., 2008. The effects of rock het-
579–590. erogeneities on dyke paths and asymmetric ground deformation: the example of
Johnson, S.E., Paterson, S.R., Tate, M.C., 1999. Structure and emplacement history of Piton de la Fournaise (Réunion Island). J. Volcanol. Geotherm. Res. 173 (3), 289–302.
multiple-center, cone-sheet-bearing ring complex: the Zarza Intrusive Complex, Lin, G., Amelung, F., Lavallee, Y., Okubo, P.G., 2014. Seismic evidence for a crustal magma
Baja California, Mexico. Geol. Soc. Am. Bull. 111 (4), 607–619. reservoir beneath the upper east rift zone of Kilauea volcano, Hawaii. Geology http://
Johnson, S.E., Schmidt, K.L., Tate, M.C., 2002. Ring complexes in the Peninsular Range dx.doi.org/10.1130/G35001.1.
Batholith, Mexico and the USA: magma plumbing systems in the middle and upper Lipman, P.W., 1997. Subsidence of ash-flow calderas: relation to caldera size and magma-
crust. Lithos 61, 187–208. chamber geometry. Bull. Volcanol. 59, 198–218.
Johnson, D.J., Eggers, A.A., Bagnardi, M., Battaglia, M., Poland, M.P., Miklius, A., 2010. Lipman, P.W., 2007. Incremental assembly and prolonged consolidation of Cordilleran
Shallow magma accumulation at Kīlauea Volcano, Hawaii, revealed by microgravity magma chambers: Evidence from the Southern Rocky Mountain volcanic field.
surveys. Geology 38 (12), 1139–1142. Geosphere 3 (1), 42–70.
Jónsson, S., 2009. Stress interaction between magma accumulation and trapdoor faulting Lister, J.R., 1990. Buoyancy-driven fluid fracture: the effects of material toughness and of
on Sierra Negra volcano, Galápagos. Tectonophysics 471, 36–44. http://dx.doi.org/10. low-viscosity precursors. J. Fluid Mech. 210, 263–280.
1016/j.tecto.2008.08.005. Lister, J.R., Kerr, R.C., 1991. Fluid‐mechanical models of crack propagation and their
Karlstrom, L., Dufek, J., Manga, M., 2009. Organization of volcanic plumbing through mag- application to magma transport in dykes. J. Geophys. Res. Solid Earth 96 (B6),
matic lensing by magma chambers and volcanic loads. J. Geophys. Res. 114, B10204. 10049–10077.
http://dx.doi.org/10.1029/2009JB006339. Lloyd, A.S., Ruprecht, P., Hauri, E.H., Rose, W., Gonnermann, H.M., Plank, T., 2014.
Karlstrom, L., Dufek, J., Manga, M., 2010. Magma chamber stability in arc and continental NanoSIMS results from olivine-hosted melt embayments: magma ascent rate during
crust. J. Volcanol. Geotherm. Res. 190 (3), 249–270. explosive basaltic eruptions. J. Volcanol. Geotherm. Res. 283, 1–18.
Karlstrom, L., Wright, H.M., Bacon, C.R., 2015. The effect of pressurized magma chamber Lowenstern, J.B., Wallmann, P.C., Pollard, D.D., 1991. The west Mageik lake sill complex as
growth on melt migration and pre-caldera vent locations through time at Mount an analogue for magma transport during the 1912 eruption at the valley of Ten
Mazama, Crater Lake, Oregon. Earth Planet. Sci. Lett. 412, 209–219. Thousand Smokes, Alaska. Geophys. Res. Lett. 18 (8), 1569–1572.
Kauahikaua, J.P., 1993. Geophysical characteristics of the hydrothermal systems of Kilauea Maaløe, S., 1987. The generation and shape of feeder dykes from mantle sources. Contrib.
Volcano, Hawaii. Geothermics 22, 271–299. Mineral. Petrol. 96 (1), 47–55.
Kauahikaua, J.P., Hildenbrand, T., Webring, M., 2000. Deep magmatic structures of Hawaiian Maaløe, S., Scheie, Å., 1982. The permeability controlled accumulation of primary magma.
volcanoes, imaged by three-dimensional gravity models. Geology 28, 883–886. Contrib. Mineral. Petrol. 81 (4), 350–357.
Kavanagh, J.L., Menand, T., Sparks, R.S.J., 2006. An experimental investigation of sill forma- Maccaferri, F., Rivalta, E., Keir, D., Acocella, V., 2014. Off-rift volcanism in rift zones deter-
tion and propagation in layered elastic media. Earth Planet. Sci. Lett. 245, 799–813. mined by crustal unloading. Nat. Geosci. 7, 297–300. http://dx.doi.org/10.1038/
Kennedy, B., Stix, J., 2007. Magmatic processes associated with caldera collapse at Ossipee NGEO2110.
ring dyke, New Hampshire. Geol. Soc. Am. Bull. 119 (1–2), 3–17. MacLeod, N.S., Sherrod, D.R., Chitwood, L.A., Jensen, R.A., 1995. Geologic map of Newberry
Kervyn, M., Ernst, G.G.J., van Wyk, B., de Vries, L., Mathieu, P. Jacobs, 2009. Volcano load Volcano, Deschutes, Klamath, and Lake Counties, Oregon: U.S. Geological Survey
control on dyke propagation and vent distribution: insights from analogue modeling. Miscellaneous Investigations Series Map I-2455, 2 sheets, scale 1:62,500, pamphlet,
J. Geophys. Res. 114, B03401. http://dx.doi.org/10.1029/2008JB005653. 23 p.
Kissel, C., Laj, C., Sigurdsson, H., Guillou, H., 2010. Emplacement of magma in Eastern Magee, C., Stevenson, C., O'Driscoll, B., Schofield, N., McDermott, K., 2012. An alternative
Iceland dikes: insights from magnetic fabric and rock magnetic analyses. J Volcanol emplacement model for the classic Ardnamurchan cone sheet swarm, NW Scotland,
Geoth Res 191, 79–92. involving lateral magma supply via regional dykes. J. Struct. Geol. 43, 73–91.
Klausen, M.B., 2004. Geometry and mode of emplacement of the Thverartindur cone Mahood, G.A., Hildreth, W., 1986. Geology of the peralkaline volcano at Pantelleria, Strait
sheet swarm, SE Iceland. J. Volcanol. Geotherm. Res. 138, 185–204. of Sicily. Bull. Volcanol. 48, 143–172.
Klausen, M.B., 2006. Geometry and mode of emplacement of dike swarms around the Malthe-Sørenssen, A., Planke, S., Svensen, H., Jamtveit, B., 2004. Formation of saucer
Birnudalstindur igneous centre, SE Iceland. J. Volcanol. Geotherm. Res. 151 (4), shaped sills. In: Breitkreuz, C., Petford, N. (Eds.), Physical Geology of High-level Mag-
340–356. matic Systems: Geological Society. Special Publication, London, pp. 215–227.
