THERMO-MECHANICAL EFFECTS OF
MAGMA CHAMBERS AND CALDERA FAULTS
A thesis submitted for the degree of
Doctor of Philosophy at the
University of London
John Browning
Rock Fractures and Fluid Flow research group
Department of Earth Sciences
Royal Holloway, University of London
December 2015
Declaration
I, John Browning, hereby declare that this thesis and the work presented in it is entirely
my own, unless otherwise stated. The main body of thesis, Chapters 3-6, forms a
collection of co-authored papers which are either published, in press or in prep for
publication. I am responsible for the data collection throughout this thesis, for primary
authorship of all four of the included papers. Statements of my contribution are included
on the cover page of each paper.
Signed
Dated
II
Abstract
Shallow magma chambers influence a range of crustal processes at active volcanoes.
For example many and perhaps most dyke fed eruptions originate from shallow magma
storage regions. Magma chamber failure, resulting in the initiation of caldera faults or
magma filled fractures (dyke) is likely governed by a complex interplay between
regional and local mechanical stresses and thermal effects. This study utilises a
multitude of techniques to decipher salient thermo-mechanical processes occurring as a
result of magma stored at shallow (<10 km) depth. A new model to forecast magma
chamber rupture and dyke initiation is proposed. The analytical solutions presented are
applied to field data from Santorini, Greece, and combine poro-elastic material
constraints with geodetic data to estimate both magma volumes stored beneath the
caldera and internal excess pressure generated during periods of magmatic recharge.
Predicting the path or propagation of magma once it has left a shallow magma
chamber is an important but so far unachievable goal in volcanology. Caldera faults
offer pathways for magma and often develop ring-dykes. A previously unreported
mechanism for the formation of ring-dykes is the capture of inclined sheets at caldera
fault boundaries. Geological field observations of an exceptionally well-exposed ringfault at Hafnarfjall in Western Iceland were used as input to the finite element method
numerical modelling software COMSOL multiphysics to infer a mechanism of principal
stress rotation within a fault damage zone. The same modelling technique was then used
to estimate the far-field crustal displacements resulting from the failure and collapse of
a shallow magma chamber roof. This study is framed within the context of the 2014-15
Bardarbunga-Holuhraun (Iceland) dyke injection and eruptive episode, and
hypothesises that significant ice subsidence was not solely associated with crustal
subsidence but instead related to ice flow within Bardarbunga caldera generated by the
dyke emplacement. Thermal stresses resulting from hot magma emplacement and
gradual cooling likely combine to weaken volcanic edifices. For example, field evidence
suggests many normal faults nucleate from cooling joints. A suite of thermal stressing
experiments finds that cooling and contraction produces larger and more abundant
micro-cracks when compared with heating expansion. This is an important result when
considering that almost all previous studies concerned with thermal stressing focused
on the heating cycle.
III
Acknowledgements
First and foremost, my thanks go to Agust Gudmundsson and Philip Meredith as
without their support and guidance this thesis would not have been possible. I thank
Agust for introducing me to the world of volcano-tectonics, and allowing me substantial
academic freedom such that I was able to develop a project and research framework that
matched my interests. I thank Phil for agreeing to be a part of this project, and by doing
so providing an unrivalled enthusiasm for experimental studies that has definitely
caught on. Thank you both for the time devoted to helping me through this process.
Thanks to the examiners Philip Benson and Valerio Acocella who contributed to
improving the final thesis and providing interesting insights and ideas during my VIVA.
Large components of this project involved fieldwork and for assistance in Iceland I
thank Zoe Barnett (RHUL) and Hannah Reynolds (University of Iceland), and for
making Greece my second home I thank Sandy Drymoni (RHUL). Much of this work
benefitted from collaboration and discussions with numerous academics, of whom there
are far too many to mention here but special thanks are owed to Yan Lavallée, Felix
Von Aulock (University of Liverpool), Hugh Tuffen (Lancaster University), Tom
Mitchell and Rick Wall (UCL). For assistance with designing and building our
experimental apparatus I thank Neil Hughes, and I am indebted for almost constant
assistance from Steve Boon. Jim Davey deserves a special mention for assistance with
thin sections and SEM. I benefited greatly from attending several training courses and
events organised by the EU network NEMOH in the early stages of my PhD. I am very
happy to be considered an ‘honorary NEMOH’, and I am eternally grateful to have met
many wonderful colleagues and friends from around the world during NEMOH events.
Thanks to those who read, and re-read various sections of my thesis, especially
Nathaniel Forbes-Inskip, Mohsen Bazargan, Jon Pownall, Ozgur Karaoglu and Hannah
Reynolds (again). Thanks to my friends and family for supporting me through the last
three years (and more). Thanks to Mike Murphy, all of my office workmates at RHUL
and UCL and the residents of 61 St Judes Road. I feel at this point that I must apologise
to my band for disappearing for large periods of time, normally when away on
fieldwork, not always! In the time of completing this PhD we also finished a new record
and for that I am equally proud. Finally, huge thanks to my closest friend Prof. Resheed
for offering support, encouragement, ambition and distraction, always.
IV
V
Contents
Abstract .......................................................................................................................... III
Acknowledgements ........................................................................................................ IV
Contents .......................................................................................................................... V
List of figures ............................................................................................................... VII
Chapter 1: Introduction ..................................................................................................... 1
1.2
Magma chambers ............................................................................................... 1
1.3
Dykes and inclined sheets .................................................................................. 2
1.4
Calderas and caldera faults ................................................................................. 3
1.5
Ring dykes .......................................................................................................... 5
1.6
Field sites ............................................................................................................ 6
1.6.1 Iceland .............................................................................................................. 6
1.6.2 Santorini ........................................................................................................... 9
1.6.3 Tenerife .......................................................................................................... 11
1.7
Contributions .................................................................................................... 13
1.8
Thesis roadmap ................................................................................................. 14
Chapter 2: Methodologies ............................................................................................... 17
2.1
Numerical modelling methodology..................................................................... 17
2.2
Experimental methodology .................................................................................. 27
2.2.2
Introduction ............................................................................................... 27
2.2.3
Sample selection and preparation ............................................................. 28
2.2.4
Sample characterisation ............................................................................ 29
2.2.5
Thermo-mechanical measurements ........................................................... 33
2.2.6
Experimental apparatus ............................................................................. 36
2.2.7
Temperature calibration setup ................................................................... 40
2.2.8
Thermal gradients and timescales of thermal equilibration ...................... 44
2.2.9
Acoustic emission (AE) ............................................................................ 47
2.2.10
Jig generated AE ....................................................................................... 48
2.2.11
Wave-guide and sensor modifications ...................................................... 50
VI
2.3
Field methods ...................................................................................................... 52
Chapter 3 Forecasting magma-chamber rupture ............................................................. 53
Chapter 4 An alternative mechanism of ring-dike formation ......................................... 54
Chapter 5 Cooling dominated cracking in volcanic rocks .............................................. 55
Chapter 6 Surface displacements resulting from magma-chamber roof collapse ........... 56
Chapter 7: Discussion, critical evaluation and future directions..................................... 58
7.2 Forecasting worldwide magma chamber failure conditions ............................. 58
7.3 Towards a characterisation of caldera fault damage zones ............................... 60
7.4 Assessing the influence of thermo-mechanical damage accumulation ............. 63
7.5 What is a ‘realistic’ magma chamber host rock rheology? ............................... 64
7.6 Towards a method of quantifying crack annealing in volcanic rocks ............... 64
7.7 The ‘gravity problem’ ....................................................................................... 65
7.9 The unknowns of magma volume expansion and compressibility ................... 66
7.10 The ‘state-of-the-art’ ....................................................................................... 67
References ...................................................................................................................... 69
Appendices ..................................................................................................................... 84
List of figures
Chapter 1 Introduction
1
Figure 1.1: Traditional model for the formation of ring-dykes........................................ 6
Figure 1.2: Active volcanoes of Iceland .......................................................................... 8
Figure 1.3: Tectonic setting of Santorini in the Aegean arc .......................................... 11
Figure 1.4: Geologic map of the Anaga and Teno massifs of Tenerife ......................... 12
VII
Chapter 2 Methodologies
17
Figure 2.1: Components of stress in two and three dimensions. ................................... 19
Figure 2.2: Typical COMSOL numerical model set-up ................................................ 21
Figure 2.3: Typical triangular mesh setup in an asymmetric model type. ..................... 22
Figure 2.4: Typical COMSOL model result output. ...................................................... 24
Figure 2.5: Results of a Discrete Element Method (DEM) ........................................... 26
Figure 2.6: Apparatus layout for benchtop seismic velocity measurements. ................ 31
Figure 2.7: Schematic of a helium pycnometer setup .................................................... 32
Figure 2.8: Simplified schematic of the Netzsch Hyperion 402. ................................... 34
Figure 2.9: Hot-stage microscope setup. ........................................................................ 36
Figure 2.10: Thermal stressing test experimental setup................................................. 38
Figure 2.11: Thermal stressing test jig design and assembly ......................................... 39
Figure 2.12: Basalt core modified. ................................................................................. 41
Figure 2.13: Thermal jig setup for acoustic emissions .................................................. 41
Figure 2.14: Programmed and actual heating and cooling profiles ............................... 42
Figure 2.15: Outer sample temp against temperature surface temp ............................... 43
Figure 2.16: Natural sample cooling rate. ...................................................................... 44
Figure 2.17: Temperatures at the sample and furnace ................................................... 45
Figure 2.18: Heating and cooling rate of a basalt sample .............................................. 46
Figure 2.19: Individual AE hits produces a seismic coda .............................................. 47
Figure 2.20: Acoustic emissions generated from the steel jig ....................................... 49
Figure 2.21: Acoustic emissions generated from a basalt sample ................................. 50
Figure 2.22: Thermal stressing test setup for in-situ P-wave velocity measurements ... 51
Figure 2.23: (a) Original method of attaching the P-wave transducer ........................... 51
Chapter 3 Forecasting magma chamber rupture
VIII
53
Figure 1: Simplified geological map of Santorini
Figure 2: Orientation and thickness of dykes at Santorini
Figure 3: Model of Santorini volcanic system and pressure dynamics
Chapter 4 An alternative mechanism of ring-dike formation
54
Figure 1: Geological map of Hafnarfjall central volcano
Figure 2: Sheets and dikes at Hafnarfjall
Figure 3: Caldera ring fault outcrop
Figure 4: Dikes occupying the ring-fault
Figure 5: Conditions for fracture propagation
Figure 6: COMSOL model setup
Figure 7: Model results indicating sheet propagation through the fault
Figure 8: Model results indicating sheet deflection or arrest
Figure 9: Model results indicate principal stress rotation
Figure 10: Model results showing a stiff central fault layer
Figure 11: Model results showing effects of various boundary condition
Figure 12: Background stresses
Figure 13: Conceptual resurgent caldera model
Chapter 5 Cooling induced cracking in volcanic rocks
Figure 1: Cooling joint examples from the field
Figure 2: Schematic of thermal stresses
Figure 3: Optical light microscope images
Figure 4: Thermo-mechanical analysis
IX
55
Figure 5: Thermal strain and resultant material softening
Figure 6: Schematic diagram of experimental arrangement
Figure 7: Programmed and actual heating and cooling rates
Figure 8: Standard AEs dataset
Figure 9: AEs hit energy
Figure 10: NKD AEs energy and events for a ~1 minute hold period
Figure 11: NKD AEs energy and events for a ~30 minute hold period
Figure 12: NKD AEs energy and events for a ~2 hour high T period
Figure 13: Radial P-wave velocities
Figure 14: SEM images and crack analysis for IB sample
Figure 15: Comparison of crack frequency in SEM
Figure 16: AE energy and seismic b values
Figure 17: Macro-fractures in IB
Figure 18: Tensile thermal stress as a function of temperature
Figure 19: Kaiser ‘temperature memory’ effect test results
Chapter 6 Surface displacements from magma chamber roof subsidence
Figure 1: Simplified geological map of Iceland
Figure 2: Exposed sections of extinct volcanoes in Iceland
Figure 3: Sub-glacial caldera occupied by a caldera lake
Figure 4: Stress fields favouring caldera ring-fault formation
Figure 5: Sketch of model setup
Figure 6: ‘Mogi’ model theoretical results
Figure 7: Crustal surface and ice-surface displacements from underpressure
Figure 8: Vertical and horizontal displacement in homogeneous crust
Figure 9: Effect of deep-seated reservoir doming
X
56
Figure 10: Displacements from a model including a weak fault
Figure 11: Displacements in a heterogeneous crust
Figure 12: Displacement along a vertical fault
Figure 13: Displacements resulting from differential roof subsidence
Figure 14: Profile of sub-glacial caldera
Chapter 7 Discussion, critical evaluation and future work
58
Figure 7.1: Models of relationship between damage and core zones ............................ 62
Figure 7.2: Schematic diagram of thesis contributions .................................................. 62
List of tables
Chapter 5 Cooling induced cracking in volcanic rocks
55
Table 1: Whole rock XRF results
Table 2: AE hit and energy totals and rates
Table 3: P-wave velocities
Chapter 6 Surface displacements from magma chamber failure
Table 1: Model parameters
Table 2: Selected model results
XI
56
Chapter 1: Introduction
Chapter 1: Introduction
Eruptions from caldera volcanoes represent one of only a few potential processes for
major (natural) extinction events. As such it is vital to have as complete an
understanding as possible, of the processes governing mechanisms of magma transport
and storage in and around caldera settings. This thesis investigates various volcanotectonic processes with the aim of increasing our understanding of caldera volcanoes. In
this first chapter I provide a brief overview and introduction to the relevant fields of
study and suggest areas that require further or better understanding.
1.2 Magma chambers
Magma stalls and accumulates in the crust at various depths below the ground surface
(Marsh, 1989). Our predominant understanding of magma storage conditions, in terms
of storage depths and host rock interactions, come from structural field studies (e.g
Gudmundsson, 1986a; Menand, 2011), petrological analyses of fluid inclusions in
crystals (e.g Druitt et al., 2012; Walker, 1960), statistical studies of ground deformation
(e.g Mogi, 1958) and analyses of magma derived gases (e.g Stevenson and Blake,
1998). Mechanisms controlling the formation of magma chambers are still widely
debated (Menand, 2011), but it seems clear that many chambers form from the
accumulation of sills (Barnett and Gudmundsson, 2014; Gudmundsson, 2011a). It is
apparent from geodetic and field studies that most magma chambers form an
approximate sill like geometry (Burchardt and Gudmundsson, 2009; Burchardt et al.,
2013; Camitz et al., 1995; Gudmundsson, 2012; Yun et al., 2006), although other
geometries such as spheres and ellipsoids are plausible (Gudmundsson, 1986a; Hickey
et al., 2013; Lipman, 1997; Menand, 2011; Parks et al., 2012). In fact magma chambers
are most commonly approximated as spheres, or as small points of pressure, when
modelling volcano deformation (Anderson, 1937; Mogi, 1958; Parks et al., 2012;
Sturkell et al., 2006b). Such models are used to provide 1st order estimates of magma
chamber pressure and volume changes as well as constrain depth ranges, although the
model assumptions are inherently erroneous (Manconi et al., 2007; Masterlark, 2007).
These crude approximations are considered valid partly because the crust acts primarily
1
Chapter 1: Introduction
as a linear elastic material and therefore any significant pressure changes cause the
surrounding rocks to become stressed and undergo strain which is then represented
instantaneously as ground deformation at the surface (Segall, 2013). As well as
inducing stresses in the crust, magma chambers commonly intrude into cooler host rock
and therefore the magma and host rock can undergo many cycles of thermal stressing.
Some authors challenge the notion that magma chambers exist as significant volume
reservoirs, and instead favour a more complex model of packages of melt (Cashman and
Giordano, 2014). However, in order to form large sustained eruptions and potentially
collapse calderas, magma must be present at shallow depth and in sufficient volume
(Gudmundsson, 2012). The precise pressure and stress conditions required to form
eruptions are less well known, and these may be significantly influenced by thermal
degradation (Annen, 2011; Gregg et al., 2012) and mechanical heterogeneity
(Gudmundsson and Philipp, 2006) in crustal host rocks.
1.3 Dykes and inclined sheets
A magma chamber can fail in one of several ways, namely through dyke or inclined
sheet injection, gravity driven roof collapse, or a combination of the two producing
ring-dykes. Dykes, or ‘dikes’ (US spelling), are magma driven fractures which
propagate in a direction perpendicular to the minimum compressive (maximum tensile)
principal stress σ3 (Figure 1). The shape of an underlying shallow magma chamber and
the regional tectonic controls (boundary conditions) dictate the orientation of stresses in
shallow volcanic systems. As a consequence magma, as dykes, does not propagate
vertically in many cases, instead favouring an inclined trajectory toward the surface
(Burchardt et al., 2011; Gautneb and Gudmundsson, 1992), Figure 1.2. Dykes that dip
less than around 75 are commonly referred to as inclined sheets. It is possible in some
cases to estimate the location of a shallow magma chamber based on the attitude of a
collection of inclined sheets (e.g Burchardt et al., 2013; Gautneb and Gudmundsson,
1992), assuming that the direction of magma propagation does not change significantly
away from the magma chamber. Whilst statistically this may be a valid treatment, it has
been shown that stress fields are complex ahead of dyke tips (Gudmundsson and Phillip,
2006). As a consequence of these complex stress fields, which tend to arrest dykes,
most dykes recorded geodetically or in the field are non-feeders (Geshi et al., 2010;
2
Chapter 1: Introduction
Gudmundsson, 2011a; Hooper et al., 2011; Pollard, 2010), i.e. they stall inside the
volcano to form intrusions rather than provided magma for an eruption (Gudmundsson
and Philipp, 2006). The complexity of volcanic stratigraphy and tectonic relations
probably affects dyke propagation paths to a greater degree than generally recognised.
Therefore, new models or findings that provide insights into magma propagation in
various settings are useful and important to the volcanological community.
1.4 Calderas and caldera faults
Calderas are typically bounded by circumferential faults, or ring-faults. In piston and
trapdoor type calderas, ring-faults, or part ring-faults, represent the calderas structural
boundary. It is not possible to directly observe the structural boundary of many
calderas; the faults are often obscured or eroded by subsequent water emplacement
(caldera lakes) or mass wasting (landslides) (Lipman, 1997). The boundary observed at
such calderas is a topographic boundary, which is largely related to the structural
boundary although the main subsidence accommodating faults may be situated several
tens (Hartley and Thordarson, 2012) to hundreds of meters (Wilson et al., 1994) away
from the topographic collar. Therefore, caldera fault geometry must be largely inferred
from geophysical (e.g Saunders, 2001), analogue (e.g Acocella, 2007) and numerical
(e.g Gray and Monaghan, 2004) studies. In rare circumstances it is possible to observe
caldera faults in geologic outcrops at deeply eroded volcanoes. In fact much of our early
understanding of caldera processes comes from studies of ancient and eroded systems in
the USA (Lipman, 1984), Iceland (Gudmundsson, 1987a), and Scotland (Branney,
1995).
The mechanics governing caldera ring-fault formation are reasonably well studied,
although studies considering the effects of pre-existing faults on caldera dynamics are
less numerous. A combination of geodetic observations, analogue and numerical models
and field work have shown that calderas are formed from a combination of inward
(normal) and outward (reverse) dipping faults. For example, the 650 m deep and 2.5 km
wide caldera formed during the 1991 silicic eruption of Pinatubo was shown to subside
along predominantly outward dipping ring-faults inferred predominantly from
seismicity (Mori et al., 1996). Similar outward dipping structures have also been
3
Chapter 1: Introduction
inferred from seismicity recorded at Rabaul caldera (Mori and McKee, 1987) and Mt St
Helens (Scandone and Malone., 1985), although normal faults
seem to control
subsidence at Bardarbunga caldera in Iceland (Fichtner and Tkalčić, 2010). Most
analogue models of caldera formation suggest that collapse occurs predominantly along
reverse faults, with a contribution from peripheral normal faults (see the review by
Acocella, 2007). Whereas numerical modelling seems to suggest both inward
(Gudmundsson and Nilsen, 2006) and outward (Holohan et al., 2015) dipping faults,
depending on the modelling technique implemented (Geyer and Marti, 2014). There
have been several studies investigating or observing the influence of caldera faults on
deformation patterns at active calderas (Bathke et al., 2015; De Natale and Pingue,
1993; Hutnak et al., 2009; Masterlark, 2007), but relatively few that investigate how
magma transport is effected by caldera structures (Jónsson, 2009; Saunders, 2004,
2001).
There have been only four, possibly five events involving caldera formation or slip on
caldera faults in the past century. Those events occurred in 1968 at Fernandina in the
Galapagos (Filson et al., 1973), in 1991 during the ignimbrite forming eruption of
Pinatubo (Mori et al., 1996), in 2000 at the Miyakejima volcano in Japan (Geshi et al.,
2002), in 2007 at Piton de la Fournaise at La Reunion (Michon et al., 2009; Peltier et
al., 2009), and possibly in 2014 at Bardarbunga in Iceland (Sigmundsson et al., 2014).
Apart from large volume eruptions in Iceland and the Philippines, most of the calderas
formed in association with moderate to small volume eruptions. The two best monitored
collapses in terms of geodetic data were Bardarbunga and Piton de la Fournaise,
although ideas regarding the mechanism of caldera deformation at both remain
controversial. It is still not clear if a caldera actually slipped during the 2014
Bardarbunga event, visual inspection is limited because the volcano lies beneath
approximately 800 m of glacial ice. The best evidence for collapse comes from the
location and type of earthquakes surrounding the volcano (the inferred ring-fault), and
over 60 m of ice surface subsidence (Sigmundsson et al., 2014) directly above the
caldera. All of the aforementioned collapse events were monitored by various
combinations of seismometers and GPS, but only Bardarbunga offers data on the
ground deformation effects of a caldera collapse from a great distance (>10 km) away
the centre of collapse, although the situation was made more complex by the intrusion
of at least one major dyke. All of the other caldera forming collapses have occurred on
4
Chapter 1: Introduction
small islands where the coverage of GPS is limited predominantly to the summit region,
and while this is not the case at Pinatubo there are no published GPS data in the farfield. As such there remains a substantial gap in our knowledge of expected ground
surface deformation signals associated with collapse caldera formation.
1.5 Ring dykes
An important process concerning the timing and dynamics of caldera fault formation is
the emplacement of ring-dykes. Anderson’s (1937) model is the most widely used to
explain the occurrence of ring-dykes (Figure 1.1). In this model, ring-dykes (and ringfaults) form as a consequence of a pressure decrease in a magma reservoir, the
fundamental basis for the under-pressure or withdrawal of magmatic support model for
the formation of collapse calderas. Gravity driven failure of the magma chamber
follows magma withdrawal, with the general principle being that dykes are then
squeezed out of the chamber through shear (Mode II-III) ring fractures, which appears
to be supported by field evidence in some cases (e.g Kokelaar, 2007). As is well known
however, dykes are predominantly mode I fractures which require a pressure in excess
of the tensile strength of the host rock to propagate. The simple Anderson (1936) model
begs the question: how does a dyke remain open and supply magma out of a chamber
which contains magma at a pressure less than lithostatic (i.e Gudmundsson and Nilsen,
2006)?
One solution is that ring-dykes form from magma chamber overpressure
conditions, such that the caldera collapse is the cause of large eruptions rather than
being the consequence (Gudmundsson, 2015, 1998).
5
Chapter 1: Introduction
Figure 1.1: Traditional model for the formation of ring-dykes (outward-dipping, yellow
lines) based on Anderson’s (1937) analytical solution. Trajectories of maximum tensile
principal stress, σ3, as dashed white lines, and maximum compressive principal stress,
σ1, as solid red lines. Those red lines indicate the likely direction of magma propagation
in this simple example, modified after Gudmundsson and Nilsen, 2006.
1.6 Field sites
Three main field areas were chosen as sites of either data collection or sites of rock
sample collection. The following section describes volcanological and tectonic aspects
of each field area in respect to the project aims. All sites were selected based on the
suitability of outcrops and exposures of volcano-tectonic features.
1.6.1 Iceland
Many of the fundamental ideas from the field of volcano-tectonics were created by
investigating processes at active and extinct volcanoes in Iceland (Gudmundsson,
6
Chapter 1: Introduction
1986b, 1987a, 1986b; Walker, 1960). Iceland is one of the best monitored volcanic
fields in the world and home to many active caldera volcanoes, as well as many more
extinct caldera volcanoes. Utilisation of various monitoring techniques at active
volcanoes and along the divergent margin has shown that spreading occurs at a rate of
around 10 mm/yr, and that Iceland’s volcanoes deform in numerous ways
(Sigmundsson, 2006). Krafla, Askja and Hengill (shown in Figure 1.2) have for more
than 30 years been subsiding (deflating) at relatively steady rates (Miller et al., 1998;
Sturkell and Sigmundsson, 2000; Sturkell et al., 2006a, 2006b), whereas
Eyjafjallajokull, Grimsvotn and Bardarbunga have all experienced both local inflation
and regional doming with intermittent periods of rapid subsidence (Sigmundsson,
2006). The signals represent a complex interplay between magma transfer from depths
greater than 10 km to shallow storage regions (Gudmundsson, 1987b), at depths around
2 to 5 km, where this shallow magma also cools and contracts. Cooling shallow magma
chambers are inferred to be the main cause of ground surface deflation at many of
Iceland’s volcanoes (Sturkell and Sigmundsson, 2000), although crustal spreading also
likely contributes to the signal (Camitz et al., 1995; De Zeeuw-van Dalfsen et al., 2012).
The 2014-15 unrest and rifting episode at Bardarbunga caldera was one example of an
incredibly well monitored volcano-tectonic episode (Gudmundsson et al., 2014;
Sigmundsson et al., 2014). The event raised a number of questions regarding dyke
propagation, stress accumulation at central volcanoes and caldera deformation. For
example, the rifting episode appeared to confirm a long held hypothesis that dykes can
propagate laterally over a distance of more than 40 km before coming to the surface to
produce eruptions (Sigmundsson et al., 2014), an idea which originates from the
discovery of Bardarbunga signature lavas near the Torfajokull complex (McGarvie,
1984), and numerous fissure eruptions North and South of the Askja volcanic system
(Hjartardóttir et al., 2009; Key et al., 2011). Many of these eruptions were probably fed
by deeper reservoirs (Hartley and Thordarson, 2013), and this appears to be the case for
the 2014 Holuhraun eruption (Haddadi et al., 2015), rather than lateral flow of magma
from a shallow source. Whilst the dynamics of magma propagation have been very well
studied in Iceland (Gudmundsson, 2011b; Sigmundsson et al., 2014), there is still no
comprehensive model to predict where and when an eruption is likely to take place.
7
Chapter 1: Introduction
Figure 1.2: Active volcanoes of Iceland, many have calderas and some are associated
with very recent dyke intrusions at Holuhraun, Gjalp and Eyjafjallajokull (marked). All
of the volcanoes and associated calderas in Iceland are subject to overall regional tensile
stress, exhibited as plate divergence at a rate of approximately 9.7 mm/yr, modified
after (Sturkell et al., 2006a).
The ancient and eroded, ‘Tertiary’ volcanics of Eastern and Western Iceland offer
insights into the dynamics of magma plumbing systems and processes occurring at
currently active volcanoes. Many of the eroded central volcanoes host dyke swarms
(Walker, 1974), for example Geitafell in the East (Burchardt and Gudmundsson, 2009)
and radial and concentric dykes such as at Hafnarfjall (Gautneb et al., 1989). In some
locations the extent of erosion is such that it is possible to observe the tops of plutons
and extinct magma chambers (Gudmundsson, 2012), perhaps the best exposed of such
sections is the Slaufrudalur pluton and central volcano, which hosts a granophyre body
extending roughly 8 km in length and 2 km in diameter (Burchardt et al., 2010). At the
top of the granophyre much of the host rock into which it was emplaced still exists. At
these localities it is possible to observe the sharp contact between the once felsic magma
and mafic host. A number of dykes can also be seen to cut through the roof of the
chamber, indicating that it probably acted as a magma supply for eruptions. Outside of
8
Chapter 1: Introduction
central and shield volcanoes in ‘Tertiary volcanics’ the attitude and dimensions of dykes
change such that they become generally thicker and steeper (Gudmundsson, 1990;
Gautneb and Gudmundsson, 1992), these type of dykes are referred to as regional
dykes. There exists a clear relationship between the occurrence of dyke swarms (and
inclined sheets) and magma chambers and the occurrence of regional dykes generally
away from shallow magma centres. This finding suggests that dyke sheets and swarms
are controlled by local stress fields generated by shallow magma bodies, whereas
regional stress fields associated with divergent plate movements and deeper and larger
magma reservoirs control the emplacement of regional dykes (Gudmundsson, 1983;
Gudmundsson et al., 2014).
1.6.2 Santorini
Santorini is a very well-studied caldera volcano located in the southern Aegean of
Greece. Arc related volcanism in this region began in the early Pliocene (Fytikas et al.,
1984), forming the islands of Santorini, Nisyros, Milos, Kos, Aegina, Methana and
Poros. All volcanism exhibits a strong tectonic control, related to NE-SW lithospheric
normal-fault rupture zones (Papazachos and Panagiotopoulos, 1993) (Figure 1.3).
Therefore whilst volcanism is clearly a consequence of subduction related processes, the
volcanic arc has formed in an overall extensional stress regime. Magmas in the Aegean
arc are commonly andesite to dacite, only Santorini produces abundant basaltic
successions (Druitt et al., 1999; Fytikas et al., 1984). Santorini is a stratigraphically and
chemically complex stratovolcano, made up of many hundreds to thousands of distinct
lava and ignimbrite successions (Druitt et al., 1999). Several of the ignimbrites are
associated with catastrophic caldera formation (Druitt and Francaviglia, 1992; Druitt
and Sparks, 1982; Roche and Druitt, 2001). There have been four well documented
caldera forming events in the geologic record (Druitt et al., 1999). Structurally, this
means that the present day volcano contains abundant ring-fracture and fault systems
(Giannopoulos et al., 2015; Heiken and Mccoy, 1984). As well as the caldera faults,
arguably the two most important volcano-tectonic features are the Coloumbo and
Kameni lines which constitute a series of sub-parallel NE-SW trending normal faults
which run through the northern and central sectors of the present day caldera (Druitt et
9
Chapter 1: Introduction
al., 1999; Konstantinou et al., 2013; Papoutsis et al., 2013). Associated with the
Coloumbo line in the northern part of the caldera is a 1-2 km thick dyke swarm made up
of around 60 individual dykes (Browning et al., in press), only around 5% of which feed
eruptions at the Megalo Vouno shield complex (Druitt et al., 1999). Present day
volcanism within the caldera is located along the Kameni line in the form of
predominantly dacitic domes and a lava bearing shield complex named the Kameni
islands (Nomikou et al., 2014). Recent geodetic studies have postulated the existence of
a growing shallow magma chamber (Parks et al., 2014) that probably feeds recent
eruptions at Nea Kameni (Druitt et al., 2012). Of greatest interest to this study are the
are the post Minoan caldera lavas which are described in detail by Barton and
Huijsmans (1986). All lava flows at Nea Kameni are dacite in composition and exhibit
glassy textures with numerous vesicles. It is estimated that the magma that supplied
these lavas is compositionally zoned and contains approximately 3-4 wt % H2O, of
which most (2.5-3.7 %) is lost during eruptive degassing (Barton and Huijsmans, 1986;
Druitt et al., 2012). Samples of the lava were collected from the margin of each flow for
use in thermo-mechanical testing (see Chapter 5). Whilst the geochemistry of individual
Nea Kameni lavas, and in fact many of Santorini’s deposits, are well studied, very little
attention has been given to the geomechanical properties of these rock units. A
combined understanding of geochemical and geomechanical properties is essential to
estimating eruption hazards at Santorini volcano. It is especially important here because
the caldera forming eruptions, whilst largely overestimated in their destructive potential
(Dominey-Howes and Minos-Minopoulos, 2004), can produce significant damage to the
islands inhabitants and tourists. Examples of the destructive power can be seen
throughout Santorini as ignimbrite covered ruins, associated with the Bronze age
demise of the Minoan civilisation (Bond and Sparks, 1976; Druitt, 2014; Pyle, 1997).
10
Chapter 1: Introduction
Figure 1.3: Tectonic setting of Santorini in the Aegean arc, after Friedrich (2009)
1.6.3 Tenerife
Whilst the volcanic island of Tenerife was not studied directly in this thesis, rock
samples were collected and used in experimental studies in Chapter 5. Therefore a brief
description of salient volcano-tectonic features and previous studies is appropriate.
Tenerife is the largest of the Canary Islands covering 2058 km2, with a roughly
triangular or ‘mercedes star’ shape (Marti et al., 1994). The present day island is
characterized by a summit caldera (Las Canades) and central edifice (Teidi), but was
originally built by the conglomeration of three volcanic shields; in the North West,
11
Chapter 1: Introduction
Teno, in the Northeast, Anaga and the South, Roque del Conde. Several distinct steep
sided rift-zone or ‘arms’ periodically become unstable and produce huge volume
landslides (Garcia-piera et al., 2000; Hu et al., 1999). Whilst many agree that the
summit caldera is a result of the vertical collapse of the Las Canadas magma chamber,
possibly over three separate cycles (Coppo et al., 2008; Marti and Gudmundsson, 2000),
others believe the caldera is a giant landslide scarp (Carlos et al., 1999; Carracedo,
1999; Watts, 1998). A block was collected from one dyke in a swarm of dykes near the
town of Taganana in the Anaga province of Tenerife. These dykes have been wellstudied structurally (Marinoni and Gudmundsson, 2000) and chemically (HernándezPacheco, 1996). Most of the dykes are altered and mafic to phonolite in composition,
they form from a minimum compressive stress field orientated NE-SW and WNW-ESE.
Figure 1.4: Geologic map of the Anaga and Teno massifs of Tenerife, showing the
approximate location of a sampled dyke, modified after Marinoni and Gudmundsson
(2000).
12
Chapter 1: Introduction
1.7 Contributions
In this thesis I use a range of techniques to study problems in volcano-tectonics, volcano
deformation and fracture mechanics with specific application to areas of Iceland and
Greece. The scientific contributions of this work can be distinguished by three
categories. First are contributions to volcano research where applied forward models are
proposed to aid the temporal forecasting of magma chamber failure and estimate ground
surface changes resulting from roof collapse. Second are contributions to ongoing
research activities at the locations mentioned previously, these are of interest to those
who study the regions or similar areas. And third are contributions to fracture
mechanics through the design of new experimental methods and applications, which are
of interest to those studying rock mechanics and commercial geothermal activities.
The research reported within this thesis is not focused on solving any one large
geophysical problem. Instead a number of smaller problems are tackled, that may
sometimes seem only loosely connected to one another. Here is a summary of the main
scientific contributions within this thesis.
1. We estimate, using poro-elastic analytical models, continuum mechanics and
fracture mechanics that the volume of shallow magma stored at Santorini
volcano is approximately 120 km3. We postulate a range of magma chamber
rupture conditions based on magma recharge from a deep seated source. Our
findings suggest that the magma chamber came close to rupture during
2011-2012 magmatic unrest.
2. We report on a previously unrecognised mechanism concerning the
propagation of dykes and inclined sheets at caldera faults. New field
observations and data collected from a well exposed inward-dipping caldera
fault suggest that magma is captured within caldera faults due to changes in
the mechanical properties of the rocks surrounding and within the fault.
3. We present a new experimental apparatus and method for detecting acoustic
emissions in material undergoing cooling and contraction from high
temperature.
13
Chapter 1: Introduction
4. The experimental design and method are used to infer that cooling
contraction produces a greater abundance of larger micro-fractures when
compared to thermal expansion related fractures. This finding is important
as most previous studies of thermal cracking concentrated only on damage
accumulation during heating.
5. We find that magma chamber roof collapse is an unlikely mechanism for the
generation of ice surface subsidence related to the Bardarbunga-Holuhraun
rifting event in 2014-15. Instead we postulate that the ice subsided as a
consequence of dyke induced stress changes within the steep caldera walls,
which is an entirely novel and unreported mechanism.
1.8 Thesis roadmap
Each chapter of this thesis can be considered as an independent study with some
chapters presented as peer-reviewed and published papers and others which are in
preparation or review for publication in scientific journals. It is possible, therefore, for
the reader to examine chapters of interest without having to lookup background material
in previous chapters, aside from Chapter 2 which contains a thorough description and
analysis of the methods throughout, and Chapter 7 which critically evaluates and
discusses the main findings as well as suggesting possible routes for future research. As
three of the four results chapters have been published or accepted for publication, they
are presented as stand-alone works with independent page numbers. Chapter 4 is
presented using the journals printed format, it is expected that three chapters in total will
be formatted by journal specifications when the final thesis is complete.
In Chapter 2 the main methodological frameworks that govern each independent study
are presented. This chapter is divided into three main sections, relating to the techniques
used throughout, namely numerical and analytical studies, field data collection and
laboratory experiments. All of the methods are critically evaluated, with suggestions for
improvements or erroneous or incomplete results discussed.
14
Chapter 1: Introduction
In Chapter 3 we present a simple analytical technique for monitoring the pressure
evolution within a shallow magma chamber during periods of magma recharge and
resultant unrest, based on fracture mechanics and continuum mechanics principles. Field
data of dykes observed and measured within the caldera wall at Santorini are presented
and used to inform the analytical model results. We apply the results to the well
documented unrest period at Santorini volcano in 2012, and suggest that future work
could seek to generate similar forecasts at other well monitored volcanoes.
In Chapter 4 the field campaign switches to Western Iceland where an exceptionally
well-exposed caldera ring-fault outcrops at the Hafnarfjall volcano. Field data is used to
constrain finite element method numerical models built using COMSOL Multiphysics.
Using a combination of those methods we present a novel interpretation for ring-dike
formation. In this work we find that inferred mechanical heterogeneities across a caldera
fault zone influence magma propagation to the extent that inclined sheets can become
deflected and captured within the fault. The model is used as an explanation for the high
frequency of volcanic eruptions and edifices around the margins of calderas.
We depart from field studies and numerical modelling in Chapter 5, to examine volcanic
rocks in the laboratory. Here we investigate the role of thermal stresses in generating
micro-cracks in a range of differentially emplaced igneous rocks. A new apparatus is
used for detecting in-situ micro crack development during both heating and cooling by
recording acoustic emissions. Contemporaneous data acquisition is complimented by a
suite of micro-crack analyses under SEM and P-wave velocity determination. All rock
types undergo thermo-mechanical analysis to determine their respective mechanical
response to high temperature treatment.
Chapter 6 presents a series of numerical models designed to better understand the
surface deformation associated with magma chamber roof failure and subsidence. The
results are applied to the Bardarbunga-Holuhraun eruptive episode which exhibited
signals of a caldera collapse origin (Sigmundsson et al., 2014). Our findings suggest an
offset between roof subsidence as a function of depth below the surface and total ground
subsidence. We suggest that roof collapse generating ring-fractures was an unlikely
cause of ice subsidence observed at Vatnajokull, which instead may relate to
compressive stresses being formed within the caldera as a result of regional dyke
emplacement.
15
Chapter 1: Introduction
Chapter 7 draws all of the key findings together and evaluates in a critical manner the
methods and techniques used throughout. How the work can be expanded and insights
into future research directions are also discussed.
16
Chapter 2: Methodologies
Chapter 2: Methodologies
2.1
Numerical modelling methodology
2.1.2
Introduction
Numerical models are used to simulate physical problems and solve complex equations
when analytical solutions become too complex
Detailed guides explaining the
computational domain of finite element method modelling already exist (e.g Pryor,
2011) and are therefore not repeated here. In addition the software used (COMSOL
Multiphysics versions 4.1 to 5.1) is extremely well tested and benchmarked (e.g Hickey
& Gottsmann, 2014) for solving problems in volcanology and structural engineering.
Within this section I provide the constitutive equations that are solved numerically and
discuss their limitations and assumptions. Stress solution comparisons are discussed
with respect to previous studies to show the (in)consistency of results throughout
different finite element method software.
2.1.3
Software
All numerical modelling within the papers presented as part of this thesis were
conducted using the commercial finite element code within COMSOL Multiphysics
versions 4.1 to 5.1 (www.comsol.com). All models use the ‘Structural Mechanics’
module which solves Navier-Cauchy equations for linear elastic stress and displacement
as a result of load application. As the software is extremely well tested, relatively
simple to use, and has many published instructional manuals I do not provide a detailed
step by step guide to model creation in COMSOL, but only discuss points salient to
those models created in the studies within this thesis.
2.1.4
Theory and constitutive equations
All numerical models assume linear elasticity through Hooke’s law, i.e
17
Chapter 2: Methodologies
σ = Eϵ
(2.1)
where σ is stress, force over area, E is Young’s modulus and ϵ is strain, change in length
over original length of an object. A linear, isotropic elastic solid simply means that
when an applied force is removed from the material it will return to its original shape.
The prinicipal stress and strain axes in this material coincide such that the coordinate
system can be conveniently written as (Turcotte and Schubert, 2002):
1 ( 2G) 1 2 3
(2.2)
2 1 ( 2G) 2 3
(2.3)
3 1 2 ( 2G) 3
(2.4)
where λ, Lame’s constant, and the modulus of rigidity G are the Lame’s parameters.
The state of stress at any point in a solid can be calculated by providing σxx, σyy, σzz, σxy,
σxz and σyz, or the value of the principal stress and the orientation of the principal axes,
highlighted in Figure 2.1a. The normal stresses on planes perpendicular to the principal
axes are the principal stresses, denoted as σ1, σ2 and σ3. By convention the meaning of
the maximum and minimum principal stress in physics and engineering is the opposite
of standard geological meanings, as COMSOL is primarily an engineering tool, it
utilises engineering definitions. It is simple to reverse the notation so that the standard
convention is chosen such that σ1 ≥ σ2 ≥ σ3. Therefore, σ1 is the maximum principal
compressive stress, σ2 is the intermediate principal stress and σ3 is the minimum
principal stress (Turcotte and Schubert, 2002).