132 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Manconi, A., Longpré, M.-A., Walter, T.R., Troll, V.R., Hansteen, T.H., 2009. The effects of Miura, D., 1999. Arcuate pyroclastic conduits, ring faults, and coherent floor at Kumano
flank collapses on volcano plumbing systems. Geology 37 (12), 1099–1102. http:// caldera, southwest Honshu, Japan. J. Volcanol. Geotherm. Res. 92 (3), 271–294.
dx.doi.org/10.1130/G30104A.1. Miura, S., Ueki, S., Sato, T., Tachibana, K., Hamaguchi, H., 2000. Crustal deformation asso-
Mandl, G., 1988. Mechanics of Tectonic Faulting: Models and Basic Concepts. Elsevier, ciated with the 1998 seismo-volcanic crisis of Iwate Volcano, Northeastern Japan, as
Amsterdam (407 pp.). observed by a dense GPS network. Earth Planets Space 52, 1003–1008.
Marcotte, S.B., Klepeis, K.A., Clarke, G.L., Gehrels, G., Hollis, J.A., 2005. Intra-arc Mkweli, S., Kamber, B., Berger, M., 1995. Westward continuation of the craton–Limpopo
transpression in the lower crust and its relationship to magmatism in a Mesozoic Belt tectonic break in Zimbabwe and new age constraints on the timing of the thrust-
magmatic arc. Tectonophysics 407, 135–163. ing. J. Geol. Soc. 152 (1), 77–83.
Marra, F., 2001. Strike-slip faulting and block rotation: a possible triggering mechanism Moore, I., Kokelaar, P., 1998. Tectonically controlled piecemeal caldera collapse: a case
for lava flows in the Alban Hills? J. Struct. Geol. 23, 127–141. study of Glencoe volcano, Scotland. Geol. Soc. Am. Bull. 110, 1448–1466.
Marsh, B., 1989. On the convective style and vigour in sheet-like magmac hambersd. Moore, J.G., Normark, W.R., Holcomb, R.T., 1994. Giant Hawaiian landslides. Annu. Rev.
Petrol 30, 479–530. Earth Planet. Sci. 22, 119–144.
Marsh, B.D., 1982. On the mechanics of igneous diapirism, stoping, and zone-melting. Am. Mori, J., McKee, C., 1987. Outward-dipping ring-fault structure at Rabaul Caldera as shown
J. Sci. 282, 808–855. by earthquake locations. Science 235, 193–195. http://dx.doi.org/10.1126/science.
Marsh, B.D., 2000. Magma chambers. Encyclopedia of volcanoes 1, 191–206. 235.4785.193.
Marsh, B.D., 2004. A magmatic mush column Rosetta Stone: the McMurdo dry valleys of Muirhead, J.D., Airoldi, G., Rowland, J.V., White, J.D., 2012. Interconnected sills and in-
Antarctica. EOS Trans. Am. Geophys. Union 85, 497–508. clined sheet intrusions control shallow magma transport in the Ferrar large igneous
Martì, J., Geyer, A., 2009. Central vs. flank eruptions at Teide–Pico Viejo twin stratovol- province, Antarctica. Geol. Soc. Am. Bull. 124 (1–2), 162–180.
canoes (Tenerife, Canary Islands). J. Volcanol. Geotherm. Res. 181, 47–60. Muller, O.H., 1986. Changing stresses during emplacement of the radial dike swarm at
Martí, J., Mitjavila, J., Roca, E., Aparicio, A., 1992. Cenozoic magmatism of the Valencia Spanish Peaks, Colorado. Geology 14 (2), 157–159.
trough (western Mediterranean): relationship between structural evolution and vol- Muller, O.H., Pollard, D.D., 1977. The stress state near Spanish Peaks, Colorado, deter-
canism. Tectonophysics 203 (1), 145–165. mined from a dike pattern. Pure Appl. Geophys. 115, 69–86.
Massonnet, D., Sigmundsson, F., 2000. Remote sensing of volcano deformation by radar Murase, T., Mcbirney, A.R., 1973. Properties of some common igneous rocks and their
interferometry from various satellites. Geophys. Monogr. Ser. 116, 207–221. melts at high temperatures. Geol. Soc. Am. Bull. 84, 3563–3592.
Mastin, L.G., Pollard, D.D., 1988. Surface deformation and shallow dike intrusion processes Nairn, I.A., McKee, C.O., Talai, B., Wood, C.P., 1995. Geology and eruptive history of the
at Inyo Craters, Long Valley, California. J. Geophys. Res. 13221–13235. Rabaul caldera area, Papua New Guinea. J. Volcanol. Geotherm. Res. 69 (3–4),
Mathieu, L., van Wyk de Vries, B., 2009. Edifice and substrata deformation induced by in- 255–284.
trusive complexes and gravitational loading in the Mull volcano (Scotland). Bull. Nairn, I.A., Kobayashi, T., Nakagawa, M., 1998. The ~10 ka multiple vent pyroclastic erup-
Volcanol. 71 (10), 1133–1148. http://dx.doi.org/10.1007/s00445-009-0295-5. tion sequence at Tongariro Volcanic Centre, Taupo Volcanic Zone, New Zealand. Part
Mattia, M., Patane, D., Aloisi, M., Amore, M., 2007. Faulting on the western flank of Mt Etna 1. Eruptive processes during regional extension. J. Volcanol. Geotherm. Res. 86,
and magma intrusions in the shallow crust. Terra Nova 19, 89–94. 19–44.
Mazzuoli, R., Tortorici, L., Ventura, G., 1995. Oblique rifting in Salina, Lipari and Vulcano Nakada, S., Nagai, M., Kaneko, T., Nozawa, A., Suzuki-Kamata, K., 2005. Chronology and
islands (Aeolian islands, southern Italy). Terra Nova 7, 444–452. products of the 2000 eruption of Miyakejima Volcano, Japan. Bull. Volcanol. 67,
McCall, G.J.H., Bristow, C.M., 1965. An introductory account of Suswa volcano, Kenya. Bull. 205–218.
Volcanol. 28, 333–367. Nakamura, K., 1977. Volcanoes as possible indicators of tectonic stress orientation:
McGuire, W.J., Pullen, A.D., 1989. Location and orientation of eruptive fissures and principle and proposal. J. Volcanol. Geotherm. Res. 2, 1–16.
feederdykes at Mount Etna; influence of gravitational and regional tectonic stress re- Nakamura, K., Uyeda, S., 1980. Stress gradient in arc-back arc regions and plate subduc-
gimes. J. Volcanol. Geotherm. Res. 38.3, 325–344. tion. J. Geophys. Res. 85, 6419–6428.