The stresses and displacements around circular holes can be solved analytically using
the Kirsh solution as shown below (Eq.2.5-2.9) and in Figure 2.1b:
rr
a2
P
(1 K )1 2
2
r
r
a2
a4
(1 K )1 4 2 3 4 cos 2
r
r
a2
P
a2
(
1
)
1
(
1
)
1
3
K
K
2
r2
r 2
P
a2
a4
(1 K )1 2 2 3 4 sin 2
2
r
r
18
cos 2
(2.5)
(2.6)
(2.7)
Chapter 2: Methodologies
a2
Pa 2
Ur
(1 K ) (1 K ) 4(1 ) 2
4Gr
r
U
cos 2
Pa 2
a2
K
(
1
)(
2
(
1
2
)
sin
2
4Gr
r2
(2.8)
(2.9)
Refer to Figure 2.1b for explanation of the geometrical terms used, P and K relate to
applied pressure, a is radius and r distance from the model edge. The elastic behaviour
of a material can be characterised by supplying either λ and G, or E and ν. All of the
models used in this study are created by specifying Young’s modulus E and the
Poisson’s ratio ν. All models were created in a two-dimensional regime; however,
Figure 2.1 highlights the potential for extension of results to three-dimensions.
Figure 2.1: Components of stress in two and three dimensions and relation to
displacements (b), modified from Turcotte and Schubert (2002).
19
Chapter 2: Methodologies
2.1.5
Model set-up and design
All models presented throughout utilise a two dimensional (2d) domain which can be
selected in COMSOL as either an axisymmetric or symmetric plane. Essentially the
physics governing either symmetrical or axisymmetrical models are the same, however
certain studies such as ground surface deformation have traditionally been displayed
using axisymmetry type displays (e.g Mogi 1958), and I therefore attempt to follow
those conventions for clarity when dealing with ground deformation (e.g Browning and
Gudmundsson, in press). All models are created with sufficient length and width to
ensure that edge effects become negligible (Figure 2.2). The full range of model details,
including descriptions of input parameters are given in Chapters 4 and 6. Initial preprocessing involves the input of geologically relevant geometries which are defined and
imported into the model. Model geometries in Chapter 4 (Browning and Gudmundsson,
2015) are informed by geological field measurements, whereas the geometries in
Chapter 6 are inferred predominantly from previously published geodetic data (e.g
Fichtner & Tkalčić 2010). Mechanical and physical properties are taken directly from
relevant laboratory and field studies.
Boundary conditions are prescribed as either pressure with the units Pa or displacement
field with the value given as the amount of total displacement in metres. Boundary
conditions are commonly applied to the edges of each model to simulate far-field
tectonic stresses, for example in the 2D symmetric model shown in Figure 2.2, the
hypothetical tectonic situation is that of regional extension as may occur in a rift zone.
Typical values of stress for such areas are around 5 MPa (Gudmundsson, 2011b). The
corners or edges of models are fixed in order to avoid solid-body rotation. Roller
constraints are occasionally used when a model interface or edge is to be allowed to
slide. Finally, infinite element domains are normally placed around the inner boundary
of each model in order to reduce computational processing requirements, and therefore
reduce the time taken to solve each model. A free surface condition is applied to the
upper edge of each model to simulate the ground surface, or more specifically the
contact with air or water, i.e. a zone free from shear stresses. None of the studies within
20
Chapter 2: Methodologies
this thesis model the effect of topographic loading, such as a volcanic edifice, there are
some consequences related to this admission which are further discussed in Chapter 7.
Figure 2.2: Typical COMSOL numerical model set-up, a) in 2d axi-symmetric and b)
in 2d symmetric. Features such as elliptical cavities are placed with a rectangular box of
horizontal and depth dimensions of sufficient size to ensure minimal boundary effects.
In addition boundaries are divided by infinite element domains to reduce computational
21
Chapter 2: Methodologies
requirement and boundary effects. At the outer edges of each model a range of different
boundary effects are applied.
2.1.6
Model meshing
The next step in a COMSOL model preparation is mesh definition and construction. It is
upon the elements (or lines) which make up a computational mesh that the relevant
simultaneous equations are solved in order to produce a suitable solution (of resultant
stresses, strains, displacements etc). Mesh elements vary in size throughout models, but
typically provide centimetre scale resolution near an area of interest, transitioning into
coarser metre scale resolution in distal areas (Figure 2.3). COMSOL offers various
geometrical arrangements for mesh elements, the most commonly used are triangular
elements, and whilst changing geometry does not significantly alter results a triangular
mesh is used throughout the studies that comprise this thesis.
Figure 2.3: Typical triangular mesh setup in an asymmetric model type. In this example
the mesh has been created to be finest (highest resolution) around the tip of a
22
Chapter 2: Methodologies
pressurised cavity (red inset box), at this location individual elements may be as short as
0.05 m. The maximum length of any element in this case is 15.5 m (green triangle). The
example shown is part of a larger model, and therefore the edges of the image do not
represent the edges of the model.
2.1.7
Model outputs and interpretation of results
COMSOL outputs a variety of graphic tools which can be used in order to assess, or
estimate the location of rock failure or the distribution of strain and resultant
displacement. Here those most relevant to the studies presented in this thesis are
discussed, with a brief explanation of how each output is interpreted in a geological
sense. Firstly the minimum principal compressive stress or maximum tensile stress (σ3)
is plotted and interpreted alongside the trajectories of maximum principal compressive
stress (Figure 2.4a&c). All results plotted as two dimensional stress maps or 1
dimensional ‘stress magnitude over lateral distance’ using the convention that tensile
stress is positive (Figure 2.4d). In addition von Mises shear stress, i.e σ1-σ3 is usually
presented in the same format (Figure 2.4b). The direction of principal stress axes are
plotted as stress trajectories, since we are normally concerned with the propagation of
dykes or inclined sheets, the maximum principal stress (σ1) is plotted as cones or
arrows. It is important to note that whilst COMSOL outputs give the impression of
stress directionality, the direction is purely aesthetic and has no physical meaning. The
orientation rather than the specific direction of stress trajectories is of interest in all
models.
23
Chapter 2: Methodologies
Figure 2.4: Typical COMSOL model result output generated from an symmetric model
setup where the dimensions of the x and y co-ordinate are in kilometres. a) a stress map
showing the magnitude and distribution of minimum principal compressive stress,
termed tensile stress. b) a stress map showing the magnitude and distribution of von
Mises shear stress, i.e σ1-σ3. c) trajectories or orientations of the maximum principal
compressive stress. d) magnitude of stress along a horizontal plane, in this case the
plane is that of the free surface (upper edge of the model), both shear and tensile stress
values are plotted.
2.1.8
Limitations of the FEM method
A limitation of the majority of numerical modelling methods that calculate stress and
resultant strains and displacements is that they are essentially static models which
provide a snapshot of conditions as a function of the model geometry and
boundary/loading conditions. That does not mean of course that many models cannot
24
Chapter 2: Methodologies
simulate time-dependant processes such as visco-elastic strain, but all FEM models rely
on a geometrical setup that does not alter with each time-step. Take the example of a
propagating fracture, like the ones (dykes) shown in chapter 4. In order to characterise
the propagation pathway it is necessary to interpret the stress signal (or stress map) and
then re-build the model based on the previous results.
A comparison of FEM models to standard analytical benchmarks suggests that the
maximum percentage error is ≪ 1% (e.g Grosfils 2007). This translates to absolute
differences in strain and displacement fields of 10-4 to 10-5 m (Grosfils, 2007; Hickey
and Gottsmann, 2014), and absolute stress differences of 10-3 MPa near an area of
interest (Grosfils, 2007; Hobbs, 2011). Grosfils (2007) has proposed that FEM models
which do not apply additional gravity loading give misleading results, a claim which
has been refuted by Gudmundsson (2012). The topic is covered in further detail in
Chapter 7.
2.1.9
Alternative software and numerical methods
Discrete element method numerical software is growing in popularity in structural
studies. The method differs from boundary element and finite element in that the model
mesh is made from many spherical nodes with the surrounding edge containing a
cohesion parameter which is used to input values of E and other salient parameters. The
main advantage of this technique is that the bonds between individual nodes can be
broken, and therefore dynamic fracturing processes can be directly observed rather than
interpreted from stress snapshots, as shown in Figure 2.5 (Holohan et al., 2015). The
problem with this method, and the predominant reason why it was not utilised, is that
spherical nodes are not a good representation of crustal materials. The method provides
a similar problem as often experienced in analogue models which use sand as the crustal
material. When scaled often the individual sand particles become much larger than any
possible fractures or spaces at depths below 1 km. It is perhaps convenient that many of
the advocates of the discrete element method are those who use analogue type sandbox
models, often the results are supportive which comes as little surprise when considering
the mesh and material physics.
25
Chapter 2: Methodologies
Figure 2.5: Results of a Discrete Element Method (DEM) model showing stress and
strain evolution during a ‘piston type’ magma chamber failure. With increasing magma
chamber depletion (from top to bottom) it is possible to track the development of faults
from kinematic relations as well as stress concentrations. The method can also be
utilised to estimate the propagation path of new fractures from principal stress
orientations, after Holohan et al (2015).
Results obtained from COMSOL and other FEM modelling techniques, i.e ANSYS and
boundary element techniques, i.e BEASY compare very well (Barnett and
Gudmundsson, 2014; Gudmundsson and Philipp, 2006; Gudmundsson, 2012, 2007).
2.1.10 Displacement mode of cracks
Throughout this thesis we consider three displacement modes of cracks, namely, mode
I, mode II and mode III. Mode I cracks, also referred to as tensile cracks, occur when
the crack walls move directly apart from each other. Mode II and Mode III cracks refer
to sliding, or tearing of the crack walls. In simple analytical models of an extension
26
Chapter 2: Methodologies
fraction such as a dyke, the appropriate model is a mode I crack (Gudmundsson, 2011).
Similarly a large strike-slip fault can be modelled as a mode III crack, and a dip-fault (a
caldera fault) can be modelled as a mode II or mode III crack. Many fractures are best
modelled as hybrid, or mixed-mode cracks.
2.2 Experimental methodology
2.2.2
Introduction
Cooling induced micro-cracking is an important but poorly understood process in
volcanic systems. The suite of experimental methods detailed in this chapter aim to
provide insights into the types of fractures induced by thermal contraction in volcanic
rocks. A new experimental setup is introduced, and a new technique for quantifying
crack annealing suggested. The vast majority of studies on thermal cracking have
concentrated on those fractures produced as a sample expands during heating, and is
therefore subjected to overall compressive stress. Our study seeks to understand the
fractures generated as a sample contracts during cooling, and is therefore subjected to
overall tensile stress.
Experimental studies of cooling induced fractures are complicated because it is difficult
to ascertain at which stage fractures were formed during the heating and cooling cycle.
To overcome this difficulty we investigated a method of annealing fractures by forming
melt at high temperature. We attempted to capture measurements of P-wave velocity
(Vp) throughout the thermal stressing tests. The hypothesis was that Vp would decrease
during heating as micro-cracks formed from thermal stresses, but would then increase at
maximum temperature as the viscous melt produced relaxed into open fractures. As the
experimental methodology developed it became clear that measurements of P-wave
velocities were not achievable during thermal stressing tests with the current setup.
In order to characterise fractures during thermal stressing tests we measure acoustic
emission (AE) output. A common method for providing insights into fractures during
rock deformation experiments is the acquisition of acoustic emission (AE). This
technique requires AE transducers to be in either direct contact with the sample surface,
27
Chapter 2: Methodologies
or in contact with a sample loaded wave-guide. In high temperature experiments AE
transducers would become damaged if they were in direct contact with the sample
surface inside a furnace, and therefore a wave-guide of sufficient length was needed to
position transducers outside of the furnace. The jig which hosts that wave guide had to
be capable of accommodating sample expansion as well as sample contraction. As such,
high force springs were attached to either end of the jig to ensure constant loading and
parallel contact between wave guide and sample throughout heating and cooling cycles.
Measurements of P-wave velocities provide an indication of the relative amount of pore
space (fractures) contained within a sample and were measured and compared prior to
and post heat treatment. These measurements were complemented by image analysis of
micro-fractures using Scanning Electron Microscopy. Attempts to characterise fracture
annealing at high temperature eventually proved unfruitful. All sample materials were
characterised prior to heat treatment and following heat treatment using a range of
methods highlighted within this chapter.
2.2.3
Sample selection and preparation
Three rock types are utilised throughout this study, an intrusive basalt from Seljadur in
Iceland (IB), previously described by Vinciguerra et al (2005), an intrusive phonolite
dyke from Anaga in Tenerife (AP) and an extrusive dacite lava from Nea Kameni,
Santorini (NKD). All calibration and initial experiments were conducted using the
Icelandic basalt, this follows for a number of reasons. 1) There was an abundant supply
of the material, 2) many of the rock’s mechanical properties have been previously
investigated and are therefore quite well known, 3) previous attempts have been made to
understand thermal cracking processes using this rock (e.g Vinciguerra et al., 2005), 4)
the melting point of the rock was understood to be sufficiently high as not to melt in the
furnace during high temperature calibrations.
Samples were cored into 25 mm diameter cylinders from their respective blocks using a
diamond core drill at University College London (UCL). The ends of each core were
then ground using a surface grinder to ensure parallelism and smoothness of the two end
surfaces. Sample lengths ranged from 70 mm to 45 mm. It is important that the sample
28
Chapter 2: Methodologies
ends were ground as smooth and even as possible to ensure a flush contact with the
wave guide cones.
Length and diameter of each sample were taken using a digital calliper. Three
independent measurements were taken at 120o intervals, in order to fully encompass any
variation in sample size. Any samples which have a difference in length of +/- 0.1 mm
were re-ground so that maximum area contact was achieved when clamped into the
experimental jig. Average radial length discontinuity was +/- 0.05 mm.
2.2.4
Sample characterisation
In order to characterise each sample type prior to and following heat treatment we
undertook the following analysis:
1) Optical light microscopy and scanning electron microscopy (SEM) to investigate preexisting crack populations and rock mineralogy.
2) Benchtop P-wave velocity analysis to investigate sample isotropy and initial void
space
3) Porosity measurements as a proxy for the amount of pre-existing void space
4) Thermo-mechanical analysis to determine the materials elastic and plastic response to
heating and cooling, as well as determine the thermal expansion co-efficient (α) of each
sample type.
5) X-ray fluorescence (XRF) in order to understand the chemistry of each sample type
6) Hot-stage microscopy to gain insights into sample softening temperatures
2.2.4.1
Optical light and scanning electron microscopy
Initially cylindrical samples were cut into thin sections along axis and observed using an
optical microscope. Whilst this technique was useful for constraining petrological
29
Chapter 2: Methodologies
aspects of the samples, it gave little information on the number and size of fractures.
Against the dark groundmass of the sample matrix it is difficult, perhaps impossible in
most cases to see any fractures. In fact, the only fractures that can be confidently
measured are those which occur within crystal planes. These fractures are likely related
to crystal formation however, and represent cleavage planes normally associated with
the crystals observed. This method was therefore deemed unsuitable for analysing crack
morphologies, and instead was used only to determine the petrology of each rock type
based upon standard igneous petrographic techniques (Shelley 1993). Petrologic data
are used to complement other analytical methods (TGA-DSC, Dilatometry) to inform
about possible phase transitions and estimate likely melting points.
The scanning electron microscope technique provided much greater detail with respect
to fractures. In all sample types and sections fractures can be observed in clear detail
with a resolution and measurement uncertainty of approximately +/- 5 μm, derived from
repeat measurements.
2.2.4.2
Benchtop P-wave velocity measurements
Preliminary ultrasonic wave velocity measurements were made on all sample types at
ambient conditions. An Agilent Technologies 1.5 GHz ‘Infiniium’ digital oscilloscope
was connected to a JSR DPR300 35MHz Ultrasonic pulser/receiver which was used to
excite one Panametrics V103 P-wave transducer at 1 MHz resonance frequency (Fig.
2.6). Waveforms captured from an identical transducer, both of which have 12.7 mm
diameter piezoelectric elements, were pre-amplified and then displayed on the digital
oscilloscope (Figure 6). Radial measurements were taken at 20 increments around the
circumference of each sample. Pre-heat treated samples were re-measured using the
same technique following each thermal cracking test. Wave velocities are calculated as
a function of the time taken for the induced wave to travel (t) through a measured length
of sample (d). P-wave travel times were measured by picking the signal break from a
stacked signal which was relatively noise free and therefore possible to pick the first
deviation from zero. All measurements of P-wave travel times take account of travel
time through the transducer’s tungsten carbide face-plates, calibrated previously as 0.14
μs for the P-wave transducer.
30
Chapter 2: Methodologies
Figure 2.6: Apparatus layout for benchtop seismic velocity measurements. P-wave
transducers held in an assembly clamp are connected to the rock surface. A P-wave is
transmitted from a pulser which travels through the rock and back to a receiver. The
resultant wave can then be analysed on an oscilloscope.
2.2.4.3
Density and Porosity measurements
Bulk density, ρ, of all materials were measured using cylindrical samples. Volume was
calculated using caliper measurements of the diameter and length with an error of ±
0.02 mm. All measurements were repeated three times, with the mean diameter and
length calculated. The mass of each dry sample was measured using digital scales with
an error of ± 0.02 g.
31
Chapter 2: Methodologies
Porosity is the proportion of bulk rock volume Vbulk that is not occupied by solid
material. If the volume of solid Vs and the volume of pore space is Vp = Vbulk – Vs, then
porosity ϕ can be defined as:
Vp
Vbulk
Vbulk Vs
Vbulk
(2.8)
Porosity was measured using an AccuPyc II 1340 Helium pyconometer, which involves
the injection of Helium gas into a chamber containing a rock of known volume (Figure
2.7). The Helium penetrates the pore-space of the rock, and the volume of gas which has
been injected is calculated using Boyle’s law:
pV k
(2.9)
where p is the systems pressure, V is the volume of gas, and k is a constant. Porosity is
then calculated from the volume of gas injected into the sample and expressed as a
percentage.
Figure 2.7: Schematic of a helium pycnometer setup where a sample chamber is filled
with a volume of gas, this volume is then released through a valve into an expansion
chamber and the volume difference calculated.
32
Chapter 2: Methodologies
2.2.5
Thermo-mechanical measurements
The maximum temperature and potential annealing timescale τ set for each thermal
stressing test was determined from thermal mechanical analysis. A Netzsch TMA 402
F1/F3 Hyperion thermo-mechanical analyser (TMA) at University of Liverpool (Figure
2.8) was used to determine linear thermal expansion (α) and strain (dL/Lo). A sample
load of 3 N was applied to a linear displacement transducer (LVDT) and the measured
change in sample length was obtained with a resolution of 0.125 nm (Siratovich et al.,
2015). Each materials volumetric thermal expansion coefficient can be defined as
(Simmons and Cooper, 1978):
v v / T .
(2.10)
where αv is the coefficient of volumetric thermal expansion, αϵv is the change in strain,
and T is the change in temperature. In order to calculate thermal expansion
experimentally the change in strain and change in temperature have to be changed from
a volumetric to a one-directional (length) solution (Siratovich et al., 2015) by:
L
1 L
L T
(2.11)
In this solution, αL is the linear expansion coefficient where L is the reference length of
a sample at initial temperature To and L is the difference in length of the sample
induced by a change in temperature T .
33
Chapter 2: Methodologies
Figure 2.8: Simplified schematic of the Netzsch Hyperion 402 thermo-mechanical
analyser used within this study, modified after Siratovich et al (2015).
Thermogravimetric measurements (TG) and differential scanning calorimetry (DSC)
were carried out using Netzsch STA 449 simultaneous thermal analysis equipment at
University of Liverpool. Small chips of sample around 50 mg were heated in a crucible
at a rate of 10˚C/min to a maximum temperature of 800-1100˚C depending on the
material used. Measurements were used to inform about 1) the range of temperatures
over which the glass transition (Tg) occurs in each material and 2) help characterize
bulk melting and re-crystallisation.
34
Chapter 2: Methodologies
2.2.4.4
Whole rock geochemical analysis
Bulk powered samples were analysed for whole rock geochemistry using X-ray
fluorescence (XRF) on fused glass discs. All analyses were undertaken on the Philips
PW1480 at Royal Holloway using the techniques of Thirlwall et al. (2000).
2.2.4.5
Hot-stage microscopy
Hot-stage microscopy was used as an initial test to investigate the temperature range
over which each sample would become plastic or partially molten. A Linkam TS1500
hot-stage connected to a Zeiss Axioscope at Lancaster University was used as the
apparatus for these heating experiments (Fig. 2.9), for further details see the description
of Applegarth et al. (2013). Thin wafers (<100 μm) of Icelandic basalt and Anaga
Phonolite were double polished and heated to 1150˚C. Images captured at a rate of 1/s
were compared at the highest temperature to check for fracture annealing or signs of
plasticity. Results proved largely inconclusive as individual fractures could not be
discerned using this method. However partial melt relaxation was inferred at the highest
temperature run in the Icelandic Basalt, the same test showed no obvious relaxation in
Anaga Phonolite.
35
Chapter 2: Methodologies
PC
Axioscope
Heated stage
Figure 2.9: Hot-stage microscope setup at the Lancaster University thermal laboratory.
A thin (100 m) sample wafer is placed inside a ceramic furnace, temperature is set by a
Linkam TS1500 heated stage mounted on a Zeiss Axioscope. In-situ observations are
made using a connected PC.
2.2.6
Experimental apparatus
Samples were placed inside a Carbolite tube furnace using a specially designed jig that
is capable of withstanding high temperatures and allows the transmission of ultrasonic
waves through a central wave guide (Figure 2.10). The sample is loaded by high force
springs which were located, outside the furnace at either end of the jig. Transducers,
connected to a PC were used to measure ultrasonic waves and placed on cones at the far
end of the waveguide. Temperature was measured using a K-type thermocouple placed
on the sample surface during all experiments. Initial temperature calibration tests were
conducted using two K-type thermocouples, one placed on the sample surface, and one
placed in the sample centre, as well as a thermocouple in-built to the furnace. During
36
Chapter 2: Methodologies
these calibration tests, temperature was measured digitally using a thermocouple
conditioner connected to a PC and also manually by noting the temperature display on
each conditioner and the furnace temperature display.
A new stainless steel jig capable of resisting oxidation up to a maximum temperature of
1100˚C was designed and built at the UCL rock physics laboratory (Figure 2.11). The
steel, an alloy 310 (UNS S31000) is an austenitic stainless steel developed specifically
for use in high temperature environments. It was important that the jig remained
relatively free from oxidation to ensure the free movement of key parts and to ensure
contact surfaces (cones) on the wave guide remained clean and with a smooth surface.
Previous similar steel jigs have undergone significant oxidation and resultant corrosion
during high temperature tests. A rock sample (30 mm to 75 mm in length) was held in
between two cones, connected by 1.1m long wave guide using a yoke system which
could move longitudinally independent from the rest of the jig. The yoke system
comprised a set of two high force springs which were located near the ends of the jig.
The springs ensured constant sample loading throughout the full heating (sample
expansion) and cooling (sample contraction) cycle. When assembled the entire jig was
slid into a Carbolite tube furnace so that the sample was in the centre of the furnace.
Acoustic emissions were recorded by transducers placed on cones at the end of the
central waveguide.
37
Chapter 2: Methodologies
Figure 2.10: Thermal stressing test experimental setup. Rock samples are held within a
tube furnace using a specially designed steel jig, which acts as a waveguide. During
calibration tests, as the setup is shown here, two K-type thermocouples were placed at
the rock interface inside the furnace. AE transducer signal and temperature is recorded
on a connected PC.
38
Chapter 2: Methodologies
Figure 2.11: Thermal stressing test jig design and assembly, original drawing by
N.Hughes, personal communication (2015).
39
Chapter 2: Methodologies
2.2.7
Temperature calibration setup
The Carbolite CTF12/75/700 tube furnace contains an in-built thermocouple and
Eurotherm 808 temperature control unit. The control unit allows temperature to be set
with the option of two programmable heating and cooling profiles. A set of temperature
calibration experiments were carried out with two main aims, 1) to ensure that the
temperature reading on the furnace controller could be matched to a known temperature
at the sample interface, and 2) to calculate the temperature gradient within a typical
sample during differential heating and cooling rates. Each temperature calibration
utilised a maximum hold temperature of 1100 ˚C, this temperature is well below the
calculated softening point of typical basalt and therefore crack annealing was not
expected or tested for.
A basalt core was specifically manufactured to allow a thermocouple to be placed
within the sample centre during heat treatment (Figure 2.12). This method made it
possible to calculate a temperature gradient, as an additional thermocouple was placed
on the outside of the sample. In total three thermocouples and therefore three sets of
temperature measurements were taken, one built-in to the furnace and connected to the
Eurotherm 808 controller unit, a second placed on the outside of the sample within the
furnace and connected to a thermocouple controller, and a third placed within the
sample core and connected to a second thermocouple controller (Figure 2.13). Both of
the thermocouples placed on or in the sample were also connected, through the
controller unit; to a PC which allowed the simultaneous recording of temperature, time
and acoustic emissions data. Unfortunately there was no provision for connecting the
furnace thermocouple to the PC to allow continuous recording of temperature, so
temperature was recorded manually at 5 to 10 minute intervals.
40
Chapter 2: Methodologies
25 mm
Figure 2.12: Basalt sample modified to allow temperature to be measured in the centre
of the sample during heat treatment. A) The basalt sample core was cut in half with a
1.5 mm grove cut into the centre and another thin grove cut at a ~45 angle
perpendicular to the centre grove to allow for a thermocouple to be placed inside, as
shown in (B).
Figure 2.13: Thermal jig setup for acoustic emissions capture during initial temperature
calibration experiments.
The slowest heating and cooling rate used was 1˚C/min, and the fastest heating rate was
10˚C/min, in Figure 2.14 the temperature recorded at the sample surface and furnace is
shown as well as the programmed furnace temperature. Sample and furnace temperature
profiles are consistent aside from an initial heating phase up to ~400˚C (Figure 2.15).
This is in contrast to the cooling phase where the fastest possible cooling rate was
~4˚C/min, and therefore the furnace program and measured temperature deviate
significantly during tests with a set cooling rate >1˚C. Cooling rate was limited by
atmospheric cooling conditions, and therefore it was not possible to cool faster than
41
Chapter 2: Methodologies
~4˚C/min, a rate which is reached for only a small fraction of the total cooling cycle.
The slowest cooling rate of 1˚C/min was used as this limited the total length of
individual experiments whilst capturing the range of temperatures over which salient
annealing and fracturing processes were likely occur (350 to 1150˚C). As such we
define a cut-off limit for all cooling that occurs below the 1C/min rate, indicated as a
green line in Figure 14 and corresponding to a temperature of 350. The natural rate of
sample cooling is approximated using a 2nd order polynomial (Figure 2.16), and is
consistent among all cooling calibration tests.
Figure 2.14: Programmed and actual heating and cooling profiles for a range of
temperature ramp rates, A) 1˚C/min, B) 4C/min and C) 8˚C/min. Temperatures were
measured at the sample surface. A green line indicates the temperature at which cooling
rates decrease below 1˚C/min, the slowest programmed rate of cooling.
42
Chapter 2: Methodologies
Figure 2.15: Outer sample temperature against temperature recorded inside the sample
and at the furnace thermocouple for three differential heating rates.
43
Chapter 2: Methodologies
Figure 2.16: Natural sample cooling rate, where maximum cooling occurs at a rate of
approximately 4C/min. All samples cool slower than 1C/min below 350 independent
of programmed cooling rate. A 2nd order polynomial model is fitted with a R2
correlation of 0.99.
2.2.8
Thermal gradients and timescales of thermal equilibration
It is important to know how long a typical sample takes to thermally equilibrate as a
function of heating and cooling rate. In the calibration tests described in this section, the
furnace is set to complete a ‘ramp and hold’ program in which a heating and cooling
rate is set with temperatures held for 30 minutes, Figure 2.17. Sample temperature, both
surface and core is consistently higher than the recorded furnace temperature, except
during the initial furnace ramp where a delay in sample heating is observed. During
cooling, sample and furnace temperature are more closely matched.
44
Chapter 2: Methodologies
Figure 2.17: Temperatures at the sample and furnace recorded during a 10˚C/min ramp
and hold cycle.
The experiment was used to test the timescale of thermal equilibration within a test
sample, thermal equilibration is said to be complete when the thermocouple on the
inside of the sample (core) gives a reading within error of the thermocouple on the
sample surface. The offset or error is temperature dependant (Figure 2.18c) but on
average is around 5˚C. When the prior condition is met it is assumed that temperature
within the sample is uniform, i.e there exists no radial or longitudinal thermal gradient.
Sample surface temperature is consistently higher than sample core temperature during
heating and perhaps more surprisingly, also during cooling. At each hold cycle it is
possible to observe the timescale with which thermal energy ceases to be input or output
from the sample, as recorded by a zero heating/cooling rate (Figure 2.18). This scale is
recorded during heating as the time between maximum heating rate and zero heating
rate. This timescale is commonly between 20 and 30 minutes (Figure 2.18b). The same
timescales are observed during a cooling cycle. The same method can be conducted
when considering the internal sample thermal gradient. The temperature difference
between the inner and outer sample is observed and the timescale of thermal
equilibrium calculated when the difference reaches zero. However there exists an offset
45
Chapter 2: Methodologies
between the two thermocouples and temperature controllers used, so there is always a
difference of ~5˚C which increases slightly with increasing temperature (Figure 2.18c).
Figure 2.18: Heating and cooling rate of a basalt sample centre during a ramp and hold
calibration test. a) heating and cooling rate as a function of temperature , b) heating and
cooling rate as a function of time in minutes and c) temperature difference between
sample surface and sample core. Red line indicates temperature recorded during heating
and the blue line is temperature recorded during cooling Thermal equilibration is
defined by the point at which heating or cooling rate decreases to zero, or the difference
between recorded sample surface and sample core temperature is zero. In this case there
is a temperature offset between the two thermocouples used to measure surface and core
temperatures, which increases slightly with increasing temperature.
46
Chapter 2: Methodologies
2.2.9
Acoustic emission (AE)
Acoustic emissions are elastic strain waves generated as a result of energy release
during crack interactions and propagation in materials (Lockner et al., 1992). The
energy release is a manifestation of transient waves which propagate from the area of
failure. Each discrete AE event over time is termed throughout as an AE hit. Within any
one AE hit it is possible to distinguish the relative size of a corresponding fracture based
on the signal amplitude (dB) and duration (μs) (Nelson and Glasar, 1992). The stored
waveform from any discrete AE hit can be used to calculate relative energy, which in
the experiments described always has arbitrary units (au). All energy is calculated using
the same method and therefore is directly comparable. The energy of any AE event is
calculated digitally by summing the envelope of the AE waveform (Fig. 2.19), (see Cox
& Meredith 1993 for a detailed description of the AE recording methodology). Acoustic
emissions can be measured simply as a rate of emission over time, which provides a
proxy for relative damage in a rock body. The emissions can also be characterized by
the frequency response (Figure 2.19) and energy released from each event. Here a full
seismic response study is not attempted or performed, this is because we are not
interested in the precise location or waveform of each discrete event but merely wish to
carry out a comparison of the relative ‘size’ and number of events.
Figure 2.19: Individual AE hits produces a seismic coda such as that shown. An event
such as this is likely generated due to micro-crack development within the rock. Time
(seconds) is along the x-axis and signal size, in mV on the y-axis
47
Chapter 2: Methodologies
AE hit and corresponding energy rate were averaged per ten AE events in order to
reduce the size of the often very large (>10,000 events) AE database, making data
interpretation more manageable. AE event rates ( AE ) were calculated as:
AE
AE f AE i
t f ti
(2.12)
Where subscripts i and f indicate the initial and final AE events over time (t) within the
10 hit interval specified, and the dot refers to derivative with respect to time.
Acoustic emissions were recorded on two separate hardware devices, 1) the Vallen
AMSY-5 and 2) the AEWin Pac kit. All experiments utilised one Panametrics V103
piezoelectric P-wave transducer located at one end of the wave-guide. As previously
described, AE transducers were located outside of the furnace to ensure that the Curie
point, whereby piezoelectric properties are limited or lost due to temperature exposure,
was not reached. All AE was filtered at a minimum cut-off of 35 dB to avoid the
presence of excessive noise.
2.2.10
Jig generated AE
In Figure 2.20 the acoustic response of the steel jig undergoing a typical thermal
stressing cycle is shown, no rock sample is loaded in these tests. A dummy sample
made of the same 310 alloy steel used for the waveguide construction was loaded in
place of a rock specimen. Acoustic emissions were recorded during a heating and
cooling cycle using a maximum temperature differential. It is shown that the apparatus
does produce AE but the vast majority occurs below approximately 300˚C and
predominantly during the cooling cycle. The test results are repeated through three
identical stressing experiments. It is likely that AE is generated by slip and movement
on the steel apparatus and springs during contraction, why the jig should be so ‘noisy’ at
low temperatures is not clear.
48
Chapter 2: Methodologies
Figure 2.20: Acoustic emissions generated from the steel jig heated and cooled to and
from 800 ˚C at a rate of 8˚C/min. Most AE’s occur below 300˚C during the cooling
cycle. Those tend to have small to moderate amplitudes, and low energy.
In comparison to a typical thermal stressing experiment with a rock sample loaded, as in
Figure 2.21, it can be seen that significant moderate to high amplitude and energy AE is
produced throughout heating and cooling at all temperatures. This indicates that over
the temperature range of interest, >300˚C, we can be confident that those AE recorded
are associated with sample cracking as oppose to an artefact or noise associated with
movement on the jig and waveguide.
49
Chapter 2: Methodologies
Figure 2.21: Acoustic emissions generated from a basalt sample heated and cooled to
and from 800˚C at a rate of 4˚C/min.
2.2.11
Wave-guide and sensor modifications
As the wave guide needed to be sufficiently long to ensure that transducers and springs
were positioned outside of the furnace, elastic wave attenuation became a significant
issue. As the experimental methodology developed, the steel wave-guide underwent a
number of modifications in order to attempt to improve acoustic wave transmission.
This was especially important for the acquisition of P-wave velocities, as the generated
pulse needed to pass through the entire length of the wave guide and the rock sample,
but less so for the acquisition of AE where the elastic waves only passed through half of
the wave guide. Despite many attempts and much improvement to the efficiency of
wave transmission through the various interfaces of the wave-guide, through for
example smoothing contact surfaces, it was not possible to detect a pulsed P-wave
through the entire wave-guide length. Whilst a determination of P-wave arrival times
and calculated velocities would therefore have been very useful in determining the
50
Chapter 2: Methodologies
presence of crack annealing, it was not possible to do so using the current experimental
setup, shown in Figure 2.22.
Figure 2.22: Thermal stressing test setup for in-situ P-wave velocity measurements
In the first high temperature experiment (>1100˚C) we attempted to track dynamic wave
velocities using two P-wave transducers glued to the ends of the wave-guide. At the end
of the experiments, which did not successfully record the transmission of P-waves, the
transducers were terminally damaged when being removed from the wave-guide.
Transducers at this stage were applied to the wave-guide with adhesive, and removed
simply by applying a sharp object in between the contact of the transducer plate and the
steel. The transducer holding mechanism was redesigned in order to ensure that the new
transducers would not be damaged upon removal. The new design used a plate which
was screwed to the outer most ring of the jig in order to apply a force to the transducer,
no additional adhesive was needed or used (Figure 2.23).
Figure 2.23: (a) Original method of attaching the P-wave transducer at the end of a
51
Chapter 2: Methodologies
cone on the steel wave guide using adhesive, and (b) new method utilising a back-plate
and studding to avoid the use of adhesive
2.3
Field methods
All field measurements were taken using standard geologic techniques and conventions
as described in many structural geology textbooks (e.g Fossen, 2010). All field
measurements utilised a compass clinometer to record dip and strike of faults, fractures
and dykes. Strike is here defined as the trend of the plane or body being measured. Dip
is the angle between the plane of the structure and a horizontal plane. Together strike
and dip define the attitude of a structure, written using the convention strike, dip and dip
direction (i.e 005/60E). Where possible and in order to reduce human measurement
error, multiple measurements of the same structures where taken with the repeat
measurements averaged. Measurements of attitude commonly possessed an error of ± 5
degrees. A standard tape measure was used to record fracture openings (dilation), and
when filled with magma or minerals; this is can be referred to as paleoaperture and
measured as thickness. Observations of dykes and inclined sheets represent minimum
values as strictly they are partly molten features that have cooled and contracted over
time. This shrinkage is normally considered to represent approximately 10% of the total
dyke thickness known from considerations of the density differences between magma
and host rock (Gudmundsson, 2011b). Measurements of dyke thickness taken remotely,
from distance have a larger uncertainty which was crudely estimated by taking repeat
measurements. Where possible, photos taken in the field contain a scale, otherwise a
scale-bar is drawn near the subject of interest.
52
Chapter 3: Forecasting magma chamber rupture
Chapter 3
Nature Scientific Reports
Forecasting magma-chamber rupture at Santorini volcano, Greece
Browning, J, Drymoni, K., and Gudmundsson, A
Scientific reports, 2015, 5, DOI 10.1038/srep15785
Statement of contribution:
All authors contributed to original idea
Analytical solutions were developed by JB from earlier work of AG. Specifically, novel
modifications were made to allow the estimation of pressure increase and failure.
Fieldwork conducted by all authors
Data interpreted by JB
1st draft of all figures by JB, with later input from all authors
1st manuscript draft by JB
53
www.nature.com/scientificreports
OPEN
Received: 18 June 2015
Accepted: 01 October 2015
Forecasting magma-chamber
rupture at santorini volcano,
Greece
John Browning1, Kyriaki Drymoni1,2 & Agust Gudmundsson1
Published: 28 October 2015
How much magma needs to be added to a shallow magma chamber to cause rupture, dyke injection,
and a potential eruption? Models that yield reliable answers to this question are needed in order
to facilitate eruption forecasting. Development of a long-lived shallow magma chamber requires
periodic influx of magmas from a parental body at depth. This redistribution process does not
necessarily cause an eruption but produces a net volume change that can be measured geodetically
by inversion techniques. Using continuum-mechanics and fracture-mechanics principles, we
calculate the amount of magma contained at shallow depth beneath santorini volcano, Greece. We
demonstrate through structural analysis of dykes exposed within the Santorini caldera, previously
published data on the volume of recent eruptions, and geodetic measurements of the 2011–2012
unrest period, that the measured 0.02% increase in volume of santorini’s shallow magma chamber
was associated with magmatic excess pressure increase of around 1.1 MPa. This excess pressure
was high enough to bring the chamber roof close to rupture and dyke injection. For volcanoes with
known typical extrusion and intrusion (dyke) volumes, the new methodology presented here makes it
possible to forecast the conditions for magma-chamber failure and dyke injection at any geodetically
well-monitored volcano.
Santorini is a volcanic island built predominantly by lava effusion and dome-forming eruptions1, periodically interrupted by catastrophic ignimbrite-forming eruptions2. The most recent caldera-forming event
occurred approximately 3650 years ago (at 3.6 ka) and is commonly referred to as the Minoan eruption.
Since that eruption Santorini has experienced primarily effusive activity, located centrally in the caldera
complex, which over time has formed the Kameni islands1. Nea Kameni (Fig. 1) produced at least three
well documented eruptive episodes during the 20th century1. The volume of magma extruded during
each individual event is estimated from the subaerial shapes and sizes of the lava flows and domes1.
Volumes of older submarine eruptions have also be estimated using bathymetric data1,3. The Kameni
islands lie along the Kameni line4, a tectonic lineament which may influence magma emplacement and
caldera faulting5. The average volume of magma issued during each individual effusive eruption is around
0.06 km3. This is much smaller than the estimated volume of magma involved in Santorini’s caldera
forming events at 3.6 ka and ~26 ka with the dense-rock equivalent (DRE) volumes of 20–30 km3 2. Whilst
spectacular and impressive, the 20th century eruptions posed little risk to the majority of Santorini’s
inhabitants. However, the islands are now a major tourist destination with a summer population in excess
of 50,0006. Even a small future eruptive event coupled with caldera-wall instabilities could therefore have
negative consequences.
In January 2011 Santorini volcano entered a period of unrest, meaning that the ground surface began
inflating3,7–9 and the magnitude and frequency of earthquakes increased3–5,7,8 . This period lasted until
April 2012 when signs of unrest ceased. The unrest was triggered by magma being transported as a dyke
1
Department of earth Sciences, Royal Holloway University of London, egham, tW20 0eX, United Kingdom.
Department of Mineralogy and Petrology, national and Kapodistrian University of Athens, Greece. correspondence
and requests for materials should be addressed to J.B. (email:
[email protected] or A.G. (email:
[email protected])
2
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
1
www.nature.com/scientificreports/
Figure 1. Simplified geological map of Santorini. Showing two main tectonic elements: the Kameni and
Coloumbo lines, the inferred Skaros caldera rim, and the approximate location of dykes within the northern
caldera wall. All the exposed dykes are located along the northernmost extent of the Skaros caldera wall
and the island of Therasia; some are marked in the photographs with red arrows. Most dyke measurements
were taken from a boat along the profile A–A′ . The stratigraphy of the caldera is complex, being made up
of many different types and ages of deposits. Many dykes within the wall are arrested, i.e. are non-feeders.
Santorini geological map is modified from29. Photos: John Browning.
(a fluid-driven fracture) from great depths (> 10 km) below the surface to a much shallower (~4 km deep)
magma chamber3,7–9. Using geodetic techniques, it is estimated that a combined volume of approximately
0.021 km3 (21 million cubic metres) of magma entered the shallow magma chamber, presumably in
two main phases, in just over one year3. None of the geodetic or seismic signals indicate that magma
rose from the shallow chamber as a dyke towards the surface, suggesting that increased pressure in the
shallow chamber due to the volume of new magma was insufficient to rupture the chamber roof. But
how close to rupture was the chamber? To answer that question for Santorini and other well-monitored
volcanoes, we provide a model to calculate the excess pressure in the chamber following the receipt of
new parental magma.
Results
In the simplest terms, a magma chamber roof will rupture when10–12
pl + pe = σ3 + T0
(1)
where pl is the lithostatic or overburden pressure (due to the weight of the overlying rocks), pe is the
magmatic excess pressure within the chamber, σ3 is the local minimum compressive principal stress and
T0 is the local tensile strength of the host rock. Since σ3 is the local stress, at the margin of the chamber, stress-concentration effects due to magma-chamber shape and loading are automatically taken into
account in Eq. (1)11,12. Typical values of solid-rock tensile strengths range from 0.5 to 6 MPa, the most
common values being 3–4 MPa11,13. It follows from Eq. (1) that for a part of the chamber to fail in tension
the local value of pe must during an unrest period reach To. At any other time the chamber is considered
to be in lithostatic equilibrium with the surrounding host rock, in which case the excess pressure pe is
zero (this assumption is discussed in the section Methods). Evidence for the mechanism of chamber rupture comes from fracture mechanics principles and field observations of extinct and now fossil magma
chambers, in Iceland and elsewhere, some of which have the well-exposed roofs dissected by dykes12.
Common intrusive (dyke) volumes at Santorini volcano.