McGuire, W.J., Pullen, A.D., Saunders, S.J., 1990. Recent dyke-induced large-scale block Nakamura, K., Jacob, K.H., Davies, J.N., 1977. Volcanoes as possible indicators of tectonic
movement at Mt. Etna and potential slope failure. Nature 343, 357–359. stress orientation — Aleutinian and Alaska. Pure Appl. Geophys. 115, 87–112.
McGuire, W., Murray, B., Pullen, A.D., Saunders, S.J., 1991. Ground deformation monitoring Nakamura, K., Plakfer, G., Jacob, K.H., Davies, J.N., 1980. A tectonic stress trajectory map of
at Mt. Etna; evidence for dyke emplacement and slope instability. J. Geol. Soc. London Alaska using information from volcanoes and faults. Bull. Earthquake Res. Inst. 55,
148, 577–583. 89–100.
McKenzie, D., 1984. A possible mechanism for epeirogenic uplift. Nature 307 (5952), Neri, G., Barberi, G., Orecchio, B., Mostaccio, A., 2003. Seismic strain and seismogenic stress
616–618. regimes in the crust of the southern Tyrrhenian region. Earth Planet. Sci. Lett. 213,
McKenzie, D., McKenzie, J.M., Saunders, R.S., 1992. Dike emplacement on Venus and on 97–112. http://dx.doi.org/10.1016/S0012-821X(03)00293-0.
Earth. J. Geophys. Res. 97, 15,977–15,990. Newman, A.V., Dixon, T.H., Gourmelen, N., 2006. A four-dimensional viscoelastic deforma-
McLeod, P., Tait, S., 1999. The growth of dykes from magma chambers. J. Volcanol. tion model for Long Valley caldera, California, between 1995 and 2000. J. Volcanol.
Geotherm. Res. 92 (3), 231–245. Geotherm. Res. 150, 244–269. http://dx.doi.org/10.1016/j.jvolgeores.2005.07.017.
McNulty, B.A., Tong, W., Tobisch, O.T., 1996. Assembly of a dyke-fed magma chamber: the Nicholson, R., Pollard, D.D., 1985. Dilation and linkage of echelon cracks. J. Struct. Geol. 7,
Jackass Lakes pluton, central Sierra Nevada, California. Geol. Soc. Am. Bull. 108 (8), 583–590.
926–940. Nicolas, A., Boudier, F., Ildefonse, B., 1994. Dyke patterns in diapirs beneath oceanic ridges:
Memeti, V., Paterson, S.R., Matzel, J., Mundil, R., Okaya, D., 2010. Magmatic lobes as the Oman Ophiolite. In: Ryan, M. (Ed.), Magmatic Systems. Academic Press, New
“snapshots” of magma chamber growth and evolution in large, composite batholiths: York, pp. 77–95.
an example from the Tuolumne Intrusion, Sierra Nevada, CA. GSA Bull. http://dx.doi. Nunn, J., 1996. Buoyancy-driven propagation of isolated fluid-filled fractures: implica-
org/10.1130/B30004.1. tions for fluid transport in the Gulf of Mexico. J. Geophys. Res. 101, 2963–2970.
Menand, T., 2008. The mechanics and dynamics of sills in layered elastic rocks and their Odé, H., 1957. Mechanical analysis of the dike pattern of the Spanish Peaks area, Colorado.
implications for the growth of laccoliths and other igneous complexes. Earth Planet. Geol. Soc. Am. Bull. 68, 567–576.
Sci. Lett. 267 (1), 93–99. http://dx.doi.org/10.1016/j.epsl.2007.11.043. O'Driscoll, B., Troll, V.R., Reavy, R.J., Turner, P., 2006. The Great Eucrite intrusion of
Menand, T., 2011. Physical controls and depth of emplacement of igneous bodies: a re- Ardnamurchan, Scotland: reevaluating the ring-dike concept. Geology 34 (3),
view. Tectonophysics 500, 11–19. 189–192.
Menand, T., Tait, S.R., 2002. The propagation of a buoyant liquid-filled fissure from a Olivier, P., Ameglio, L., Richen, H., Vadeboin, F., 1999. Emplacement of the Aya Variscan
source under constant pressure: an experimental approach. J. Geophys. Res. 107 granitic pluton (Basque Pyrenees) in a dextral transcurrent regime inferred from a
(B11), 2306. combined magneto-structural and gravimetric study. J. Geol. Soc. Lond. 156,
Menand, T., Daniels, K.A., Benghiat, P., 2010. Dyke propagation and sill formation in 991–1002.
a compressive tectonic environment. J. Geophys. Res. Solid Earth (1978–2012) Ozerov, A.Yu., Ariskin, A.A., Kyle, P., Bogoyavlenskaya, G.E., Karpenko, S.F., 1997. Petrological–
115 (B8). geochemical model for genetic relationships between basaltic and andesitic magmatism
Meriaux, C., Jaupart, C., 1998. Crack propagation through an elastic plate. J. Geophys. Res. of Klyuchevskoi and Bezymyannyi volcanoes, Kamchatka. Petrology 5, 550–569.
103 (B8), 18295–18314. Parsons, T., Sleep, N.H., Thompson, G.A., 1992. Host rock rheology controls on the
Meriaux, C., Lister, J.R., 2002. Calculation of dike trajectories from volcanic centers. emplacement of tabular intrusions: implications for underplating of extended crust.
J. Geophys. Res. 107 (B4), 2077. http://dx.doi.org/10.1029/2001JB000436. Tectonics 11, 1348–1356.
Michaut, C., 2011. Dynamics of magmatic intrusions in the upper crust: theory and appli- Pasquarè, F.A., Tibaldi, A., 2003. Do transcurrent faults guide volcano growth? The case of
cations to laccoliths on Earth and the Moon. J. Geophys. Res. Solid Earth (1978–2012) NW Bicol Volcanic Arc, Luzon, Philippines. Terra Nova 15 (3), 204–212.
116 (B5). Pasquarè, F.A., Tibaldi, A., 2007. Structure of a sheet-laccolith system revealing the inter-
Michon, L., Saint-Ange, F., Bachelery, P., Villeneuve, N., Staudacher, T., 2007. Role of the play between tectonic and magma stresses at Stardalur Volcano, Iceland. J. Volcanol.
structural inheritance of the oceanic lithosphere in the magmato-tectonic evolution Geotherm. Res. 161, 131–150.
of Piton de la Fournaise volcano (La Réunion Island). J. Geophys. Res. 112, B04205. Pasquarè, G., Francalanci, L., Garduno, V.H., Tibaldi, A., 1993. Structure and geological
http://dx.doi.org/10.1029/2006JB004598. evolution of the Stromboli volcano, Aeolian islands, Italy. Acta Volcanol. Pisa 3,
Miller, C.F., Miller, J.S., 2002. Contrasting stratified plutons exposed in tilt blocks, Eldorado 79–89.
Mountains, Colorado River rift, Nevada, USA. Lithos 61, 209–224. Paterson, S.R., Miller, R.B., 1998. Mid-crustal magmatic sheets in the Cascades Mountains,
Miller, R.B., Paterson, S.R., 1999. In defense of magmatic diapirs. J. Struct. Geol. 21, Washington: implications for magma ascent. J. Struct. Geol. 20, 1345–1363.
1161–1173. Paterson, S.R., Vernon, R.H., 1995. Bursting the bubble of ballooning plutons: a return to
Miller, R.B., Paterson, S.R., 2001. Construction of mid-crustal sheeted plutons: examples nested diapers emplaced by multiple processes. GSA Bull. 107, 1356–1380.
from the north Cascades, Washington. Geol. Soc. Am. Bull. 113 (11), 1423–1442. Paterson, S.R., Tobisch, O.T., Morand, V.J., 1990. The influence of large ductile shear zones
Mitjavila, J., Marti, J., Soriano, C., 1997. Magmatic evolution and tectonic setting of the Ibe- on the emplacement and deformation of the Wyangala Batholith, SE Australia.
rian Pyrite Belt volcanism. J. Petrol. 38 (6), 727–755. J. Struct. Geol. 12 (5/6), 639–650.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 133

Paterson, S.R., Memeti, V., Putirka, K., Paige, M., 2010. Volcano-plutonic plumbing sys- Roman, D.C., Gardine, M.D., 2013. Seismological evidence for long-term and rapidly accel-
tems: a comparison of the Jurassic Guadalupe Igneous Complex to the Cretaceous erating magma pressurization preceding the 2013 eruption of Redoubt Volcano,
Tuolumne batholith, central Sierra Nevada, California. GIC Field Trip Guide (41 pages). Alaska. Earth Planet. Sci. Lett. 371–372, 226–234.
Pedersen, R., Sigmundsson, F., 2004. InSAR based sill model links spatially offset areas of Roman-Berdiel, T., Gapais, D., Brun, J.P., 1995. Analogue models of laccolith formation.
deformation and seismicity for the 1994 unrest episode at Eyjafjallajokull volcano, J. Struct. Geol. 17 (9), 1337–1346.
Iceland. Geophys. Res. Lett. 31, L14610. http://dx.doi.org/10.1029/2004GL020368. Roni, E., Westerman, D.S., Dini, A., Stevenson, C., Rocchi, S., 2014. Feeding and growth of a
Peltier, A., Ferrazzini, V., Staudacher, T., Bachelery, P., 2005. Imaging the dynamics of dyke dyke–laccolith system (Elba Island, Italy) from AMS and mineral fabric data. J. Geol.
propagation prior to the 2000–2003 flank eruptions at Piton de La Fournaise, Reunion Soc. 171 (3), 413–424.
Island. Geophys. Res. Lett. 32. http://dx.doi.org/10.1029/2005GL023720. Roper, S.M., Lister, J.R., 2005. Buoyancy-driven crack propagation from an overpressure
Peltier, A., Bachèlery, P., Staudacher, T., 2009. Magma transport and storage at Piton de La source. J. Fluid Mech. 536, 79–98.
Fournaise (La Reunion) between 1972 and 2007: a review of geophysical and geo- Rosenberg, C.L., 2004. Shear zones and magma ascent: a model based on a review of the
chemical data. J. Volcanol. Geotherm. Res. 184, 93–108. http://dx.doi.org/10.1016/j. Tertiary magmatism in the Alps. Tectonics http://dx.doi.org/10.1029/2003TC001526.
jvolgeores.2008.12.008. Rossetti, F., Storti, F., Salvini, F., 2000. Cenozoic noncoaxial transtension along the western
Petford, N., Atherton, M.P., 1992. Granitoid emplacement and deformation along a major shoulder of the Ross Sea, Antarctica, and the emplacement of McMurdo dyke arrays.
crustal lineament: the Cordillera Blanca, Peru. Tectonophysics 205, 171–185. Terra Nova 12, 60–66.
Petford, N., Gallagher, K., 2001. Partial melting of mafic (amphibolitic) lower crust by pe- Rubin, A.M., 1993. Dykes vs. diapirs in viscoelastic rock. Earth Planet. Sci. Lett. 119,
riodic influx of basaltic magma. Earth Planet. Sci. Lett. 5983, 1–17. 641–659.
Petford, N., Cruden, A.R., Mccaffrey, K.J.W., Vigneresse, J.L., 2000. Granite magma forma- Rubin, A.M., 1995a. Propagation of magma-filled crack. Annu. Rev. Earth Planet. Sci. 23,
tion, transport and emplacement in the Earth's crust. Nature 408, 669–673. 287–336.
Phillips, W.J., 1974. The dynamic emplacement of cone sheets. Tectonophysics 24, Rubin, A.M., 1995b. Getting granite dikes out of the source region. J. Geophys. Res. 100
69–84. (B4), 5911–5929. http://dx.doi.org/10.1029/94JB02942.
Pinel, V., Jaupart, C., 2004. Magma storage and horizontal dyke injection beneath a volca- Rubin, A.M., Pollard, D.D., 1987. Origins of blade-like dikes in volcanic rift zones. In:
nic edifice. Earth Planet. Sci. Lett. 221, 245–262. Decker, R.W., Wight, T.L., Stuffer, P.H. (Eds.), Volcanism in Hawaii. US Geological
Piochi, M., Kilburn, C.R.J., Di Vito, M.A., Mormone, A., Tramelli, A., Troise, C., De Natale, G., Survey Professional Papers 1350, pp. 1449–1470.
2014. The volcanic and geothermally active Campi Flegrei caldera: an integrated mul- Rubin, A.M., Pollard, D.D., 1988. Dike-induced faulting in rift zones of Iceland and Afar.
tidisciplinary image of its buried structure. Int. J. Earth Sci. 103 (2), 401–421. Geology 16 (5), 413–417.
Pitcher, W.S., 1978. The anatomy of a batholith. J. Geol. Soc. Lond. 135, 157–182. Ruprecht, P., Plank, T., 2013. Feeding andesitic eruptions with a high-speed connection
Pitcher, W.S., 1979. The nature, ascent and emplacement of granitic magmas. J. Geol. Soc. from the mantle. Nature 500, 68–72. http://dx.doi.org/10.1038/nature12342.
136, 627–662. Russo, G., Giberti, G., 2000. Mechanical stability of Mt. Vesuvius volcano effects of
Pitcher, W.S., Berger, A.R., 1972. The Geology of Donegal: A Study of Granite Emplacement asymmetries on the stress field. Surv. Geophys. 21, 407–421.
and Unroofing. Wiley, New York (435 pp.). Russo, G., Giberti, G., Sartoris, G., 1996. The influence of regional stresses on the mechan-
Planke, S., Rasmussen, T., Rey, S.S., Myklebust, R., 2005. Seismic characteristics and distri- ical stability of volcanoes: Stromboli (Italy). In: McGuire, W.J., Jones, A.P., Neuberg, J.
bution of volcanic intrusions and hydrothermal vent complexes in the Vøring and (Eds.), Geological Society Special Publication. Volcano Instability on the Earth and
Møre basins. In: Doré, A.G., Vining, B.A. (Eds.), Petroleum Geology: North-West Other Planets 110, pp. 65–75.