Geological exposures at Santorini also
offer insights into the dynamics of magma movement within the volcano over time. At least 63 dykes
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
2
www.nature.com/scientificreports/
Figure 2. (a) Orientation and (c) thickness of 63 dykes in (b) the northern caldera wall of Santorini. Most
dykes are less than 1.5 m thick and strike dominantly NE-SW; those dykes which strike NW-SE generally
tend to be thicker and composed of felsic magmas. The average thickness of dykes measured is 1 m, the
minimum being 0.1 m and the maximum 5 m. For visualisation purposes the thickest dyke shown is 3 m.
(c) Dykes thicknesses plotted as cumulative frequency distributions follow an exponential trend (blue bars).
Individual dyke measurements plotted as a histogram with bin size 0.1 m are shown as red bars28. Photo:
John Browning.
(frozen or solidified magma-filled fractures) can be observed cutting the scalloped caldera wall in the
northernmost part of the island of Thera (Fig. 1). The dykes range from andesite to trachydacite in composition14 and are primarily exposed over a narrow section of the caldera wall at around 3.5 km east of
the town of Oia and south of Finikia. The caldera wall is accessible by boat, and abseiling in some parts,
making it possible to measure the thickness (roughly the palaeo-aperture or dyke-fracture opening)
and the strike and dip (attitude) of the dykes. Dykes strike dominantly NE-SW (Fig. 2), matching the
inferred strike of the Coloumbo line, a tectonic lineament which connects the Santorini volcanic complex
to the nearby Coloumbo volcano15,16. Fifteen dykes strike NW-SE. These dykes tend to be thicker and
lighter in colour, indicating a more evolved (felsic) composition. The thickest of the NE-striking dykes
is 2 m, the average thickness being 0.7 m. By contrast, the thickest of the NW-striking dykes is about
5 m, the average thickness being 1.7 m. Dyke thicknesses fit an exponential scaling law when plotted as
a cumulative frequency distribution (Fig. 2). Alternatively, the finding may reflect the relatively small
dataset or indicate two power-law sets with different scaling exponents—larger data sets normally suggest
dyke thicknesses following power-laws11. The dykes are predominantly sub-vertical, dipping on average
around 80°. The dip of individual dykes, however, varies considerably, indicating local stress variations
in the host rock17.
In order to estimate the volume of magma contained within any one individual dyke generated from a
shallow magma chamber at Santorini caldera we assume a dyke-length (along strike or strike dimension)
to thickness ratio of 150010,11. This ratio, a common value based on measurements of dykes worldwide, is
used because it is not possible to measure the lateral extent of any dykes within the caldera at Santorini.
The assumed ratio also takes into account that dykes tend to become longer at greater depths because of
general increase in Young’s modulus (E) with crustal depth11. Many dyke tips are seen, suggesting that
most of the dykes within the caldera wall are non-feeders, i.e. did not supply magma to an eruption
but rather became arrested at contacts between dissimilar layers within the volcano. Arrested dykes,
non-feeders, are commonly observed in well-exposed outcrops such as caldera walls and cliffs18,19, as in
Santorini, indicating that magma-chamber rupture and dyke injection is no guarantee for an eruption.
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
3
www.nature.com/scientificreports/
In particular, Santorini has a complex geologic stratigraphy made up of many rock units and layers with
contrasting mechanical properties1,17 (Fig. 1) whose contacts tend to arrest dykes11,17.
There is little difference between the thicknesses of the feeder-dykes and non-feeders (arrested dykes)18.
Since dyke thickness is linearly related to the dyke strike and dip dimensions11, we use an average dyke
dimension when calculating the volume of magma transported out of the chamber during common
eruptions. Using an average dyke thickness of 1 m, then, based on the length/thickness ratio above, the
average length or strike-dimension is 1500 m. Similarly, based on the geodetically determined depth to
the present magma chamber (about 4000 m), the average dyke depth or dip-dimension is 4000 m. Using
these dimensions and a thickness of 1 m, the average dyke volume is 0.006 km3. This average dyke volume
can then be combined with the known average volume of material erupted during the Santorini’s 20th
century eruptions to estimate the volume of the shallow source chamber and the necessary added volume
needed to rupture the chamber and inject a new dyke.
Estimating the volume of Santorini’s magma chamber.
The total volume Vm of a shallow chamber located within host rock of average compressibility βr and tensile strength T0 is related to the total
volume Ve of magma leaving (being squeezed out of) the chamber to produce the eruptive materials and/
or the injected dyke through the equation10,12,20
Vm =
Ve
T0 (β m + βr )
(2)
where βm is the magma compressibility. Using a typical shallow-crustal compressibility of 3 × 10−11 Pa−1 12,
an average in-situ tensile strength of 3.5 MPa6, and magma compressibility of 1.25 × 10−10 Pa−1 10, then
Eq. (2) reduces to
Vm = 1850 Ve
(3)
Here we use an average value taken from experimentally derived ranges for compressibility of various
magmas and compressibility and tensile strength of host rocks6,10,12, assuming a totally molten magma
chamber. Many magma chambers may be partly compartmentalised with zones of differential volatile
concentrations and crystal mushes, in which case they should be modelled as poro-elastic. These and
related topics are discussed further in the section Methods.
Using the estimated average volume of a typical individual dyke within the Santorini caldera,
0.006 km3, and the average measured volume of magma erupted for a typical individual eruptive phase
on the Kameni islands, 0.06 km3, then Ve in Eq. (3) becomes 0.066 km3. It follows then from Eq. (3)
that the total volume Vm of the shallow chamber active during these eruptions is about 122 km3. For a
penny-shaped or sill-like chamber, as are common12, and based on the dimensions of the three caldera
structures which make up Santorini, the chamber radius would be about 4 km and the thickness about
2 km. The geometry may, of course, be different. We do not aim to constrain the precise chamber geometry, since it is not needed for the present purpose. The main points are to assess the trade of between
radius and thickness and to show that, for the estimated volume, the chamber must be so large as to
encompass a significant area of the present-day caldera.
Magma-chamber rupture during recharge. Since the excess pressure at the time of magma-chamber
rupture is normally equal to the local tensile strength at the rupture site (Eq. 1), we can substitute pe for
T0 in Eq. (2). Also, assuming that the volume added to the chamber before rupture ∆ Vm is roughly equal
to the magma volume leaving the chamber following the rupture Ve, we can rewrite Eq. (2) as
pe =
∆Vm
Vm (β m + βr )
(4)
Here it is assumed that before the new magma of volume ∆Vm entered the chamber (from a deeper
source or reservoir), the chamber was in lithostatic equilibrium with the host rock and its excess pressure
pe thus zero. This is a normal assumption for periods of quiescence and follows partly because unrest
(e.g., inflation and earthquakes) would be expected in case of rising pe (pe > 0) whereas quiescence periods are characterised by the absence of unrest signals10,12.
During the 2011–12 unrest period in Santorini, the volume of new magma that entered the shallow
chamber ∆ Vm is estimated at around 0.021 km3 [Ref. 7]. Substituting this in Eq. (4) and using the above
values for the size of the chamber and the compressibilities, the corresponding excess pressure pe in the
chamber increased from zero to 1.1 MPa during the unrest period. Our results indicate that whilst the
total amount of new magma which entered the shallow chamber during the 2011–2012 unrest period at
Santorini represents a very small fraction (~0.02%) of the estimated total magma stored, the excess pressure increase within that shallow chamber came close to the surrounding host rock’s tensile strength10,
and therefore close to rupturing the chamber boundary and injecting a dyke (Fig. 3). For completeness
we also consider the slow inflation episode of 1994–1999 where the volume of new magma that entered
a chamber to the north of the caldera was estimated at around 0.78 × 10−5 km3 [Ref. 21]. For the five
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
4
www.nature.com/scientificreports/
Figure 3. (a) Simplified 3D model of Santorini volcanic system based on geodetic3,4,7–9 and geological data.
A deep reservoir feeds magma into a shallow system at around 4 km depth; this shallow chamber has a
current total volume of approximately 122 km3. The volume is estimated using the average volume of dykes
(0.006 km3) and the average volume of 20th century eruptions (0.06 km3) together with fracture-mechanics
and continuum-mechanics principles. The exact nature of the Kameni and Colombo tectonic lineaments is
unclear, but here both are drawn as normal faults. The box is drawn between 25.3–25.5° E and 36.3–36.5°
N to a depth of 15 km below the surface. (b) Excess pressure (pe) within the shallow magma chamber at
Santorini as a function of the volume of new magma (∆ Vm) entering the chamber from a deeper source
over time. Here the results are applied to the shallow chamber of Santorini based on the estimated size of
the chamber. The method, however, can be applied to any active central volcano for which (1) there exist
extrusion (lava and pyroclastic flows) and intrusion (primarily dyke) volume estimates and (2) geodetic
data as to inflation volumes. Rupture probability statements based on increasing excess pressure within the
shallow chamber allow forecasts of dyke formation to be made in real time during magma recharge events.
The model has been applied to the inflation episodes of 1994–1999 (red star) and
2011–2012 (purple star).
year period we estimate the excess pressure increase within the shallow chamber as about 0.3 MPa. In all
the unrest episodes, even if the chamber boundary ruptures and injects a dyke, the local stresses within
the edifice ultimately govern whether the dyke becomes arrested or, alternatively, reaches the surface to
supply magma to an eruption17 at Nea Kameni or elsewhere in Santorini.
Most models used to explain periods of unrest at Santorini simulate one shallow magma chamber
pressure centre north of the Kameni islands3,7–9. Other models, however, relate the unrest to two shallow
magma sources4,21,22, some citing the anomalous distribution of seismicity along the Kameni line and
a separate pressure source at a depth of 1 km, or possibly 5.5 km further north in association with the
1994–1999 inflation. Our calculated volume constraints easily incorporate the area of the proposed two
magma sources4,21. Two chambers are thus not needed in our model—a single, moderately large and
partly compartmentalised12 chamber is sufficient—but our results certainly do not rule out that possibility. Focussed seismicity on the Kameni line during periods of unrest may be related to its mechanical
properties being different from those of the surrounding crust, resulting in stress concentration along
the line, or deep-seated reservoirs. Further considerations of that topic, however, are outside the scope
of the present paper.
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
5
www.nature.com/scientificreports/
Eruptions at Santorini volcano are mostly with volumes ≪ 0.1 km3. However, much larger eruptions,
with volumes > 30 km3, occur occasionally and presumably from the same magma chamber. For a chamber with a volume of some 122 km3, a large fraction (about one-fourth) of its magma must be squeezed
out to generate such a large eruption. Ordinary elastic and poro-elastic models of the type described
here cannot explain such large magma removal from the chamber. The forced chamber volume reduction during piston-like caldera collapse, however, is apparently able to squeeze out a high proportion
of the magma in the chamber, thereby explaining occasional large eruptions from moderately sized
chambers. Then the large-volume eruptions are not the cause but rather the consequence of the caldera
collapse23. Combining the ordinary poro-elastic mechanism with the collapse-driven mechanism, the
estimated moderately large shallow chamber at Santorini volcano can supply magma to both small and
large eruptions.
Conclusion
The methodology presented here and applied to Santorini volcano can be used alongside real-time geodetic observations to help forecast magma chamber rupture at any geodetically well-monitored volcano.
This new method, therefore, represents a valuable first-order tool for volcano observatories during periods of volcanic unrest. Further steps must be taken in order to better constrain the local stresses within
the shallow parts of volcanic edifices, as these provide primary control on dyke propagation paths. As
yet no comprehensive model exists to ascertain whether a dyke injected from a ruptured magma chamber will reach the surface and supply magma to an eruption. Even so, estimating the volume of magma
stored at shallow depths and the conditions required to mobilise that magma are important steps in the
development of reliable volcano-tectonic models for forecasting volcanic eruptions.
Methods
Dyke measurements at Santorini were conducted during a five day field campaign in April 2014. Dykes
dissecting the northern caldera wall were measured directly on land as well as from a boat. Outcrops of
dykes are mostly limited to the northern caldera wall and parts of Therasia. Dyke attitudes (strike and
dip) were measured using a compass clinometer and thicknesses and morphological data of some dykes
were measured directly in the field, but mostly spotted from the boat at a distance of around 10–15 m.
Lava flow volumes are taken directly from previous studies1,7. Here we average all of the known lava
flow volumes to obtain the individual eruption average of 0.06 km3. Maximum and minimum lava volumes are given in the Supplementary Data.
Excess pressure (pe) is derived from the difference between total fluid pressure (pt) within the chamber
and the lithostatic stress (pl) where
pe = pt − pl
(5)
For lithostatic equilibrium, an assumed condition when the chamber is not undergoing unrest, all the
principal stresses at the chamber boundary are equal (σ1 = σ2 = σ3) and equal to the lithostatic stress
(pl). It then follows from Eq. (1) that pe = T0, which is used to derive Eq. (4). The assumption of lithostatic equilibrium is valid because any pressure deviation from lithostatic results in stress concentration
in the host rock of the chamber, and thus in volcanic unrest as would be reflected in e.g., geodetic
changes and seismicity. Even relatively small-scale unrests are detected on well-monitored volcanoes,
such as the creep-like inflation during 1994–1999 at Santorini22. When the chamber ruptures and injects
a dyke the overpressure or driving pressure po in that dyke is given by11
po = pe + (ρr − ρm ) gh + σd
(6)
where pe is the excess magma pressure in the magma chamber, ρr is the average host rock density, ρm is
the average dyke-magma density, g is acceleration due to gravity (9.81 ms−2), h is the height of the dyke
above its contact with the chamber (or the dip dimension of the dyke), and σd is differential stress, i.e. the
difference between vertical stress and the minimum principal horizontal stress in the rock layer where
the dyke overpressure is calculated. Note that this formulation includes the effects of gravity. The opening
displacement of dykes and the depth to magma chamber intersection can also be calculated analytically24
but is not within the scope of the current work.
To calculate the ratio of erupted material to volume of the shallow magma chamber we assumed the
chamber to be totally molten. This is the standard assumption used in the inversion of geodetic data to
infer the depths to magma chambers associated with inflation and deflation (unrest) periods25. However,
many chambers contain volatiles and crystals and therefore may be closer to a poro-elastic material, in
which case Eq. (4) becomes modified to10,12,26
pe =
∆Vm
Vm (β m + β p )
(7)
where βp is the pore compressibility of the chamber, i.e. the fractional change in pore volume (magma
fraction) of the chamber for unit change in the excess pressure. In this case, when new magma is received
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
6
www.nature.com/scientificreports/
by the chamber (from the deeper source), the new magma is partly accommodated through compression
of the old magma and partly by expanding chamber pore space. Compression of old magma leads to an
increase in magmatic pressure (+ pe), whereas pore expansion leads to a decrease in magmatic pressure
(− pe). The excess pressure increase as new magma is added to the chamber then depends on the values
of the pore and magma compressibilities. The magma compressibility, however, is generally much higher
than either the host-rock compressibility (Eq. 4) or the pore compressibility (Eq. 7). It follows that the
calculated excess pressure for a given addition of new magma to the chamber depends primarily on the
magma compressibility, and the results are similar when using Eqs (7 and 4) for shallow magma chambers. Our model assumes that magma compressibility remains constant throughout an unrest period and
is homogeneously distributed. More data is needed on magma compressibilities and their variations, and
until such data become available the present assumption has to be made, as is the case in most deformation studies3,7. Also, we focus on the magma chamber compressibility as a whole. Therefore, whilst
variations in compressibility almost certainly exist in compartmentalised chambers12, and will influence
aspects of associated localised volume changes, when the chamber is treated as a single homogeneous
system our assumptions are justified.
More specifically, there are no doubt significant uncertainties or errors in the estimated compressibilities of the rocks and the magmas used in eqs (2–4 and 7). The calculated compressibilities are based
on earlier data provided by Murase and McBirney20. However, no uncertainties are provided for these
original data, so that the standard propagation of uncertainties or errors estimates, whereby the uncertainties or errors add in quadrature27, cannot be made. In contrast to the compressibilites, which may
vary considerably, the in-situ tensile strengths may be regarded as close to constant. The most common
values are 3–4 MPa11, so that the average value of 3.5 MPa, used here, has an uncertainty or error of about
0.5 MPa, or less than 15%. A rough estimate of the total error in the excess pressure estimates, based on
the assumptions used, would suggest an uncertainty of perhaps 50%.
Dyke thicknesses are plotted on a histogram (Fig. 2) with bin widths of 0.1 m together with cumulative frequency distributions where the probability P(x) that x has a value greater than or equal to x, is
given by28,
P (x ) =
∫x
∞
p (x ′) dx ′
(8)
References
1. Nomikou, P. et al. The emergence and growth of a submarine volcano: The Kameni islands, Santorini (Greece). GeoResJ 1–2,
8–18 (2014).
2. Druitt, T. H. & Francaviglia, V. Caldera formation on Santorini and the physiography of the islands in the late Bronze Age. Bull.
Volcanol. 54, 484–493 (1992).
3. Parks, M. M. et al. From quiescence to unrest: 20 years of satellite geodetic measurements at Santorini volcano, Greece. J.
Geophys. Res. 120, 1309–1328 (2014).
4. Saltogianni, V., Stiros, S. C., Newman, A. V., Flanagan, K. & Moschas, F. Time-space modeling of the dynamics of Santorini
volcano (Greece) during the 2011–2012 unrest. J. Geophys. Res. 119, 8517–8537 (2014).
5. Konstantinou, K. I., Evangelidis, C. P., Liang, W. T., Melis, N. S. & Kalogeras, I. Seismicity, Vp/Vs and shear wave anisotropy
variations during the 2011 unrest at Santorini caldera, southern Aegean. J. Volcanol. Geotherm. Res. 267, 57–67 (2013).
6. Dominey-Howes, D. & Minos-Minopoulos, D. Perceptions of hazard and risk on Santorini. J. Volcanol. Geotherm. Res. 137,
285–310 (2004).
7. Parks, M. M. et al. Evolution of Santorini Volcano dominated by episodic and rapid fluxes of melt from depth. Nat. Geosci. 5,
749–754 (2012).
8. Newman, A. V. et al. Recent geodetic unrest at Santorini Caldera, Greece. Geophys. Res. Lett. 39, 1–5 (2012).
9. Papoutsis, I. et al. Mapping inflation at Santorini volcano, Greece, using GPS and InSAR. Geophys. Res. Lett. 40, 267–272 (2013).
10. Gudmundsson, A. Formation and mechanics of magma reservoirs in Iceland. Geophys. J. R. Astr. Soc. 91, 27–41 (1987).
11. Gudmundsson, A. Rock fractures in geological processes. (Cambridge University Press, Cambridge, 2011).
12. Gudmundsson, A. Magma chambers: Formation, local stresses, excess pressures, and compartments. J. Volcanol. Geotherm. Res.
237–238, 19–41 (2012).
13. Amadei, B. & Stephansson, O. Rock stress and its measurement. (Chapman Hall, New York, 1997).
14. Puchelt, H., Hubberton, H. W., Stellrecht, R. The geochemistry of the radial dykes of the Santorini Caldera and its implications,
Thera and the Aegean World III Pro. 3rd Int. Cong., Santorini, Greece, 1989 (D. A. Hardy et al., eds) 229–236 (Thera Found,
London 1990).
15. Nomikou, P. et al. Submarine volcanoes of the Kolumbo volcanic zone NE of Santorini Caldera, Greece. Glob. Planet. Change
90–91, 135–151 (2012).
16. Konstantinou, K. I. & Yeh, T. Y. Stress field around the Coloumbo magma chamber, southern Aegean: Its significance for
assessing volcanic and seismic hazard in Santorini. J. Geodyn. 54, 13–20 (2012).
17. Gudmundsson, A. & Philipp, S. L. How local stress fields prevent volcanic eruptions. J. Volcanol. Geotherm. Res. 158, 257–268
(2006).
18. Geshi, N., Kusumoto, S., Gudmundsson, A. Effects of mechanical layering of host rocks on dyke growth and arrest. J. Volcanol.
Geotherm. Res. 223–224, 74–82 (2012).
19. Geshi, N., Kusumoto, S., Gudmundsson, A. The geometric difference between non-feeders and feeder dykes. Geology 38, 195–198
(2010).
20. Murase, T. & McBirney, A. R., Properties of some common igneous rocks and their melts at high temperatures. Geol. Soc. Am.
Bull. 84, 3563–3592 (1973).
21. Saltogianni, V. & Stiros, S. C. Modeling the Mogi magma source centre of the Santorini (Thera) volcano, Aegean Sea, Greece,
1994-1999, based on a numerical-topological approach. Stud. Geophys. Geod. 56, 1037–1062 (2012).
22. Stiros, S. C., Psimoulis, P., Vougioukalakis, G. & Fyticas, M. Geodetic evidence and modeling of a slow, small-scale inflation
episode in the Thera (Santorini) volcano caldera, Aegean Sea. Tectonophysics 494, 180–190 (2010).
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
7
www.nature.com/scientificreports/
23.
24.
25.
26.
27.
28.
29.
Gudmundsson, A. Collapse-driven large eruptions. J. Volcanol. Geotherm. Res. 304, 1–26 (2015).
Becerril, L., Galindo, I., Gudmundsson, A. & Morales, J. M. Depth of origin of magma in eruptions. Sci. Rep. 3, 6 (2013).
Dzurisin, D., Volcano `deformation: new geodetic monitoring techniques. (Springer, 2007).
Bear, J. Dynamics of fluids in porous media. (Elsevier, New York, 1972).
Taylor, J. R. An introduction to error analysis. (University Science Books, 1997).
Mohajeri, N. & Gudmundsson, A. The evolution and complexity of urban street networks. Geogr. Anal. 46, 345–367 (2014).
Druitt, T. H., Edwards, L. & Mellors, R. Santorini volcano. Geol. Soc. Lond. Mem. 19, pp. 1–165. (1999).
Acknowledgements
J.B. is grateful for a Kirsty Brown memorial fund grant which enabled fieldwork. We thank Valerio
Acocella and two anonymous reviewers for comments that improved the manuscript. Thanks also to
Nathaniel Forbes-Inskip for discussions throughout.
Author Contributions
J.B., A.G. and K.D. contributed equally to the development of ideas and collection of data. Analytical
calculations were completed by J.B. and A.G. and the manuscript was compiled and written by J.B. with
input from A.G.
Additional Information
Supplementary information accompanies this paper at http://www.nature.com/srep
Competing financial interests: The authors declare no competing financial interests.
How to cite this article: Browning, J. et al. Forecasting magma-chamber rupture at Santorini volcano,
Greece. Sci. Rep. 5, 15785; doi: 10.1038/srep15785 (2015).
This work is licensed under a Creative Commons Attribution 4.0 International License. The
images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in the credit line; if the material is not included under the
Creative Commons license, users will need to obtain permission from the license holder to reproduce
the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/
Scientific RepoRts | 5:15785 | DOi: 10.1038/srep15785
8
Chapter 4: An alternative mechanism of ring-dike formation
Chapter 4
Bulletin of Volcanology
Caldera faults capture and deflect inclined sheets: an alternative mechanism of
ring dike formation
Browning, J and Gudmundsson, A
Bulletin of Volcanology 2015,77, 4. DOI 10.1007/s00445-014-0889-4
Statement of contribution
Collection of primary field data by JB with support from co-author and field assistants
All field photographs by JB
JB Conceived and built all numerical models and resulting figures
Complete 1st draft of manuscript and figures by JB
Revisions and subsequent drafts made with co-author input
Interpretation of field and numerical data along with final model conducted by JB with
support from co-author
54
Bull Volcanol (2015) 77:4
DOI 10.1007/s00445-014-0889-4
RESEARCH ARTICLE
Caldera faults capture and deflect inclined sheets: an alternative
mechanism of ring dike formation
John Browning & Agust Gudmundsson
Received: 27 July 2014 / Accepted: 3 December 2014
# Springer-Verlag Berlin Heidelberg 2015
Abstract The subsurface structures of caldera ring faults are
often inferred from numerical and analog models as well as
from geophysical studies. All of these inferred structures need
to be compared with actual ring faults so as to test the model
implications. Here, we present field evidence of magma
channeling into a caldera ring fault as exhibited at Hafnarfjall,
a deeply eroded and well-exposed 5-Ma extinct volcano in
western Iceland. At the time of collapse caldera formation,
over 200 m of vertical displacement was accommodated along
a ring fault, which is exceptionally well exposed at a depth of
approximately 1.2 km below the original surface of the volcano. There are abrupt changes in the ring fault attitude with
depth, but its overall dip is steeply inward. Several inclined
sheets within the caldera became arrested at the ring fault;
other sheets became deflected up along the fault to form a
multiple ring dike. We present numerical models showing
stress fields that encourage sheet deflection into the
subvertical ring fault. Our findings provide an alternative
mechanical explanation for magma channeling along caldera
ring faults, which is a process likely to be fundamental in
controlling the location of post-caldera volcanism.
Keywords Calderas . Inclined sheets . Numerical modeling .
Ring-dikes . Ring-faults
Editorial responsibility: G. Giordano
Electronic supplementary material The online version of this article
(doi:10.1007/s00445-014-0889-4) contains supplementary material,
which is available to authorized users.
J. Browning (*) : A. Gudmundsson
Department of Earth Sciences, Royal Holloway, University of
London, Egham TW20 0EX, UK
e-mail:
[email protected]
Introduction
Many ring faults around the world are intruded by dikes (e.g.,
Smith and Bailey, 1968; Johnson et al., 2002). These dikes are
believed to have been emplaced either during the injection of
magma during collapse caldera formation (Anderson, 1936;
Sparks, 1988; Walter, 2008) or incrementally through many
injections along the ring fault (Saunders, 2001). The common
assumption, based on Anderson’s (1936) model, is that ring
dikes are injected directly into the ring fault at its contact with
the magma chamber.
Dikes and sheets commonly intrude pre-existing weaknesses such as joints (Delaney et al., 1986) and faults
(Gudmundsson, 2011; Magee et al., 2013; Bedard et al.,
2012). While there are numerous examples of well-studied
ring faults, for example Glencoe caldera (Clough et al., 1909;
Kokelaar, 2007) and Hannegan caldera (Tucker et al., 2007) as
well as many others (see Lipman 1984), observations of wellexposed ring faults and ring dikes in the same vertical cross
section at depth are extremely rare. As such, the mechanics of
magmatic interaction with caldera faults is still poorly
understood.
Many calderas experience post-collapse resurgence which
may culminate in eruptive activity. Commonly, post-caldera
volcanism concentrates spatially above the vertical extent of a
caldera ring fault (Geyer and Martí 2008; Saunders, 2004).
Therefore, understanding the subsurface structure of caldera
ring faults is important for identifying the location and timing
of renewed volcanic activity within active calderas. Caldera
ring faults have traditionally been studied using analog (e.g.,
Acocella et al., 2000; Acocella, 2007; Geyer et al., 2006;
Holohan et al., 2005; Kennedy et al., 2004), numerical (e.g.,
Gudmundsson, 1998; Hardy, 2008), and analytical methods
(e.g., Gudmundsson, 1998), with the subsurface structure
often inferred from such models, as well as from seismicity
(e.g., Ekstrom, 1994), geodetic studies (e.g., Jonsson, 2009),
4
Page 2 of 13
Bull Volcanol (2015) 77:4
and near-surface observations (e.g., Geshi et al., 2002; Troll et al.
2002). Caldera ring faults are primarily subvertical dip-slip shear
fractures, although in some cases, the faults accommodate an
oblique slip (Holohan et al., 2013). Whether a caldera ring fault
dips inward or outward from the center of subsidence is a longdebated and contentious issue (Gudmundsson and Nilsen, 2006;
Burchardt and Walter, 2010; Geyer and Marti, 2014). For example, Branney (1995) suggests that most ring faults dip outward,
whereas observations from the collapse of Miyakejima, Japan, in
2000 (Geshi et al., 2002; Burchardt and Walter, 2010) and Piton
de la Fournaise, La Reunion, in 2007 (Michon et al., 2009)
indicate both inward- and outward-dipping ring faults. The use
of near-surface observations of ring faults (e.g., Michon et al.,
2009) may be misleading as caldera walls are subject to mass
wasting and erosion (Lipman, 1997). Furthermore, it may be
difficult to infer correctly the subsurface structure from surface
observations because the fault-generating local stresses are likely
to vary with depth in the volcanic edifice, thereby affecting the
overall geometry of the fault structure (Gudmundsson, 2011).
Here, we present field data and numerical models which
show that dikes and sheets can become deflected at and along
a ring fault, the deflection being primarily due to the difference in
material properties between (and within) the fault zone and the
host rock. These observations and associated modeling provide
an alternative mechanism for the formation of ring dikes.
Fig. 1 Geological map of
Hafnarfjall central volcano,
located in western Iceland,
approximately 50 km northwest
of the active volcanic zone (AVZ).
The ring fault of the 7.5×5-km
NW-SE elongated caldera is
shown. Study area is marked with
a yellow star. Modified after
Franzson (1978)
A well-exposed caldera ring fault in western Iceland
Hafnarfjall is an inactive and deeply eroded 5-Ma-old central
volcano (stratovolcano with a caldera) in western Iceland. The
volcano is composed of a predominantly basaltic lava pile
overlain by brecciated andesite and andesitic lava, as
described in detail by Franzson (1978) (Fig. 1). The volcano
originally formed in the southwest volcanic zone of Iceland
but subsequently drifted, through crustal spreading, 40–50 km
(Gautneb et al., 1989) to the west-northwest of the rift zone.
Hafnarfjall therefore offers the opportunity to study a caldera
formed in a divergent plate boundary setting. We estimate that
glacial erosion has removed the uppermost parts of the volcano based on the assumptions of Walker (1960) and
Johannesson (1975) who used zonation of amygdale minerals
to estimate the level of erosion in a nearby area. Hafnarfjall
volcano contains numerous inclined sheets, predominantly
basaltic, which dip on average at around 65°, trend NE, and
have thicknesses that are commonly about 1 m or less
(Gautneb et al., 1989). The thickest sheets, however, reach
about 10 m and tend to be composed of rhyolite. Many of
these intrusions in Hafnarfjall are highly altered, and it is often
difficult to discern characteristic intrusive features such as
chilled selvages and cooling jointing. In Fig. 2, we show
several dikes and inclined sheets ranging in thickness from
o
66 N
Borgarnes
o
6433' N
AVZ
Study area
o
64N
100 km
o
22 W
o
18 W
10
o
6430' N
o
6427' N
0
30
0
70
0
3 km
Route 1
o
Legend
221' W
o
2151'
W
o
2141'
W
Caldera fault
Rhyolite lava
Tholeiite lava
Basaltic breccia
Olivine Tholeiite lava
Gabbro
Brecciated Rhyolite and tuff
Andesite lava
Landslide
Bull Volcanol (2015) 77:4
Page 3 of 13 4
2.5 m (Fig. 2a) to 3 cm (Fig. 2c). They are all clearly intrusions
which are discordant to bedding; however, very few display
characteristic chilled selvages or discernible horizontal fracture patterns.
At around 4.6 Ma, Hafnarfjall experienced a major collapse
event, forming part of a NW-SE elongated caldera approximately 7.5×5 km in diameter (Franzson, 1978). The most
striking evidence of this collapse is the exposed ring fault in
a gulley, oriented roughly NW-SE, at the northernmost margin
of the caldera, as shown in Fig. 3. At this locality, an E-W
trending segment of the ring fault can be observed in a vertical
cross section for 200 m along dip and around 700 m along
strike (Fig. 3a). Here, the ring fault cuts a 300-m-thick lava
pile composed of 2–5-m-thick flows of basaltic tholeiite.
Around the outer caldera margin, lavas dip 15° S, whereas
the dip of lavas which constitute the caldera block increases
to ∼35° S. This implies that the caldera block has tilted during
faulting. Dips of intra-caldera lavas also increase with depth
toward the center of the caldera, i.e., individual lava layers at
the outcrop base exhibit a steeper dip than those near the
present-day surface. The caldera fault can be traced in several
localities as mapped by Franzson (1978) and shown in Fig. 1
as a single fault plane with displacement in excess of 100 m.
The exposure described in detail here indicates a throw in
excess of 200 m, although the normal fault offset is considered
a minimum and is based on the inability to trace individual
lava flow layers horizontally, across the entire vertical section.
As mentioned, the vertical section is 300 m, although individual lava layers can only be obviously discerned for approximately 200 m of this section.
The studied vertical section of the ring fault does not display
a constant attitude. This finding is in contrast to those of many
Fig. 2 Examples of sheets and
dikes found in the Hafnarfjall lava
pile. a View southeast toward the
caldera ring fault, indicated as a
dashed line. b Dikes and sheets
are discernible as discordant
linear features within the lava
pile, in this example a pair of
cross-cutting dikes, the largest
and youngest of which is
approximately 2.5 m in thickness
and dips 65–75° S. Other dikes
and sheets are indicated by red
arrows. c, d A 3-cm-thick
arrested dike exhibiting a narrow
and pointed dike tip. Due to
weathering and alteration, it is
difficult to observe any chilled
selvage or horizontal fracture
pattern within the dike; this is
common in many of the intrusions
in Hafnarfjall
a
models on ring fault formation that often predict a simple
inward- or outward-dipping trend (Acocella, 2007). Instead,
the fault alternates in dip between 85 N–90° and 85 S–90°,
suggesting that a number of stress perturbations occurred during fracture propagation. Such subtle changes in fault attitude
are unlikely to be detected in models unless a heterogeneous, in
particular, a layered, edifice is considered. On average, the fault
has a normal trend and dips steeply inward, at ∼85° S.
Steep-sided slopes surrounding the fault exposure limit
observations to the base of two gullies and, from a distance,
to a parallel topographic high (Fig. 3a, b). At the base of the
caldera fault (Fig. 4) are five thin (<0.7 m) basaltic, but highly
altered, dikes. Thin (0.5–1 cm) mineral veins separate some of
the individual intrusions, as shown in Fig. 4c. Surprisingly, no
breccia or fault gouge is found along the main fault plane. We
interpret the dikes within the fault plane as ring dikes and now
refer to them as such throughout. In Fig. 4b, two ring dikes can
be clearly observed within the fault plane; the northernmost
dike strikes N086° E, and it becomes offset around 25 m
vertically in the pile. The strike of the second dike is variable
between approximately 095 and 115 and appears to follow the
fault plane. In the intra-caldera lava pile, thin (<1 m) inclined
sheets dip between 45 and 75° S, and upon contact with the
fault, the sheets either become arrested or change attitude and
deflect vertically into the fault. No sheets can be traced from
inside of the caldera margin, across the fault, and to the outside
of the caldera.
On the intra-caldera margin, synclinal drag folding indicates a normal sense of dip-slip shear and displacement (see
Fig. 3a). The zone of folding extends for approximately 10–
15 m horizontally toward the caldera center. In this region, it is
difficult to distinguish between individual lava flows, as their
b
Caldera fault
c
d
Dike tip
4
Page 4 of 13
Fig. 3 a Caldera ring fault
exhibits a general inward-dipping
normal trend, although subtle
variations in attitude occur
throughout. The height from the
base to the top of the fault is
∼300 m. Lavas on the inner
caldera margin dip more steeply
than those outside, and generally
dip increases with depth. Several
markers are used to infer synclinal
drag folding, perhaps the most
prominent is a light white tuff
which clearly bends into the fault.
Displacement is greater than the
vertical section, so no horizontal
markers can be traced across the
fault plane. b Many individual
lava layers with thicknesses
between 1 and 2 m can be traced
to the fault contact on the outer
margin; however, most lavas on
the inner margin are sufficiently
deformed and not discernible. c, d
Section of the upper observable
part of the fault. Individual dikes
shown and numbered are around
1 m thick. (Location 64° 30′ 01″
N 21° 52′ 39″ W)
Bull Volcanol (2015) 77:4
a
~20 m
~5m
c
b
Sub-vertical
ring-dikes within
the fault plane
2
1
d
~ 10 m
characteristic scoria margins have been sufficiently deformed. There is no indication of reverse-sense displacement or motion on the fault, which is a commonly
interpreted mechanism during caldera unrest due to magma
intrusion (Acocella et al., 2000; Walter and Troll 2011;
Jonsson, 2009). This is important to note, as our later numerical models simulate an overpressured condition, which
is of course a requirement for propagating the original
sheets described throughout.
Factors influencing the propagation of dikes and sheets
Stress barriers
In order for a sheet to successfully propagate, tensile stress
magnitudes should exceed the in situ host rock tensile
~5m
Sheet deflection at fault contact
strength, which is generally between 0.5 and 6 MPa
(Amadei and Stephansson, 1997; Gudmundsson, 2011).
The direction of propagation is based on Anderson’s
(1936) theory of sheet and ring dike propagation and
on numerous field observations of dikes and sheets
which suggest that magmatic fractures will propagate
in a direction parallel to the trajectories σ1 and perpendicular (or normal) to σ3 (cf. Gudmundsson, 2011). A
stress barrier is a layer or unit where the local stress
field is unfavorable for the propagation of a particular
type of fracture. For example, for a vertically propagating extension fracture, a stress barrier would be a layer
where the maximum compressive stress flips to horizontal, a situation which favors dike arrest (Gudmundsson
and Phillip, 2006). Both stiff (high Young’ modulus)
and soft (low Young’s modulus, compliant) layers can act as
stress barriers (Gudmundsson and Phillip, 2006).
Bull Volcanol (2015) 77:4
Page 5 of 13 4
Fig. 4 Base of the outcrop of the
ring fault, where four thin ring
dikes (<0.7 m) of basaltic
composition occupy the
subvertical segment (a). Two of
these dikes are observable in the
vertical section (b, c): one can be
traced inside the caldera margin
(d), several inclined sheets strike
E-W to NE-SW and dip between
65 and 80° S; at least two of the
sheets meet the fault contact
higher in the pile
a
b
1
3
2
4
2
Ring-dikes
1m
1
c
~5m
d
~2m
Elastic mismatch
The second mechanism responsible for dike deflection is
related to the difference in material properties of the
layers hosting, and directly in front of, a propagating
fracture and the associated contacts. Dikes and sheets
Gtotal ¼ GI þ GII þ GIII ¼
ð1−v2 ÞK I 2 ð1−v2ÞK II 2 ð1 þ vÞK III 2
þ
þ
E
E
E
GI–III are the energy release rates (Jm−2) of ideal mode I–III
cracks (Anderson, 2005; Gudmundsson, 2011), E is the
Young’s modulus, v is the Poisson’s ratio, and KI–III are the
stress intensity factors. When the rock layer which hosts the
dike or sheet has the same or similar mechanical properties to
a rock layer above, then the strain energy release rate for a
mode I crack, Gp, reaches a value suitable for fracture extension, which is equal to the material toughness of the layer, ΓL.
Therefore, from Eq. (1), the condition becomes
ð1−v2 ÞK I 2
¼ ΓL
Gp ¼
E
are extension fractures or mode I cracks. However,
when such fractures meet and become deflected into a
contact or discontinuity, they temporarily become
mixed-mode (He and Hutchinson, 1989; Xu et al.,
2003). Consider the total strain energy release rate Gtotal
in mixed-mode fracture propagation:
ð2Þ
ð1Þ
However, the dike or sheet will deflect into the discontinuity if the strain energy release rate becomes the same as the
material toughness of the discontinuity, ΓD. Deflection at the
discontinuity then occurs when
Gd ¼
ð1−v2 Þ
K I 2 þ K II 2 ¼ Γ D
E
ð3Þ
It follows that a dike or sheet will continue on the same
trajectory, through a discontinuity if (He and Hutchinson,
1989)
Page 6 of 13
4
Bull Volcanol (2015) 77:4
Gd
Γ D ðΨ Þ
<
Gp
ΓL
ð4Þ
or become deflected at the discontinuity if
Here, E denotes the plane strain Young’s modulus, and the
subscripts 1 and 2 relate to the moduli of the rock above and
hosting the dike, respectively. The ratio below the curve in
Fig. 5 indicates areas where the deflection of a dike or sheet is
favored, whereas those areas above the curve indicate continued propagation with no deflection.
Cook-Gordon mechanism
Gd Γ D ðΨ Þ
≥
Gp
ΓL
ð5Þ
where Ψ denotes the measure of the relative proportion of
mode II to mode I:
.
Ψ ¼ tan−1 K II K I
ð6Þ
In Fig. 5, the ratio of Gd/Gp is plotted as a function of α,
which represents the Dundurs elastic mismatch parameter and
can be presented in the following form (He and Hutchinson,
1989):
Experiments on crack propagation have shown that CookGordon debonding is a common mechanism in the delamination of composite materials (Xu et al., 2003; Wang and Xu
2006). It has been shown that the tensile strength ahead of a
propagating dike can open up a contact ahead of the dike tip
(Gudmundsson, 2011). Such a mechanism is important in
where there is an abrupt change in the mechanical properties
or rocks across an interface or a contact, particularly when the
contact is clearly defined and mechanically weak (with a low
tensile strength). However, in a fault zone such as that described previously, the contact between the fault and host rock
is not clearly defined. As such, we do not consider this
mechanism as important for capturing sheets within the fault,
at least initially.
Numerical model setup
α¼
E 1 −E 2
E1 þ E2
ð7Þ
Relative energy release rate Gd / Gp
2
Layer 2
Contact
1.5
a) arrest
Layer 1
1
b) penetration
Fracture
c) deflection
Fracture penetration
0.5
Fracture deflection
-1
-0.5
0
0.5
Dundurs Elastic Mismatch parameter
1
Fig. 5 Conditions for fracture propagation: upon meeting a contact
between two layers with contrasting material properties, a fracture will
either (a) arrest, (b) penetrate the contact, or (c) deflect at the contact. The
ratio of strain energy release rate for fracture deflection (Gd) against
fracture penetration (Gp) is plotted as a function of the Dundurs elastic
mismatch parameter (α; see text for details). Modified after He et al.
(1994)
To test the proposal that sheets can become arrested at or
deflected into a caldera fault, we made several numerical
models. In all the models, we calculate the stress field around
a 1-m-thick dike subject to an internal magmatic excess pressure of 5 or 10 MPa as the initial loading. The weight of
overlying host rock or the overburden pressure is included in
the lithostatic stress (Jaeger and Cook, 1979) and is therefore
taken into account when considering loading as excess pressure in the sheet (Gudmundsson, 2011).