Europe and Global Perspectives. Proceedings of the 6th Petroleum Geology Confer- Russo, G., Giberti, G., Sartoris, G., 1997. Numerical modeling of surface deformation and
ence. Geological Society, London, pp. 833–844. mechanical stability of Vesuvius volcano, Italy. J. Geophys. Res. 102, 24785–24800.
Poland, M.P., Fink, J.H., Tauxe, L., 2004. Patterns of magma flow in segmented silicic dikes Ryan, M.P., 1987. Neutral buoyancy and the mechanical evolution of magmatic systems.
at Summer Coon volcano, Colorado: AMS and thin section analysis. Earth Planet. Sci. In: Mysen, B.O. (Ed.), Magmatic Processes: Physicochemical Principles. Spec. Publ.
Lett. 219, 155–169. Geochem. Soc. 1, pp. 259–287.
Poland, M.P., Moats, W.P., Fink, J.H., 2008. A model for radial dike emplacement in com- Ryan, M.P., 1988. The mechanics and three-dimensional internal structure of active mag-
posite cones based on observations from Summer Coon volcano, Colorado, USA. matic systems: Kilauea Volcano, Hawaii. J. Geophys. Res. Solid Earth 93 (B5),
Bull. Volcanol. 70, 861–875. http://dx.doi.org/10.1007/s00445-007-0175-9. 4213–4248.
Pollard, D.D., 1987. Elementary fracture mechanics applied to the structural interpretation Saint Blanquat, M., Tikoff, B., Teyssier, C., Vigneresse, J.L., 1998. Transpressional kinematics
of dykes. Mafic Dyke Swarms 34, 5–24. and magmatic arcs. In: Holdsworth, R.E., Strachan, R.A., Dewey, J.F. (Eds.), Continental
Pollard, D.D., Johnson, A.M., 1973. Mechanics of growth of some laccolithic intrusions in Transpressional and Transtensional Tectonics. Special Publications vol. 135. Geologi-
the Henry Mountains, Utah. II. Bending and failure of overburden layers and sill for- cal Society of London, pp. 327–340.
mation. Tectonophysics 18, 311–354. Sanchez, J.J., Wyss, M., Mcnutt, S.R., 2004. Temporal–spatial variations of stress at Redoubt
Pollard, D.D., Muller, O.H., Dockstader, D.R., 1975. The form and growth of fingered sheet volcano, Alaska, inferred from inversion of fault plane solutions. J. Volcanol.
intrusions. Geol. Soc. Am. Bull. 86 (3), 351–363. Geotherm. Res. 130, 1–30.
Pollard, D.D., Delaney, P.T., Duffield, W.A., Endo, E.T., Okamura, A.T., 1983. Surface defor- Sartoris, G., Pozzi, J.P., Phillippe, C., Mouel, J.L.L., 1990. Mechanical stability of shallow
mation in volcanic rift zones. Tectonophysics 94, 541–584. magma chambers. J. Geophys. Res. 95 (B4), 5141–5151. http://dx.doi.org/10.1029/
Polteau, S., Mazzini, A., Galland, O., Planke, S., Malthe-Sorenssen, A., 2008. Saucer-shaped JB095iB04p05141.
intrusions: occurrences, emplacement and implications. Earth Planet. Sci. Lett. 266 Saunders, S.J., 2005. The possible contribution of circumferential fault intrusion to caldera
(1–2), 195–204. resurgence. Bull. Volcanol. 67 (1), 57–71.
Porreca, M., Acocella, V., Massimi, E., Mattei, M., Funiciello, R., De Benedetti, A.A., 2006. Saunders, A.D., Jones, S.M., Morgan, L.A., Pierce, K., Widdowson, M., Xu, Y.G., 2007. Region-
Geometric and kinematic features of the dike complex at Mt. Somma, Vesuvio al uplift associated with continental large igneous provinces: the roles of mantle
(Italy). Earth Planet. Sci. Lett. 245, 389–407. http://dx.doi.org/10.1016/j.epsl.2006. plumes and the lithosphere. Chem. Geol. 241 (3), 282–318.
02.027. Scarrow, J.H., Vaughan, A.P.M., Leat, P.T., 1997. Ridge-trench collision-induced switching
Preston, R.J., 2001. Composite minor intrusions as windows into subvolcanic magma res- of arc tectonics and magma sources: clues from Antarctic Peninsula mafic dykes.
ervoir processes: mineralogical and geochemical evidence for complex magmatic Terra Nova 9 (5–6), 255–259.
plumbing systems in the British Tertiary Igneous Province. J. Geol. Soc. 158, 47–58. Schirnick, C., van den Bogaard, P., Schmincke, H.-U., 1999. Cone sheet formation and in-
Quareni, F., Ventura, G., Mulargia, F., 2001. Numerical modelling of the transition from fis- trusive growth of an oceanic island — the Miocene Tejeda complex on Gran Canaria
sure to central-type activity on volcanoes: a case study from Salina Island, Italy. Phys. (Canary Islands). Geology 27 (3), 207–210.
Earth Planet. Inter. 124, 213–221. Schmidt, C.J., Smedes, H.W., O'neill, J.M., 1990. Syncompressional emplacement of the
Quick, J.E.S., Sinigoi, S., Mayer, A., 1994. Emplacement dynamics of a large mafic intrusion Boulder and Tobacco Root Batholiths (Montana, USA) by pull-apart along old fault
in the lower crust, Ivrea–Verbano zone, Northern Italy. J. Geophys. Res. 99, zones. Geol. J. 25, 305–318.
21559–21573. Schofield, N.J., Brown, D.J., Magee, C., Stevenson, C.T., 2012. Sill morphology and comparison
Ramelow, J., Riller, U., Romer, R.L., Oncken, O., 2006. Kinematic link between episodic of brittle and non-brittle emplacement mechanisms. J. Geol. Soc. 169 (2), 127–141.
trapdoor collapse of the Negra Muerta Caldera and motion on the Olacapato–El Secor, D., Pollard, D., 1975. On the stability of open hydraulic fractures in the Earth's crust.
Toro Fault Zone, southern central Andes. Int. J. Earth Sci. (Geol. Rundsch.) 95, 529–541. Geophys. Res. Lett. 2, 510–513.
Ranalli, G., 1995. Rheology of the Earth. 2nd ed Chapman Hall, London, p. 413. Self, S., Rampino, M.R., 1981. The 1883 eruption of Krakatau. Nature 294, 699–704.
Richey, J.E., Thomas, H.H., 1930. The Geology of Ardnamurchan, North-west Mull and Sewell, R.J., Tang, D.L., Campbell, S.D.G., 2012. Volcanic–plutonic connections in a tilted
Coll. Memoir of the Geological Survey of Great Britain, Sheet 51 and 52, Scotland nested caldera complex in Hong Kong. Geochem. Geophys. Geosyst. 13 (1). http://
(393 pp.). dx.doi.org/10.1029/2011GC003865.
Rickwood, P.C., 1990. The anatomy of a dyke and the determination of propagation and Shaw, H.R., 1965. Comments on viscosity, crystal settling and convection in granitic sys-
magma flow directions. In: Parker, A.J., Rickwood, P.C., Tucker, D.H. (Eds.), Mafic tems. Am. J. Sci. 263, 120–152.
Dykes and Emplacement Mechanisms. Balkema, Rotterdam, pp. 81–100. Shaw, H.R., 1980. Fracture mechanisms of magma transport from the mantle to the sur-
Riller, U., Petrinovic, I., Ramelow, J., Strecker, M., Oncken, O., 2001. Late Cenozoic tecto- face. In: Hargreaves, R.B. (Ed.), Physics of Magmatic Processes. Princeton University
nism, collapse caldera and plateau formation in the central Andes. Earth Planet. Sci. Press, Princeton, pp. 201–264.
Lett. 188, 299–311. Shimura, T., 1992. Intrusion of granitic magma and uplift tectonics of the Hidaka meta-
Rittmann, A., 1962. Volcanoes and Their Activity. Interscience Publishers. morphic belt, Hokkaido. J. Geol. Soc. Jpn. 98 (1), 1–20.
Robson, G.R., Barr, K.G., 1964. The effect of stress on faulting and minor intrusions in the Sigmundsson, F., 2006. Magma does the splits. Nature 442, 251–252.
vicinity of a magma body. Bull. Volcanol. 27, 315–330. Simkin, T., Howard, K.A., 1970. Caldera collapse in the Galapagos Islands, 1968. Science
Rodríguez-Losada, J.A., Martinez-Frias, J., 2004. The felsic complex of the Vallehermoso 169, 429–437.
Caldera: interior of an ancient volcanic system (La Gomera, Canary Islands). Sinton, J.M., Detrick, R.S., 1992. Mid‐ocean ridge magma chambers. J. Geophys. Res. Solid
J. Volcanol. Geotherm. Res. 137 (4), 261–284. Earth 97.B1 (1978–2012), 197–216.
134 A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135

Smith, R.P., 1987. Dyke emplacement at Spanish Peaks, Colorado. In: Halls, H.C., Fahrig, Tobisch, O.T., Paterson, S.R., 1990. The Yarra granite: an intradeformational pluton associ-
W.F. (Eds.), Mafic Dyke Swarms. Geological Association of Canada Special Paper 34, ated with ductile thrusting, Lachlan Fold Belt, southeastern Australia. Geol. Soc. Am.
pp. 47–54. Bull. 102, 693–703.
Smith, R.L., Bailey, R.A., 1968. Resurgent cauldrons. Geol. Soc. Am. Mem. 116, 613–662. Toyoshima, T., Komatsu, M., Shimura, T., 1994. Tectonic evolution of lower crustal rocks in
Sparks, R.S.J., 1988. Petrology and geochemistry of the Loch Ba ring-dyke, Mull (NW an exposed magmatic arc section in the Hidaka metamorphic belt, Hokkaido, north-
Scotland): an example of the extreme differentiation of tholeiitic magmas. Contrib. ern Japan. Island Arc 3, 182–198.
Mineral. Petrol. 100, 446–461. Troll, V.R., Walter, T.R., Schmincke, H.-U., 2002. Cyclic caldera collapse: piston or piece-
Sparks, R.S.J., Huppert, H.E., Turner, J.S., Sakuyama, M., O'Hara, J., 1984. The fluid dynamics meal subsidence? Field and experimental evidence. Geology 30, 135–138.
of evolving magma chambers [and discussion]. Philos. Trans. R. Soc. London, Ser. A Upton, B.G.J., Wright, J.B., 1961. Intrusions of gabbro and granophire in the Snaefelsness,
310, 511–534. western Iceland. Geol. Mag. 98 (6), 488–492.
Spence, D.A., Sharp, P., 1985. Self-similar solutions for elastohydrodynamic cavity flow. Valentine, A.G., Krogh, K.E.C., 2006. Emplacement of shallow dikes and sills beneath a
Proc. R. Soc. Lond. A400, 289–313. small basaltic volcanic center — the role of pre-existing structure (Paiute Ridge,
Spence, D.A., Sharp, P., Turcotte, D.L., 1987. Buoyancy-driven crack propagation: a mech- southern Nevada, USA). Earth Planet. Sci. Lett. 246, 217–230.
anism for magma migration. J. Fluid Mech. 174, 135–153. van Wyk de Vries, B., Francis, P.W., 1997. Catastrophic collapse at stratovolcanoes induced
Stachel, T., Brey, G., Stanistreet, I.G., 1994. Gross Brukkaros — the unusual intracaldera by gradual volcano spreading. Nature 387 (6631), 387–390.
sediments and their magmatic components. Commun. Geol. Surv. Namibia 9, 23–41. van Wyk de Vries, B., Matela, R., 1998. Styles of volcano-induced deformation: numerical
Steven, T.A., Lipman, P.W., 1976. Calderas of the San Juan volcanic field, southwestern models of substratum flexure, spreading and extrusion. J. Volcanol. Geotherm. Res.
Colorado. U. S. Geol. Surv. Prof. Pap. 95 (35 pp.). 81, 1–18.
Stevens, B., 1911. The laws of intrusions. Bull. Am. Inst. Min. Eng. 41, 1–23. Vigneresse, J.L., Bouchez, J.L., 1997. Successive granitic magma batches during pluton em-
Stevenson, C.T., O'Driscoll, B., Holohan, E.P., Couchman, R., Reavy, R.J., Andrews, G.D., 2008. placement: the case study of Cabeza de Araya, Spain. J. Petrol. 38, 1767–1776.
The structure, fabrics and AMS of the Slieve Gullion ring-complex, Northern Ireland: Voight, B., Elsworth, D., 1997. Failure of volcano slopes. Geotechnique 47, 1–31.
testing the ring-dyke emplacement model. Geol. Soc. Lond., Spec. Publ. 302 (1), Walker, G.P.L., 1958. Geology of the Reydarfjordur Area, Eastern Iceland. Q. J. Geol. Soc.
159–184. 114 (1/4), 367–391.