In the numerical models presented here (Figs. 7, 8, 9, 10,
and 11), the focus is on the mechanical properties of the
caldera fault zone, namely the stiffness or Young’s modulus
(E) of the layers that constitute the damage zone and the core
of the fault, and the more gently dipping layers through which
the sheet propagates. Density and Poisson’s ratio are kept
constant in all model runs, with values of 2500 kg m−3 and
0.25, respectively. We estimate that the damage zone surrounding the ring fault is approximately 15 m thick and occurs
predominantly on the southern, down-throw (“hanging wall”)
side of the fault. The damage zone is qualitatively estimated
based on the ability to discern individual lava flows, which on
the northern wall can be traced to the fault contact but on the
southern wall are highly altered and deformed. This deformation zone is observed laterally south for around 15 m from the
fault core, which contains the ring dikes, to a point where
Bull Volcanol (2015) 77:4
Page 7 of 13 4
individual lava flow characteristics are discernible. Fault core
and damage zones have been recognized and measured on
different scales in many areas of Iceland (Gudmundsson et al.
2011). The fault zone is modeled simply as three individual
layers of similar thickness but with differing mechanical properties. The models presented are designed to test how the local
stress field changes as an inclined sheet (a) approaches a ring
fault and (b) becomes captured by the fault.
The sheet is situated perpendicular to the fault zone (see the
model setup in Fig. 6). A section of the ring fault is modeled in
two dimensions as a series of subvertical layers decreasing in
stiffness toward a fault core, replicating the variation in stiffness of the fault damage zone (e.g., Gudmundsson, 2011). In
this setup, the softest layer is the fault core which is characterized by the lowest value of Young’s modulus or stiffness
(0.1 GPa); this is surrounded by a fault damage zone which
stiffens gradually approaching that of the host rock (40 GPa).
For simplicity, here, the damage zone stiffness is assumed
constant at any particular time. Fault zone stiffness, however,
was varied over time in separate model runs to incorporate the
dynamic nature of fault development both syn- and postcollapse. Temporal changes reflect initial fault growth, and
subsequent healing and intrusion by dikes. All models simulate snapshots of the magnitudes and directions of principal
stress around the pressurized sheet; the likely propagation path
is then inferred based on the trajectories of the maximum
X
X
X
X
X
Fault zone
E = 0.1 - 30 GPa
E4
X
E3
E2
θ
X
X
Inclined
Sheet
X
Pe= 5 - 10 MPa
Host rock
E1 = 40 GPa
X
X
X
E1
X
X
Fig. 6 General COMSOL model setup used throughout. In all runs, the
models are fastened at the outer boundary using a fixed constraint; this is
coupled with an infinite element domain on the inner model margin to
ensure the fixed boundary effects do not influence results. The host rock
and fault zone are modeled using different values for Young’s modulus
(E) as specified in the “Model results” section. An inclined sheet is
modeled as a cavity with an excess pressure (Pe); sheet angle is varied
throughout model runs
principal compressive stress (σ1). Boundary conditions were
used to test the model’s response to a simulated vertical
normal-fault dip-slip displacement (Fig. 11), modeled as a
compressive stress, of 5 MPa applied to the collapsing block,
and normal-fault dilation (Fig. 11), modeled as a tensile stress,
of 5 MPa applied perpendicular to the fault. Trajectories of the
maximum principal compressive stress (σ1) and the magnitude of the minimum principal compressive stress (σ3) are
shown in all model runs.
Model results
Several numerical models were run with different fault zone
setups, generated by varying the mechanical properties
(Young’s modulus) across the fault. A homogeneous setting,
whereby the fault zone shared the same mechanical properties
as the host rock, was initially modeled and provided a reference to compare other model results. In this setup, the fault has
little or no effect on sheet propagation. Additional model runs
are displayed throughout (Figs. 7, 8, 9, 10, and 11).
Several geometrical setups were run by varying the sheet
dip angle (θ), to confirm the relative effects of fault zone
mechanical heterogeneity against intrusion angle. We observe
that shallow-dipping sheets, those with dips less than 45°, do
not alter the local fault stress significantly enough to promote
deflection or arrest, unless additional boundary conditions are
applied (see Fig. 11). In all models, the maximum tensile
stress (σ3) occurs at the sheet tip, and the maximum compressive stress (σ1), shown as white cones, are used to interpret the
likely fracture pathway. All models show stress contours in the
range of 0.5–6 MPa, the upper end of this range being the
most likely to induce fracturing (Amadei and Stephansson,
1997).
When a mechanical contrast between the fault and host
rock is modeled, a rotation of principal stress occurs near the
sheet tip and a temporary change in principal stress orientation
at the contact between the fault and host rock. In order for the
sheet to become deflected into a subvertical dike along the
ring fault, as observed in Hafnarfjall, the values of Young’s
modulus must be sufficiently different between the individual
modeled fault zone layers, in order to induce an elastic mismatch. Figure 7 highlights a situation whereby the layers
which constitute the fault damage zone are not significantly
different from those of the host rock and thus promote neither
elastic mismatch nor principal stress rotation. This condition,
namely similar elastic properties of the fault zone and the host
rock, favors transverse sheet propagation through the fault. In
contrast, models shown in Figs. 8 and 9 simulate a more
mechanically heterogeneous fault zone, with Young’s modulus differing by orders of magnitude. In these situations, sheet
deflection is favored due to the mechanical mismatch between
the individual layers. The fracture propagation path at the
4
Page 8 of 13
Bull Volcanol (2015) 77:4
Fig. 7 Young’s modulus in the
fault zone ranges between 10 and
30 GPa. Sheet dips shown are
45°, 65°, and 75°, and sheet
overpressure is 10 MPa. In all
models, tensile stress concentrates
through the fault plane, indicating
that any fracture would likely
propagate through the fault.
Rotation of the maximum
principal compressive stress
(white cones) occurs in line with
the fault dip for those sheets that
dip 65° and greater, suggesting
that a fracture may align
preferentially with the fault. Large
stress shadows are created ahead
of fracture tip, as indicated by the
absence of stress trajectories
sheet tip is inferred from the trajectories of σ1 (which the sheet
follows), surrounding the sheet and fault zone. It is only when
these trajectories rotate to subvertical that the sheet will deflect
into a dike within the fault. Such a situation is much more
likely in models which simulate a sheet with an initially steep
dip (i.e., >65°).
To model the effect of a previously intruded dike in the ring
fault, a stiff layer is added to the center of the fault zone
(Fig. 10). Such a situation is likely if a lower part of the fault
was intruded by a ring dike in the conventional manner
described by Anderson (1936) and Walter (2008) or if inclined
Fig. 8 Young’s modulus in the
fault zone ranges between 0.1 and
10 GPa, and sheet overpressure is
10 MPa in all models. In all model
runs, tensile stress concentrates
within the 10-GPa layer closest
the sheet. Softer layers suppress
most of the tensile stress, and
principal stress rotation within the
fault core favors vertical sheet
deflection. This rotation effect is
much less pronounced for those
sheets which dip below 60–70°;
for example, the 65° model
shown displays inclined σ1
trajectories throughout the
vertical fault cross section
sheets were previously captured and deflected (see Appendix 1). The stiff layer creates a clear stress barrier and zone
of elastic mismatch, indicating that transverse sheet propagation
in this scenario is unlikely. Instead, any sheet would likely
become deflected or arrested at the contact between the weak
damage zone and the stiff dike (Fig. 10).
Boundary conditions were applied to the model edge to
simulate (1) extension across the fault and (2) block subsidence. A compressional stress of 5 MPa was applied to the
intra-caldera block (Fig. 11). In this model run, the modeled
principal stress axis rotates to subvertical with little
Bull Volcanol (2015) 77:4
Page 9 of 13 4
Fig. 9 Fault zone Young’s
modulus ranges from 1 to 30 GPa,
and sheet overpressure is 10 MPa
in all models. In models with
sheets that dip greater than ∼45°,
a significant rotation of the
maximum compressive stress is
observed. c, d The trajectories of
σ1 rotate to vertical within the
fault, whereas in b, trajectories are
initially vertical but become
inclined higher up the fault
dependence on the angle of the sheet; this situation would
favor sheet propagation into and along the plane of the ring
fault, thereby allowing magma to propagate in a manner
consistent with that interpreted from field observations, although principal stress rotation further up dip of the fault may
induce fracture arrest later on. The second boundary condition
simulates an extensional force acting perpendicular to the
fault, which would likely encourage fault dilation. In these
situations, it is much easier for a low-angle sheet to deflect into
the fault; in Fig. 11, we show a 35° dipping sheet.
Fig. 10 A stiff layer (40 GPa) is
added to the center of the fault
zone to simulate a previous dike
intrusion. Principal stress rotation
at the sheet tip and stress
concentration within the stiff layer
favor fault capture, most likely
along the nearside edge of the
previous intrusion. Sheet
approach angle has little effect on
the stresses within the fault zone
above 35°
Finally, a model is included to assess the background
boundary effects of a magma chamber subject to pressure in
excess of lithostatic, a necessary requirement for the propagation of a sheet. In this model (Fig. 12), an 8-km-wide and 1km-thick sill-like magma chamber is situated in a homogeneous crustal segment at 5-km depth. The magnitude and
orientation of stresses highlight the potential for inclined sheet
propagation, particularly at the chamber outer margins. Several softer layers were added to the model; these were placed
vertically to simulate the caldera fault zones previously
4
Page 10 of 13
Bull Volcanol (2015) 77:4
Fig. 11 Models showing the
effect of various boundary
conditions. Here, a tensile stress
(5 MPa) is applied to the far right
of the model (a) and a
compressive stress (5 MPa) to the
upper part of the intra-caldera
block in the model (b)
stress rotation and elastic mismatch parameters favor dike
propagation over continuation of an inclined sheet.
discussed and horizontally to simulate, in a simple way, mechanical layering in a volcanic edifice. Stresses clearly concentrate within the fault margins in the caldera block, indicating that faults act as stress barriers. There is little principal
stress rotation observed at the fault contact, indicating that if
the tensile stress was high enough (0.5–6 MPa) to generate a
fracture, then the likely propagation path would be through the
fault. However, the required stress concentration is rarely met
as the inner (core) part of fault tends to suppress stresses.
Once a fault captures an inclined sheet and deflects it into
the first subvertical dike, it is generally much easier for subsequent sheets to deflect at the contact between the lava pile
and the first dike (Fig. 3). This follows because the stiffness of
a vertical intrusion is much greater than that of the area
surrounding the fault damage zone, and therefore, principal
Discussion and conclusions
It is well known that normal faults can alter the propagation
pathway of subvertical dikes (Valentine and Krogh, 2006;
Gaffney et al., 2007; Ziv and Rubin, 2000) and sills (Magee
et al., 2013). Several assumptions and inferences have been
made regarding the role of caldera ring faults in channeling
magma (Anderson, 1936; Saunders 2001; Jonsson, 2009); the
most common mechanical explanation offered to decipher this
process relies on magma chamber underpressure (Anderson,
1936). This model suggests that many ring dikes form during
Free surface
X
X
Pe = 5 MPa
20 km
Fault
zone
E = 1 GPa
E = 10 GPa
Approximate area of interest
X
Pe = 5 MPa
X
a
c
40 km
X
E = 40 GPa
X
6
5
4
3
2
1
b
Fig. 12 Background stresses associated with a modeled a 8-km-wide and
1-km-thick sill-like magma chamber at 5-km depth with an overpressure of 5 MPa. b Principal stresses become oriented in a manner which
favors inclined sheet propagation from the chamber margins. c Model
setup includes a peripheral fault zone, with a lower Young’s modulus than
MPa
d
the surrounding host rock, as well as two soft horizontal layers at 1- and 2-km
depth. d Results indicate that the fault acts as a significant stress barrier.
Furthermore, the soft horizontal layers act in a similar manner, also
exhibiting significant principal stress rotation at the boundary of the
uppermost and softest layer
Bull Volcanol (2015) 77:4
Page 11 of 13 4
large explosive eruptions, leading to chamber roof collapse
and the flow of evolved magmas up into the ring faults.
Alternative mechanical models are hampered by the poor
availability of data on the structure of ring faults at depth
and the lack of detailed observations of the interactions
between ring faults and magma.
Here, we present an alternative mechanism of ring dike
formation which suggests that some ring dikes do not channel
magma directly from the margins of a magma chamber but
rather form through inclined sheets being captured and
deflected along the ring fault (Fig. 13). This mechanism is in
agreement with observations of restless calderas (e.g.,
Saunders, 2001) and the location of numerous dike-fed eruptive centers, located on ring fault margins (Walker, 1984;
Geyer and Martí 2008). However, several caldera volcanoes
experience eruptive activity located outside of their ring fault
margins, for example the fissure eruptions of Fernandina in
1995 (Chadwick et al., 2011). Our model supports the interpretation that the sheet which fed these fissures had a shallow
dip (12–14°) at a depth of ∼1 km, near the chamber margin
(Chadwick et al., 2011). This follows because sheet deflection
is unlikely at dip angles lower than ∼45° (see Figs. 7, 8, 9, and
10). Note, also, that many mafic calderas have complex ring
fault structures controlled by post-collapse subsidence, for
example Colli Albani (Giordano et al., 2006).
Dikes can become arrested and deflected at contacts
between layers with mechanical mismatch (contrasting mechanical properties) such as stiff lava flows and compliant
(or soft) tuff layers (Gudmundsson, 2011; see Fig. 12). It
has been suggested that elastic mismatch and associated
local stresses partly control the frequency with which dikes
reach the surface to feed eruptions (Gudmundsson and
Phillip, 2006). In this view, it is only when the state of
stress in a volcanic edifice becomes roughly homogeneous
that a dike can propagate to the surface, a condition more
likely to be reached in edifices composed of mechanically
similar layers. Similarly, if a caldera fault develops a fault
core, it may act as a single vertical and roughly homogeneous layer, thereby promoting stress field homogenization
and providing a pathway for magma channeling toward the
surface.
A large proportion of the world’s volcanism occurs within
or around active calderas (Newhall and Dzurisin, 1988), and
therefore, understanding the control of caldera structures on
magma movement is vital for predicting the location and
timing of eruptive activity. The caldera fault at Hafnarfjall is
one of the best exposures of its kind in the world. It represents
a segment of the ring fault, visible as a 200-m vertical exposure. The fault has variable attitude, but overall, it is a steeply
inward-dipping normal fault.
Many faults offer potential pathways for magma. However,
very steeply dipping faults, such as the ring fault in
Hafnarfjall, are perhaps particularly favored paths because
the normal stress on steep faults in rifting environments tends
to be comparatively small (Gudmundsson, 2011). In this
particular case, the ring fault has deflected and acted as a
channel for inclined sheet propagation; some of the sheets
may have reached the surface to supply magma for eruptions.
In addition, the ring fault has captured and deflected inclined
sheets to form a ring dike, a mechanism of ring dike formation
that has not been reported earlier. These results further underline the importance of understanding magma-fault interaction
in relation to volcanic hazards. It is likely that the process of
sheet deflection and ring dike formation, described here, provides a major control on the location of resurgent caldera
volcanism.
Inclined sheets captured in a
ring fault eventually become
feeder dikes
Arrested
dikes
Ring fault
Alternating stiff
and soft layers
Ring dike
Magma chamber
Shallow
dipping sh
eet
Outline of original chamber
Fig. 13 Conceptual resurgent caldera model. In the model, an edifice is
made of layers of contrasting mechanical properties, e.g., stiff lavas and
soft tuffs and sediments. A number of dikes fail to reach the surface and
become arrested at layer contacts. Inclined sheets propagating from the
shallow magma chamber become captured by the ring fault. The ring fault
deflects many sheets into subvertical dikes, some of which may have
propagated to the surface. Shallow-dipping sheets may penetrate the fault
to force eruptions outside of the ring fault margin
4
Page 12 of 13
Acknowledgments Fieldwork of JB was partly funded by the geologists’ association Baker-Arber Fund. We would like to thank Adelina
Geyer and an anonymous reviewer for their very helpful comments. We
also thank the editor Guido Giordano for his helpful suggestions. We are
grateful to Hjalti Franzson for providing his geological map of the area
and advice on field exposures, Hannah Reynolds and Zoe Barnett for
field assistance, and Jonathan Pownall for fruitful discussion.
References
Acocella V, Cifelli F, Funiciello R (2000) Analogue models of
collapse calderas and resurgent domes. J Volcanol Geotherm
Res 104:81–96
Acocella V (2007) Understanding caldera structure and development: an
overview of analogue models compared to natural calderas. Earth
Sci Rev 85:125–160
Amadei B, Stephansson O (1997) Rock stress and its measurement.
Chapman and Hall, New York
Anderson EM (1936) The dynamics and formation of cone-sheets, ringdikes, and cauldron-subsidence. R Soc Edinb Proc 128–157
Anderson TL (2005) Fracture mechanics: fundamentals and applications,
3rd edn. Taylor & Francis, London, p 621
Bedard JH, Naslund HR, Nabelek P, Winpenny A, Hryciuk M,
Macdonald W, Hayes B, Steigerwaldt K, Hadlari T, Rainbird R,
Dewing K, Girard E (2012) Fault-mediated melt ascent in a
Neoproterozoic continental flood basalt province, the Franklin sills,
Victoria Island, Canada. Geol Soc Am Bull 124:723–736
Branney MJ (1995) Downsag and extension at calderas: new perspectives
on collapse geometries from ice-melt, mining, and volcanic subsidence. Bull Volcanol 57:303–318
Burchardt S, Walter TR (2010) Propagation, linkage, and interaction of
caldera ring-faults: comparison between analogue experiments and
caldera collapse at Miyakejima, Japan, in 2000. Bull Volcanol 72:
297–308
Chadwick WW, Jonsson S, Geist D, Poland M, Johnson DJ, Batt S,
Harpp KS, Ruiz A (2011) The May 2005 eruption of Fernandina
volcano, Galapagos: the first circumferential dike intrusion observed
by GPS and InSAR. Bull Volcanol 73:679–697
Clough CTH, Maufe HB, Bailey EB (1909) The cauldron subsidence of
Glen Coe and the associated igneous phenomena. Q J Geol Soc
Lond 65:611–678
Delaney PT, Pollard DD, ZIony JI, McKee EH (1986) Field relations
between dikes and joints: emplacement processes and paleostress
analysis. J Geophys Res 91:4920–4938
Ekstrom G (1994) Anomalous earthquakes on volcano ring-fault structures. Earth Planet Sci Lett 128:707–712
F r a n z s o n H ( 1 9 7 8 ) S t r u c t u r e a n d p e t r o c h em i s t r y o f t h e
Hafnarfjall_Skardsheidi central volcano and the surrounding basalt
succession, W-Iceland [Ph.D. thesis]: Edinburgh, Scotland, university of Edinburgh 264 pp
Gaffney ES, Damjanac B, Valentine GA (2007) Localization of volcanic
activity: 2. Effects of pre-existing structure. Earth Planet Sci Lett
263:323–338
Gautneb H, Gudmundsson A, Oskarsson N (1989) Structure, petrochemistry and evolution of a sheet swarm in an Icelandic central volcano.
Geol Mag 126:659–673
Geyer A, Folch A, Martí J (2006) Relationship between caldera collapse
and magma chamber withdrawal: an experimental approach. J
Volcanol Geotherm Res 157:375–386
Geyer A, Marti J (2014) A short review of our current understanding of
the development of ring faults during collapse caldera formation.
Front Earth Sci 2:22. doi:10.3389/feart.2014.00022
Bull Volcanol (2015) 77:4
Geyer A, Martí J (2008) The new worldwide collapse caldera database
(CCDB): a tool for studying and understanding caldera processes. J
Volcanol Geotherm Res 175:334–354
Geshi N, Shimano T, Chiba T, Nakada S (2002) Caldera collapse during the
2000 eruption of Miyakejima volcano, Japan. Bull Volcanol 64:55–68
Giordano G, De Benedetti AA, Diana A, Diano G, Gaudioso F, Marasco
F, Miceli M, Mollo S, Cas RAF, Funiciello R (2006) The Colli
Albani mafic caldera (Roma, Italy): stratigraphy, structure and petrology. J Volcanol Geotherm Res 156:49–80
Gudmundsson A (1998) Formation and development of normal-fault
calderas and the initiation of large explosive eruptions. Bull
Volcanol 60:160–170
Gudmundsson A, Nilsen K (2006) Ring-faults in composite volcanoes:
structures, models and stress fields associated with their formation.
Geol Soc Lond, Spec Publ 269:83–108
Gudmundsson A, Berg SS, Lyslo KB, Skurtveit E (2011) Fracture networks
and fluid transport in active fault zones. J Struct Geol 23:343–353
Gudmundsson A (2011) Rock fractures in geological processes.
Cambridge University Press, Cambridge. doi:10.1017/
CBO9780511975684
Gudmundsson A, Phillip SL (2006) How local stress fields prevent
volcanic eruptions. J Volcanol Geotherm Res 158:257–268
Hardy S (2008) Structural evolution of calderas: insights from twodimensional discrete element simulations. Geology 36:927
He MY, Hutchinson JW (1989) Crack deflection at an interface between
dissimilar elastic materials. Int J Solids Struct 31:3443–3455
He MY, Evans AG, Hutchinson JW (1994) Crack deflection at an
interface between dissimilar elastic materials. Int J Solids Struct
25:1053–1067
Holohan EP, Troll VR, Walter TR, Münn S, McDonnell S, Shipton ZK
(2005) Elliptical calderas in active tectonic settings: an experimental
approach. J Volcanol Geotherm Res 144:119–136
Holohan, EP., Walter, TR., Schöpfer, MPJ., Walsh, JJ., van Wyk de Vries,
B. and Troll, VR. (2013). Origins of oblique-slip faulting during
caldera subsidence. J Geophys Res Solid Earth, No. 2, p. n/a–n/a
Jaeger JC, Cook NGW (1979) Fundamentals of rock mechanics.
Chapman and Hall, London
Johannesson H (1975) Structure and petrochemistry of the Reykjadalur
central volcano and surrounding areas, midwest Iceland. PhD
Thesis, University of Durham, Durham, 273 pp
Johnson SE, Schmidt KL, Tate MC (2002) Ring complexes in the
Peninsula Ranges Batholith, Mexico and the USA: magma
plumbing systems in the middle and upper crust. Lithos 61:
187–208
Jonsson S (2009) Stress interaction between magma accumulation and
trapdoor faulting on Sierra Negra volcano, Galapagos.
Tectonophysics 471:36–44
Kennedy B, Stix J, Vallance JW, Lavallée Y, Longpré MA (2004)
Controls on caldera structure: results from analogue sandbox modeling. Geol Soc Am Bull 116:515
Kokelaar P (2007) Friction melting, catastrophic dilation and breccia
formation along caldera superfaults. J Geol Soc 164:751–754
Lipman, PW (1984) The roots of ash flow calderas in Western North
America: Windows into the tops of Granitic batholiths. J Geophys
Res 89:8801–8841
Lipman PW (1997) Subsidence of ash-flow calderas: relation to caldera
size and magma chamber geometry. Bull Volcanol 59:198–218
Magee C, Jackson CAL, Schofield N (2013) The influence of normal
fault geometry on igneous sill emplacement and morphology.
Geology 41:407–410
Michon L, Villeneuve N, Catry T, Merle O (2009) How summit calderas
collapse on basaltic volcanoes: new insights from the April 2007
caldera collapse of Piton de la Fournaise volcano. J Volcanol
Geotherm Res 184:138–151
Newhall CG, Dzurisin D (1988) Historical unrest at large calderas of the
world. US Geol Sur Bull 72:85–100
Bull Volcanol (2015) 77:4
Saunders ST (2001) The shallow plumbing system of Rabaul caldera: a
partially intruded ring fault? Bull Volcanol 63:406–420
Saunders ST (2004) The possible contribution of circumferential fault
intrusion to caldera resurgence. Bull Volcanol 67:57–71
Smith RL, Bailey RA (1968) Resurgent cauldrons. Geol Soc Am Mem
116:613–662
Sparks RSJ (1988) Petrology and geochemistry of the Loch Ba ring-dike,
Mull (NW Scotland): an example of the extreme differentiation of
theolitic magmas. Contrib Mineral Petrol 100:446–461
Troll V, Walter TR, Schmincke HU (2002) Cyclic caldera collapse: piston
or piecemeal subsidence? Field and experimental evidence. Geology
30:135–138. doi:10.1130/0091-7613(2002)030<0135
Tucker D, Hildreth W, Ullrich T, Friedman R (2007) Geology and
complex collapse mechanisms of the 3.72 Ma Hannegan caldera,
North Cascades, Washington, USA. Geol Soc Am Bull 119:329–
342. doi:10.1130/825904.1
Valentine GA, Krogh KEC (2006) Emplacement of shallow dikes and
sills beneath a small basaltic volcanic center—the role of pre-
Page 13 of 13 4
existing structure (Paiute Ridge, southern Nevada, USA). Earth
Planet Sci Lett 246:217–230
Walker GPL (1960) Zeolite zones and dike distribution in relation to the
structure of the basalts of eastern Iceland. J Geol Soc 68:515–527
Walker GPL (1984) Downsag calderas, ring faults, caldera sizes, and
incremental growth. J Geophys Res 89:8407–8416
Walter TR (2008)Facilitating dike intrusions into ring-faults. In:
Gottsmann J, Joan Martí (eds) Caldera volcanism: analysis,
modelling and response, vol 10. Elsevier, Heidelberg, pp 351–374
Walter TR, Troll VR (2011) Formation of caldera periphery faults: an
experimental study. Bull Volcanol 63:191–203
Wang P, Xu LR (2006) Dynamic interfacial debonding initiation induced
by an incident crack. Int J Solids Struct 43(21):6535–6550
Xu LR, Huang YY, Rosakis AJ (2003) Dynamic crack deflection and
penetration at interfaces in homogeneous materials: experimental
studies and model predictions. J Mech Phys Solids 51:461–486
Ziv A, Rubin AM (2000) Stability of dike intrusion along pre-existing
fractures. J Geophys Res 105
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Chapter 5
Manuscript in preparation for journal submission
Cooling dominated cracking in thermally stressed volcanic rocks
Browning, J, Meredith, P.G., and Gudmundsson, A
Statement of contribution
Initial idea for experiments from all authors
Experimental apparatus designed by all authors with input from technicians at UCL
JB conducted all experiments with initial assistance from PGM
JB completed 1st draft of manuscript
JB made all figures and images
Data interpretation with input from co-authors
JB collected and interpreted all sample characterisation data
55
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Cooling dominated cracking in thermally stressed
volcanic rocks
John Browning1&2*, Philip Meredith1, Agust Gudmundsson2
1
Department of Earth Sciences, University College London, London
2
Department of Earth Sciences, Royal Holloway University of London, Egham,
TW20 0EX, United Kingdom
*e-mail:
[email protected]
Abstract
Several hypotheses have been proposed regarding the role of thermo-mechanical
contraction in producing cracks and joints during cooling of volcanic rocks.
Nevertheless, most studies of thermally-induced cracking to date have focused on the
generation of cracks formed during heating and thermal expansion. In this latter case,
the cracks are formed under an overall compressional regime. By contrast, cooling
cracks are formed under an overall tensile regime. Therefore, both the nature and
mechanism of crack formation during cooling are hypothesised to be different from
those for crack formation during heating. Furthermore, it remains unclear whether
cooling simply reactivates pre-existing cracks, induces the growth of new cracks, or
both.
We present results from experiments based on a new method for testing ideas on
cooling-induced cracking. Cored samples of volcanic rock (basaltic to dacitic in
composition) were heated at varying rates to different maximum temperatures inside
a tube furnace. In the highest temperature experiments samples of both rocks were
raised to the softening temperature appropriate to their composition, determined
using thermal mechanical analysis, forcing melt interaction and crack annealing. We
present in-situ acoustic emission data, which were recorded throughout each heating
and cooling cycle. It is found consistently that the rate of acoustic emission is much
1
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
higher during cooling than during heating. In addition, acoustic emission events
produced during cooling tend to be significantly higher in energy than those
produced during heating. We therefore suggest that cracks formed during cooling are
significantly larger than those formed during heating. Seismic velocity comparisons
and crack morphology analysis of our cyclically heated samples provide further
evidence of contrasting fracture morphologies. These new data are important for
assessing the contribution of cooling-induced damage within volcanic structures and
layers such as sills and lava flows. Our observations may also help to constrain
evolving ideas regarding the formation of columnar joints.
1. Introduction
Crustal segments hosting magma chambers experience complex stress regimes,
generated by the often combined effects from regional tectonics and local magmatic
intrusions. Magma residing within a chamber at depth exerts stress within an edifice
due to changes in pressure as a consequence of bulk volume change due to magma
supply from a deeper source (Gudmundsson, 1998) or due to volatile exsolution
(Turner et al., 1983). An additional but perhaps less well understood inputs are the
effect of thermal expansion and contraction of rocks hosting magmatic intrusions.
Any thermal stressing produces damage in rocks (David et al., 1999), but in volcanic
systems these stresses, like mechanical stresses are likely to be generated cyclically
(Heap et al., 2013b) by repeat intrusion and extrusion of magma. The cyclic process
may combine to produce an additive stress thereby contributing to instability, as well
as influencing the location and pathway of magmatic flow (Chouet, 1996) but there
may be annealing and healing of fractures at high temperature in between events.
Normal faults for example, are thought to commonly nucleate from cooling joints
(Acocella et al., 2000). When magma is intruded into a cooler host rock, its time of
solidification is proportional to the thickness of the intrusion in the 2nd power
(Gudmundsson, 2011), with the cooling surface generally located perpendicular to
the direction of propagation. Dykes and inclined sheets will tend to intrude preexisting weaknesses such-as cooling joints (Figure. 1); and therefore understanding
the mechanics of cooling induced fracturing and joint formation is important in
forecasting magma pathways. Targeted injection of cool fluids is a technique
2
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
employed by the geothermal industry to force rapid contraction of the host rock
around a borehole and force pre-existing cracks to dilate and re-open or to induce the
formation of new tensile cracks (Axelsson et al., 2006; Brudy and Zoback, 1999;
Kitao et al., 1990; Tarasovs and Ghassemi, 2012). The aim is to increase
permeability and fracture surface area and thereby enhance the efficiency of the
geothermal system (Axelsson et al., 2006). Non-double couple earthquakes have
been related to cooling and tensile fractures induced by cooling fluid injection at
geothermal sites (Julian et al., 2010), and by cooling contraction of magma chambers
(Miller et al., 1998). Field relations can indicate the mechanisms of cooling-induced
fracturing but it is often difficult to characterise the dynamic processes involved. For
example many aspects concerning the formation of columnar joints still remain
unclear (Hetényi et al., 2012). In addition to deep sub-surface processes, it is likely
that cooling related fractures play a key role in the degassing of magma in shallow
conduits (Tuffen and Dingwell, 2005; Tuffen et
al, 2003) as well as at the surface
in viscous domes and lava flows (Cabrera et al, 2011). In all these cases it is clear
that understanding volcanic systems requires knowledge of the physical properties of
volcanic rocks as a function of stress, pressure and temperature.
When subjected to a change in temperature a rock mass will experience fracture
when the thermal stresses caused by expansion or contraction of individual grains in
contact with other grains become high enough to exceed the localised tensile or shear
strength (Figure 2). These thermal stresses are generated by two main mechanisms,
1) the mismatching thermal expansion co-efficient of mineral grains, and 2)
differential mineral grain orientation and anisotropy. Consider a mono-mineralic
rock in which the individual grains are randomly oriented, here stress builds up
between the grains as they expand, through heating, in direction directions. In a polymineralic rock the situation above still likely applies but now each of the grains
expands through thermal expansion at different rates and therefore contributes
further strain and resultant stress. A less important mechanism with respect to this
study is that of thermal shock whereby rocks are subjected to a very high rate of
heating or cooling. All the experiments reported here heat and cool the rocks
significantly slowly as to not induce thermal shock.
3
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
A salient previous study with respect to this work presented is that conducted by
Vinciguerra et al (2005), in which the seismic properties of an intrusive basalt from
Iceland were compared with a basalt from a lava flow on Mt Etna. The study found
that P-wave velocities decreased in both rock types as a consequence of thermal
stressing; in this case velocity decrease was associated with the growth of microcracks. Velocity decrease was noticeably higher for the Icelandic basalt (around 2.0
km s-1) in comparison to the Etnean basalt (negligible), indicating that faster cooled
extrusive Etna basalt contained significantly more crack damage than the slow
cooled intrusive basalt prior to thermal stressing (Vinciguerra et al, 2005). The
results are used to explain the low seismic velocities (approximately 3-4 km s-1)
often
found
in
basaltic
volcanic
edifices.
Cracking
was
monitored
contemporaneously by acoustic emission output during the heating phase.
A difficulty with studying cooling induced cracking in the laboratory is that any rock
must first be heated, and therefore any end member produced will show a composite
pattern of cracks generated during both the heating and cooling cycle. We therefore
provide contemporaneous measurements of acoustic emissions (AE) which act as a
proxy for the number and relative size of micro-cracks. This dataset is
complemented by static measurements of ultrasonic wave velocities and micro-crack
image analysis of rocks both pre and post heat-treatment.
4
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 1. A) Large cooling joints formed during cooling contraction of a sill in
Stiflisdalsvatn, South-West Iceland. B) A sub-vertical dyke intrudes a layer of subvertically oriented cooling joints at Brekkufjall, West Iceland. C) A Sub-vertical
dyke intrudes a Hyalaclastite unit in Anaga, Tenerife, principle cooling direction is
shown as blue arrows, perpendicular to the direction of emplacement.
Figure 2. Individual grains expand during heating, when the grains come into
contact the resultant stress induces fracture and the propagation of mixed mode
cracks. During cooling, the individual grains contract which is likely to induce mode
I tensile cracks although the process is not fully understood.
5
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Here we present a series of experiments using a new apparatus designed to acquire
acoustic emissions during expansion and contraction of thermally treated cores of
volcanic material. These experiments are designed to test ideas on the magnitude and
frequency of cracking as a consequence of thermal stresses, with a direct comparison
between expansive and contractive stresses. Acoustic emissions, ultrasonic wave
velocities and micro-crack analyses are used as proxies to understand likely fracture
modes.
2.1 Sample characterisation and preparation
The type of material used in any high temperature experiment that induces melting
must be carefully chosen. The three rocks selected for this study are all fine grained
igneous rocks. Icelandic basalt (IB) has been widely used in rock physics studies (e.g
Vinciguerra et al. 2005) and so the chemical and mechanical properties of this rock
type are well known. IB is an intrusive tholeitic basalt dominated by an intergranular
matrix of plagioclase, granular pyroxene and iron oxides. Partially oriented
plagioclase is found along with a rare abundance of augite, olivine and an interstitial
glass phase. IB has a density of 2900 kg/m3 ± 10 kg/m3 and an initial porosity of
around 4 %. A section of a phonolitic dyke from the Anaga province of Tenerife
(AP) was additionally selected as an ideal material for comparison with the mafic
Icelandic basalt. AP has a trachytic and porphyritic texture dominated by amphibole,
feldspar and sanidine with minimal abundance of glass phase. AP has a density of
2300 kg/m3 and an initial porosity of around 7 %. The final rock type is a silica rich
(Table 1) Dacite lava from the 1939-40 Reck flow on Nea Kameni (NKD), Santorini
(Pyle and Elliott, 2006). The dacite lavas are petrophysically well studied (e.g Barton
and Huijsmans, 1986), however remarkably little work has been conducted on the
mechanical properties of Santorini’s recent lavas. NKD is a glass bearing dacite with
a density of 2200 kg/m3 ± 10 kg/m3 and an initial porosity of around 10%. By nature
the lavas are extrusive and have therefore undergone a significantly contrasting
cooling history to those previously described rocks. NKD has a porphyritic texture
dominated by plagioclase, pyroxene, iron oxides and rare sanidine.
Samples were cut into ~65 mm in length and 25 mm in diameter cores with the
surfaces ground using a surface grinder to ensure parallelism and smoothness of the
two end surfaces.
6
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 3. Optical light microscope images in ppl (left) and xpl (right). Samples are
IB, AP and NKD from top to bottom.
Table 1. Whole rock x-ray fluorescence results for materials used within this study.
7
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
2.2 Thermal characterisation
As we are heating each rock to high temperature with the aim of partly annealing
fractures then it is important to have an understanding of each materials baseline
thermal properties. We determined thermal expansion co-efficients (α) and material
softening by thermo-mechanical analysis (TMA) using a Netzsch TMA 402 thermomechanical analyser. In Fig. 4b we show the total thermal expansion plotted as
change in length of each sample type and thermal expansion co-efficient (α) values
as a function of temperature (Fig. 4a) and time (Fig. 4c). Generally there is very little
variation in α between heating and cooling, provided that the temperature throughout
each sample is homogeneous. α is similar in all three rocks and ranges commonly
from around 15 x 10-6/˚C to 25 x 10-6/˚C apart from a deviation and increase at
around 800˚C in NKD and 1100˚C in IB and AP (Figure 4). TMA was used to
characterise each materials softening point (Fig. 5), which is in turn defined as a
point at time and temperature where the material begins to show negative
expansivity as a result of a small force (3 N) applied to each end. We take this point
to indicate when the material is beginning to form melt phases and behave plastically
or as viscoelastically. A sharp softening point is easily observable in NKD at 800˚C,
which corresponds well with the measured temperature of Tg (glass transition) from
separate differential scanning calorimetry measurements (Appendix A). The precise
softening points of IB and AP are less well defined, in fact AP appears to exhibit two
softening points, and IB may well too, however in order to avoid bulk melting and
potential damage to the equipment tests were concluded at 1130˚C. The apparent
stiffening of AP may be related to expansion due to volatile exsolution and resultant
pressure increase. For all standard thermal stressing tests (section 3.1) we take the
softening point of each rock type to be the maximum hold temperature used, that is
1100˚C for IB and AP, and 800˚C for NKD.
Once each material’s softening point was calculated we then estimate the time taken
for any melt within the samples to relax into pre-existing cracks or those cracks
formed during heating. The timescale of viscous relaxation (τ) of melt into cracks
and pores was calculated from (Dingwell and Webb, 1989):
G
(2)
8
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
where η is melt viscosity and G∞ is infinite frequency elastic shear modulus,
normally given as 1010±0.5 GPa for silicate melt (Dingwell and Webb, 1989; Hess et
al., 2008). Melt viscosity is calculated using ViscosityCalc (Giordano et al., 2008)
with a standard value of 0.5 wt% H2O (Barton and Huijsmans, 1986) and whole
rock chemical composition obtained from XRF analysis (Table 1). We obtain a
viscosity of 107.53 Pa s for dacite at 800˚C containing 0.5 wt% H2O, which gives a
relaxation time of ~0.003 s. To allow significant relaxation and therefore the best
chance for annealing, we use the common convention that a sample should be left at
temperature for ≥103τ, which in this case gives ~30 minutes. The model of Giordano
et al (2008) falls down when using comparatively low silica content rocks such as
basalt and to a lesser extent phonolite, where we obtain unrealistic melt viscosities of
101.64 Pa s and 103.77 respectively at the softening point for each material, previously
measured as 1130˚C.
Figure 4. Thermo-mechanical analysis results. (a) Temperature profiles for each
material tested, (b) thermal strain during heating and cooling at 8˚C/min, (c) thermal
expansion co-efficient (α) as a function of time and resultant temperature.
9
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 5. Thermal strain and resultant material softening, indicated at ~800˚C for
NKD, and 1100˚C for AP and IB, although AP appears to exhibit an earlier lower
temperature softening.
2.3 Experimental apparatus
Temperature profiles were controlled using a two-stage programmable Eurotherm
808 controller attached to a tube furnace. Samples were heat treated inside the
Carbolite CTF12/75/700 tube furnace, which is capable of reaching temperatures up
to 1200 oC. Sample cores were held within a purpose built jig manufactured from a
310 steel alloy capable of sustaining temperatures up to 1100 oC without significant
corrosion. The jig is 1.1 m in length and comprises a series of rods and springs to
hold the sample under constant end-load within the central, uniform temperature
section of the furnace, as illustrated in Figure 7. The central rods act as waveguide
which had to be of sufficient length so that the transducers and springs were held
outside of the furnace to remain cool. Those springs allow the central rods to move
in response to sample expansion and contraction during heating and cooling and
therefore maintain a flush contact throughout the experiments. All experimental runs
are conducted at 1 atm, as the setup was not in its current state, capable of adding a
confining pressure.
10
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 6. Schematic diagram of the experimental arrangement used for our thermal
stressing experiments.
2.4 Methodology
In all thermal stressing tests samples were heated at set constant control rates of 1, 4
and 8˚C/min in order to investigate the effect of differential heating rate. The
samples were then held at a maximum temperature determined from TMA for length
τ, in most standard tests this is around 30 minutes, although for targeted tests the
hold time varies. Samples were then cooled to room temperature at a maximum of
4˚C/min, but cooling rates varied. This is because the sample assembly did not make
it possible to cool faster than a natural cooling rate (Figure 7). Two rates of cooling
were performed, one constant control rate of 1˚C/min and one variable natural
cooling rate.
11
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 7. Programmed and actual heating and cooling profiles for a range of
temperature ramp rates, A) 1˚C/min, B) 4C/min and C) 8˚C/min. Temperatures
were measured at the sample surface. A green line indicates the temperature at which
cooling rates decrease below 1˚C/min, the slowest programmed rate of cooling.
In order to capture dynamic crack growth and nucleation events during our thermal
stressing tests we record contemporaneous acoustic emissions (AE). One
Panametrics V103 piezoelectric P-wave transducer was attached at the end of the
steel wave guide; the signal was passed through a preamplifier and recorded using a
Vallen AMSY-5 connected to a PC. Each discrete AE event is termed an AE hit.
Within any one AE hit the recovered waveform can be used to measure the
amplitude (dB) and duration (μm) to calculate discrete AE energy by summing the
envelope of the AE waveform (see Cox and Meredith 1999 for a detailed description
of the AE recording method) which acts as a proxy for the relative size of fracture
events.
12
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
3. 0 Results
We report on two main types of experiments that record contemporaneous AE during
thermal stressing tests, and then show results from static image analysis and
ultrasonic wave velocity tests conducted on pre- and post- heat treated end member
samples. The first set of AE tests are systematic or standard tests which heat each
rock type to a maximum temperature defined from earlier TMA measurements, held
for 30 minutes and then cooled at a set rate. The second set of tests are targeted to
report on the effect of hold times and maximum temperature predominantly in NKD.