Tait, S., Jaupart, C., Vergniolle, S., 1989. Pressure, gas content and eruption periodicity of a Walker, G.P.L., 1960. Zeolite zones and dike distribution in relation to the structure of the
shallow, crystallizing magma chamber. Earth Planet. Sci. Lett. 92, 107–123. http://dx. basalts of eastern Iceland. J. Geol. 68, 515–528.
doi.org/10.1016/0012-821X(89)90025-3. Walker, G.P.L., 1974. The structure of Eastern Iceland. In: Kristjánsson, L. (Ed.),
Takada, A., 1990. Experimental study on propagation of liquid-filled crack in gelatin: Geodynamics of Iceland and the North Atlantic Area. Reidel, Dordrecht, pp. 177–188.
shape and velocity in hydrostatic stress conditions. J. Geophys. Res. 95, 8471–8481. Walker, G.P.L., 1975. Intrusive sheet swarms and the identity of Crustal Layer 3 in Iceland.
Takada, A., 1997. Cyclic flank-vent and central-vent eruption patterns. Bull. Volcanol. 58, J. Geol. Soc. Lond. 131, 143–161.
539–556. Walker, G.P.L., 1984. Downsag calderas, ring faults, caldera sizes, and incremental caldera
Takada, A., 1999. Variations in magma supply and magma partitioning: the role of tecton- growth. J. Geophys. Res. 89B, 8407–8416.
ic settings. J. Volcanol. Geotherm. Res. 93 (1), 93–110. Walker, G.P.L., 1999. Volcanic rift zones and their intrusion swarms. J. Volcanol.
Thomson, K., 2007. Determining magma flow in sills, dykes and laccoliths and their impli- Geotherm. Res. 94, 21–34.
cations for sill emplacement mechanisms. Bull. Volcanol. 70 (2), 183–201. Wallmann, P.C., Pollard, D.D., Hildreth, W., Eichelberger, J.C., 1990. New structural limits
Tibaldi, A., 1995. Morphology of pyroclastic cones and tectonics. J. Geophys. Res. 100 on magma chamber locations at the Valley of Ten Thousand Smokes, Katmai National
(B12), 24,521–24,535. Park, Alaska. Geology 18, 1240–1243.
Tibaldi, A., 1996. Mutual influence of diking and collapses at Stromboli volcano, Aeolian Walter, T.R., 2008. Facilitating dike intrusions into ring-faults. Dev. Volcanol. 10, 351–374.
Arc, Italy. Geol. Soc. Spec. Pub. 110, 55–63. Walter, T.R., Schmincke, H.-U., 2002. Rifting, recurrent landsliding and Miocene structural
Tibaldi, A., 2001. Multiple sector collapses at Stromboli volcano, Italy: how they work. reorganization on NW-Tenerife (Canary Islands). Int. J. Earth Sci. 91, 615–628.
Bull. Volcanol. 63 (2/3), 112–125. Walter, T.R., Troll, V.R., 2003. Experiments on rift zone evolution in unstable volcanic ed-
Tibaldi, A., 2003. Influence of volcanic cone morphology on dykes, Stromboli, Italy. ifices. J. Volcanol. Geotherm. Res. 127, 107–120.
J. Volcanol. Geotherm. Res. 126, 79–95. Walter, T.R., Klügel, A., Münn, S., 2006. Gravitational spreading and formation of new rift
Tibaldi, A., 2004. Major changes in volcano behaviour after a sector collapse: insights from zones on overlapping volcanoes. Terra Nova 18, 26–33.
Stromboli, Italy. Terra Nova 16, 2–8. Warpinski, H.R., 1985. Measurement of width and pressure in a propagating hydraulic
Tibaldi, A., 2008. Contractional tectonics and magma paths in volcanoes. J. Volcanol. fracture. J. Soc. Petrol. Eng. 46–54.
Geotherm. Res. 176, 291–301. Watanabe, T., Koyaguchi, T., Seno, T., 1999. Tectonic stress controls on ascent and em-
Tibaldi, A., Groppelli, G., 2002. Volcano-tectonic activity along structures of the unstable placement of magmas. J. Volcanol. Geotherm. Res. 91, 65–78.
NE flank of Mt. Etna (Italy) and their possible origin. J. Volcanol. Geotherm. Res. Watson, J.V., 1984. The ending of the Caledonian orogeny in Scotland. J. Geol. Soc. Lond.
115 (3), 277–302. 141, 193–214.
Tibaldi, A., Pasquarè, F.A., 2008. A new mode of inner volcano growth: the “flower intru- Wauthier, C., Cayol, V., Poland, M., Kervyn, F., d'Oreye, N., Hooper, A., Samsonov, S.,
sive structure”. Earth Planet. Sci. Lett. 271, 202–208. Tiampo, K., Smets, B., 2013. Nyamulagira's magma plumbing system inferred from
Tibaldi, A., Romero, Leon J., 2000. Morphometry of Late Pleistocene–Holocene faulting in 15 years of InSAR. Geol. Soc. Lond., Spec. Publ. 380 (1), 39–65.
the southern Andes of Colombia and volcano-tectonic relationships. Tectonics 19 (2), Weertman, J., 1971. Theory of water-filled crevasses in glaciers applied to vertical magma
358–377. transport beneath oceanic ridges. J. Geophys. Res. 76, 1171–1183.
Tibaldi, A., Vezzoli, L., 1998. The space problem of caldera resurgence: an example from Weertman, J., 1980. The stopping of a rising, liquid-filled crack in the Earth's crust by a
Ischia Island, Italy. Geogr. Rundsch. 87, 53–66. freely slipping horizontal joint. J. Geophys. Res. 85 (B2), 967–976.
Tibaldi, A., Lagmay, A.M.F., Ponomareva, V.V., 2005. Effects of basement structural and Weinberg, R.F., Sial, A.N., Mariano, G., 2004. Close spatial relationship between plutons
stratigraphic heritages on volcano behaviour and implications for human activities and shear zones. Geology 32, 377–380.
(the UNESCO/IUGS/IGCP project 455). Episodes. J. Int. Geosci. 28 (3), 1–13. Wenzel, F., Brun, J.P., 1991. A deep reflection seismic line across the Northern Rhine
Tibaldi, A., Corazzato, C., Kozhurin, A., Lagmay, A.F.M., Pasquaré, F.A., Ponomareva, V., Rust, Graben. Earth Planet. Sci. Lett. 104 (2–4), 140–150.
D., Tormey, D., Vezzoli, L., 2008a. Influence of substrate tectonic heritage on the evo- Westerman, D.S., Rocchi, S., Dini, A., Farina, F., Roni, E., 2015. Rise and fall of a multi-sheet
lution of volcanoes: predicting sites of flank eruptions, lateral collapses, and erosion. intrusive complex, Elba Island, Italy. Advs in Volcanology. Springer http://dx.doi.org/
Glob. Planet. Chang. 61, 151–174. 10.1007/11157_2014_5.
Tibaldi, A., Vezzoli, L., Pasquarè, F.A., Rust, D., 2008b. Strike-slip fault tectonics and the em- White, A.J.R., Chappell, B.W., 1977. Ultrametamorphism and granitoid genesis.
placement of sheet–laccolith systems: the Thverfell case study (SW Iceland). J. Struct. Tectonophysics 43, 7–22.