Targeted tests are also reported in the discussion where we consider the Kaiser
‘thermal memory’ effect. In Figure 8 we show the temperature profile and AE
generated during a standard test on IB which has been heated at a rate of 8˚C/min
held at 1100˚C for 30 minutes and then cooled at a natural cooling rate. We show the
AE hit rate (Fig. 8a) which is averaged per 10 AE hits and given as discrete AE
hits/h, although not shown here, the same method can be applied to calculate AE
energy rate. Instead we show discrete AE amplitude (Fig. 8c) and discrete AE energy
(Fig. 8e), amplitude has also been separated to show the number of AE hits that
occur as a function of amplitude (Fig. 8d). Finally AE hits and energy are given as
cumulative plots (Fig. 8b and f). In this initial test we find that the maximum
amplitude, energy and rate of AE hits occurs during the cooling cycle. A sharp
inflection on both cumulative plots indicates an increase in the total number (Fig. 9b)
and relative size (Fig. 8f) of AE hits at around 800˚C during the cooling cycle.
Throughout we choose to use AE energy as the best proxy for cracking. Whilst
previous studies have shown AE hit events and amplitudes (e,g Vinciguerra et al.,
2005), in many of our studies AE is characterised by an increased rate of AE hits and
higher amplitude hits, in which case the energy output during this period is
significantly higher, so therefore AE energy gives the best indications.
13
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 8. Standard acoustic emissions dataset for a basalt sample heated at 4˚C/min
and cooled at a natural rate. a) discrete acoustic emissions hit rate, b) cumulative
number of acoustic emissions hits, c) discrete acoustic emission amplitude (dB), d)
number of acoustic emission hits binned as a function of amplitude (dB), e) discrete
acoustic emission energy (arbitrary units), f) cumulative acoustic emissions energy.
3.1 Standard tests
In Fig. 9 we plot AE energy as function of time and temperature for all materials and
at two heating and cooling rates, 1C/min and 8C/min, for clarity 4˚C/min is given
in Appendix A. In all tests, the relative size and rate of AE hits are notably higher
during the cooling cycle of thermal treatment than the heating cycle. The onset of
sustained high AE hit energy is very similar during cooling in all experimental runs
and occurs at around 800C, apart from in the slowest cooled basaltic sample where
the onset occurs around 650C (Figure. 9). AE hit onset, rate and energies are not
14
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
significantly different between the two cooling rates tested in all rock types.
Although IB and AP produce a larger magnitude and frequency of AE events as a
consequence of faster heating, the effect is less apparent in NKD. The lowest AE hit
rate consistently occurs during the maximum temperature hold period, which
suggests thermal equilibration and therefore minimal thermal stress occurs at this
point.
Figure 9. Acoustic emission hit energy as a function of temperature for three rock
types (Basalt, IB, Phonolite, AP, Dacite, NKD) and two different heating (1C/min –
and 8C/min) and cooling rates (black dashed line). Onset of high energy bursts
occur at the onset of cooling (blue lines) with most energy generated around 800˚C
upon cooling. Time in hours is scaled differently for each heating and cooling rate
test to encompass the full range of results. Similarly AE hit energy is shown on
different scales for the same reason.
15
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
In table 2 we compare the total number of acoustic emission hits and resultant energy
for each rock type during a 1˚C/min heating and cooling cycle. As both the heating
and cooling cycle in these tests have the same duration, it is useful to compare direct
totals. A direct comparison of the number of AE hits is not possible in those tests
where a natural cooling rate is induced as the duration of cooling is much greater
than during heating, in this case we prefer to look at the AE hit rate or AE energy
rate. In table 2 AE hit and energy rate are values averaged over the full heating and
cooling duration. In all rock types it is clear to see that there are a significantly
greater number of AE generated during cooling.
Table 2. AE hit and energy totals and rates for a 1C/min
thermal stressing experiment.
3.2 Targeted tests
To test the effect of high temperature residence timescales and degree of potential
fracture annealing we perform three tests of samples of NKD. This sample type was
chosen as it is characterised by the lowest temperature softening point. In the first
test a sample of NKD was heated to 800˚C, the experimentally determined softening
point, and held for ~1 minute (Figure 10). The AE from this test where then
compared with results from the AE hits occurring when a different sample of NKD
was held for ~ 2 hours at the softening point (Figure 11). AE hit energy are similar in
both tests, thereby indicating that either the relaxation timescale in both tests was
sufficient to anneal a similar amount of fractures, or that the melt was still too
viscous at this temperature range and therefore the annealing timescale was too small
16
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
in both cases. As the time taken to accumulate events in each cycle is dependent on
the temperature differential we also consider the number of hits and amount of AE
energy as a function of time, i.e a rate of AE hits (Fig 10b and Fig 11b). A noticeable
decrease in AE output is observed at high temperatures (>800˚C), corresponding to
either a) the material starting to behave plastically or more likely b) the reduction of
thermal stress due to sufficient thermal equilibration as previously noted in the
standard tests. When the cumulative number of AE events or AE energy is plotted
against time (Fig. 10c and Fig. 11c) we notice a consistent plateau during the
temperature hold phase, consistent with thermal stress reduction.
Figure 10. Acoustic emission energy and events during a heating and cooling cycle
in NKD, held at 800˚C for ~1 min. a) discrete AE hit energy as a function of time
and temperature. b) AE hit energy rate. c) cumulative number of AE hits and d)
frequency of AE hits as a function of amplitude.
17
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 11. Acoustic emission energy and events during a heating and cooling cycle
in NKD, held at 800˚C for ~2 hours. a) discrete AE hit energy as a function of time
and temperature. b) AE hit energy rate. c) cumulative number of AE hits and d)
frequency of AE hits as a function of amplitude.
In Fig. 12 we show the results of a final annealing test whereby NKD was heated to
200˚C above the determined softening point, producing a melt viscosity which
ranges from ~108 Pa s to ~105 Pa s over a period of approximately 2 hours (Fig. 12c
inset). Upon cooling we note very high energy bursts of AE occurring around 600˚C,
whilst the general pattern of AE hit amplitude, energy and rate is similar to previous
tests we do note a number of slightly higher amplitude events and an increased
frequency of AE hits. At viscosities of the order of ~ 105 Pa s we would expect the
melt to be on the cusp of flowing and therefore infer that some fracture annealing is
highly likely. We therefore conclude that because the AE hit response in this final
test is similar to that of the previous two targeted tests, then sufficient fracture
annealing took place even with the smallest length of high temperature hold.
18
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 12. Acoustic emission energy and events generated during a heating and
cooling cycle, held at 1000˚C for 30 minutes. NKD lies within an annealing window
(c) for approximately 2 hours in which time melt viscosity continually decreases
(inset) thereby increasing the likelihood of viscous annealing. a) discrete AE hit
energy as a function of time and temperature. b) AE hit energy rate. c) cumulative
number of AE hits and d) frequency of AE hits as a function of amplitude.
3.3 Acoustic wave velocities
Radial P-wave velocities as a function of azimuth are reported in Figure 13. Values
for non-heat treated (NHT) Icelandic basalt (IB) range from 5.42 to 5.55 km/s,
Anaga Phonolite (AP) between 4.38 to 4.58 km/s and Nea Kameni Dacite (NKD)
from 5.27 to 5.31 km/s. Within individual samples, anisotropy value (A) ranges from
0.01 to 0.09 indicating that grain and crack populations are relatively isotropic.
Samples were heat treated to the maximum temperature determined from TMA, i.e
1100˚C for IB and AP, and 800˚C for NKD. The degree of internal sample isotropy
remained following heat treatment, but P-wave velocities decreased in all samples.
Velocity decreases are most significant in the intrusive IB and AP rocks where V p
drops by around 45%, whereas the change whilst still significant, is less so in NKD
19
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
where Vp drops by 21%. With an initial P-wave velocity of 5.29 km/s in NKD, an
extrusively cooled lava, this is unusually faster than the intrusively cooled AP.
Although the drop in P-wave velocity is much less in NKD than the other rock types,
we would expect the initial value to be substantially less too, owing to an abundance
of pre-existing micro-cracks and pores.
Table 2. P-wave velocities in Icelandic basalt (IB), Anaga
phonolite (AP) and Nea Kameni dacite (NKD) tested with no heat
treatment (NHT) and following heat treatment (HT).
Figure 13. Radial P-wave velocities for all materials no heat treatment (NHT) (solid
line) and heat treated (HT) (dashed line). In all cases P-wave velocity decreases
following heat treatment, but the effect is less apparent in NKD. All samples are
relatively isotropic.
3.4 Crack sizes and morphologies
Crack morphologies observed under scanning electron microscope were analysed
and mapped using a MATLAB measuring code (see Mitchell & Faulkner 2008 for
20
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
further details). Individual fractures identified from SEM images were imported into
Adobe illustrator. Fracture lengths and orientations are manually recorded and then
counted using the MATLAB code. The method does not provide a fracture density
but it allows a quantitative description of the relative total number and size of
fractures across comparable sample types.
In order to reduce bias in user
measurements, all images for fracture analysis and counting were chosen randomly.
When quoting numbers of micro-cracks we are referring to the number found within
the randomly chosen square. The size of the square changes between each sample
type, but remains the same between samples of the same composition and can
therefore be used for direct comparison. Although it is important to note that the
analysis of heat treated and non-heat treated rocks is made on separate samples, and
therefore whilst we draw comparisons and parallels between the two, the method is
subjective and prone to issues regarding sample heterogeneity. The technique can
therefore be considered at best, complimentary to the previously described results
and techniques.
SEM images of end member pre- and post- heat treatment samples were analysed in
terms of the numbers of micro-cracks. Non-heat treated IB and AP samples exhibit
an abundance of micro-cracks evenly distributed throughout the samples. The cracks
range in size from approximately 25 μm to 150 μm in length. Following heat
treatment it is found that the length of fractures in these sample types does not
increase significantly but the total number of fractures observed increases markedly.
In Fig. 14 we show a crack map indicating the location of all micro-cracks counted
in a sample of heat treated IB. Cracks are isotropically oriented and distributed
throughout. The total number of cracks recorded in this sample was 420, which was
compared to a randomly picked section of a non-heat treated basalt sample where
282 cracks were measured. This represents a 42 % increase in the number of microcracks observed. It should be noted that crack length increases only within the
measurement error, rising from 78 μm in NHT IB to 82 μm in HT IB. We estimate
the error in individual measurements to be approximately 5 μm found through repeat
measurements. A comparable level of micro-crack frequency increase is found in AP
were the total number of fractures rises 70% following heat treatment, but less so in
NKD where the increase is around 20%. Micro-cracks in NHT AP tend to be slightly
longer than those in HT AP (Figure 15). NKD initially contains very few micro21
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
cracks (<61), but a much greater abundance of pores prior to heat treatment, those
fractures which are present tend to be quite large (> 200 μm) and orientated
preferentially along an azimuth at approximately 90˚. It is important to also consider
that in SEM we are studying a two-dimensional plane, we consider that the same
effects can be extrapolated into the third-dimension and therefore the micro-cracks
recorded represent a minimum.
Figure 14. SEM images and crack analysis of a heat treated basalt sample. Total
number of cracks measured is 418 with an average length of 82 μm ± 5 μm. a)
original SEM image with scale shown, b) manually recorded crack map produced
from MATLAB, c) a close-up section of the image of the above image, d) fracture
numbers with respect to orientation (azimuth).
22
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 15. Comparison of crack sizes in non-heat treated (left) and heat-treated
(right) samples. From top to bottom materials are basalt, phonolite and dacite.
Whilst highly subjective, it is possible to draw comparisons between orientation of
cracks and the increase in number of cracks observed pre- and post- heat treatment
and the reduction in P-wave velocities, and P-wave isotropy. The two methods
therefore largely confirm the contemporaneous AE data which shows significantly
more cracking occurring during cooling.
23
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
4.0 Discussion
The vast majority of studies dealing with thermal cracking have concentrated
primarily on the heating cycle (Meredith et al., 2001; Richter and Simmons, 1974;
Simmons and Cooper, 1978; Vinciguerra et al., 2005). Although some previous
studies of thermal cracking in poly-crystalline materials have presented anecdotal
evidence of increased cracking during cooling (Heap et al., 2013; Mollo et al., 2013).
Our results suggest that low energy AE on heating within an all-round compressive
stress regime are likely associated with small increments of pre-existing micro-crack
extension; possibly with some new mixed-mode cracks forming. By contrast high
energy AE on cooling within an all-round tensile stress regime are associated with
larger increments of crack growth; and formation of new mode I cracks. Our
contemporaneous AE results are supported by complimentary static measurements
on pre- and post- heat treated rocks indicating substantially reduced seismic
velocities related to thermal cracking and increased number of cracks in SEM
analysed images. We suggest that those studies which have focused on thermal
cracking only within the heating and compressive regime (e.g Vinciguerra et al.,
2005) substantially underestimate the role of thermal stressing.
4.1 Seismic b value
In Figure 16 we show the same data as per Figure 9, but now include seismic b
values. b values were obtained for all experimental runs using a MATLAB script (P.
Benson, personal communication, 2015) that calculates the maximum likelihood
method (Aki, 1965), these values were then plotted alongside acoustic emission
energy for each rock type at the fastest and slowest rates of heating and cooling. The
b value provides an indication of the frequency and relative size of cracks, in the
case where b value decreases this generally indicates the presence of a greater
number of larger cracks. Interestingly we note much higher and varied values of b
obtained for NKD, ranging between 0.5 to 3, than IB and AP which range from
around 0.4 to 1. Generally, b values drop sharply at around 800C in the cooling
cycle corresponding with increased AE as previously described. In those cases where
it is possible to obtain b values during the heating cycle, we find that the values tend
to be higher than during the cooler cycle.
24
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 16. Acoustic emission energy and seismic b values as a function of
temperature for IB, AP and NKD heated and cooled at two different rates (1˚C/min
and 8˚C/min). b value was calculated using Aki’s maximum likelihood method for
200 hits at 100 hit intervals.
4.2 Fracture annealing
Fracture annealing is a relatively poorly studied topic, but the mechanics of the
process in silicic material relies on the ability for melt to fill the fracture voids and
then overcome fracture surface tension (Vassuer et al., 2013). This is obviously only
possible where melt can be produced in a rock type. The three chosen rock types
have vastly differing characteristics in terms of melt production at high temperature.
Generally, glass is the first phase to melt, at the glass transition temperature (Tg).
Neither IB nor AP contain an abundant glass phase and therefore any melting will be
associated with crystal phase or iron oxide melting. Tg was best constrained in NKD
at around 800˚C. This material therefore offered the best chance for characterising
25
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
fracture annealing and resultant fracture generation upon cooling. Two-main issues
that make the problem of fracture annealing more complicated, one concerns the
practicality of traditional annealing timescales in complex partially solid silicate
systems, and the other concerns the development of melt driven overpressure within
partially solidified rocks.
Thermal mechanical analysis shows that all test materials experience large volume
expansion at the highest temperatures, a process which may counter the ability to
anneal fractures. Each material will have undergone various phases changes, partial
melting during heating and re-crystallisation during cooling. Therefore the bulk
material may be quite different from the starting material upon cooling from the
maximum hold temperature. It has not been possible to ascertain those precise petrophysical alterations, and it is beyond the scope of this study to do so. However, it is
possible to note that thermal expansivity (α) upon heating and contraction upon
cooling are within a similar α range. Therefore increased rates and size of acoustic
emission hits during cooling are created as a result of the overall tensile stresses
created by thermal contraction.
As well as a greater number and size of cracks distributed throughout the sample
volume, some samples of IB indicate evidence of melt textures post heat-treatment
(Figure. 17). Samples of IB held above 1130˚C exhibit macro fractures, up to 3 mm
in width and > 1cm in length on the sample surface (Figure. 17) aligned
predominantly along axis. Highly vesicular melt nodules emanate from many of the
sample surface fractures, indicating that melt was generated and either filled
previously formed fractures or played a role in generating new fractures. Melt over
this temperature range is most likely formed due to phase changes in Fe oxides or a
hydrous mineral such as amphibole. Melt-like textures can be observed in thin
section and SEM analysis surrounding the much larger tension cracks (Figure 17). In
this case the melt textures are shown as high contrast areas of spherical to elongated
material randomly dispersed between a plagioclase and pyroxene matrix. Individual
melt nodules which emanate from the basaltic samples at high temperature are highly
vesicular, indicating significant viscous relaxation and degassing. It is possible to
consider that the macro-fractures on the sample surface are a type of hydraulic
fracture, formed primarily due to magma overpressure rupturing the clearly
26
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
solidified and brittle outer crust of the sample. If this is the case then those melt
nodules essentially represent basaltic dykes. How and why they form in this instance
remains poorly known, particularly why for instance parts on the inner sample can
form melt whereas the sample surface remains with a solid brittle crust. The features
may be related to excess pressure generated from melt volume expansion at high
temperature. In all rocks types we find sharp increases in expansivity at highest
temperatures, as shown in Fig. 4. However, melt nodules are not formed in either AP
or NKD at the maximum hold temperatures. Instead of producing melt rupture
features at the highest hold temperatures, samples of NKD behave plastically
shortening in length and expanding in width substantially. Samples of AP exhibit
macro evidence of melting along planes of pre-existing veins but no whole scale
melting such as observed in the form of melt nodules in IB. More work is needed to
understand the role of melt volume expansion in generating excess pressure. It is
likely that more information is needed on the volatile content of each rock type,
which is beyond the scope of this study.
Figure 17. Macro-fracture development in a heat treated sample of IB. a) SEM
image of non-heat treated IB and b) photograph of starting material. c) SEM image
of a sample of IB heat treated to 1100˚C, d) corresponding photograph of endmember material. The axial tension fracture indicated by the red circle is shown in
the SEM image on the left, micro-scale (<10 μm) melt textures observed under SEM
27
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
can be observed around the edge of those tension fractures. In many cases melt
nodules are observed emanating from tension fractures on the surface of the sample
cores, as shown in e).
Thermo-elastic stress can be simplified and calculated from the following relation
(Timoshenko and Goodier, 1970):
t
ET
(1 )
(1)
where σt is tensile thermal stress (MPa), E is Young’s modulus (MPa), ∆T is the
temperature difference (˚C) and ν is the Poisson’s Ratio. Upon heating, stress is
generated by thermal expansion, whereas upon cooling by negative thermal
expansion or thermal contraction. The thermal expansion coefficient (α) of any
material, which has previously been shown to be similar during heating or cooling
(Fig. 4), is an important factor in controlling thermally generated cracks (Siratovich
et al., 2015). In Figure 16 we show the thermal tensile stress σt generated in the bulk
rock, an olivine and clinopyroxene mineral as a function of temperature difference.
The models assume that each material has a Young’s modulus of 10 GPa, and a
Poisson’s ratio of 0.25, the former is certainly a minimum value. This assumption
may be erroneous and therefore provide misleading magnitudes of tensile stress but
the model helps to show the relative contribution of each mineral phase in generating
thermal stress. Upper and lower values of α were obtained from previous TMA
measurements for each bulk rock and range from 15 × 10-6/˚C to 25 × 10-6/˚C. Upper
and lower values of α for each mineral phase were taken from Clark (1966) and
range from 30 × 10-6/˚C to 35 × 10-6/˚C for olivine and 25 × 10-6/˚C to 30 × 10-6/˚C
for pyroxene. Other minerals tend to fall within the ranges previously given and so
are not shown. We show that the first minerals that begin to produce stresses
sufficient to generate tensile (mode I) fractures, assuming a tensile strength To of 0.56 MPa (Amadei and Stephansson, 1997), are olivine and pyroxene with a
temperature difference of approximately 20˚C. As we previously state, the precise
value of tensile stress should be treated with caution. All values of α for bulk rock
are lower because the rocks contain pre-existing interval void space, i.e cracks and
28
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
pores. During heating, cooling or mechanical loading these voids can close or open
internally so the effects measured are reduced. Single mineral crystals do not have
any void space and therefore give higher values of α. This reason also indicates that
the value of Young’s modulus used is too low, although as previously stated we do
not provide the precise range of failure, but indications as to which mineral phase
contributes.
Figure 18. Tensile thermal stress σt as a function of temperature difference ∆T.
Increasing thermal expansion co-efficient α and ∆T have the effect of increasing σt.
As the mineral phases have larger values of α they generate the highest values of σt
within the lowest temperature change.
Each rock type has a similar range of thermal expansion and contraction co-efficient
values during both heating and cooling and therefore likely generates similar
magnitudes of stress during both temperature cycles. We have found that more
cracks are formed during cooling, which can be presumably explained as To ≈ 0.1σc,
where To is tensile strength and σc is compressive stress. Therefore, as is well known,
it is easier for rocks to fail in tension (Gudmundsson, 2011).
4.3 Kaiser ‘temperature-memory’ effect test
The mechanical Kaiser ‘stress memory’ effect is a relatively well understood process
(Heap et al., 2009; Lockner, 1993), where new cracking growth only once the
29
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
previous state of maximum stress has been exceeded. It is not clear of the effects or
even existence of a Kaiser ‘temperature-memory’ effect, although some studies have
shown such an effect to exist ( Choi et al., 2005; Heap et al. 2013). In order to test
ideas surrounding Kaiser ‘temperature-memory’ effects, we conducted a series of
ramp and hold experiments. A sample of IB was heated at between 4C/min and
8C/min to maximum cyclic hold temperatures at 100C intervals, which are then
exceeded on the following cycle. Samples are cooled at a natural cooling rate
(<4C/min) to 300C, and then re-heated once thermally equilibrated. Our tests
proved inconclusive to the existence of a Kaiser ‘temperature memory’ effect in
basalt, for the most part indicating that such an effect does not exist. In Figure 19 we
show the various heating and cooling cycles performed, the resultant AE energy and
corresponding b values. Acoustic emissions only commence during cycle 3, at a
temperature exceedance of 600˚C, in the following cycle acoustic emissions
commence around 650˚C and then again at around 800˚C, very few emissions are
recorded above this temperature even though each cycle signifies the exceedance of
the previous temperature. AE output is relatively low in all heating parts of each
cycle. Interestingly, the maximum AE during cooling is produced only when highest
hold temperature (900˚C) is reached, indicating that before this temperature cooling
contraction does not form new cracks in basalt but simply reactivates pre-existing
ones in a manner similar as during heating. This implies that some fracture annealing
was completed before or around 900˚C in Icelandic Basalt. Whilst our data, namely
the lack of cyclic AE, can be used to suggest the absence of a Kaiser ‘temperature
memory’ effect, it is difficult to conclude the process does not occur by the absence
of signals. However, consider that the Kaiser ‘stress memory’ effect is based upon
the notion of stress loading and unloading, in the case of thermal stressing most of
the unloading as we have shown occurs during cooling. Therefore any stress loading
that is created upon heating and expansion is effectively reset during cooling
contraction.
30
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Figure 19. Kaiser ‘temperature memory’ effect test results. IB was heated at a rate of
4C/min and held at various temperatures and then cooled at a natural rate to 300C
in five separate cycles, in the final cycle a ramp rate of 8˚C/min was used. Very low
energy bursts of AE are indicated by grey lines during the heating cycles 3,4 and 5
but generally there is only substantial AE produced during cooling in cycles 4, 5 and
6, indicating very little evidence for a Kaiser ‘temperature memory’ effect.
5.0 Conclusions
Our findings suggest that heating and resultant thermal expansion in volcanic rocks
acts to propagate, through small increments, previously existing micro-cracks or
encourage new mixed mode micro-crack formation. Whereas cooling contraction
encourages larger crack growth increments, and the formation of longer
predominantly mode I micro-cracks.
All the evidence suggests that there is far more cracking during cooling than
during heating. This is reflected by contemporaneous acoustic emissions that
show substantial increase in hit rate, hit amplitude and hit energy during
cooling from high temperature. These increases in AE occur immediately on
cooling but commonly peak around 800˚C in all rock types studied.
Microstructural analysis of heat treated samples supports this view also.
31
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Evidence from ultrasonic wave velocities and microstructural observation
suggests that thermal cracking is isotropically distributed. This is expected
because thermal stresses are by nature isotropic.
Although it has not been possible to directly observe or measure crack
annealing at high temperature, all samples were held at a high enough
temperature and for a long enough time to infer some cracks healed.
We suggest that a previously reported Kaiser ‘temperature-memory’ effect
does not exist. This seems a reasonable assumption as the mechanical Kaiser
effect, of which the principles of the temperature memory effect are based
upon, relies on unloading of mechanical stress. Whereas in temperature tests
most cracking occurs during cooling, or the unloading cycle, therefore
increased cracking during cooling overrides any stresses that are generated by
each new heating cycle.
Our findings our important for the study of cooling volcanic rocks in
different settings. It is already well known that extrusive lava flows such as
those on Mt Etna contain substantially more crack damage than intrusive
rocks, as noted from lower P-wave velocities. It is clear now that most crack
damage in volcanic rocks is generated during the cooling contraction cycle of
emplacement.
6.0 References
Acocella, V., Gudmundsson, A., and Funiciello, R. 2000. Interaction and linkage of
extension fractures and normal faults: examples from the rift zone of Iceland.
Journal of Structural Geology, 22, 1233-1246.
Aki, K., 1965. Maximum Likelihood Estimate of b in the Formula logN=a-bM and
its Confidence Limits, 43, 237–239.
32
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Axelsson, G., Thorhallsson, S., Bjornsson, G., 2006. Stimulation of geothermal wells
in basaltic rock in Iceland. Enhanced geothermal innovation network for Europe
Workshop, Zurich Switzerland, 3,
Amadei, B., and Stephenson, O., 1997. Rock stress and its measurement. Chapman
and Hall, London.
Barton, M., and Huijsmans, J. P. (1986). Post-caldera dacites from the Santorini
volcanic complex, Aegean Sea, Greece: an example of the eruption of lavas of nearconstant composition over a 2,200 year period. Contributions to Mineralogy and
Petrology, 94, 472-495.
Brudy, M. and Zoback, M.., 1999. Drilling-induced tensile wall-fractures:
implications for determination of in-situ stress orientation and magnitude.
International Journal of Rock Mechanics and Mining Sciences, 36, 191–215.
Cabrera, A., R. F. Weinberg, H. M. N. Wright, S. Zlotnik, and R. A. F.Cas 2011,
Melt fracturing and healing: A mechanism for degassing and origin of silicic
obsidian, Geology, 39, 67–70.
Choi, N.K., Kim, T.W., Rhee, K.Y. 2005 Kaiser effects in acoustic emission from
composites during thermal cyclic-loading. NDT&E International. 38, 268-274
Chouet, B.A. 1996 Long-period volcano seismicity: its source and use in eruption
forecasting. Nature, 380, 309-316
Clark, Sydney Procter, ed. 1966 Handbook of physical constants. 97. Geological
Society of America.
Cooper, H.W and Simmons, G. 1977. The effect of cracks on the thermal expansion
of rocks. Earth and Planetary science letters. 36, 404-412
Cox, S.J.D., and Meredith, P.G., 1993. Microcrack formation and material softening
in rock measured by monitoring acoustic emissions. Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. 30, 11–24.
33
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
David, C., Menendez, B., and Darot, M. 1999. Influence of stress-induced and
thermal cracking on physical properties and microstructure of La Peyratte granite.
International Journal of Rock Mechanics and Mining Sciences, 36, 433-448.
Dingwell, D.B. and Webb, S.L., 1989. Structural relaxation in silicate melts and nonNewtonian melt rheology in geologic processes. Physics and Chemistry of Minerals,
16, 508–516.
Fredrich, J. T., and Wong, T. 1986. Micromechanics of thermally induced cracking
in three crustal rocks. Journal of Geophysical research, 91.
Giordano, D., Russell, J.K., Dingwell, D.B., 2008. Viscosity of magmatic liquids: A
model. Earth Planet. Sci. Lett. 271, 123–134.
Gudmundsson, A. 1998, Formation and development of normal-fault calderas and
the initiation of large explosive eruptions, Bulletin of Volcanology, 60, 160–170.
Gudmundsson, A. 2011. Rock fractures in geologic processes. Cambridge university
press, Cambridge.
Heap, M.J., Vinciguerra, S. & Meredith, P.G., 2009. The evolution of elastic moduli
with increasing crack damage during cyclic stressing of a basalt from Mt. Etna
volcano. Tectonophysics, 471, 153–160
Heap, M. J., Faulkner, D. R., Meredith, P. G., & Vinciguerra, S. 2010. Elastic
moduli evolution and accompanying stress changes with increasing crack damage:
implications for stress changes around fault zones and volcanoes during deformation.
Geophysical Journal International, 183, 225–236.
Heap, M.J., Lavallée, Y., Laumann, A., Hess, K.U., Meredith, P.G., Dingwell, D.B.,
Huismann, S., Weise, F., 2013. The influence of thermal-stressing (up to 1000 °c) on
the physical, mechanical, and chemical properties of siliceous-aggregate, highstrength concrete. Constr. Build. Mater. 42, 248–265.
Heap, M. J., Silvio Mollo, Sergio Vinciguerra, Yan Lavallée, K-U. Hess, Donald B.
Dingwell, Patrick Baud, and Gianluca Iezzi. 2013b Thermal weakening of the
carbonate basement under Mt. Etna volcano (Italy): implications for volcano
instability. Journal of volcanology and geothermal research 250, 42-60.
34
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Hetényi, G., Taisne, B., Garel, F., Médard, É., Bosshard, S., Mattsson, H.B., 2012.
Scales of columnar jointing in igneous rocks: Field measurements and controlling
factors. Bull. Volcanol. 74, 457–482.
Julian, B.R., Foulger, G.R., Monastero, F.C., Bjornstad, S., 2010. Imaging hydraulic
fractures in a geothermal reservoir. Geophys. Res. Lett. 37, 1–5.
Kitao, K., Aiki, K., Watanabe, H., Wakita, K., 1990. Cold-water well stimulation
experiments in the Sumikawa Geotheral field, Japan. Geotherm. Resourc. Counc.
Trans 14, 1219–1224.
Kranz, R. L. 1983. Microcracks in rocks: A review. Tectonophysics, 100, 449–480.
Lockner, D. 1993. The role of acoustic emission in the study of rock fracture.
International journal of rock mechanics. 30, 833-899
Meredith, P. G., Knight, K. S., Boon, S. A., & Wood, I. G. 2001. The microscopic
origin of thermal cracking in rocks: An investigation by simultaneous time‐of‐flight
neutron diffraction and acoustic emission monitoring. Geophysical research letters,
28, 2105-2108.
Miller, A D., Julian, B.R. & Foulger, G.R., 1998. Three-dimensional seismic
structure and moment tensors of non-double-couple earthquakes at the HengillGrensdalur volcanic complex, Iceland. Geophysical Journal International, 133, 309–
325.
Mitchell, T.M. and Faulkner, D.R., 2008. Experimental measurements of
permeability evolution during triaxial compression of initially intact crystalline rocks
and implications for fluid flow in fault zones. Journal of Geophysical Research:
Solid Earth, 113, pp.1–16.
Mollo, S., Heap, M.J., Dingwell, D.B., Hess, K.U., Iezzi, G., Masotta, M., Scarlato,
P., Vinciguerra, S., 2013. Decarbonation and thermal microcracking under magmatic
P-T-fco2 conditions: The role of skarn substrata in promoting volcanic instability.
Geophys. J. Int. 195, 369–380.
35
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
Pyle, D.M. and Elliott, J.R., 2006. Quantitative morphology, recent evolution, and
future activity of the Kameni Islands volcano, Santorini, Greece. Geosphere, 2, 253–
268.
Richter, D and Simmons, G. 1974. Thermal expansion behaviour of Igneous Rocks.
International journal of rock mechanics and mineral science, 15. 145-148
Simmons, G., and Cooper, H. W. 1978. Three Igneous Rocks. International Journal
of Rock Mechanics, Mineral Science and Geomechanics, 15, 145–148.
Siratovich, P. A., von Aulock, F.W., Lavallée, Y., Cole, J.W., Kennedy, B.M.,
Villeneuve, M.C., 2015. Thermoelastic properties of the Rotokawa Andesite: A
geothermal reservoir constraint. J. Volcanol. Geotherm. Res. 301, 1–13.
Tarasovs, S. and Ghassemi, A., 2012. On the Role of Thermal Stress in Reservoir
Stimulation. Thirty-Seventh Workshop on Geothermal Reservoir Engineering.
Timoshenko, S.P and Goodier, A, 1970. Theory of Elasticity. 3rd ed. McGraw-Hill,
New York, NY.
Turner, J.S., Huppert, H.E., Sparks, S.J., 1983. An experimental investigation of
volatile exsolution in evolving magma chambers. J. Volcanol. Geotherm. Res. 16,
263-277
Tuffen, H., and D. B. Dingwell. 2005, Fault textures in volcanic conduits: Evidence
for seismic trigger mechanisms during silicic eruptions, Bull. Volcanol., 67, 370–387
Tuffen, H., D. B. Dingwell, and H. Pinkerton. 2003, Repeated fracture and healing of
silicic magma generate flow banding and earthquakes?, Geology, 31, 1089–1092.
Vasseur, J., F. B. Wadsworth, Y.,Lavallee, K.-U. Hess, and D. B. Dingwell. 2013,
Volcanic sintering: Timescales of viscous densification and strength recovery,
Geophys. Res. Lett., 40, 5658–5664
Vinciguerra, S., Trovato, C., Meredith, P.G. and Benson, P.M. 2005, Relating
seismic velocities , thermal cracking and permeability in Mt . Etna and Iceland
36
Chapter 5: Cooling dominated cracking in thermally stressed volcanic rocks
basalts, International Journal of
Rock Mechanics and Mineralogical Science
42,900–910.
37
Chapter 6: Surface displacements resulting from magma-chamber roof subsidence
Chapter 6
Journal of Volcanology and Geothermal research
Surface displacements resulting from magma-chamber roof subsidence, with
application to the 2014 Bardarbunga-Holuhraun episode
Browning, J and Gudmundsson, A
Journal of Volcanology and Geothermal Research, 2015, 308, 82-98
DOI 10.1016/j.jvolgeores.2015.10.015
Statement of contribution
Original idea conceived by both authors was inspired by events at Bardarbunga in
2014 following modelling work of JB
All numerical models and figures created by JB
Model interpretation and comparisons by JB
1st draft of manuscript by JB
Revisions and subsequent drafts with input from co-author
Significant input from co-author discussion and interpretation to Bardarbunga
episode
56
Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Contents lists available at ScienceDirect
Journal of Volcanology and Geothermal Research
journal homepage: www.elsevier.com/locate/jvolgeores
Surface displacements resulting from magma-chamber roof subsidence,
with application to the 2014–2015 Bardarbunga–Holuhraun
volcanotectonic episode in Iceland
John Browning ⁎, Agust Gudmundsson
Royal Holloway University of London, Department of Earth Sciences, Egham TW20 0EX, United Kingdom
a r t i c l e
i n f o
Article history:
Received 22 March 2015
Accepted 7 October 2015
Available online 19 October 2015
Keywords:
Surface deformation
Surface stresses
Magma chambers
Caldera collapse
Bardarbunga–Holuhraun
a b s t r a c t
The conditions which lead to caldera collapse are still poorly constrained. As there have only been four, possibly
five, well-documented caldera forming events in the past century, the geodetic signals produced during chamber
roof subsidence, or chamber volume reduction (shrinkage) in general, are not well documented or understood.
In particular, when two or more geodetic sources are operating and providing signals at the same time, it is important to be able to estimate the likely contribution of each. Simultaneous activities of different geodetic sources
are common and include pressure changes in magma chambers/reservoirs occurring at the same time as dyke
emplacement. Here we present results from numerical models designed to simulate the subsidence of a
magma-chamber roof, either directly (chamber shrinkage) or through ring-fault displacement, and the induced
surface deformation and crustal stresses. We consider chamber depths at 3 km, 5 km, and 7 km below the crustal
surface, using both non-layered (isotropic) and layered (anisotropic) crustal models. We also model the effects of
a caldera lake and of a thick ice cover (ice sheet) on top of the caldera. The results suggest that magma-chamber
roof subsidences between 20 m and 100 m generate large (tens of centimetres) vertical and, in particular, horizontal displacements at the surfaces of the ice and the crust out to distances of up to tens of kilometres from the
caldera/chamber centre. Crustal layering tends to reduce, but increasing chamber depth to enlarge, the horizontal and vertical surface displacements. Applying the results to the ice subsidence in the Bardarbunga Caldera during the 2014–2015 Bardarbunga–Holuhraun volcanotectonic episode indicates that the modelled ice
displacements are less than those geodetically measured. Also, the geodetically measured crustal displacements
are less than expected for a 60 m chamber-roof subsidence. The modelling results thus suggest that only part of
the ice subsidence is due to chamber-roof subsidence, the other part being related to flow in and down-bending
of the ice. We show that such a flow is likely within the caldera as a result of the stress induced by the 45-km-long
regional dyke emplaced (primarily in vertical magma flow) during the episode. This conclusion is further supported by the model results suggesting that the ring-fault (piston-like) displacements must have been much
less than the total 60 m ice subsidence, or else faults with tens-of-metres displacements would have cut through
the ice (these are not observed). We suggest that the ring-fault subsidence was triggered by small doming of the
volcanic field and system hosting the Bardarbunga Caldera and that this doming occurred as a result of magma
inflow and pressure increase in a deep-seated reservoir. The doming is confirmed by GPS measurements and
supported by the seismicity results. The magmatic pressure increase in the reservoir was, in terms of the present
model, responsible for the regional dyke emplacement, the Holuhraun eruption, and part of the stress concentration around, and displacement of, the Bardarbunga Caldera.
© 2015 Elsevier B.V. All rights reserved.
1. Introduction
Caldera collapses are a common occurrence in the evolution of major
volcanic systems (Fig. 1). While many of these events are catastrophic
and associated with the expulsion of large volumes of magma and
ignimbrite formation (Druitt and Sparks, 1984), perhaps the more prevalent situation involves relatively small or no magma expulsion
⁎ Corresponding author.
E-mail address:
[email protected] (J. Browning).
http://dx.doi.org/10.1016/j.jvolgeores.2015.10.015
0377-0273/© 2015 Elsevier B.V. All rights reserved.
(MacDonald, 1965). Well-documented caldera collapses occurred in
2000 and 2007 at the summits of Miyakejima (Geshi et al., 2002) and
Piton de la Fournaise (Peltier et al., 2008). These events have been
referred to as periodic (Geshi et al., 2002; Michon et al., 2011) or slow
collapses. These terms relate to the total caldera growth occurring
over periods of perhaps as much as one month (Geshi et al., 2002).
Much of the longer-period caldera growth was due to mass wasting, a
process which likely also shaped lake Öskjuvatn (Iceland) following
the 1875 caldera forming eruption (Hartley and Thordarsson, 2012). A
mechanism of ‘slow caldera collapse’ has also been suggested as an
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
83
general, are here applied to the 2014–15 Bardarbunga–Holuhraun
volcano-tectonic episode.
2. Stress fields controlling caldera formation
Fig. 1. Simplified geological map of Iceland, showing the main active and inactive volcanoes
that contain collapse calderas. Many of these calderas are fully or in part sub-glacial, that is
located underneath a body of ice. Glacier outlines are highlighted in this map as white areas.
explanation for the measured ice subsidence during the 2014–2015
Bardarbunga episode (Riel et al., 2015; Sigmundsson et al., 2015).
The timescale of deformation at calderas ranges from events of hours
to days (Stix and Kobayashi, 2008) to longer events taking months or
years (Hartley and Thordarsson, 2012) as well as cyclic inflation and
deflation over tens, hundreds and probably thousands of years
(Phillipson et al., 2013). Collapse may occur along pre-existing
structures, such as regional faults or earlier-formed ring-faults (Fig. 2),
but the shape and size of collapse are significantly influenced by the
depth, size, and shape of an underlying magma chamber (Cole et al.,
2005; Acocella, 2007).
The movement of large crustal segments, as occurs during the formation or reactivation of collapse calderas, must produce significant
crustal deformation. However, the magnitude and type of the deformation are poorly constrained. This is partly due to the lack of geophysical
measurements syn-collapse, the exceptions being Piton de la Fournaise
(Peltier et al., 2008) and Miyakejima (Geshi et al., 2002), although measurements at these locations were predominantly limited to the central
edifice and vent area. Therefore, understanding the far-field effects of
crustal subsidence due to caldera formation or chamber shrinkage is
useful for constraining geophysical observations at volcanoes where
the summit region cannot be observed, either due to cloud cover,
inaccessibility, or ice cover. The last point is salient because many, if
not most, of the central volcanoes in Iceland are ice covered (Figs. 1
and 3). In addition, understanding the timing and development of
collapse is important for hazard and risk estimation, partly because
many calderas are associated with the formation of ring-dikes (Fig. 2)
(Browning and Gudmundsson, 2015) and give rise to large eruptions
(Gudmundsson, 2015).
When magma leaves or flows out of a chamber/reservoir during an
eruption and/or dyke injection, the volume of the chamber/reservoir decreases. The same may happen during caldera collapse (Gudmundsson,
2014, 2015). The volume decrease or shrinkage affects the crustal segment hosting the chamber, primarily through changes in stress and associated displacement and strain. The effects of chamber shrinkage are
most easily detected through surface deformation. The aim of this work
is to understand better (1) how the surface deformation associated
with chamber shrinkage, in particular during roof subsidence, is reflected
in horizontal and vertical displacements (and stresses) at the surface of
the hosting crustal segment (as well as at the surface of the ice cover),
(2) how the surface deformation changes with distance from the
chamber, and (3) how much surface deformation can feasibly be accommodated in an elastic crust before ring-faults will form or reactivate,
resulting in a normal caldera collapse. The results, while completely
Many analogue models of caldera collapse indicate initial ground
surface slumping (Lavallée et al., 2004) followed by the formation of peripheral faults that ultimately control the majority of vertical subsidence
(Acocella et al., 2000; Kennedy et al., 2004; Holohan et al., 2005; Geyer
et al., 2006; Acocella, 2007). As many of these models use dry sand or
other similar granular materials to simulate the crust, it is often impossible to determine surface displacements far from the deformation
centre. This follows partly because a dry sand pack lacks cohesion
(which corresponds to rock tensile strength) and normally does not
transmit tensile stresses as solid linear elastic material. By contrast,
the crust behaves approximately as linear elastic solid material with a
non-zero tensile strength. More specifically, the range of in-situ tensile
strength of solid rocks is 0.5–9 MPa, the most common values being
2–4 MPa (Gudmundsson, 2011). Numerical models which simulate an
elastic crustal segment hosting a magma chamber therefore provide a
reasonable approximation of surface ground deformation (De Natale
and Pingue, 1993; Hickey and Gottsmann, 2014).
In order for a caldera to form, or for slip to occur on a pre-existing
ring-fault, there must be suitable state of stress within the crust. The initiation of sub-vertical, normal ring-faults depends on three stress field
conditions which must be satisfied simultaneously (Gudmundsson,
1998; Folch and Marti, 2004).
(1) The minimum value of σ3, the maximum tensile (minimum
compressive) principal stress, must be at the surface.