Geol. 30, 274–290. White, R.S., Smith, L.K., Roberts, A.W., Christie, P.A.F., Kusznir, N.J., Roberts, A.M., Healy, D.,
Tibaldi, A., Corazzato, C., Apuani, T., Pasquaré, F.A., Vezzoli, L., 2008c. Geological–structural Spitzer, R., Chappell, A., Eccles, J.D., Fletcher, R., Hurst, N., Lunnon, Z., Parkin, C.J.,
framework of Stromboli Volcano, past collapses, and the possible influence on the Tymms, V.J., 2008. Lower-crustal intrusion on the North Atlantic continental margin.
events of the 2002–03 crisis. In: Calvari, S., Inguaggiato, S., Puglisi, G., Ripepe, M., Nature 452 (7186), 460–464. http://dx.doi.org/10.1038/nature06687.
Rosi, M. (Eds.), The Stromboli Volcano: An Integrated Study of the 2002–2003 Erup- Whitehead, J.A., Luther, D.S., 1975. Dynamics of laboratory diapir and plume models.
tion. Geophysical Monograph Series 182. AGU. ISBN: 978-0-87590-447-4, pp. 5–17. J. Geophys. Res. 80, 705–717.
Tibaldi, A., Corazzato, C., Gamberi, F., Marani, M., 2009. Subaerial-submarine evidence of Wickham, S.M., 1987. Segregation and emplacement of granitic magmas. J. Geol. Soc.
structures feeding magma to Stromboli Volcano, Italy, and relations with edifice Lond. 144, 281–297.
flank failure and creep. Tectonophysics 469, 112–136. Wicks, C.W., Thatcher, W., Dzurisin, D., Svarc, J., 2006. Uplift, thermal unrest and magma
Tibaldi, A., Pasquarè, F.A., Tormey, D., 2010. Volcanism in reverse and strike-slip fault set- intrusion at Yellowstone caldera. Nature 440. http://dx.doi.org/10.1038/nature04507.
tings. In: Cloetingh, S., Negendank, J. (Eds.), New Frontiers in Integrated Solid Wiebe, R.A., Collins, W.J., 1998. Depositional features and stratigraphic sections in granitic
Earth Sciences. Springer-Verlag, pp. 315–348 http://dx.doi.org/10.1007/978-90-481- plutons: implications for the emplacement and crystallization of granitic magma.
2737-5. J. Struct. Geol. 20, 1273–1289.
Tibaldi, A., Pasquarè, A.F., Rust, D., 2011. New insights into the cone sheet structure of the Wigand, M., Schmitt, A.K., Trumbull, R.B., Villa, I.M., Emmermann, R., 2004. Short-lived
Cuillin Complex, Isle of Skye, Scotland. J. Geol. Soc. 168, 689–704. magmatic activity in an anorogenic subvolcanic complex: 40Ar/39Ar and ion micro-
Tibaldi, A., Bonali, F.L., Pasquaré, F.A., Rust, D., Cavallo, A., D'Urso, A., 2013. Structure of re- probe U–Pb zircon dating of the Erongo, Damaraland, Namibia. J. Volcanol. Geotherm.
gional dykes and local cone sheets in the Midhyrna–Lysuskard area, Snaefellsnes Res. 130 (3–4), 285–305.
Peninsula (NW Iceland). Bull. Volcanol. 75, 764. http://dx.doi.org/10.1007/s00445- Williams, S.N., Stoiber, R.E., 1983. “Masaya-type caldera” redefined as the mafic analogue
013-0764-8. of “Krakatau-type caldera”. EOS Trans. Am. Geophys. Union 64 (45), 877.
Tibaldi, A., Bonali, F.L., Corazzato, C., 2014. The diverging volcanic rift system. Tectonophysics Wright, T.L., 1971. Chemistry of Kilauea and Mauna Loa lava in space and time. U.S. Geol.
611, 94–113. Surv. Prof. Pap. 735, 1–40.
A. Tibaldi / Journal of Volcanology and Geothermal Research 298 (2015) 85–135 135

Wright, T.J., Ebinger, C., Biggs, J., Ayele, A., Yirgu, G., Keir, D., Stork, A., 2006. Magma- Zellmer, G.F., 2009. Petrogenesis of Sr-rich adakitic rocks at volcanic arcs: insights from
maintained rift segmentation at continental rupture in the 2005 Afar dyking episode. global variations of eruptive style with plate convergence rates and surface heat
Nature 442, 291–294. http://dx.doi.org/10.1038/nature04978. flux. J. Geol. Soc. 166 (4), 725–734.
Wright, H.M.N., Bacon, C.R., Vazquez, J.A., Sisson, T.W., 2012. Sixty thousand years of mag- Zellmer, G.F., Annen, C., 2008. An introduction to magma dynamics. In: Annen, C., Zellmer,
matic volatile history before the caldera-forming eruption of Mount Mazama, Crater G.F. (Eds.), Dynamics of Crustal Magma Transfer, Storage and Differentiation. Special
Lake, Oregon. Contrib. Mineral. Petrol. 164, 1027–1052. Publications 304. Geological Society, London, pp. 1–13.
Wylie, J.J., Helfrich, K.R., Dade, B., Lister, J.R., Philips, K., 1999. Flow localization in fissure Zenzri, H., Keer, L.M., 2001. Mechanical analyses of the emplacement of laccoliths and
eruptions. Bull. Volcanol. 60, 432–440. http://dx.doi.org/10.1007/s004450050243. lopoliths. J. Geophys. Res. Solid Earth (1978–2012) 106 (B7), 13781–13792.
Yamaoka, K., Kawamura, M., Kimata, F., Fujii, N., Kudo, T., 2005. Dike intrusion associated Zhao, C., Hobbs, B.E., Ord, A., Peng, S., 2008. Particle simulation of spontaneous crack gen-
with the 2000 eruption of Miyakejima Volcano, Japan. Bull. Volcanol. 67, 231–242. eration associated with the laccolithic type of magma intrusion processes. Int.
Yew, C.H., 1997. Mechanics of Hydraulic Fracturing. Gulf Publishing, Houston TX. J. Numer. Methods Eng. 75 (10), 1172–1193.
Zak, J., Paterson, S.R., 2006. Roof and walls of the Red Mountain Creek pluton, eastern Ziv, A., Rubin, A.M., Agnon, A., 2000. Stability of dyke intrusion along preexisting fractures.
Sierra Nevada, California (USA): implications for process zones during pluton em- J. Geophys. Res. 105 (B3), 5940–5961.
placement. J. Struct. Geol. 28, 575–587.
Zellmer, G.F., 2008. Some first-order observations on magma transfer from mantle wedge
to upper crust at volcanic arcs. Geol. Soc. Lond., Spec. Publ. 304 (1), 15–31.

Вам также может понравиться