(2) The maximum value of (σ1–σ3)/2, the shear stress, must occur
above the outer margins or lateral edges of the magma chamber,
that is, in a zone extending from the lateral edge of the chamber
to the surface and within which the ring-fault forms (or slips).
(3) The maximum tensile stress at the surface must peak at a radial
distance approximately equal to the lateral dimension, the
diameter, of the magma chamber.
These stress conditions are most likely to be induced by a double
magma chamber, where the shallow chamber is sill-like and (1) the
crustal segment hosting the double chamber is subject to horizontal extension, or (2) the deeper chamber, a large reservoir, is subject to slight
increase in magma pressure so as to dome the crustal segment hosting
the shallower chamber (Gudmundsson, 1998, 2007) (Fig. 3). Other predominant collapse trigger mechanisms (Marti et al., 2009) include (a)
internal magma chamber overpressures initiating roof and surface fractures (e.g., Gudmundsson, 1998, 2007; Gray and Monaghan, 2004;
Gregg et al., 2012) and (b) internal magma chamber underpressure following chamber rupture (e.g., Roche et al., 2000; Folch and Martı́, 2004;
Geyer et al., 2006; Kusumoto and Gudmundsson, 2009; Holohan et al.,
2011). Here we consider in detail a situation more compatible with
the second of these two mechanisms, namely an inferred underpressure
in the shallow chamber, particularly in view of the suggestions that the
ice subsidence during the Bardarbunga–Holuhraun episode being
related to pressure decrease in the chamber (e.g., Riel et al., 2015;
Sigmundsson et al., 2015).
As indicated above, shallow chambers within crustal segments
undergoing slight doming, regional extension, or both are the ones
most likely to generate stress concentrations favourable for ring-fault
formation (Gudmundsson, 2007). Prime examples of this type of regional settings are the volcanoes of the Eastern Volcanic Zone (EVZ) in
Iceland (Gudmundsson, 2007), the Kenyan Rift valley (Acocella, 2007)
and the Taupo Volcanic Zone (TVZ) in New Zealand (Cole, 1990).
Fig. 4 shows the stresses around a sill-like magma chamber with
negative internal pressure, an underpressure, of 5 MPa (e.g., Folch and
84
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
B
A
C
Fig. 2. Exposed sections of extinct central volcanoes in the Tertiary volcanic regions of East (A, B) and West (C) Iceland. (A, B) An extinct – and now well exposed – granophyre magma
chamber at Slaufrudalur in East Iceland. The exposure shows the contact (inset) between the granophyre plutonic rocks at the base of the picture and the basaltic lava pile at the top of the
picture, into which the chamber was originally emplaced. Shallow chambers such as this one commonly give rise to a vertical collapse, culminating in the formation of a surface caldera.
Although no evidence exists for a collapse at Slaufrudalur, many eroded central volcanoes in Iceland show clear ring-faults, perhaps the best example being (C) Hafnarfjall in West Iceland.
Marti, 2004), located at 5 km depth in a 40-km wide and 20-km thick
crustal segment but simultaneously subject to excess (doming)
pressure of 10 MPa from a deep-seated reservoir. In this example, the
doming pressure largely controls the stress/displacement fields and
the maximum tensile stress concentrates at the free surface in a zone
above the lateral margins or edge of the shallow sill-like chamber. In
addition, the maximum shear stress concentrates at the lateral margins
of the magma chamber at depth. These conditions are ideal for the
formation of, initially, tension fractures at the surface that propagate
down towards the chamber and change at a critical depth – normally
less than 0.5 km (Gudmundsson, 2011) – to normal faults
(Gudmundsson, 1998; Gray and Monaghan, 2004). If the tensile stresses
are higher at the magma chamber margin than at the free surface
above the margin, then a ring-dyke would be more likely to form
(Gudmundsson, 2007).
When a caldera forms, it is common for the depression to be filled
with water, generating in a caldera lake. Well-known examples include
Crater Lake, USA (Williams, 1941) and Askja, Iceland (Hartley and
Thordarson, 2012) amongst many others (Fig. 3). The occurrence of
sub-glacial lakes within calderas has also been noted, such as in the
Grimsvötn volcanic system in Iceland (Gudmundsson et al., 1997). A
caldera lake is important because the solid contact with water gives
rise to a free surface, that is, a surface of zero shear stress. Therefore, a
caldera lake will reduce the mechanical coupling between bedrock
and glacier which in-turn will influence stresses and displacements
within the ice.
3. Model setup
The finite element program Comsol was used to investigate the
crustal and ice-sheet response to the vertical displacement of the
magma-chamber roof (www.comsol.com; cf. Zienkiewicz, 1979; Deb,
2006). In these models the magma chamber is modelled as a cavity
(Gudmundsson, 2011; Grosfils et al., 2015). In the first model the
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Fig. 3. Sub-glacial caldera occupied by a caldera lake. Here the ring-fault is a normal fault,
as is inferred for nearly all collapse calderas in Iceland (Bjarnason, 2014), including the
Bardarbunga Caldera (Riel et al., 2015). The shallow chamber is fed by a much larger
deep-seated reservoir that undergoes periods of doming and inflation when receiving
new input of melt or magma. In this example the area of doming is much larger than
the caldera. Modified after Gudmundsson (2007).
chamber is residing within a homogeneous, isotropic elastic half-space
with a Young's modulus (E) of 40 GPa and Poisson's ratio (ν) of 0.25
(Fig. 4). In this model, the focus is on the typical conditions for
ring-fault formation or reactivation in rift-zone environment. Thus,
the loading condition is a combination of doming excess pressure of
10 MPa in the deep-seated reservoir and a horizontal tension of
85
5 MPa. The results are in agreement with earlier results suggesting
that doming, horizontal tension, or both are loading conditions that
favour the concentration of shear stress in a zone above the lateral
ends or edges of the shallow chamber (Fig. 4C). Furthermore, the
induced tensile stress peaks where these zones meet the free surface
(Fig. 4D). The zones of high shear stress are thus likely to develop
ring-faults, for the given loading conditions.
Horizontal tension, however, will not be much discussed here. In the
later part of the paper, we discuss the effects of doming by the
deep-seated reservoir. The main focus here is on the effects of
shallow-chamber roof subsidence on the associated surface deformation and stresses. The roof (upper boundary) of the chamber is supposed to subside, so that, the loading is prescribed negative vertical
(z-axis) displacement. The vertical roof displacements tested in the
models range from 20 m to 100 m. To make the models realistic, the
crustal segment hosting the chamber is also modelled as anisotropic,
that is, layered (Fig. 5). In the anisotropic models, directly above the
chamber there are six layers, each with thickness t and of varying stiffness (Young's modulus, E) but constant density (ρ = 2500 kg/m3) and
constant Poisson's ratio (ν = 0.25), simulating an anisotropic crust.
The number of layers used in the models is arbitrary as most volcanic
systems are presumably made of hundreds of layers, while many of
these may group into larger units of internally similar mechanical properties. Here we choose to include six layers or units simply to investigate
the effects of crustal anisotropy on the local stresses and displacements.
The uppermost layer of thickness (2 t) represents an elastic body of ice;
A
B
C
D
E
Fig. 4. Stress fields favouring the formation of caldera ring-faults. The stresses shown are generated around a sill-like magma chamber, 8 km wide and 2 km thick, located at 5 km depth in a
20-km-thick and 40-km-wide crustal segment subject to horizontal tensile stress of 5 MPa and doming stress (pressure) from a deep-seated reservoir (at the base of the model) of 10 MPa.
Chamber excess pressure is negative (underpressure) 5 MPa and it is hosted within a homogeneous, isotropic crustal segment of stiffness (Young's modulus) 40 GPa and Poisson's ratio of
0.25 (see Table 1). A, model configuration; B, magnitudes of the maximum principal tensile stress σ3; C, magnitudes of the von Mises shear stress τ; D, maximum principal tensile stress σ3,
and von Mises shear τ stress at Earth's free surface; E, maximum principal tensile stress σ3, and von Mises shear τ stress around the magma-chamber boundary. Figure C shows clearly that
the shear stress concentrates in subvertical zones above the lateral ends of the chamber, thereby encouraging the formation of a subvertical ring-fault.
86
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Fig. 5. Sketch of the model setup showing the geometric relationship between a shallow magma chamber within a crustal segment composed of six layers (including the glacier). In some
subsequent models (referred to as layered models), the 5 layers have different stiffnesses (Young's moduli), whereas in other models (referred to as non-layered models) all the 6 layers
have the same Young's modulus and thus function as a single, thick layer. At the surface of the crust, there is a caldera lake, providing a free surface (surface free of shear stress). In addition,
a glacier is on top of the lake and the surrounding crust, the top of the glacial layer is another free surface. All models shown are symmetric with rotation around the z-axis (axi-symmetric),
the base and vertical margins are fixed (i.e. experience zero displacement). The magma chamber roof is subject to a prescribed vertical displacement between 20 and 100 m. In addition to
the shallow magma chamber, several models include a deep-seated reservoir. (Modified after Folch and Martı́, 2004, and Kinvig et al., 2009).
namely a glacier, with a Young's modulus of 4 GPa. We use ice as the
top-most layer primarily because many volcanoes are located beneath
ice sheets, particularly in Iceland, including recently erupting volcanoes
in Iceland such as Bardarbunga (Gudmundsson et al., 2014; Riel et al.,
2015; Sigmundsson et al., 2015), Grimsvotn (Gudmundsson et al.,
1997), and Eyjafjallajökull (Gudmundsson et al., 2012).
For the purpose of this study we model the ice as a brittle layer
which behaves elastically through its entire thickness (Geyer and
Bindeman, 2011). Other studies assume that only certain parts of
an ice layer behave elastically, with the remaining parts behaving as
ductile – using, for example Glen's flow law (e.g. Paterson, 1994;
Gudmundsson et al., 2004; Schulson and Duval, 2009). Ice behaves elastically at high strain rates and comparatively low stresses or pressures
(Schulson and Duval, 2009). The brittle deformation of ice is exemplified in the formation or fractures, crevasses, as are common during subsidence associated with volcanism (Gudmundsson et al., 1997, 2004).
The assumption of linear elastic behaviour of the ice is thus reasonable
and does not significantly affect the calculated displacements and
stresses in the crust (the surface rock) itself outside the ice sheet. The
crustal layering or anisotropy is of much greater significance than the
assumed elastic behaviour of the ice as regards surface deformation
(e.g. Manconi et al., 2007; Geyer and Gottsmann, 2010). The mechanical
properties of ice are variable (Schulson and Duval, 2009). For example,
typical laboratory values of stiffness or Young's modulus (E) can range
from as high as 15 GPa (Gammon et al., 1983; Schulson and Duval,
2009) to more commonly 8–9 GPa, depending on temperature, and
grain size and orientation (Parameswaran, 1987). The stiffness values
are only moderately anisotropic (Schulson and Duval, 2009). These
are dynamic values, however. Static values are more difficult to measure
because of time-dependent deformation in ice. Estimated typical static
or field values for Young's modulus of ice are around 1 GPa (Schulson
and Duval, 2009). Poisson's ratios for ice are commonly between 0.2
and 0.4 (Schulson and Duval, 2009). In the modelling we use a Young's
modulus somewhere between typical field and (dynamic) laboratory
values, or 4 GPa. Also, we use a Poisson's ratio of 0.3 and a density of
920 kg m− 3. The general crustal and ice parameters used in the
numerical models are given in Table 1.
In all models we assume a strong coupling between glacier and
bedrock or crustal surface (except at the location of the caldera lake)
using the same assumptions as Geyer and Bindeman (2011). More
specifically, if the coupling between the ice and the bedrock is of
sufficient strength, stresses within the crust are transmitted to the ice.
Then the ice can be considered to act mechanically as part of the layered
crust. The other mechanical situation is where the ice and crust are
weakly bonded, in which case slip may occur along the weak boundary
and stresses would not be transferred from the bedrock surface to the
ice. We consider one such scenario where the ice and crust are not directly coupled, designed to represent a caldera lake. The lake depth is
0.5 t (half the thickness of a typical crustal layer) and its width is a
(the radius of the magma chamber/caldera). The lake is modelled as a
free surface at all edges. This follows because the contact between
water and the bedrock below as well as the contact with the ice above
are surfaces of zero shear stress. As previously stated, many if not
most calderas develop a caldera lake at some point, particularly those
calderas formed under ice (Fritz et al., 1990).
If the stresses within a volcano are suitable for the formation of a
caldera then displacement would be likely to occur along a bounding
ring-fault (circumferential fault). In order to incorporate the mechanical
response to ring-faulting we include in one of the models a soft (lowYoung's modulus) vertical zone directly above the magma-chamber
edge. This zone is supposed to represent a typical caldera fault, a ringfault (without a ring-dyke) consisting of a highly fractured and mechanically soft damage zone with respect to the host rock (Browning and
Gudmundsson, 2015). The magnitudes of the vertical and horizontal
displacements depend on the magma chamber size – in this case the
chamber radius a is 4 km (its horizontal diameter thus 8 km). All models
assume uniform vertical displacement of the roof, that is, a piston-like
subsidence irrespective of the absence (as in most models) or the
presence (as in one model) of the ring-fault itself.
4. Comparison of numerical and analytical solutions
Periods of unrest are often characterised by surface inflation or deflation of the volcano. This deformation signal is commonly explained
in terms of a magmatic excess pressure change (pe) in the associated
magma chamber of radius a and depth d below the surface, modelled
as a nucleus of strain. In volcanology, such a nucleus is normally referred
to as the “Mogi model” (Mogi, 1958; Fig. 6), although the nucleus-ofstrain solution with application to volcanoes was initially derived by
Anderson (1936). Mogi's analytical solution can be replicated using
the finite element method (e.g., Masterlark, 2007; Hickey and
Gottsmann, 2014). If a Poisson's ratio of 0.25 is assumed for the elastic
Table 1
Model parameters.
Material
Young's modulus (E)
(GPa)
Density (ρ)
(kg/m3)
Poisson's ratio
(v)
Crust
Glacial ice
Hyaloclastite
Ring fault
40
4
1
0.1
2500
920
2500
2500
0.25
0.3
0.25
0.25
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
87
the surface above the magma chamber, respectively. Also, pe is the
magmatic excess pressure in the chamber, a is the radius of the
chamber, μ is shear modulus, d is the depth to the centre of the chamber
below the surface of the earth (Fig. 6), and r is the radial coordinate at
the surface. At the point right about the centre of the magma chamber,
we have r = 0, and the maximum vertical displacement uz becomes
(Fig. 6):
uz ¼
Fig. 6. Point-pressure source, a nucleus of strain, referred to as the Mogi model in volcanology. Such a model is commonly used for explaining surface deformation above an
assumed spherical magma chamber. The solid curve gives the vertical surface displacement, which is maximum above the chamber centre. The dashed curve gives the horizontal surface displacements. The magma chamber, with radius (a) is subject to negative
excess pressure, that is, underpressure (−pe) and located at a depth (d) below the surface.
(cf. Eqs. 1–3).
half-space – generally a reasonable assumption – then the basic
equations of the Mogi (1958) can be presented as follows:
uz ¼
3pe a3 d
3=2
2
4μ r2 þ d
ð1Þ
ur ¼
3pe a3 r
3=2
2
4μ r 2 þ d
ð2Þ
where uz and ur are the vertical and horizontal (radial) displacements at
3pe a3
4μd
2
:
ð3Þ
Magmatic underpressure, that is, pressure less than lithostatic, is
often regarded as the condition for ring-fault and ring-dyke formation.
In fact, an underpressure or contracting nucleus-of-strain was
Anderson (1936) original model for the formation of ring-dykes and
the connection with the Mogi model is straightforward (cf. Kusumoto
and Gudmundsson, 2009).
In Fig. 7 we show the numerical results of a two-dimensional
(circular) chamber subject to an underpressure of 10 MPa, a common
underpressure value when considering ring-fault formation (Geyer
et al., 2006; Gudmundsson, 2007; Kusumoto and Gudmundsson,
2009). We model the horizontal and vertical displacement at the surface
of the ice and at the surface of the bedrock (the crust under and outside
the ice sheet) for two magma-chamber depths: 3 km and 5 km. The
modelled chamber radius is 1 km and is thus small in relation to the
chamber depth below the surface, as it should be for a “Mogi model”.
There are two basic model configurations. The first one (Fig. 7A) has
no caldera lake, but the second one (Fig. 7B) has a caldera lake between
A)
B)
Fig. 7. Crustal surface and ice-surface displacements resulting from negative excess pressure, or depressurisation, of −10 MPa in a circular chamber with a roof at 3 km and 5 km depth
below the crustal surface. The upper model (A) shows a homogeneous isotropic crust with an upper ice layer. The lower model (B) also shows a homogeneous isotropic crust but this time
incorporates a rectangular free surface, designed to replicate a caldera lake. Total displacement contours are given on the right in metres. The magnitude of displacements in the two
models are generally similar, but the displacement patterns differ somewhat, especially in the ice layer.
88
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
the bottom of the ice and the bedrock surface. The lake is included in
several of the models in this paper because, as indicated above. Such
lakes are common in the many calderas located beneath ice in Iceland
(Gudmundsson et al., 1997, 2004, 2012).
The surface displacements, vertical and horizontal, are very small
(less than 3 cm) for this type of loading (Fig. 7), suggesting that a
‘Mogi model’ is, as a rule, not very suitable for generating large (tensof-metre scale) subsidences. The geometries of the displacement curves
(Fig. 7), however, are in excellent agreement with those obtained from
the Mogi model (Fig. 6). The displacement results (Fig. 7) are shown
both for the surface of the rock (the crust under and outside the ice)
as solid lines as well as for the surface of the ice itself, as broken lines.
As is also seen in subsequent models, the caldera lake has great effects
on the displacement curves for the surface of the ice. The other main
results as regards the surface-displacement curves will be discussed in
context of the later and more realistic models, to which we turn now.
5. Roof subsidence of a sill-like magma chamber
Here we present the results of the stresses and surface displacement
induced by a given subsidence of the roof of a sill-like magma chamber.
The chamber, modelled as a cavity within a crustal segment, is given a
zero excess pressure condition at its lower boundary and prescribed a
vertical displacement at the upper boundary in all models apart from
those simulating slip on the ring-fault. Thus, in the models the chamber
is in lithostatic equilibrium with its surroundings prior to the prescribed
vertical displacement, that is, the roof subsidence. While these models
are partly “inspired” by the events in Bardarbunga 2014–15, they are
completely general and apply to all central volcanoes – collapse calderas
in particular – under ice. We explore two main types of models, namely
where the magma chamber is located (1) in a homogeneous, isotropic
crustal segment, and (2) in a layered, anisotropic, crustal segment.
Based on information from Bardarbunga, where the maximum subsidence of the ice surface is estimated at around 60 m (Hensch et al.,
2015), we explore vertical roof displacements or subsidences from
20 m to 100 m. To cover the likely shallow chamber depths, we consider
chambers with roofs at depths of 3 km, 5 km, and as an extreme shallow-chamber depth, 7 km below the crustal or rock surface.
5.1. Homogeneous crustal segment
These models are somewhat similar in set-up as the elastic halfspace or the Mogi model (Mogi, 1958; Kusumoto and Gudmundsson,
2009; Fig. 7). There are, however, three main differences between the
present models (Fig. 8) and the numerical models in Fig. 7. First, the
shallow magma chamber (cavity) has here (Fig. 8) a sill-like geometry
in contrast to the spherical or point-like Mogi source (circular in
Fig. 7). Also, here the radius of the chamber a is 4 km (Fig. 8), and
thus 4-times the radius of the previous circular chamber (Fig. 7), and
with a maximum thickness 2b of 2 km. Second, the displacements at
the surface of the bedrock (the crust) and the ice result here from
prescribed chamber-roof vertical displacement or subsidence rather
than the underpressure in the models in Fig. 7. Third, the subsequent
sill-like chamber models analyse chambers in a layered (anisotropic)
crustal segment rather than in an elastic half-space as is done in the
Mogi model (Fig. 7), and some of the sill-like models also include a
lake beneath the ice, thereby forming a free surface.
Fig. 8 shows the vertical (uz) and horizontal or radial (ur) surface
displacements of the ice and the bedrock (or crust) resulting from a
vertical chamber-roof displacement or subsidence of 100 m. Here
there is no ring-fault. The chamber roof is prior to the displacement at
different depths below the bedrock or crustal surface d, namely at
Fig. 8. Vertical and horizontal ground-surface (solid curves) and ice-surface (dashed curves) displacements in metres resulting from a maximum prescribed chamber-roof displacement of
100 m. Here the crust is non-layered. In all model runs the maximum vertical surface displacements of the ice and the crust occur over the chamber centre (in the caldera centre) whereas
the maximum horizontal displacements of ice and crust occur around 5 km away from the caldera centre. Vertical crustal surface displacements are larger than the ice displacements, but
the opposite is true for the horizontal displacements. Insets display the distances from the chamber where b1 m of vertical and horizontal displacements would be observed. Note that
significant (10–20 cm) vertical displacement occurs out to 15 to 16 km from the caldera centre. Similarly, horizontal displacements of 10–20 cm occur out to 30 to 45 km (depending
on chamber depth) from the caldera centre. Chamber radius (a) is 4 km and half-thickness (b) is 1 km, depth (d) varies between each model run, as indicated by the separate line colours.
89
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Table 2
Model results assuming a 100 m displacement of a chamber roof residing at 3 km.
Setup
Homogeneous
Heterogeneous A
Heterogeneous B
Ring-fault
Uz Max (−m)
Ur Max (−m)
x distance (km) when
Uz = 0.5 m
x distance (km) when
Ur = 0.5 m
Crust
Ice
Crust
Ice
Crust
Ice
Crust
Ice
97.21
79.12
84.14
98.5
71.75
64.14
64.68
70.18
24.75
2.3
5.78
23.09
32.22
10.26
6.92
30.31
13.8
12.2
11.6
13.5
14.4
12.3
11.6
14.1
18.6
11.8
16.9
18.4
19.1
14.7
16.3
18.9
depths of 3, 5 and 7 km. Salient model results are shown in Table 2, but
here we summarise some of the basic results (Fig. 8) as follows:
(1) The maximum vertical displacement (shown as negative
displacement or surface subsidence), both of the ice and the
crust, is above the centre of the magma chamber. The subsidence
reaches about 97 m in the bedrock/crust and about 78 m in the
ice (Table 2). The subsidence changes to uplift or doming at
distances of 15–18 km (depending on the chamber depth)
from the surface point right above the chamber or caldera centre
(Fig. 8). Unless otherwise stated, chamber/caldera centre in the
discussion that follows refers to this surface point.
(2) The horizontal displacement towards the centre (above the centre of the chamber), shown as negative, reaches its maximum at
4–5 km from chamber/caldera centre. The horizontal displacement reaches a maximum of about 25 m in the crust and about
32 m in the ice (Table 2). For the chamber at 3 km depth, however, the horizontal displacement becomes positive (movement
away from the centre) at about 25 km distance from the centre.
(3) The vertical surface displacement, both in the ice and in the
bedrock/crust, is less than that of the chamber roof. There is
thus not a one-to-one correspondence between the displacement at the surface either of the ice or the crust and the chamber
roof subsidence.
(4) The vertical and horizontal displacements extend to distances far
from the chamber/caldera centre. Thus, in both the bedrock/crust
and the ice the vertical displacement is in excess of 0.5 m out to
distances of about 14 km, whereas the horizontal displacements
are in excess of 0.5 m out to distances of about 19 km (Table 2).
Generally, significant surface displacements associated with the
chamber-roof subsidence of 100 m occur at lateral distances of up
to 40–50 km (in the ice as well as in the bedrock/crust) from the
chamber/caldera centre. For example, a chamber located at 3 km
depth produces horizontal surface displacements of 20 cm at approximately 21 km from the chamber/caldera centre, while a chamber at
7 km depth produces the same displacement at approximately 33 km
from the centre. Thus, for the imposed vertical displacement of the
chamber roof, large horizontal displacements are expected out to tens
of kilometres from the chamber, and these should be easily detected
in the ice or at the bedrock/crustal surface by geodetic measurements.
In the second set of homogeneous crustal models, we consider the
effects of a pressurised deep-seated reservoir, such as are common as
magma sources for shallow chambers in Iceland (Gudmundsson,
2012) (Fig. 9). Doming is modelled as being the effect of 10 MPa excess
magmatic pressure acting on the roof (a boundary load) of the reservoir.
The general effect of doming is to reduce the magnitude of vertical and
horizontal surface displacements but increase the surface area where
those displacements are significant. In other words, the subsidence
becomes much less concentrated at the surface immediately above the
shallow chamber.
In the third set of homogeneous crustal models, we added vertical
faults (Fig. 10). These are supposed to represent a two-dimensional
version of a caldera ring-fault. The fault is modelled as a soft elastic
inclusion, that is, as a zone with a low Young's modulus. This is because
active or recently active faults have generally lower Young's moduli
than most of the host rock because the fault is composed of a fractured
damage zone and breccia fault core (Gudmundsson, 2011; Browning
and Gudmundsson, 2015). The precise relationship between damage
and Young's modulus evolution in caldera settings is, as yet, poorly
constrained. The results (Fig. 10) are similar to those of the previous
models without a fault (Fig. 8) but differ in that surface subsidence is
concentrated within a narrower region around the chamber margin.
In addition the crust experiences significant (~ 30 cm) positive
(doming) displacement, measured as an inflation signal between
approximately 15 km and 20 km from the centre.
5.2. Layered (anisotropic) crustal segment
Two layered crustal-segment models were run (Fig. 11). One model
(A) has two soft layers in-between stiffer crustal units, while the other
model (B) has three soft layers, including the top layer. All the soft layers
have a stiffness of 1 GPa, which corresponds to the stiffnesses of soft
hyaloclastites (basaltic breccias) and of glacial sediments, such as are
common in most active volcanoes in Iceland. The layers are modelled
as soft to explore the maximum effects that sediments and soft breccias
could have on the displacement fields. Introducing mechanical heterogeneities and anisotropies through soft layers with low Young's moduli
into the model setup has the following effects:
(1) There is a general reduction in magnitudes of the far-field displacements. That is, the horizontal and vertical displacements
far from the chamber/caldera centre are smaller in the layered
models than in the non-layered models (Table 2).
(2) The maximum vertical displacements are also smaller in the
layered models than in the non-layered models. More
specifically, the maximum surface vertical displacements in the
bedrock/crust are 79–84 m in the layered models but 97–98 m
in the non-layered models (Table 2). Similarly, the maximum
surface displacement in the ice in the layered models is
64–65 m, but 70–72 m in the non-layered models.
(3) The maximum horizontal surface displacements are much smaller in the layered models than in the non-layered models. In the
layered models the maximum surface horizontal displacement
is 2–6 m but ~25 m in the non-layered models.
The general effect of layering is to reduce the displacements
measured at the surface of the bedrock/crust and the ice. The reasons
for the reductions are partly that the stresses become “dissipated” at
the contacts with the soft layers. Similar results have been obtained in
general studies of surface deformation associated with various pressure
sources, such as dykes (Gudmundsson, 2003). Crustal segments with
alternating stiff and very soft layers generally transport less stress and
deformation to the surface than non-layered segments, or segments
where all the layers have similar mechanical properties. Well-known
examples of the reducing effects of mechanically contrasting layers on
90
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
A
B
Fig. 9. Vertical and horizontal ground-surface and ice-surface displacements in metres resulting from a maximum prescribed chamber-roof displacement of 100 m, and either A) doming
overpressure of 10 MPa in a deep-seated reservoir (blue lines) or B) fixed lower boundary (red lines). Here the crust is non-layered. Doming has the general effect of reducing the
maximum horizontal displacement but increasing the radial distance or area over which significant horizontal displacement occurs. Magnitudes of vertical displacement are not greatly
affected by doming, but the area of vertical subsidence increases in the absence of a fixed boundary. Shallow chamber radius (A) is 4 km and half-thickness (B) is 1 km, depth (d) is 3 km.
Deep reservoir radius is 16 km with a half-thickness of 2 km at a depth of 10 km.
Fig. 10. Resultant displacements from a model which includes a soft (Young's modulus 1 GPa) fault zone, representing a ring-fault. In this model the crust is homogeneous and isotropic,
i.e., not layered. Results are similar to those in models without the weak fault zone, but differ in that surface subsidence is concentrated within a narrower region. Additionally, the crustal
segment experiences a positive vertical displacement (~30 cm), inflation between around 15 and 20 km from the chamber centre. Chamber radius (a) is 4 km and half-thickness (b) is
1 km, depth below the surface (d) is 5 km.
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
91
A
B
Fig. 11. Displacements generated through chamber-roof subsidence of 100 m in a layered crustal model. In the upper model (A), the layering is configured so that two soft layers (Young's
modulus E = 1 GPa) lie in-between stiffer units which have the same Young's modulus as the crust. In the lower model (B) we add three soft layers in between crustal units, where the
uppermost is in contact with the glacier. Coloured layers indicate soft hyaloclastite as material inputs shown in the model setups on the left. Graphs on the right indicate the vertical and
horizontal displacements in the crustal (solid) and ice (dashed) surface for each layer configuration. Generally, there is a reduction in far-field displacements and the magnitude of local
displacements is less in both model set-ups, compared to the previous homogeneous setups. Chamber radius (a) is 4 km and half-thickness (b) is 1 km, depth (d) is 5 km.
surface stresses and deformation/displacement relate to emplacement
of dykes and other vertically fluid-driven fractures (Gudmundsson,
2003; Philipp et al., 2013).
6. Ring-fault subsidence
We also modelled the effects of a piston-like subsidence along a
ring-fault on the surface displacement fields. In view of the results
from Bardarbunga, where inferred vertical maximum displacement in
the ice inside the collapse caldera is about 60 m (Hensch et al., 2015),
we impose 50 m vertical displacement on the ring-fault (Fig. 12). The
ring-fault, the fault zone, is modelled as a soft inclusion, with a Young's
modulus of 0.1 GPa. We tried other stiffnesses for the fault zone, such as
0.01 GPa, but the overall results remained similar. The crust itself is nonlayered in this model with the properties used in the earlier non-layered
models (Table 1).
The results (Fig. 12) show that the displacement, both the vertical
and the horizontal, becomes more concentrated at and within the caldera (the ring-fault) than in the previous roof-subsidence models without
ring-fault. The maximum subsidence of the bedrock/crust is the same as
that of the fault, namely 50 m, but the ice subsidence is greater, or 60 m.
This is because the ice can bend or subside somewhat into the caldera
lake (and/or soft sediments) at the contact between the ice and the
crust, whereas the crust clearly cannot do so. For the same reason, the
horizontal displacement (towards the centre) at the surface of the ice
also exceeds that of the crust. Both reach a maximum at the location
of the ring-fault, the vertical displacement of the ice (the fault throw)
being up to about 17 m and that of the bedrock/crust up to 10 m.
These results illustrate various aspects of the effect of ring-fault subsidence in a caldera located beneath ice, including the following:
(1) The crustal displacements, the horizontal and, in particular,
the vertical, reflect strongly the ring-fault geometry. This
means that both displacements are maximum at the caldera
fault. In fact, the vertical displacement reaches its maximum of
50 m at the fault and stays the same throughout the roof of the
chamber.
(2) The caldera lake magnifies the surface displacement of the ice.
The horizontal displacement in the ice is considerably larger
than that in the crust. And, most importantly, the total vertical
displacements in the ice exceed that imposed on the ring-fault
by about 10 m. This is because of the caldera lake beneath the
ice into which the ice can subside.
(3) Displacement of 50 m is so large that it would certainly cut
through the ice as a fault. The inferred vertical displacement at
the location of the ring-fault are about 17 m. No tensile or
shear strength is given to ice in the model, so it does not fracture.
But 50 m vertical displacement in an ice sheet of thickness, say,
800–1000 m would become very clearly through-faulted.
92
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Caldera lake
2t
Glacier
Layers
t
Vertical inclusion (ring fault)
d
0.1 GPa
Displacement
vector
b
Axis of symmetry
Subsidence (m)
t
50 m
Magma
chamber
a
Uz and Ur displacement (m)
0
-10
0
-20
-30
-0.5
Uz
-40
Ur
crust
ice
-50
-1
15
-60
35
55
-70
0
10
20
30
40
Distance (km)
Fig. 12. Piston-like vertical subsidence of 50 m along the inner edge (z-axis) of a ring-fault, a soft elastic zone which extends from the crustal surface to 3 km depth. The soft fault zone has a
Young's modulus of 0.1 GPa and the crustal segment is homogeneous and isotropic. Displacements are highly focussed within the caldera region, with maximum vertical and horizontal
displacement greatest in the ice surface. The caldera lake acts to magnify displacements within the ice, presumably because the ice is able to subside into the lake surface. Chamber radius
(a) is 4 km and half-thickness (b) is 1 km, depth below the surface (d) is 5 km.
7. Discussion: Implications for the 2014–15 Bardarbunga–Holuhraun
episode
The underpressure or withdrawal-of-magmatic-support model is
often favoured when explaining the formation of calderas, both in
analogue-model setups (Acocella et al., 2000; Holohan et al., 2011)
and for explaining geophysical observations (Peltier et al., 2008;
Kusumoto and Gudmundsson, 2009). Most recently this model has
been invoked to explain ice surface subsidence above the Bardarbunga
Caldera (Sigmundsson et al., 2015). The assumption is then that a volume of magma was removed from a chamber by lateral magma propagation, eventually forcing an eruption some 45 km from the central
volcano. Similar ideas have been offered to explain the occurrence of
lavas outside the main central volcanoes and within the active rift
zone of Iceland (Sigurdsson and Sparks, 1978), although more recent
studies have shown that alternative explanations with predominating
vertical magma propagation are equally plausible (Hartley and
Thordarson, 2012). The competing hypothesis is that the Holuhraun
lavas, and many other large and rather primitive basaltic fissure
eruptions in Iceland, are fed by regional dykes which are injected from
magma reservoirs at a much greater depths (15–25 km) than the
shallow chambers (Gudmundsson et al., 2014).
The models presented in this paper have certain implications for
volcano-tectonic processes in central volcanoes in general. Further
implications apply primarily to calderas located in ice sheets such as
many calderas in Iceland – Bardarbunga in particular. We consider
first the implication for the magnitude of the surface displacements
and the size of the area affected (the surface area showing significant
displacement). Both aspects of the deformation are very important,
particularly when trying to separate the deformation associated with a
caldera and/or a shallow magma chamber from that associated with
simultaneous dyke emplacement.
7.1. Surface displacements
Vertical surface displacements of 10–20 cm extend out to distances
of 15–16 km from the centre of the caldera, and horizontal displacements of similar magnitude to 20–30 km (Fig. 8). For the horizontal displacement, 10 cm displacements occur out to 40–50 km from the centre,
depending on the depth of the chamber. These refer to the non-layered
models and the exact distances for the displacement mentioned depend
on the depth of the chamber: the surface displacements induced by the
deepest chamber, at 7 km, extend for the greatest distances from the
centre. If the surface displacement would relate partly to a deepseated reservoir, as we propose here, say a reservoir at the depth of
15–20 km, then significant surface displacements would extend still
further from the centre. Using the same model configuration and properties but roof-subsidence varying from 20 m to 100 m, the results are
similar – significant displacements extend to 15–20 km from the centre
(Fig. 13A) and are only slightly less when a vertical ring-fault is introduced (Fig. 13B).
The layered models produce less displacements, both in magnitude
and lateral extension from the centre (Fig. 11). However, these models
are with somewhat extreme layering since the soft layers have stiffness
or Young's modulus of only 1 GPa, which is low for hyaloclastites and
sedimentary rocks. Nevertheless, there are still large surface
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
93
A
B
C
D
Fig. 13. Total vertical (A and C) and horizontal (B and D) displacements in crustal-surface (solid lines) and ice-surface (dashed lines) resulting from vertical chamber roof subsidence of
between 20 and 100 m. Upper displacement curves (A and B) are from a homogeneous crustal model, whereas the lower curves (C and D) are from models which include a soft vertical
zone, a ring-fault (C and D). The results are similar for both types of models, with slight variations in subsidence around the fault area. All models indicate significant far-field surface displacements, irrespective of the amount of roof subsidence. However, roof subsidence has a major control on the local surface displacements directly above the magma chamber. In all
model runs shown here chamber radius (a) is 4 km and half-thickness (b) is 1 km, with a chamber depth of 3 km.
displacements of 50 cm at distances of about 12 km (for the vertical
displacement) and 12–17 km (for the horizontal displacement) in the
bedrock/crust, and somewhat larger distances in the ice (Table 2).
Larger displacements are obtained from the 50-m-fault displacement
model (Fig. 12).
Overall the displacement results indicate that, for the models considered, large displacements, of the order of tens of centimetres or hundreds of millimetres, should be detected out to distances of 10–20 km,
for the vertical displacement, and 20–30 km or more for the horizontal
displacements. Even for a small roof-subsidence of 20 m, the horizontal
displacement at 10–12 km distance from the centre is still of the order of
tens of centimetres (Fig. 13). Results of this kind show clearly the effect
of nearby subsidence of a magma-chamber roof, or a collapse caldera
displacement, and should make it possible to distinguish between
displacements induced by such a subsidence and those induced by a
dyke formed in the same volcano-tectonic episode.
The displacement field associated with the subsidence of the ice in
the Bardarbunga episode in 2014–15 is educational in this respect. For
the period up to 6 September 2014 the GPS-estimated maximum
displacement or subsidence in the ice in the Bardarbunga Caldera was
about 16 m (Sigmundsson et al., 2015), and the entire cumulative
displacement during the episode 2014–15 is estimated at over 60 m
(Hensch et al., 2015). Dyke emplacement was essentially completed
by 31 August when the main eruption began (Gudmundsson et al.,
2014; Sigmundsson et al., 2015), and no significant horizontal
dyke-induced displacements were detected after 4 September 2014
(Ofeigsson et al., 2015). The horizontal displacements induced by the
dyke can thus largely be separated from those induced by the
subsidence measured in the Bardarbunga Caldera.
Our model results suggest that vertical displacement of about
16–20 m, corresponding to period up to about 6 September, should generate horizontal displacements of the order of tens of centimetres towards the Bardarbunga Caldera within 10–12 km from the centre of
the caldera. Similarly, horizontal displacements of many tens of
centimetres are expected out to distances of up to tens of kilometres,
depending on the model used – in particular, the assumed depth of
94
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
the shallow chamber and mechanical properties of the host rock and the
ring-fault. The measured displacements at the GPS stations in the crust
outside the ice, some of which are at 13–17 km from the centre of the
caldera, are significantly less than expected from the models
(Ofeigsson et al., 2015; Sigmundsson et al., 2015). The difference may
be partly related to the modelling procedure but most likely indicates
that only part of the subsidence in the ice within the Bardarbunga Caldera is actually directly related the magma-chamber roof subsidence, or
ring-fault displacement.
The last point is also of importance when interpreting the subsidence measured in the ice within the caldera. There are several remarkable features of the ice subsidence, as shown in maps by Sigmundsson
et al. (2015), including the following:
(1) The maximum subsidence is about 3 km from the northern caldera rim and about 5 km from the southern and south-eastern
rims.
(2) The subsidence at the rims, at the ring fault itself, is much smaller
than the maximum subsidence – in fact about zero in the early
stage of the subsidence.
(3) Fracture development at the surface of the ice is comparatively
small, with no major caldera-related fault cutting through the ice.
These observations and measurements, when compared with the
ring-fault subsidence model (Fig. 13), suggest the following interpretations. First, the displacement along the ring-fault is small in comparison
with the overall subsidence in the ice. In particular, displacements of the
order of 20–60 m along the ring fault would without doubt have propagated faults through the ice (say, vertical displacements of 10–20 m;
Figs. 11 and 12) – and these faults are not observed. From standard
fracture mechanics (Gudmundsson, 2011) and the mechanical
properties of a typical ice (Schulson and Duval, 2009; Table 2), a
close-to vertical normal fault (Hensch et al., 2015) with displacement
of up to tens of metres in ice of thickness of several hundred metres –
in fact, the ice thickness is only 200–300 m above part of the caldera
rims – would become a through crack, that is, reach the bottom and surface of the ice sheet. This conclusion is the same even if there is a caldera
lake beneath the ice (Fig. 13). Since normal faults with these throws are
not observed, the cumulative vertical ring-fault displacements cannot
be of the of the order of tens of metres, and is most likely of the order
of metres or less.
Second, for a porous-media chamber, as most chambers presumably
are (Gudmundsson, 2012), the maximum subsidence, if caused by
magma flow out of an underlying magma chamber, would normally
be, initially at least, close to the ‘outlet’, that is, the intersection of the
dyke or sheet transporting the magma with the boundary of the chamber and the active ring-fault. Sigmundsson et al. (2015) propose that the
subsidence of the ice is directly related to chamber roof-subsidence associated with magma flowing laterally along a dyke that dissects a
chamber along its southeast margin. It is not clear from the subsidence
data, however, why the maximum subsidence is then not at the outlet
and the active ring-fault but rather close to the northern margin of the
chamber/caldera.
In fact, the inferred segmentation of the dyke, with distances
between nearby tips of segments up to kilometres (Sigmundsson
et al., 2015), is a strong argument against lateral flow of magma from
a chamber beneath Bardarbunga and to the volcanic fissures in
Holuhraun and an argument for vertical flow of magma from a deepseated reservoir (Gudmundsson et al., 2014). The arguments against
the lateral flow between dyke segments are many, including the following. (1) There is no seismicity between the nearby ends of some of the 8
segments, particularly between segments 1 and 2 and 5 and 6
(Sigmundsson et al., 2015, Extended Data Fig. 2), suggesting that no
magma migrated laterally between them. The zones connecting many
of the segments, being highly oblique to the overall strike of the dyke,
are zones of high shear stress making it highly unlikely that a magmadriven fractures could propagate along the zones without triggering
earthquakes. Earthquakes are, in fact, used as criteria for identifying
magma paths. It follows that absence of earthquakes, that is, seismically
quiet zones, would normally mean absence of magma paths. (2) In the
unlikely event of aseismic magma-path formation at shallow depths
from segment 1 to 2, and from segment 5 to 6, then the same magma
would have to flow from the shallow depths vertically down to at
least 10 km depth in segments 2 and 6. Downward flow of magma on
this scale is not supported by any observations and does not agree
with well-established physical principles of fluid dynamics and dyke
propagation (Gudmundsson, 2011). Thus, the segments of the regional
Bardarbunga–Holuhraun dyke were presumably formed primarily
through vertical flow of magma from the proposed deep-seated
magma reservoir (Gudmundsson et al., 2014).
7.2. Mechanisms for the ice and ring-fault subsidences
If excess magma pressure decrease in a shallow chamber is not the
main cause of the subsidence in the ice in Bardarbunga, what alternative
mechanisms exist? One obvious mechanism, well known from caldera
studies, is slight doming or inflation of the volcanic field hosting a
shallow magma chamber. Doming was in fact detected at GPS stations
in a large area surrounding the Bardarbunga Caldera a few months
before the unrest (Ofeigsson et al., 2015). Such a doming, as small as
of the order of centimetres, is known to be one of the principal
mechanisms for generating caldera collapses (Gudmundsson, 2007),
particularly along normal ring-faults (Gudmundsson, 1998). Focal
mechanisms suggest that the slip on the ring-fault of Bardarbunga in
the 2014–15 episode was primarily through normal faulting
(Bjarnason, 2014; Hensch et al., 2015; Riel et al., 2015). Most of
the ring-fault seismicity occurred at shallow depths (b 3 km)
(Hjörleifsdóttir et al., 2015), in agreement with the ring-fault seismicity
being related to stress concentration above the margins of a proposed
shallow magma chamber, which is the model suggested here (cf.
Gudmundsson, 2007).
The doming or inflation of the volcanic field or system containing the
Bardarbunga Caldera is most likely related to the associated deepseated reservoir receiving new input of melt or magma. As doming
begins, stress concentration at the ring-fault of the Bardarbunga Caldera
results in subsidence – by how much we do not really know. The
subsidence and associated faulting and possible ring-dyke formation
(Gudmundsson et al., 2014) reduces the effective thickness of the
crustal segment or plate above the shallow magma chamber
(Figs. 5 and 9). The reduction in the effective plate thickness de
encourages further doming of the volcanic field hosting the
Bardarbunga caldera even if the magmatic excess pressure at the
deep-seated reservoir remains constant or decreases slightly for a
while (Gudmundsson, 1998).
7.3. Relation between dyke emplacement and subsidence in the caldera
This last point brings us to the 45-km-long regional dyke and associated eruption, and how they relate to the subsidence in the Bardarbunga
Caldera. The first thing to notice is that the strike of the dyke close to the
caldera/chamber is in perfect agreement with the local trajectories of
the maximum horizontal principal stress around a circular or slightly
elliptical cavity under tension (Savin, 1961; Gudmundsson, 2011).
Further from the chamber/caldera the regional stress field took over,
and the dyke followed the field that has existed in this part of Iceland
for at least 8–10 Ma (Gudmundsson et al., 2014). The main dyke was
injected when the excess pressure in the deep-seated reservoir reached
the conditions (Gudmundsson, 2011):
pl þ pe ¼ σ þ T 0
ð4Þ
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
where pl is the lithostatic stress or overburden pressure at the reservoir
rupture site (in the reservoir roof), pe = pt − pl is the difference between the total fluid pressure pt in the reservoir and the lithostatic
stress at the time of reservoir rupture, σ3 is the minimum compressive
or maximum tensile principal stress, and T0 the local in situ tensile
strength at the rupture site. When the dyke became injected into the
roof of the reservoir and began to propagate up into the crustal layers
above, its overpressure po changed as:
p0 ¼ pe þ ðρr −ρm Þgh þ σ d
ð5Þ
where pe is the magmatic excess pressure in the reservoir at the time of
rupture (and equal to the in-situ tensile strength of the roof at the
rupture site, T0), ρr is the host-rock density, ρm is the magma density,
g is acceleration due to gravity, h is the dip dimension or height of the
dyke above the rupture site, and σd is the differential stress at a particular depth in the crust (the depth of interest). At the magma-chamber
rupture site itself, the stress difference is included in the excess pressure
term, so that there σd = 0. Also, at the rupture site, before the dyke has
propagated and reached any significant height, we have h = 0, so that
the third term in Eq. (5), the buoyancy term, becomes zero. It follows
that the only pressure available to rupture the reservoir roof and drive
the magma out at and close to the roof contact with the magma is the
excess magmatic pressure pe. We also know that pe = T0, that is, the
excess pressure at the time of roof rupture is equal to the in-situ tensile
strength, with in-situ (field) values ranging from 0.5 to 9 MPa (Amadei
and Stephansson, 1997), the most common values being 2–4 MPa
(Gudmundsson, 2011).
It follows that during the rupture and initial propagation of the
resulting dyke, the only driving pressure is pe, of the order of several
mega-pascal. As the dip dimension (height) of the dyke increases,
however, positive buoyancy adds to the driving pressure, so long as
the average magma density is less than the average density of the rock
layers through which the dyke propagates. The magma is olivine tholeiite (Haddadi et al., 2015) so that its density may be taken as about
2700 kg m−3 (Murase and McBirney, 1973). The erupted magma originated at depths somewhere between 10 and 20 km (Bali et al., 2014;
Haddadi et al., 2015), that is, from a deep-seated magma reservoir as
have been proposed under most volcanic systems in Iceland, and the
Bardarbunga System in particular (Gudmundsson et al., 2014). Given
the crustal density in Iceland, then from Eq. (5) the magmatic overpressure p0 or driving pressure of the dyke, at different crustal depths (and
thus with different σd values) could easily have reached 10–15 MPa
(cf. Becerril et al., 2013; Gudmundsson et al., 2014).
In the model presented here, the injection of the main dyke from the
deep-seated reservoir, as well as the subsidence in the Bardarbunga Caldera, were both primarily the consequence of the same process: namely
inflow of magma into the deep-seated reservoir. This inflow may have
started many years before the 2014 episode, particularly from 2006
and onwards as indicated by seismicity (Vogfjörd et al., 2015), and
was certainly noticeable as widespread doming or uplift on GPS instruments for months before the regional dyke injection began in August
2014 (Ofeigsson et al., 2015). There may have been magma flow into
the shallow chamber associated with the caldera, and several smaller
dykes may have been emplaced during the early stages of the episode
– some from the deep-seated reservoirs, others (small radial dykes)
from the shallow chamber. The only dyke to develop into a major
dyke, however, was the 45-km-long regional dyke emplaced over a
period of 2 weeks in August 2014 (Gudmundsson et al., 2014;
Sigmundsson et al., 2015).
The regional dyke presumably came from depths of at least 15–
20 km, perhaps deeper. This is suggested partly by the chemistry of
the erupted lavas (Bali et al., 2014; Haddadi et al., 2015), partly by the
widespread doming detected in the months before the episode,
discussed above, and partly by the earthquake distribution in the area.
From 2012 there were many earthquakes north and northeast of the
95
Bardarbunga Caldera (Vogfjörd et al., 2015), where one of the earthquake swarms (and possible dyke injection) occurred during the first
days of the August 2014 episode. Even more importantly, deep earthquakes occurred in a vertical zone southeast of the Bardarbunga Caldera
from about this time and extended until August 2014 (Vogfjörd et al.,
2015) at roughly the location of the first segment of the regional dyke,
as formed in the first days of the August 2014 episode.
The regional dyke had enormous stress effects on the Bardarbunga
Caldera and, by implication, the associated shallow magma chamber
(Gudmundsson et al., 2014). The stress field induced by the dyke
around the caldera contributed to three important aspects of the 2014
episode; (1) normal faulting along the caldera ring-fault, (2) elongation
of the caldera in a roughly north-south direction, and (3) ductile deformation and flow of the ice, primarily inside the caldera.
Normal faulting is the dominating mechanism on the Bardarbunga
ring-fault during the present episode (Bjarnason, 2014; Hensch et al.,
2015; Riel et al., 2015). This is in agreement with the two main mechanisms of caldera slip proposed here, namely: (1) a combination of
stresses concentrating at the ring-fault as a consequence of slight doming due to excess pressure increase in the deep-seated reservoir
(Gudmundsson, 2007) and (2) dyke-induced stress concentration, particularly at the northern and southern sectors of the ring-fault
(Gudmundsson et al., 2014). Both encourage normal faulting on the
ring-fault itself, while the dyke-induced stresses also encourage strikeslip and reverse-faulting on differently oriented faults away from the
ring-fault (Gudmundsson et al., 2014).
The elongation of the ring-fault in the roughly north-south direction
is due to the compressive and shear stresses that concentrate in the
“breakout areas” around the caldera (Gudmundsson et al., 2014).
Elongation of collapse calderas due to “breakouts” is well known from
other areas (Bosworth et al., 2003). The elongation would encourage
flow of magma to the ring-fault in these sectors, possible ring-dyke
formation – which may partly explain the common non-double couple
earthquakes (Riel et al., 2015) – and contribute to the subsidence of the
caldera roof. The main reason why the earthquake activity along the
Bardarbunga Caldera has been so concentrated in the north and south
parts of the caldera is presumably related to the dyke-induced stresses
in these sectors (Gudmundsson et al., 2014).
While no attempts were made to measure or monitor ice flow in the
ice-sheet cover of the Bardarbunga Caldera and its vicinity during the
2014–15 episode, such flow is likely to have occurred. During the emplacement of the regional dyke east and northeast of the caldera, the
magmatic overpressure in the dyke (Eq. 5) may easily have reached
10–15 MPa. The dyke induced major displacements and thus stresses,
within the caldera, and the high mountains of the caldera rim must
have transmitted those compressive stresses (σH) from the dyke into
the ice (Fig. 14). Depending somewhat on the strain rate, ice flows at
Fig. 14. A simplified E–W profile of the sub-glacial caldera at Bardarbunga volcanic system.
Overpressure (po) from a regional dyke emplaced to the east of the caldera rim imposes
large horizontal compressive stresses (σH) on the high caldera walls. The compression
acts to squeeze ice within the caldera, making the ice behave as ductile up to the surface,
and leading to increased ice flow out of the caldera. The resulting ice flow is likely to have
contributed largely to the measured ice subsidence at Bardarbunga. The regional dyke is
not to scale in this diagram and the sub-glacial caldera topography is a representation of
the likely topographic setting of Bardarbunga, modified from Björnsson (1988).
96
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
pressures or stresses of less than 1 MPa, so that stresses of up to 10 MPa
– somewhat diminishing with distance from the dyke – would certainly
have caused flow in the ice within the caldera. The main flow would
have been within the caldera because that is where the mountains are
high – the caldera rim – and can thus most easily transmit the dykeinduced stresses to shallow levels in the ice sheet (Fig. 14). Since ice
flows from higher to lower pressure, the dyke-induced stresses would
have encouraged ice flow out of the caldera.
How much the ice flow may have contributed to the measured 60 m
subsidence in the ice is unknown. The comparatively minor fracturing at
the surface of the ice during its subsidence would suggest that the ice
was flowing right up to the surface. Flow in and down-bending of the
ice may have contributed significantly to the measured subsidence. In
this model, the flow or creep or strain rate was highest just after the emplacement of the regional dyke, and then became gradually lower, as is
typical for a creeping response to sudden load or displacement (here the
dyke emplacement). Down-bending would be encouraged by a caldera
lake and/or soft sediments (subject to earthquake shaking) existing beneath the glacier (Fig. 5).
As the excess pressure pe in the deep-seated reservoir declined, the
doming-related ring-fault displacement also gradually decreased and,
as pe approached zero, the subsidence stopped altogether. It is clear
that long before the eruption came to an end on 27 February 2015 the
earthquake activity associated with the Bardarbunga Caldera had greatly diminished. Also, the subsidence measured in the Bardarbunga Caldera ceased several weeks before the end of the eruption. These and other
observations suggest that the pressure decrease in the deep-seated reservoir was partly responsible for the ring-fault slips and associated
earthquakes.
8. Conclusions
We present general numerical models on the effects of “shallow”
magma chamber contraction at various depths, namely the result of
chamber roof subsidence at depths of 3 km, 5 km, and 7 km. In all the
models, the magma chambers are associated with a collapse caldera
which is located beneath a thick glacier or ice sheet. The models are
general, and apply to many volcanoes in Iceland and elsewhere, but
the results are here applied to the 2014–15 volcano-tectonic episode
in Bardarbunga–Holuhraun in Iceland.
Several models were tested for the shrinkage of the magma chamber
through vertical downward displacement of its roof. Some of the
models use a simple elastic crust (elastic half space) hosting the
chamber, with an ice sheet on the top. Others use layered (anisotropic)
crust above the shallow magma chamber, that is, layers with different
stiffnesses (Young's moduli). And still other models have a caldera
lake between the ice sheet and the rock or crustal surface. The simplest
loading used is 100 m vertical downward displacement of the chamber
roof. Other models include different vertical displacement of the roof (in
steps from 20 m to 100 m), as well as displacement of 50 m along a
vertical caldera fault (the ring-fault). Some of the main results are as
follows:
(1) For chamber-roof displacements in the range of 20–100 m, the
models suggest large vertical and particularly horizontal
displacements in the ice and in the bedrock/crust surface under
the ice out to distances of 10–40 km from the caldera centre,
depending on the depth of the chamber and the exact type of
modelling used. The vertical displacements in all models reach
maximum at the surface of the bedrock/crust and the surface of
the ice right above the centre of the subsiding magma-chamber
roof. The horizontal displacements at the surface, however,
reach their maximum values (maximum displacement towards
the chamber or caldera centre) at a distance of 4–5 km from
the centre.
(2) For a non-layered (isotropic) crustal model with a 100 m roof
subsidence, the vertical displacement exceeds 50 cm to a distance of 14 km from the centre and the horizontal displacement
exceeds 50 cm to a distance of 19 km from the centre. A chamber
located at 3 km depth produces horizontal displacement of
20 cm to a distance of 21 km from the centre, and for a chamber
at 7 km depth horizontal displacement of 20 cm is produced to a
distance of 33 km from the centre. Similar results are obtained if
a vertical non-slipping ring-fault is added to the model, but the
displacements show an abrupt change (a break) at the location
of the fault.
(3) The general effect of crustal layering (using mechanically layered
or anisotropic models) is to reduce the displacements measured
at the surface in comparison with those generated in the nonlayered (isotropic) models. The reasons are partly that the stresses become "dissipated" at contacts between still and soft layers.
In a model where the subsidence is related to vertical downward
piston-like displacement by 50 m of the ring-fault, the results show
that the vertical displacement in the crust/chamber roof exactly reflects
that of the ring-fault and reaches a maximum of 50 m. By contrast, the
vertical displacement in the ice follows a curve that reaches its maximum of 60 m in the centre of the caldera. This "extra" vertical displacement in the ice is partly because it can bend or subside somewhat into
the caldera lake below. Displacement of 50 m along the ring-fault is so
large that the fault would most definitely cut through the ice, forming
a through fault with displacements of up to tens of metres (which is
not observed in Bardarbunga, however).
The modelling results have several implications for the interpretation of the 2014–15 Bardarbunga–Holuhraun episode.
(1) First, the measured horizontal displacements in the surface rocks
outside the ice appear to be significantly less than expected from
modelling 60 m vertical displacement. At stations west of the
Bardarbunga Caldera, horizontal displacements towards the caldera of the order of tens of centimetres would be expected but are
not observed. This indicates that the vertical displacement in the
bedrock/crust, and thus the chamber roof-subsidence, is significantly less than the maximum of about 60 m measured in the ice.
(2) Second, a 50 or 60 m piston-like displacement along the ring-fault
is ruled out. The ring-fault would, for such a large displacement,
definitely cut through the ice to form a large and easily visible
fault, but this has not happened. By contrast, there has been
comparatively little fracturing in the ice within the Bardarbunga
Caldera during the subsidence, which suggests that the ice
behaved as ductile, was flowing, right up to its surface. The results
seem to limit the actual ring-fault (piston-like) subsidence to, at
most, a few metres.
(3) Which brings us to the third implication, namely that the 45-kmlong regional dyke generated compressive stresses in the ice within the caldera which resulted in ice flow out of the caldera, thereby
contributing to the measured subsidence in the ice. How large factor the ice flow (and possible down-bending into lake/sediments)
may have been is unknown since no measurements of the ice flow
were made. What is known, however, is that ice flows easily at low
pressures, of the order of 1 MPa, and our calculations suggest magmatic overpressure in the regional dyke of the order of 10–15 MPa.
We interpret the geochemical, seismic, and geodetic data so that the
regional dyke was injected from a large reservoir at 15–20 km depth,
perhaps deeper. Earthquake data suggest that the reservoir received
new magma over many years before the beginning (in August 2014)
of the Bardarbunga–Holuhraun episode, particularly from the year
2006 and onwards. The magma injection resulted in widespread doming (uplift), as detected by GPS instruments in the months prior to August 2014 when the dyke emplacement began (Ofeigsson et al., 2015).
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
In our interpretation, the August 2014 reservoir rupture and regional
dyke injection as well as the ring-fault displacement (caldera subsidence) are all primarily the consequence of the associated reservoir
magmatic pressure increase and doming. The conditions for reservoir
rupture, dyke injection, as well as the overpressure change with vertical
propagation of the dyke, are presented in Eqs. (4) and (5). The effects of
ring-fault formation or reactivation as a result of reservoir-pressure
increase, slight doming, and stress concentration around the chamber/
caldera are discussed in the paper with reference to earlier numerical
models (Gudmundsson, 1998, 2007), all of which suggest doming as a
main mechanism of ring-fault displacement. This mechanism is also in
agreement with the dominating normal-fault focal mechanisms of the
ring-fault earthquakes (Bjarnason, 2014; Hensch et al., 2015; Riel
et al., 2015).
We interpret the seismic and geodetic data so that, in addition to the
45-km-long regional dyke, there may have been several other dyke injections, including a northwest-striking dyke emplaced some 15 km
north of the caldera and several smaller radial dykes/inclined sheets
injected from the shallow chamber beneath the caldera. The shallow
chamber may have received magma from the reservoir during the episode before the radial dyke injection; alternatively, stress concentration
around the shallow chamber, through the external loading (doming),
can have triggered the radial dyke injection (Gudmundsson et al.,
2014).
The regional dyke induced stress concentration at the caldera/shallow
chamber, in addition to that generated by the doming. The dyke-induced
stress concentration contributed to three processes during the 2014–15
episode. First, normal-fault slip along the ring-fault. Focal mechanisms indicate dominating normal-fault slip along the ring-fault (Bjarnason,
2014; Hensch et al., 2015; Riel et al., 2015). Much of the faulting occurred
at the northern and southern sectors of the ring-fault, exactly in the areas
where numerical and analytical models suggest that dyke-induced stress
encourages normal faulting (Gudmundsson et al., 2014). Second, caldera
elongation and “breakout” mechanisms at the northern and southern
sectors of the caldera/chamber were induced by the dyke. These may
have encouraged ring-dyke emplacement. Third, ductile deformation
and flow in the ice inside the caldera. The caldera rim is composed of
tall mountains that transmitted the compressive stress induced by the
dyke to the ice, resulting in ice flow out of the caldera. The rate of flow
of ice was greatest immediately following the dyke emplacement, and
then gradually declined, as is typical of creeping material response to a
sudden load (here the dyke emplacement). How much the ice flow contributed to the measured 60-m-subsidence in the ice is as yet unknown.
As the excess pressure in the reservoir pe decreased below a certain
level, the stress concentration around the ring-fault became too small
for further significant to large slips and associated earthquakes to
occur. This follows because the main slips were through normal faulting,
so that the slips were controlled by the available driving stress at any
time. Thus both the ice flow and the ring-fault subsidence gradually
decreased with time. Significant subsidence in the caldera had
apparently stopped in early February, several weeks before the eruption
in Holuhraun came to an end on 27 February 2015. That the eruption
continued for several more weeks indicates that it ceased only when
the excess pressure in the deep-seated reservoir had vanished
completely, that is, its excess pressure pe had become zero.
Acknowledgements
We thank the Editor, Joan Marti, and two anonymous reviewers for
very helpful comments. AG thanks Ari Trausti Gudmundsson and
Thorvaldur Thordarson for many discussions about the Bardarbunga–
Holuhraun episode. The outstanding services provided by the Iceland
Meteorological Office in general, and the SIL-network group in particular, during the episode are very much appreciated. All the data from the
SIL-network referred to in the present paper, however, are from
published work.
97
References
Acocella, V., 2007. Understanding caldera structure and development: an overview of
analogue models compared to natural calderas. Earth Sci. Rev. 85, 125–160.
Acocella, V., Cifelli, F., Funiciello, R., 2000. Analogue models of collapse calderas and
resurgent domes. J. Volcanol. Geotherm. Res. 1–4, 81–96.
Amadei, B., Stephansson, O., 1997. Rock Stress and its Measurement. Chapman and Hall,
New York.
Anderson, E.M., 1936. The dynamics and formation of cone-sheets, ring dikes, and
cauldron-subsidence. R. Soc. Edinb. Proc. 128–157.
Bali, E., Gudfinnsson, G.H., Gunnarsson, H., Halldorsson, S.A., Jakobsson, S., Riishuus, M.S.,
Sigmarsson, O., Sigurdsson, G., Sverrisdottir, G., Thordarson, T., 2014. Petrology of the
new fissure eruption north of Dyngjujökull. Geoscience Society of Iceland, Autumn
Meeting. Reykjavik, pp. 5–6.
Becerril, L., Galindo, I., Gudmundsson, A., Morales, J.M., 2013. Depth of origin of magma in
eruptions. Sci. Rep. 3, 2762. http://dx.doi.org/10.1038/srep02762.
Bjarnason, I., 2014. Earthquake sequence 1973–1996 in Bardarbunga volcano: seismic
activity leading up to eruptions in the NW-Vatnajokull area. Jökull 64, 61–82.
Björnsson, H., 1988. Hydrology of Ice Caps in Volcanic Regions. University of Iceland,
Societas Scientarium Islandica.
Bosworth, W., Burke, K., Strecker, M., 2003. Effect of stress fields on magma chamber stability and the formation of collapse calderas. Tectonics 22. http://dx.doi.org/10.1029/
2002TC001369.
Browning, J., Gudmundsson, A., 2015. Caldera faults capture and deflect inclined sheets:
an alternative mechanism of ring dike formation. Bull. Volcanol. 77, 1–13. http://dx.
doi.org/10.1007/s00445-014-0889-4.
Cole, J., 1990. Structural control and origin of volcanism in the Taupo volcanic zone, New
Zealand. Bull. Volcanol. 52, 445–459. http://dx.doi.org/10.1007/BF00268925.
Cole, J., Milner, D., Spinks, K., 2005. Calderas and caldera structures: a review. Earth Sci.
Rev. 69, 1–26. http://dx.doi.org/10.1016/j.earscirev.2004.06.004.
De Natale, G., Pingue, F., 1993. Ground deformations in collapsed caldera structures.
J. Volcanol. Geotherm. Res. 57, 19–38.
Deb, D., 2006. Finite Element Method: Concepts and Applications in Geomechanics.
Prentice-Hall, New York.
Druitt, T., Sparks, R.S.J., 1984. On the formation of caldera during ignimbrite eruptions.
Nature 310, 679–681.
Folch, A., Martı́, J., 2004. Geometrical and mechanical constraints on the formation of
ring-fault calderas. Earth Planet. Sci. Lett. 221, 215–225.
Fritz, W.J., Howells, M.F., Reedman, A.J., Campbell, S.D.G., 1990. Volcaniclastic sedimentation
in and around an Ordovician subaqueous caldera, Lower Rhyolitic Tuff Formation, North
Wales. Geol. Soc. Am. Bull. 78, 1246–1256. http://dx.doi.org/10.1130/0016-7606.
Gammon, P.H., Kiefte, H., Clouter, M.J., Denner, W.W., 1983. Elastic constants of artificial
and natural ice samples by Brillouin spectroscopy. J. Phys. Chem. 29, 433–460.
http://dx.doi.org/10.1021/j100244a004.
Geshi, N., Shimano, T., Chiba, T., Nakada, S., 2002. Caldera collapse during the 2000 eruption of Miyakejima Volcano, Japan. Bull. Volcanol. 64, 55–68. http://dx.doi.org/10.
1007/s00445-001-0184-z.
Geyer, A., Bindeman, I., 2011. Glacial influence on caldera-forming eruptions. J. Volcanol.
Geotherm. Res. 202, 127–142.
Geyer, A., Gottsmann, J., 2010. The influence of mechanical stiffness on caldera deformation and implications for the 1971–1984 Rabaul uplift (Papua New Guinea).
Tectonophysics 483, 399–412. http://dx.doi.org/10.1016/j.tecto.2009.10.029.
Geyer, A., Folch, A., Martí, J., 2006. Relationship between caldera collapse and magma
chamber withdrawal: an experimental approach. J. Volcanol. Geotherm. Res. 157,
375–386.
Gray, J.P., Monaghan, J.J., 2004. Numerical modelling of stress fields and fracture around
magma chambers. J. Volcanol. Geotherm. Res. 135, 259–283.
Gregg, P.M., De Silva, S.L., Grosfils, E.B., Parmigiani, J.P., 2012. Catastrophic caldera-forming
eruptions: thermomechanics and implications for eruption triggering and maximum
caldera dimensions on Earth. J. Volcanol. Geotherm. Res. 241–242, 1–12. http://dx.
doi.org/10.1016/j.jvolgeores.2012.06.009.
Grosfils, E.B., McGovern, P.J., Gregg, P.M., Galgana, G.A., Hurwitz, D.M., Long, S.M., Chestler,
S.R., 2015. Elastic models of magma reservoir mechanics: a key tool for investigating
planetary volcanism. Geological Society of London, Special Publications http://dx.doi.
org/10.1144/SP401.2.
Gudmundsson, A., 1998. Formation and development of normal-fault calderas and the
initiation of large explosive eruptions. Bull. Volcanol. 60, 160–171. http://dx.doi.org/
10.1007/s004450050224.
Gudmundsson, A., 2003. Surface stresses associated with arrested dykes in rift zones. Bull.
Volcanol. 65, 606–619. http://dx.doi.org/10.1007/s00445-003-0289-7.
Gudmundsson, A., 2007. Conceptual and numerical models of ring-fault formation.
J. Volcanol. Geotherm. Res. 164, 142–160.
Gudmundsson, A., 2011. Rock Fractures in Geological Processes. Cambridge University
Press, Cambridge.
Gudmundsson, A., 2012. Magma chambers: formation, local stresses, excess pressures,
and compartments. J. Volcanol. Geotherm. Res. 237–238, 19–41.
Gudmundsson, A., 2014. Elastic energy release in great earthquakes and eruptions. Front
Earth Sci. 2, 10. http://dx.doi.org/10.3389/feart.2014.00010.
Gudmundsson, A., 2015. Collapse-driven large eruptions. J. Volcanol. Geotherm. Res. 304,
1–10. http://dx.doi.org/10.1016/j.jvolgeores.2015.07.033.
Gudmundsson, M.T., Sigmundsson, F., Björnsson, H., 1997. Ice–volcano interaction of the
1996 Gjálp subglacial eruption, Vatnajökull, Iceland,. Nature 389, 954–957.
Gudmundsson, M.T., Sigmundsson, F., Björnsson, H., Högnadóttir, T., 2004. The 1996 eruption at Gjálp, Vatnajökull ice cap, Iceland: efficiency of heat transfer, ice deformation,
and subglacial water pressure. Bull. Volcanol. 66, 46–65. http://dx.doi.org/10.1007/
s00445-003-0295-9.
98
J. Browning, A. Gudmundsson / Journal of Volcanology and Geothermal Research 308 (2015) 82–98
Gudmundsson, M.T., Thordarson, T., Höskuldsson, A., Larsen, G., Björnsson, H., Prada, F.J.,
Oddsson, B., Magnusson, E., Högnadottir, T., Petersen, G.N., Hayward, C.L.,
Stevenson, J.A., Jonsdottir, I., 2012. Ash generation and distribution from the April–
May 2010 eruption of Eyjafjallajökull, Iceland. Sci. Rep. 2. http://dx.doi.org/10.1038/
srep00572.
Gudmundsson, A., Lecoeur, N., Mohajeri, N., Thordarson, T., 2014. Dike emplacement at
Bardarbunga, Iceland induces unusual stress changes, caldera deformation, and
earthquakes. Bull. Volcanol. 76, 1–7. http://dx.doi.org/10.1007/s00445-014-0869-8.
Haddadi, B., Sigmarsson, O., Devidal, J.L., 2015. Determining intensive parameters through
clinopyroxene-liquid equilibrium in Grímsvötn 2011 and Bárðarbunga 2014 basalts.
Geophys. Res. Abstr. 17 (EGU2015-5791-2).
Hartley, M.E., Thordarson, T., 2012. Formation of Oskjuvatn caldera at Askja, North
Iceland: mechanism of caldera collapse and implications for the lateral flow hypothesis. J. Volcanol. Geotherm. Res. 227-228, 85–101.
Hensch, M., Cesca, S., Heimann, S., Rivalta, E., Hjörleifsdottir, V., Jonsdottir, K., Vogfjörd, K.,
Dahm, T., the SIL Earthquake Monitoring Team, 2015. Earthquake focal mechanisms
associated with the dyke propagation and caldera collapse at the Bardarbunga volcano, Iceland. Geophys. Res. Abstr. 17 (EGU2015-5854-1).
Hickey, J., Gottsmann, J., 2014. Benchmarking and developing numerical finite element
models of volcanic deformation. J. Volcanol. Geotherm. Res. 280, 126–130.
Hjörleifsdóttir, V., Jónsdóttir, K., Hensch, M., Guðmundsson, G., Roberts, M., Ófeigsson, B.,
M. T., Gudmundsson, M., 2015. Numerous large and long-duration seismic events
during the Bárðarbunga volcanic eruption in 2014: what do they tell us about the caldera subsidence? Geophys. Res. Abstr. 17 (EGU2015-8143-1EGU).
Holohan, E.P., Troll, V.R., Walter, T.R., Münn, S., McDonnell, S., Shipton, Z.K., 2005. Elliptical
calderas in active tectonic settings: an experimental approach. J. Volcanol. Geotherm.
Res. 144, 119–136.
Holohan, E.P., Schöpfer, M.P.J., Walsh, J.J., 2011. Mechanical and geometric controls on the
structural evolution of pit crater and caldera subsidence. J. Geophys. Res. 116. http://
dx.doi.org/10.1029/2010JB008032.
Kennedy, B., Stix, J., Vallance, J.W., Lavallée, Y., Longpré, M.A., 2004. Controls on caldera
structure: results from analogue sandbox modeling. Geol. Soc. Am. Bull. 116, 515.
http://dx.doi.org/10.1130/B25228.1.
Kinvig, H.S., Geyer, A., Gottsmann, J., 2009. On the effect of crustal layering on ring-fault
initiation and the formation of collapse calderas. J. Volcanol. Geotherm. Res. 186,
293–304.
Kusumoto, S., Gudmundsson, A., 2009. Magma-chamber volume changes associated with
ring-fault initiation using a finite-sphere model: application to the Aria caldera, Japan.
Tectonophysics 471, 58–66. http://dx.doi.org/10.1016/j.tecto.2008.09.001.
Lavallée, Y., Stix, J., Kennedy, B., Richer, M., Longpré, M.A., 2004. Caldera subsidence in
areas of variable topographic relief: results from analogue modeling. J. Volcanol.
Geotherm. Res. 129, 219–236.
MacDonald, G.A., 1965. Hawaiian calderas. Pac. Sci. 19, 320–334.
Manconi, A., Walter, T.R., Amelung, F., 2007. Effects of mechanical layering on volcano deformation. Geophys. J. Int. 170, 952–958. http://dx.doi.org/10.1111/j.1365-246X.
2007.03449.x.
Marti, J., Geyer, A., Folch, A., 2009. A genetic classification of collapse calderas based on
field studies and analogue and theoretical modelling. In: Thordarson, T., Self, S.
(Eds.), Volcanology: The Legacy of GPL Walker. IAVCEI-Geological Society of
London, London, pp. 249–266.
Masterlark, T., 2007. Magma intrusion and deformation predictions: sensitivities to
the Mogi assumptions. J. Geophys. Res. 112, 6419. http://dx.doi.org/10.1029/
2006JB004860.
Michon, L., Massin, F., Famin, V., Ferrazzini, V., Roult, G., 2011. Basaltic calderas: collapse
dynamics, edifice deformation, and variations of magma withdrawal. J. Geophys.
Res. 116. http://dx.doi.org/10.1029/2010JB007636.
Mogi, K., 1958. Relations between eruptions of various volcanoes and the deformations of
the ground surfaces around them. Bull. Earthquake Res. Inst. 36, 99–134.
Murase, T., McBirney, A.R., 1973. Properties of some common igneous rocks and their
melts at high temperatures. Geol. Soc. Am. Bull. 84, 3563–3592.
Ofeigsson, B.G., Hreinsdottir, S., Sigmundsson, F., Fridriksdottir, H., Parks, M., Dumont, S.,
Arnadottir, T., Geirsson, H., Hooper, A., Roberts, M., Bennett, R., Sturkell, E., Jonsson,
S., Lafemina, P., Jonsson, T., Bergsson, B., Kjartansson, V., Steinthorsson, S., Einarsson,
P., Drouin, V., 2015. Deformation derived GPS geodesy associated with Bardarbunga
2014 rifting event in Iceland. Geophys. Res. Abstr. 17 (EGU2015-7691-4).
Parameswaran, V., 1987. Orientation dependence of elastic constants for ice. Def. Sci. J. 37,
367–375. http://dx.doi.org/10.14429/dsj.37.5924.
Paterson, W.S.B., 1994. The Physics of Glaciers. 3rd ed. Pergamon/Elsevier, Kidlington
(480 pp.).
Peltier, A., Famin, V., Bachèlery, P., Cayol, V., Fukushima, Y., Staudacher, T., 2008. Cyclic
magma storages and transfers at Piton de La Fournaise volcano (La Réunion hotspot)
inferred from deformation and geochemical data. Earth Planet. Sci. Lett. 270,
180–188.
Philipp, S.L., Afsar, F., Gudmundsson, A., 2013. Effects of mechanical layering on
hydrofracture emplacement and fluid transport in reservoirs. Front Earth Sci. 1.
http://dx.doi.org/10.3389/feart.2013.00004.
Phillipson, G., Sobradelo, R., Gottsmann, J., 2013. Global volcanic unrest in the 21st century: an analysis of the first decade. J. Volcanol. Geotherm. Res. 264, 183–196.
Riel, B., Milillo, P., Simons, P., Lundgren, P., Kanamori, H., Samsonov, S., 2015. The collapse
of Bardarbunga caldera, Iceland. Geophys. J. Int. 202, 446–453. http://dx.doi.org/10.
1093/gji/ggv157.
Roche, O., Druitt, T.H., Merle, O., 2000. Experimental study of caldera formation.
J. Geophys. Res. 105, 395. http://dx.doi.org/10.1029/1999JB900298.
Savin, G.N., 1961. Stress Concentration Around Holes. Pergamon, New York.
Schulson, E.M., Duval, P., 2009. Creep and Fracture of Ice. Cambridge University Press,
Cambridge.
Sigmundsson, F., Hooper, A., Parks, M., Spaans, K., Gudmundsson, G.B., Drouin, V.,
Samsonov, S., White, R.S., Hensch, M., Pedersen, R., Bennett, R.A., Greenfield, T.,
Green, R.G., Sturkell, E., Bean, C.J., Mo, M., Femina, P.C. La, Bjo, H., Pa, F., Braiden,
A.K., Eibl, E.P.S., 2015. Segmented lateral dyke growth in a rifting event at
Barðarbunga volcanic system, Iceland. Nature 517, 191–195.
Sigurdsson, H., Sparks, S.R., 1978. Lateral magma flow within rifted Icelandic crust. Nature
274, 126–130.
Stix, J., Kobayashi, T., 2008. Magma dynamics and collapse mechanisms during four
historic caldera‐forming events. J. Geophys. Res. 113. http://dx.doi.org/10.1029/
2007JB005073.
Vogfjörd, K., Hensch, M., Hjörleifsdottir, V., Jonsdottir, K., 2015. High-precision mapping of
seismicity in the last decades at Bardarbunga volcano, Iceland. Geophys. Res. Abstr.
17 (EGU2015-13430-2).
Williams, H., 1941. Calderas and their origin, University of California Publications. Bull.
Dep. Geol. Sci. 25, 239–346.
Zienkiewicz, O.C., 1979. The Finite Element Method. McGraw-Hill, New York (787 pp.).
Chapter 6: Surface displacements resulting from magma-chamber roof subsidence
6.1 Further discussion
It is important to further consider the limitations of this modelling study. The most
significant of which concerns the assumption of linear elasticity with respect to
caldera-fault deformation. Faults are discontinuities that allow elastic movement of
adjacent walls when experiencing a required stress. However, the fault movement
can be considered as an essentially anelastic process. The finite element method
modelling used here does not model an anelastic process as the bonds that join each
mesh node cannot be broken. The deformation therefore likely represents a
maximum, although it is clear that at Bardarbunga-Holuhraun there was no
substantial development of a ring-fault as evidenced by the lack of ice surface
fractures.
Our interpretation is that subsidence within the ice surface is at least partly related to
loading from the emplaced regional dyke. This idea relies on the assumption that the
dyke loading effectively alters the physical state of the ice from a brittle material to a
partially ductile material and is thus able to flow. If the ice remained brittle
throughout then instead of subsidence we would expect a stacking or inflation signal,
which is not found anywhere on the ice surface.
57
Chapter 7: Discussion, critical evaluation and future directions
Chapter 7: Discussion, critical evaluation and future
directions
7.2 Forecasting worldwide magma chamber failure conditions
Most studies of volcano deformation concentrate on identifying the source of magma
inflation or deflation, for example importance is often concentrated on estimating the
depth and size of an underlying magma chamber (e.g. Segall, 2013; Sigmundsson,
2006). Deformation studies have also focussed attention to host rock characteristics
and how deformation is affected by varying host rock rheologies and heterogeneities
(Geyer and Gottsmann, 2010; Hickey et al., 2013; Masterlark, 2007). Very few
studies attempt to constrain conditions for the chamber wall failure and then estimate
dyke propagation paths, but these are important questions that need to be addressed
when any volcano is experiencing unrest. In chapter 3 (Browning et al., in press) we
have offered a new method for constraining magma chamber failure, and it is hoped
that the model will be expanded upon at other volcanoes, and then tested during
periods of unrest.
To build on this study it would be useful to constrain magma
chamber volumes and likely failure conditions at other well studied caldera
volcanoes, particularly those that are believed to possess a double magma chamber
feeder system, i.e many of the calderas in Iceland (Camitz et al., 1995;
Gudmundsson, 1987; Sigmundsson, 2006), Miyakejima and other volcanoes in
Japan (Geshi et al., 2002), the volcanoes of the Taupo volcanic zone (Wilson et al.,
1994) to name just a few.
The major limitation of our study on magma chamber failure concerns the simplicity
of the analytical solutions. Pre-existing faults and fractures can greatly influence
magma propagation (Browning and Gudmundsson, 2015) and ground deformation
(Browning and Gudmundsson, in press). Several well-known major fault systems
exist at Santorini caldera, for example the Kameni Line and the Coloumbo line, as
well as the caldera faults from the Minoan eruption and early events. These tectonic
features are likely to affect ground deformation and seismicity during periods of
58
Chapter 7: Discussion, critical evaluation and future directions
magmatic recharge at Santorini (Konstantinou et al., 2013; Papoutsis et al., 2013).
Therefore, whilst beyond the scope of the current study, it will be important to
incorporate fault dynamics and heterogeneities when applying the method to
forecasting magma chamber failure. This will likely involve a numerical modelling
approach as the number of parameters becomes too great to model analytically.
Firstly however, detailed information on how faults and heterogeneities affect
ground surface displacements is needed, supplementing previous work on this topic
(e.g Manconi et al., 2007; Masterlark, 2007). Much of the discussion of this thesis
describes the limitations of the models presented and suggests further studies which
may improve such methods.
All assumptions of tension related host rock failure, i.e dyke injection, assume
tensile strengths of between 0.5 and 6 MPa (Amadei and Stephenson, 1997). More
recent laboratory studies, concerned primarily with the injection of fluid driven
fractures, have estimated maximum values of tensile strength around 7-11 MPa
(Benson et al., 2012). Numerical studies of geothermal systems in New Zealand
estimate the tensile strength of the Rotakawa andesite to be in the range 10 to 24
MPa (Siratovich et al., 2015). It is clear from the literature that the range of host rock
strengths is variable, and that further studies are needed to constrain the range at
individual caldera volcanoes. Here we use relatively minimal values of tensile
strength to estimate the range over which rocks will fail as a result of an excess (or
driving) pressure, based on these assumptions.
The qualitative nature of our final forecast model (Browning et al., in press) means
that the thresholds between the various phases of chamber failure likelihood are
highly subjective, although based on data from rock mechanics experiments and
borehole data so reasonably robust. Therefore any tests of the model should aim to
constrain using statistical methods, the probability of failure occurrence over time
during the unrest, and it is hoped that our method will help inform such a
probabilistic study. In addition, future studies should look to assess how long it takes
to create forecast models. The length of time in this case will be governed by the
speed at which simple Mogi solutions for depth and pressure (volume) change can be
solved during a magma chamber inflation event. In the age of broadband
seismometers, strain meters and continuously recording and transmitting GPS
59
Chapter 7: Discussion, critical evaluation and future directions
stations, this task should be completed within a reasonable timeframe. If the model
relied on InSAR data then this could present problems for forecasting chamber
rupture, because the technique commonly relies on the repositioning of suitable
satellites, such that temporal resolution or timescale of deployment may not be
sufficient to capture the beginning of a magmatic recharge event. Similar techniques
may be hampered by meteorological conditions, or ground cover. For example much
of the early inflation at Bardarbunga in 2014 was dominated by regional doming
(Ofeigsson et al., 2015); the local subsidence signal from the caldera was not
detected until much later on, partly because of the availability of GPS stations on the
glacier. It should be noted that our study on the implications of the BardarbungaHoluhraun episode (Browning and Gudmundsson, in press) assume the presence of a
caldera lake based on studies of other sub-glacial caldera volcanoes, but may not be
the case at Bardarbunga. It is hoped that these theoretical models are useful for
understanding and delimiting various aspects of caldera deformation associated with
rapid subsidence.
Deformation studies focusing on the size and depth of magma chambers, whilst
important and valid, should aim to give more consideration to providing forecasts of
magma movement. Such forecasts should not initially be used as a mechanism to
alert at risk communities until they have been sufficiently tested at the volcano in
question. The data provide a first step in forecasting volcanic eruptions, which is one
of the ultimate goals in volcanology. Then the task of predicting how large and for
how long that eruption will last is an equally important problem.
7.3 Towards a characterisation of caldera fault damage zones
Whilst the structural and topographic boundaries of many active and ancient caldera
systems are well mapped and constrained (Geyer and Martí, 2008); very few studies
have measured the frequency of micro and macro fracture damage directly adjacent
to caldera faults. This is partly because so few calderas faults around the world are
well exposed. For example, the compilation of calderas produced by Geyer and
Marti (2008) contains information on caldera size predominantly based on
60
Chapter 7: Discussion, critical evaluation and future directions
topographical boundaries, in which case the area of a caldera fault is already
ambiguous, as is highlighted by the large error and uncertainty of areas and volume
estimates. In addition, those geological studies that directly observe caldera faults, or
fracture zones (e.g. Acocella, 2006; Branney, 1995; Browning and Gudmundsson,
2015; Lipman, 1984), are hampered often by significantly weathered and altered
rocks. This is of course a consequence of the complex geothermal and hydrothermal
systems that operate near caldera faults (Kokelaar 2007). Still, almost all of the
aforementioned studies have focused predominantly on macro-fractures surrounding
caldera faults. It would be useful and worthwhile to conduct a similar detailed study,
probably in Iceland, Western USA, New Zealand or Japan of micro-fractures within
and around caldera faults.
Many permeability studies of major fault systems have been conducted (e.g Caine et
al., 1996; Evans et al., 1997; Faulkner et al., 2003). Detailed geological studies have
shown the development of fault damage zones based on the frequency of micro and
macro fractures as a function of distance from major faults (Mitchell and Faulkner,
2012). Such studies have been used to infer a mechanism of principal stress rotation
within low angle normal faults (Faulkner et al., 2006). The quantification of damage
zones at caldera faults
are difficult for the reasons given above and perhaps
impossible because the original ring-fault has long been eroded or experienced mass
wasting (Hartley and Thordarson, 2012; Wilson et al., 1994).
Several assumptions have been made regarding the mechanical state of the rocks
within and surrounding the caldera ring-fault at Hafnarfjall, Western Iceland
(Chapter 4). Future work could try to quantify the state of damage within the fault
core and surrounding host rock using the methods of (Mitchell and Faulkner, 2012,
2008). Whilst such a study would not provide precise information about the stiffness
of rocks at the time of magma emplacement, it would give clues as to the how
caldera faults propagate and influence host rock properties. An interesting study
would be to compare caldera fault damage with data from well-studied fault systems
(Gudmundsson and Brenner, 2003; Mitchell and Faulkner, 2012). At present there is
no direct quantification of caldera fault damage zones, the only studies that
hypothesize the occurrence of such mechanical situations rely on interpretation of
numerical or dynamic wave velocity models (e.g Browning and Gudmundsson,
61
Chapter 7: Discussion, critical evaluation and future directions
2015; Giannopoulos et al., 2015). Many if not most caldera faults remain active for
many hundreds or thousands of years following their formation, as shown by
repeated seismic unrest associated with the Bardarbunga caldera faults (Fichtner and
Tkalčić, 2010; Gudmundsson et al., 2014; Konstantinou et al., 2003; Sigmundsson et
al., 2014), and Santorini caldera faults (Druitt and Francaviglia, 1992; Konstantinou
et al., 2013) making fault zone evolution is likely an important consideration.
Several fault core and damage zones may develop by repeated slip and re-activation,
as is known from studies of major tectonic related fault systems worldwide
(Faulkner et al., 2010), Figure 7.1. Estimates of damage accumulation in caldera
fault zones may be useful for understanding permeability development during
periods of volcanic unrest. Fault zone permeability models can be constrained by ice
melting at subglacial calderas (Reynolds and Gudmundsson, 2014), or degassing
(Padrón et al., 2008; Shinohara, 2008) due to increased fluid flow and geothermal
activity (Axelsson et al., 2006).
Figure 7.1: Models of relationship between damage and core zones, a) shows a
single core and associated damage zone, b) shows multiple cores and associated
damage zones modified after (Faulkner et al., 2010).
An appropriate question at this point may then be, ‘to what effect does ring-fault
damage control ground deformation during fault-slip or inflation or deflation
episodes?’. In Browning and Gudmundsson (in press) we have shown that
62
Chapter 7: Discussion, critical evaluation and future directions
significant surface subsidence occurs many tens of kilometres away from a gravity
driven roof collapse. However, a zone of weak rock, i.e. a ring-fault damage zone,
reduces the total amount of ground surface subsidence (Masterlark, 2007). It has also
been noted that the development of fracture networks can act to mask a related
subsidence inversion signal (e.g Holohan et al., 2015). For example, original
inversions of the summit collapse at Piton de la Fournaise in 2007 using standard
techniques (Mogi, 1958) presented three potentially distinct sources of deformation.
However, discrete element method models show that the deformation can be easily
explained by the deflation and collapse of one ‘sill like’ source (Holohan et al.,
2015). The difference between the ‘standard’ models and the ‘newer’ method is the
inclusion of fracture networks generated in conjunction with collapse, therefore
understanding fracture distribution, through field studies, statistical techniques and
seismic tomography are vitally important for properly constraining ground
deformation signals.
7.4 Assessing the influence of thermo-mechanical damage accumulation
It has been shown through the various chapters presented within this thesis that
fractures and other heterogeneities within rock layers can have a direct impact on
magma propagation and magma chamber failure. It is possible that part of that
damage can be created by thermal stresses generated during magmatic and host rock
cooling. Much more work is needed to understand crustal thermal stress and the
resultant effect on damage accumulation within and around shallow magma systems.
We have found that the dominant mechanism of thermal stress damage accumulation
in the crust and in eruptive products is likely to result from contraction due to
cooling rather than expansion through heating, e.g. extrusive lava domes and lava
flows where the controlling thermal regime following emplacement is cooling. In
this case cooling generates cracks which, through increased surface area, further
increase cooling potential and may act as pathways for gas escape. The two
mechanisms of cooling and gas escape combine to dramatically increase viscosity
which is a primary control on the velocity of a lava flow and also influences the lava
final flow length. In the shallow crust, both heating expansion and cooling
63
Chapter 7: Discussion, critical evaluation and future directions
contraction stresses operate, probably by similar amounts. It is though contractive,
and therefore overall tensile stress which is likely to generate the predominant
damage, simply because crustal rocks are weaker in tension.
7.5 What is a ‘realistic’ magma chamber host rock rheology?
This brings us to an important question, namely, ‘what are the conditions of rocks
surrounding magma chambers?’. Many of the models discussed within this thesis
have simulated either a linear elastic or a poro-elastic crust with various
combinations of internally pressurised cavities and differential boundary or loading
conditions. Many studies have shown that linear elasticity does not fully represent
crustal rheology at great depths and surrounding ‘hot’ magma (e.g Gregg et al.,
2012). In high temperature tests on basaltic rocks from Iceland we noted the
development of melt induced fractures on the outer surface of samples (Browning et
al., 2015). The textures observed indicate partial melt was able to generate
sufficiently high internal overpressure to overcome the tensile strength of the brittle
sample surface. Although the outcome was not expected and largely unwanted, as it
hampered crack annealing, such tests may be developed to understand more about
the high temperature host rock halo that surrounds magma chambers. Although some
authors suggest otherwise (e.g Hooper et al., 2011), it is clear from field and
laboratory evidence that dykes ‘magma filled fractures’ propagate using the same
mechanism as almost every other fluid driven fracture, namely as mode I fractures.
At several localities in Iceland, but perhaps the best exposed at Slafrudalur in the
East, felsic dykes can be observed cutting through the sharp granophyre to basalt
contact of the extinct pluton and host rock (Gudmundsson, 2012). These field
observations together with experimental observations suggest that magma chamber
envelopes can remain predominantly brittle, particularly in response to fluid
overpressure driving forces.
7.6 Towards a method of quantifying crack annealing in volcanic rocks
64
Chapter 7: Discussion, critical evaluation and future directions
Whilst the timescales of viscous relaxation are relatively well constrained (Dingwell
and Webb, 1989; Giordano et al., 2008; Hess et al., 2008), almost no work has been
done to apply that process to measure crack annealing in volcanic rocks. The most
relevant work has utilised analogue materials such as glass beads in order to
constrain viscous annealing ‘or sintering’ in volcanic ash and rheomorphic flows
(Vasseur et al., 2013; Wadsworth et al., 2014). Such studies utilise perfect glasses to
measure the time taken for the spheres to sinter, which is a function of the glass
viscosity, controlled partly by temperature, and the effects of surface tension. Whilst
the setup is suitable for deciphering processes in silicic, glass rich magma such as is
abundant in explosive ash fragments, the method is a poor analogue for intrusions
and lavas that contain abundance of crystals and groundmass, all of which seeks to
impede melt relaxation and fracture annealing. Our original method for
characterising crack annealing, based on the principle of P-wave velocity changes in
rocks due to the presence (or absence) of cracks proved ineffective. This was a
consequence of the need for a suitably long wave-guide to ensure that all moveable
parts (springs etc.), and signal transducers were located outside of the furnace hotzone. The consequence was substantially reduced P-wave transmission resulting
from wave attenuation. It is suggested that much more future work can be done to
understand fracture annealing in volcanic rocks, and ideally build on the
conventional models of viscous relaxation which really only apply to glasses and
silicates.
7.7 The ‘gravity problem’
In COMSOL Multiphysics there is an option to apply a gravity load to the model
domain. It has been speculated that omitting the loading force due to gravity
provides inaccurate results in terms of pressures associated with magma chamber
rupture and the location of rupture (see Grosfils 2007 for details). For example most
published numerical studies of pressurized ellipsoidal magma reservoirs give modest
estimates of rupture pressures (0.5-9 MPa) (e.g Jellinek & DePaolo 2003; Trasatti et
al. 2008; Gudmundsson 2012). However, those studies which add additional gravity
effects as a loading condition in the finite element model output failure related
65
Chapter 7: Discussion, critical evaluation and future directions
pressures in excess of 100 MPa (Currenti and Williams, 2014; Grosfils, 2007;
Grosfils et al., 2015). This issue can be clarified by considering initially the geostatic
state of stress, or vertical stress (σv) which can be given as:
v r gz
(7.1)
with a simple assumption that all the crustal layers have the same density, ρr. The
vertical co-ordinate z is positive downwards, i.e toward increasing crustal depth, and
g is the acceleration due to gravity. Vertical stress can also be termed ‘lithostatic
stress’ (Gudmundsson, 2012). As has been discussed in Chapter 3, throughout the
majority of a magma chambers lifespan it will reside in lithostatic equilibrium, that
is, have a magma-pressure close to the pressure of the stresses in the host rock. It is
only during periods of unrest when the host rock deforms as an immediate response
to pressure changes (e.g Mogi 1958; Dzurisin, 2006; Segall, 2010). A magma
chamber wall will rupture when the magma pressure inside the chamber exceeds the
tensile strength of the host rock (To), and then a magma filled fracture will propagate
perpendicular to the minimum principal compressive stress (σ3). The normal state of
stress in a rift zone, for example, is such that σ3 is horizontal and therefore the
propagation direction of dykes is vertical or parallel with the vertical stress σ 1. From
eq. 1 we see that the vertical stress, σ1, includes the effects of the acceleration due
gravity (g). If we consider Newton’s second law of motion, where force (F) is
defined as mass (m) times acceleration (a):
F=ma
(7.2)
Close to the Earth’s surface the acceleration a in the Earth’s gravity field can be
interchanged with g (i.e 9.81 m s-2), so that eq. 2 can be re-written as:
F=mg
(7.3)
7.9 The unknowns of magma volume expansion and compressibility
Volume expansion and compressibility of magmas and host rock is reasonably
poorly constrained with limited data available. Most work done to measure
66
Chapter 7: Discussion, critical evaluation and future directions
compressibility in magmas and volcanic rocks was completed over three decades ago
(Murase and Mcbirney, 1973), and provides generally little constraints on the error
and uncertainty of values. It is well known that magma chamber rupture occurs as a
result of volatile exsolution or magmatic recharge in shallow chamber system. What
is much less well known or investigated is the contribution of melt volume
expansion. Such a processes could occur as a result of hot mafic magma entering and
melting a partially solidified felsic chamber, with the pressure increase in that
chamber represented by 1) a new volume of mafic magma Vm, 2) gas exsolution and
bubble growth in both magmas, which is directly related to compressibility of
magma βm and 3) expansion related to the phase change from a solid to a steady state
liquid. Data for parameters 2 and 3 are much less abundant, and most data available
for parameter 1 is acquired from geodetic studies that may underestimate the
contribution of 2 and 3. This creates a problem interpreting the underlying causes of
ground deformation during periods of unrest, as such, further work is needed.
7.10 The ‘state-of-the-art’
The work that comprises this thesis has made several key contributions to
understanding many aspects of volcano-tectonics on both a regional and local scale.
The chapters are arranged such that the reader has been guided from a depth of
approximately 10 to 20 km in the Earth’s crust, where we find deep magma
reservoirs. Parental magma from these reservoirs periodically recharges shallow
magma chambers. A model has been proposed to estimate the timing of magma
chamber roof rupture associated with resultant volume and pressure increases. When
a shallow chamber does rupture a dyke is initiated, estimating the pathway of that
magma is crucial for predicting the location of an eruption. We have found that
caldera faults can significantly influence the propagation potential of favourably
oriented sheets through a mechanism of principal stress rotation and elastic
mismatch. Magma chambers, dykes and sheets are emplaced into cooler host rocks
which then heat as a result. The thermal stresses resulting from heating and cooling
throughout the emplacement cycle generate seismicity. Experimental results indicate
that the predominant micro-crack development occurs during cooling in an overall
tensile stress field. It has not been possible to relate acoustic emissions generated
67
Chapter 7: Discussion, critical evaluation and future directions
within small laboratory scale samples held at atmospheric pressure to seismicity
from magma chambers and intrusions at depth in the crust. Future work should
consider these links and investigate the potential for combing these laboratory data
with seismic data. Thick regional dykes can have significant regional stress effects,
for example by changing the flow dynamics of nearby glaciers as has been
demonstrated during the Bardarbunga-Holuhraun 2014-15 episode. Collapse of a
magma chamber roof should also encourage significant far-field surface subsidence.
Figure 7.2: Schematic diagram illustrating the main contributions of this thesis.
Parental magma recharge from a deep reservoir to a shallow chamber induces
pressure changes, inflation and eventual rupture. When the chamber roof ruptures
and initiates a sheet or dyke, the pathway is influenced by pre-existing
discontinuities such as fault zones, encouraging, in the case of calderas deflection
into ring-dykes. Shallow magma chambers and intrusions heat and cool the host rock
leading to thermal stressing and seismicity. Emplacement of substantial regional
dykes can have far reaching stress effects, for example influencing glacial flow
dynamics. Magma chamber roof collapse induces regional scale subsidence. Not to
scale.
68
Bibliography
Bibliography
Acocella, V., 2006. Regional and local tectonics at Erta Ale caldera, Afar (Ethiopia).
Journal of Structural Geology. 28, 1808-1820
Acocella, V., 2007. Understanding caldera structure and development: An overview
of analogue models compared to natural calderas. Earth-Science Reviews. 85,
125–160.
Amadei, B. & Stephenson, O., 1997. Rock stress and its measurement. Chapman and
Hall, London. 516p
Anderson, E.M., 1937. Cone-sheets and Ring-dykes : the dynamical explanation.
Bulletin of Volcanology. 1, 35-40
Annen, C., 2011. Implications of incremental emplacement of magma bodies for
magma differentiation, thermal aureole dimensions and plutonism-volcanism
relationships. Tectonophysics, 500, 3–10
Applegarth, L.J., Tuffen, H., James, M.J., Pinkerton, H., Cashman, K.V. 2013 Direct
observations of degassing induced crystallisation in basalts, Geology, 41, 243-246
Axelsson, G., Thorhallsson, S., Bjornsson, G., 2006. Stimulation of geothermal wells
in basaltic rock in Iceland. Enhanced geothermal innovation network for
Europe Workshop, Zurich Switzerland, 3,
Barnett, Z. A. & Gudmundsson, A., 2014. Numerical modelling of dykes deflected
into sills to form a magma chamber. Journal of Volcanolology Geothermal
Research. 281, 1–11.
Barton, M. & Huijsmans, J.P.P., 1986. Post-caldera dacites from the Santorini
volcanic complex, Aegean Sea, Greece: an example of the eruption of lavas of
69
Bibliography
near-constant composition over a 2,200 year period. Contributions to
Mineralogy and Petrology. 94, 472–495.
Bathke, H., Nikkhoo, M., Holohan, E.P., Walter, T.R., 2015. Insights into the 3D
architecture of an active caldera ring-fault at Tendürek volcano through
modeling of geodetic data. Earth and Planetary Science Letters, 422, 157–168.
Benson, P.M., Heap, M.J., Lavallée, Y., Flaws, A., Hess, K.-U., Selvadurai, a. P.S.,
Dingwell, D.B., Schillinger, B., 2012. Laboratory simulations of tensile fracture
development in a volcanic conduit via cyclic magma pressurisation. Earth and
Planetary Science Letters. 349-350, 231–239.
Bond, A. & Sparks, R.S.J., 1976. The Minoan eruption of Santorini, Greece. Journal
of the Geological Society London. 132, 1–16.
Branney, M.J., 1995. Downsag and extension at calderas: new perspectives on
collapse geometries from ice-melt, mining, and volcanic subsidence. Bulletin of
Volcanology 57, 303–318.
Browning, J., Meredith, P. G., Gudmundsson, A., Lavallée, Y., Drymoni, K. 2015.
Are magma chamber boundaries brittle or ductile? Rheological insights from
thermal stressing experiments. European General Assembly, Vienna.
Browning, J., Drymoni, K., Gudmundsson, A. in press. Forecasting magma chamber
rupture at Santorini volcano, Greece. Scientific Reports.
Browning, J. & Gudmundsson, A., 2015. Caldera faults capture and deflect inclined
sheets: an alternative mechanism of ring dike formation. Bulletin of
Volcanology, 77
Browning, J. & Gudmundsson, A. in press. Surface displacements resulting from
magma chamber failure and roof collapse, with application to the 2014
Bardarbunga-Holuhraun episode. Journal of volcanology and geothermal
research.
70
Bibliography
Burchardt, S. & Gudmundsson, A., 2009. The infrastructure of the Geitafell
Volcano, Southeast Iceland. Special Publication IAVCEI, 349–369.
Burchardt, S., Tanner, D.C., Krumbholz, M., 2010. Mode of emplacement of the
Slaufrudalur Pluton, Southeast Iceland inferred from three-dimensional GPS
mapping and model building. Tectonophysics 480, 232–240.
Burchardt, S., Tanner, D.C., Troll, V.R., Krumbholz, M., Gustafsson, L.E., 2011.
Three-dimensional geometry of concentric intrusive sheet swarms in the
Geitafell and the Dyrfjall volcanoes, eastern Iceland. Geochemistry, Geophys.
Geosystems 12, 1–21.
Burchardt, S., Troll, V.R., Mathieu, L., Emeleus, H.C., Donaldson, C.H., 2013.
Ardnamurchan 3D cone-sheet architecture explained by a single elongate
magma chamber. Scientific Reports, 3, 2891.
Caine, J.S., Evans, J.P., Forster, C.B., 1996. Fault zone architecture and permeability
structure. Geology 24, 1025–1028.
Camitz, J., Sigmundsson, F., Foulger, G., Jahn, C.H., Völksen, C., Einarsson, P.,
1995. Plate boundary deformation and continuing deflation of the Askja
volcano, North Iceland, determined with GPS, 1987-1993. Bulletin of
Volcanology, 57, 136–145.
Carlos, J., Day, S.J., Perez, F.J., 1999. Giant Quaternary landslides in the evolution
of La Palma and El Hierro, Canary Islands. Journal of Volcanology and
Geothermal Research, 94, 169-190
Carracedo, J.C., 1999. Growth, structure, instability and collapse of Canarian
volcanoes and comparisons with Hawaiian volcanoes. Journal of Volcanology
and Geothermal Research. 94, 1–19.
Cashman, K. V. & Giordano, G., 2014. Calderas and magma reservoirs. Journal of
Volcanology and Geothermal Research. 288, 28–45.
71
Bibliography
Coppo, N., Schnegg, P.-A., Heise, W., Falco, P., Costa, R., 2008. Multiple caldera
collapses inferred from the shallow electrical resistivity signature of the Las
Cañadas caldera, Tenerife, Canary Islands. Journal of Volcanology and
Geothermal Research. 170, 153–166.
Cox, S.J.D. & Meredith, P.G., 1993. Microcrack formation and material softening in
rock measured by monitoring acoustic emissions. International Journal of Rock
Mechanics and Mining Sciences & Geomechanics Abstracts, 30, 11–24.
Currenti, G & Williams, C. A., 2014. Numerical modeling of deformation and stress
fields around a magma chamber: Constraints on failure conditions and
rheology. Physics of the Earth and Planetary Interiors, 226, 14–27.
De Natale, G., Pingue, F., 1993. Ground deformations in collapsed caldera
structures. Journal of Volcanology and Geothermal Research. 57, 19–38.
De Zeeuw-van Dalfsen, E., Pedersen, R., Hooper, A., Sigmundsson, F., 2012.
Subsidence of Askja caldera 2000-2009: Modelling of deformation processes at
an extensional plate boundary, constrained by time series InSAR analysis.
Journal of Volcanology and Geothermal Research. 213-214, 72–82.
Dingwell, D.B. & Webb, S.L., 1989. Structural relaxation in silicate melts and nonNewtonian melt rheology in geologic processes. Physics and Chemistry of
Minerals. 16, 508–516.
Dominey-Howes, D. & Minos-Minopoulos, D., 2004. Perceptions of hazard and risk
on Santorini. Journal of Volcanology and Geothermal Research. 137, 285–310.
Druitt, T.H., 2014. New insights into the initiation and venting of the Bronze-Age
eruption of Santorini (Greece), from component analysis. Bulletin of
Volcanology. 76, 1–21.
Druitt, T.H., Costa, F., Deloule, E., Dungan, M., Scaillet, B., 2012. Decadal to
monthly timescales of magma transfer and reservoir growth at a caldera
volcano. Nature, 482, 77–80.
72
Bibliography
Druitt, T.H., Edwards, L., Mellors, R.., Pyle, D.., Sparks, R.S.J., Lanphere, M.,
Davies, M., Barriero, B., 1999. Santorini Volcano, Geological Society, London,
Memoirs, 19
Druitt, T.H. & Francaviglia, V., 1992. Caldera formation on Santorini and the
physiography of the islands in the late Bronze Age. Bulletin of Volcanology. 54,
484–493.
Druitt, T.H. & Sparks, R.S.J., 1982. A proximal ignimbrite breccia facies on
santorini , Greece. Journal of Volcanology and Geothermal Research. 13, 147–
171.
Faulkner, D.R., Jackson, C. A. L., Lunn, R.J., Schlische, R.W., Shipton, Z.K.,
Wibberley, C. A J., Withjack, M.O., 2010. A review of recent developments
concerning the structure, mechanics and fluid flow properties of fault zones.
Journal of Structural Geology. 32, 1557–1575.
Faulkner, D.R., Lewis, A. C., Rutter, E.H., 2003. On the internal structure and
mechanics of large strike-slip fault zones: Field observations of the Carboneras
fault in southeastern Spain. Tectonophysics 367, 235–251.
Faulkner, D.R., Mitchell, T.M., Healy, D., Heap, M.J., 2006. Slip on “weak” faults
by the rotation of regional stress in the fracture damage zone. Nature, 444, 922–
925.
Fichtner, A. & Tkalčić, H., 2010. Insights into the kinematics of a volcanic caldera
drop: Probabilistic finite-source inversion of the 1996 Bárdarbunga, Iceland,
earthquake. Earth and Planetary Science Letters. 297, 607–615.
Filson, J., Simkin, T., Leu, L., 1973. Seismicity of a caldera collapse: Galapagos
Islands 1968. Journal of Geophysical Research 78, 8591.
Fossen, H. 2010. Structural Geology. Cambridge University Press, p480
Friedrich, W.., 2009. Santorini. Aarhus University Press, p380
73
Bibliography
Fytikas, M., Innocenti, F., Manetti, P., Peccerillo, A., Mazzuoli, R., Villari, L., 1984.
Tertiary to Quaternary evolution of volcanism in the Aegean region. Geological
Society London, Special Publications. 17, 687–699.
Garcia-Piera, J.O., Ledesma, A., Hu, M., 2000. Causes and mobility of large
volcanic landslides : application to Tenerife , Canary Islands 103, 121-134
Gautneb, H. & Gudmundsson, A., 1992. Effect of local and regional stress fields on
sheet emplacement in West Iceland. Journal of Volcanology Geothermal
Research, 51, 339–356.
Gautneb, H., Gudmundsson, A., Oskarsson, N., 1989. Structure , petrochemistry and
evolution of a sheet swarm in an Icelandic central volcano. Geologic magazine
126, 659–673.
Geyer, A. & Martí, J., 2008. The new worldwide collapse caldera database (CCDB):
A tool for studying and understanding caldera processes. Journal of
Volcanology Geothermal Research. 175, 334–354.
Geyer, A. & Gottsmann, J., 2010. The influence of mechanical stiffness on caldera
deformation and implications for the 1971-1984 Rabaul uplift (Papua New
Guinea). Tectonophysics, 483, 399–412.
Geyer, A. & Marti, J., 2014. A short review of our current understanding of the
development of ring faults during collapse caldera formation. Frontiers in Earth
Sciences. 2, 1–13.
Geshi, N., Kusumoto, S., Gudmundsson, A., 2010. Geometric difference between
non-feeder and feeder dikes. Geology 38, 195–198.
Geshi, N., Shimano, T., Chiba, T., Nakada, S., 2002. Caldera collapse during the
2000 eruption of Miyakejima Volcano, Japan. Bulletin of Volcanology. 64, 55–
68.
Giannopoulos, D., Sokos, E., Konstantinou, K.I., Tselentis, G. A., 2015. Shear wave
splitting and VP/VS variations before and after the Efpalio earthquake
74
Bibliography
sequence, western Gulf of Corinth, Greece. Geophysical Journal International
200, 1436–1448.
Giordano, D., Russell, J.K., Dingwell, D.B., 2008. Viscosity of magmatic liquids: A
model. Earth and Planetary Science Letters, 271, 123–134.
Gray, J.P. & Monaghan, J.J., 2004. Numerical modelling of stress fields and fracture
around magma chambers. Journal of Volcanology and Geothermal Research,
135, 259–283.
Gregg, P.M., De Silva, S.L., Grosfils, E.B., Parmigiani, J.P., 2012. Catastrophic
caldera-forming eruptions: Thermomechanics and implications for eruption
triggering and maximum caldera dimensions on Earth. Journal of Volcanology
and Geothermal Research. 241-242, 1–12.
Grosfils, E.B., 2007. Magma reservoir failure on the terrestrial planets: Assessing the
importance of gravitational loading in simple elastic models. Journal of
Volcanology and Geothermal Research, 166, pp.47–75.
Grosfils, E.B., McGovern, P.J., Gregg, P.M., Galgana, G. A., Hurwitz, D.M., Long,
S.M., Chestler, S.R., 2015. Elastic models of magma reservoir mechanics: a key
tool for investigating planetary volcanism. Geological Society London, Special
Publications, 401, 239-267
Gudmundsson, A., 1983. Form and Dimensions of dykes in Eastern Iceland,
Tectonophysics, 95, 295–307.
Gudmundsson, A., 1986a. Possible effect of aspect ratios of magma chambers on
eruption frequency. Geology, 14, 991.
Gudmundsson, A., 1986b. Mechanical Aspects of Postglacial Volcanism and
Tectonics of the Reykjanes Peninsula, Southwest Iceland. Journal of
Geophysical Research, 91, 12711–12721.
75
Bibliography
Gudmundsson, A., 1987a. Lateral magma flow, caldera collapse, and a mechanism
of large eruptions in Iceland. Journal of Volcanology and Geothermal
Research. 34, 65–78.
Gudmundsson, A., 1987b. Formation and mechanics of magma reservoirs in Iceland.
Geophysical Journal International, 91, 27-41
Gudmundsson, A., 1990. Emplacement of dikes, sills and crustal magma chambers
at divergent plate boundaries. Tectonophysics, 176, 257–275.
Gudmundsson, A., 1998. Formation and development of normal-fault calderas and
the initiation of large explosive eruptions. Bulletin of Volcanology, 60, 160–
170.
Gudmundsson, A. & Brenner, S.L., 2003. Loading of a seismic zone to failure
deforms nearby volcanoes: A new earthquake precursor. Terra Nova. 15, 187–
193.
Gudmundsson, A. & Philipp, S.L., 2006. How local stress fields prevent volcanic
eruptions. Journal of Volcanology and Geothermal Research. 158, 257–268.
Gudmundsson, A. & Nilsen, K., 2006. Ring-faults in composite volcanoes:
structures, models and stress fields associated with their formation. Geological
Society of London, Special Publication, 269, 83–108.
Gudmundsson, A., 2011. Deflection of dykes into sills at discontinuities and magmachamber formation. Tectonophysics 500, 50–64.
Gudmundsson, A., 2011b. Rock Fractures in Geological Processes. Cambridge
University Press, Cambridge. 592p
Gudmundsson, A., 2012. Magma chambers: Formation, local stresses, excess
pressures, and compartments. Journal of Volcanology and Geothermal
Research. 237-238, 19–41.
76
Bibliography
Gudmundsson, A., Lecoeur, N., Mohajeri, N., 2014. Dike emplacement at
Bardarbunga , Iceland, induces unusual stress changes, caldera deformation,
and earthquakes. Bulletin of Volcanology, 76 1–7.
Gudmundsson, A., 2015. Collapse-driven large eruptions. Journal of Volcanology
and Geothermal Research. 1–26.
Haddadi, B., Sigmarsson, O., Devidal, J. L. 2015. Determining intensive parameters
through clinopyroxene-liquid equilibrium in Grímsvötn 2011 and Bárðarbunga
2014 basalts. Geophysical Research Abstracts, 17, EGU2015-5791-2
Hartley, M.E. & Thordarson, T., 2012. Formation of Oskjuvatn caldera at Askja,
North Iceland: Mechanism of caldera collapse and implications for the lateral
flow hypothesis. Journal of Volcanology and Geothermal Research. 227-228,
85–101.
Hartley, M.E. & Thordarson, T., 2013. The 1874-1876 volcano-tectonic episode at
Askja, North Iceland: Lateral flow revisited. Geochemistry, Geophys.
Geosystems 14, 2286–2309.
Heiken, G., Mccoy, F., 1984. Caldera development during the minoan eruption,
Thira, Cyclades, Greece. Journal of Geophysical Research. 89, 8441–8462.
Hernández-Pacheco, A., 1996. Geología y estructura del Arco de Taganana
(Tenerife, Canarias), Rev Soc Geol Espaæa, 9, 169–181.
Hess, K.U., Cordonnier, B., Lavallée, Y., Dingwell, D.B., 2008. Viscous heating in
rhyolite: An in situ experimental determination. Earth and Planetary Science
Letters, 275, 121–126.
Hickey, J., Gottsmann, J., Del Potro, R., 2013. The large-scale surface uplift in the
Altiplano-Puna region of Bolivia: A parametric study of source characteristics
and crustal rheology using finite element analysis. Geochemistry, Geophys.
Geosystems 14. 540-555
77
Bibliography
Hickey, J. & Gottsmann, J., 2014. Benchmarking and developing numerical Finite
Element models of volcanic deformation. Journal of Volcanology and
Geothermal Research, 280, 126–130.
Hjartardóttir, Á.R., Einarsson, P., Sigurdsson, H., 2009. The fissure swarm of the
Askja volcanic system along the divergent plate boundary of N Iceland. Bulletin
of Volcanology. 71, 961–975.
Holohan, E., Schöpfer, M., Walsh, J., 2015. Stress evolution during caldera
collapse : a Distinct Element Method perspective. Earth Planetary Science
Letters, 15, 2208.
Hobbs, T, 2011. Stress induced seismic anisotropy around magma chambers. PhD
thesis, University of Bristol, UK
Holohan, E., Schöpfer, M. & Walsh, J., 2015. Stress evolution during caldera
collapse : a Distinct Element Method perspective. Earth and Planetary Science
Letters, 15, 2208.
Hooper, A., Ófeigsson, B., Sigmundsson, F., Lund, B., Einarsson, P., Geirsson, H.,
Sturkell, E., 2011. Increased capture of magma in the crust promoted by ice-cap
retreat in Iceland. Nature Geoscience. 4, 783–786.
Hu, M., Ledesma, A., Marti, J., 1999. Conditions favouring catastrophic landslides
on Tenerife (Canary Islands), Terra Nova, 11, 106-111
Hutnak, M., Hurwitz, S., Ingebritsen, S.E., Hsieh, P. a., 2009. Numerical models of
caldera deformation: Effects of multiphase and multicomponent hydrothermal
fluid flow. Journal of Geophysical Research, 114, 1–11.
Jellinek, A. M. & DePaolo, D.J., 2003. A model for the origin of large silicic magma
chambers: precursors of caldera-forming eruptions. Bulletin of Volcanology. 65,
363–381.
Jónsson, S., 2009. Stress interaction between magma accumulation and trapdoor
faulting on Sierra Negra volcano, Galápagos. Tectonophysics, 471, 36–44.
78
Bibliography
Key, J., White, R.S., Soosalu, H., Jakobsdóttir, S.S., 2011. Multiple melt injection
along a spreading segment at Askja, Iceland. Geophysical Research Letters 38,
1–5.
Kokelaar, P., 2007. Friction melting, catastrophic dilation and breccia formation
along caldera superfaults. Journal of Geological Society of London. 164, 751–
754.
Konstantinou, K.I., Kao, H., Lin, C.H., Liang, W.T., 2003. Analysis of broad-band
regional waveforms of the 1996 September 29 earthquake at Bardarbunga
volcano, central Iceland: Investigation of the magma injection hypothesis.
Geophysical Journal International. 154, 134–145.
Konstantinou, K.I., Evangelidis, C.P., Liang, W.T., Melis, N.S., Kalogeras, I., 2013.
Seismicity, Vp/Vs and shear wave anisotropy variations during the 2011 unrest
at Santorini caldera, southern Aegean. Journal of Volcanology and Geothermal
Research. 267, 57–67.
Lipman, P.W., 1984. The Roots of Ash Flow Calderas in Western North America ’
Windows Into the Tops of Granitic Batholiths Journal of Geophysical
Research, 89, 8801–8841.
Lipman, P.W., 1997. Subsidence of ash-flow calderas: relation to caldera size and
magma-chamber geometry. Bulletin of Volcanology. 59, 198–218.
Lockner, D. A., Byerlee, J. D., Kuksenko, V., Ponomarev, A., & Sidorin, A. 1992.
Observations of quasistatic fault growth from acoustic emissions. International
Geophysics, 51, 3-31.
Manconi, A., Walter, T.R., Amelung, F., 2007. Effects of mechanical layering on
volcano deformation. Geophysical Journal International, 170, 952–958.
Marinoni, L.B. & Gudmundsson, A., 2000. Dykes, faults and palaeostresses in the
Teno and Anaga massifs of Tenerife (Canary Islands). Journal of Volcanology
and Geothermal Research. 103, 83–103.
79
Bibliography
Marsh, B. D. 1989. Magma chambers. Annual Review of Earth and Planetary
Sciences, 17, 439-474.
Marti, J. & Gudmundsson, A., 2000. Ä adas caldera ( Tenerife , Canary Islands ):
The Las Can an overlapping collapse caldera generated by magma-chamber
migration, Journal of Volcanology and Geothermal Research, 103, 161–173.
Marti, J., Mitjavilaf, J., Aranaj, V., 1994. Stratigraphy , structure and geochronology
of the Las Canadas caldera (Tenerife , Canary Islands), Geological Magazine,
131, 715–727.
Masterlark, T., 2007. Magma intrusion and deformation predictions: Sensitivities to
the Mogi assumptions. Journal of Geophysical Research, 112, 1–17.
McGarvie, D.W., 1984. Torfajokull: a volcano dominated by magma mixing.
Geology. 12, 685-688
Menand, T., 2011. Physical controls and depth of emplacement of igneous bodies: A
review. Tectonophysics, 500, 11–19.
Michon, L., Villeneuve, N., Catry, T., Merle, O., 2009. How summit calderas
collapse on basaltic volcanoes: New insights from the April 2007 caldera
collapse of Piton de la Fournaise volcano. Journal of Volcanology and
Geothermal Research. 184, 138–151.
Miller, A.D., Julian, B.R., Foulger, G.R., 1998. Three-dimensional seismic structure
and moment tensors of non-double-couple earthquakes at the HengillGrensdalur volcanic complex, Iceland. Geophysical Journal International, 133,
309–325.
Mitchell, T.M., Faulkner, D.R., 2008. Experimental measurements of permeability
evolution during triaxial compression of initially intact crystalline rocks and
implications for fluid flow in fault zones. Journal of Geophysical Research,
113, 1–16.
80
Bibliography
Mitchell, T.M. & Faulkner, D.R., 2012. Towards quantifying the matrix permeability
of fault damage zones in low porosity rocks. Earth and Planetary Science
Letters, 339-340, 24–31.
Mogi, K., 1958. Relations between the eruptions of various volcanoes and the
deformations of the ground surfaces around them. Bulletin of the Earthquake
Research Institute, 36, 99-134
Mori, J., McKee, C., 1987. Outward-dipping ring-fault structure at rabaul caldera as
shown by earthquake locations. Science 235, 193–195.
Mori, J., White, R.A., Harlow, D.H., Okubo, P., Power, J.A., Hoblitt, R.P., Bautista,
B.C., 1996. Volcanic earthquakes following the 1991 climactic eruption of
Mount Pinatubo: Strong seismicity during a waning eruption. Fire Mud
eruptions lahars Mt. Pinatubo, Philipp. USGS publications, 339–350.
Murase, T. & McBirney, A.R., 1973. Properties of some common igneous rocks and
their melts at high temperatures. Geological Society of America Bulletin, 84,
3563-3592
Nelson, P.P. & Glasar, S.D., 1992. Acoustic Emissions Produced by Discrete
Fracture in Rock. International Journal of Rock Mechanics and Mining
Sciences, 29, 237–251.
Nomikou, P., Parks, M.M., Papanikolaou, D., Pyle, D.M., Mather, T. a., Carey, S.,
Watts, a. B., Paulatto, M., Kalnins, M.L., Livanos, I., Bejelou, K., Simou, E.,
Perros, I., 2014. The emergence and growth of a submarine volcano: The
Kameni islands, Santorini (Greece). GeoResJ, 1-2, 8–18.
Ofeigsson, B.G., Hreinsdottir, S., Sigmundsson, F., Fridriksdottir, H., Parks, M.,
Dumont, S., Arnadottir, T., Geirsson, H., Hooper, A., Roberts, M., Bennett, R.,
Sturkell, E., Jonsson, S., Lafemina, P., Jonsson, T., Bergsson, B., Kjartansson,
V., Steinthorsson, S., Einarsson, P., and Drouin, V., 2015. Deformation derived
GPS geodesy associated with Bardarbunga 2014 rifting event in Iceland.
Geophysical Research Abstracts, 17, EGU2015-7691-4.
81
Bibliography
Padrón, E., Hernández, P.A., Toulkeridis, T., Pérez, N.M., Marrero, R., Melián, G.,
Virgili, G., Notsu, K., 2008. Diffuse CO2 emission rate from Pululahua and the
lake-filled Cuicocha calderas, Ecuador. Journal of Volcanology and
Geothermal Research. 176, 163–169.
Papazachos, B.C. & Panagiotopoulos, D.G., 1993. Normal faults associated with
volcanic activity arc. Tectonophysics 220, 301–308.
Papoutsis, I., Papanikolaou, X., Floyd, M., Ji, K.H., Kontoes, C., Paradissis, D.,
Zacharis, V., 2013. Mapping inflation at Santorini volcano, Greece, using GPS
and InSAR. Geophysical Research Letters, 40, 267–272.
Parks, M.M., Biggs, J., England, P., Mather, T. a., Nomikou, P., Palamartchouk, K.,
Papanikolaou, X., Paradissis, D., Parsons, B., Pyle, D.M., Raptakis, C.,
Zacharis, V., 2012. Evolution of Santorini Volcano dominated by episodic and
rapid fluxes of melt from depth. Nature Geosciences, 5, 749–754.
Parks, M.M., Moore, J.D.P., Papanikolaou, X., Biggs, J., Mather, T.A., Pyle, D.M.,
Raptakis, C., Paradissis, D., Hooper, A., Parsons, B., Nomikou, P., 2015. From
quiescence to unrest: 20 years of satellite geodetic measurements at Santorini
volcano, Greece Journal of Geophysical Research, 120, 1309-1328
Peltier, A., Staudacher, T., Bachèlery, P., Cayol, V., 2009. Formation of the April
2007 caldera collapse at Piton de La Fournaise volcano: Insights from GPS
data. Journal of Volcanology and Geothermal Research. 184, 152–163.
Pollard, M., 2010. Learning to recognise volcanic non-eruptions. Geology, 38, 287–
288.
Pryor, R.., 2011. Multiphysics modelling using COMSOL?: A first principles
approach, Jones and Bartlett education.
Pyle, D.M., 1997. The global impact of the minoan eruption of Santorini, Greece.
Environmental Geology, 30, 59–61.
82
Bibliography
Reynolds, H.I. & Gudmundsson, M.T., 2014. Variations in geothermal heat flux at
Grímsvötn, Iceland, 16, EGU General Assembly. Vienna.
Roche, O., Druitt, T.H., 2001. Onset of caldera collapse during ignimbrite eruptions.
Earth and Planetary Science Letters, 191, 191–202.
Saunders, S.J., 2001. The shallow plumbing system of Rabaul caldera: A partially
intruded ring fault? Bulletin of Volcanology, 63, 406–420.
Saunders, S.J., 2004. The possible contribution of circumferential fault intrusion to
caldera resurgence. Bulletin of Volcanology. 67, 57–71.
Scandone, R. & Malone., S.D., 1985. Magma supply, magma discharge and
readjustment of the feeding system of Mount St. Helens during 1980. Journal of
Volcanology and Geothermal Research, 3, 239–262.
Segall, P., 2013. Volcano deformation and eruption forecasting. Geological Society
London, Special Publications, 380, 85–106.
Shelley, D., 1993. Igneous and Metamorphic Rocks Under the Microscope:
Classification, Textures, Microstructures and Mineral Preferred Orientation,
Springer.
Shinohara, H., 2008. Excess degassing from volcanoes and its role on eruptive and
intrusive activity. Review of Geophysics. 46.
Sigmundsson, F., 2006. Iceland geodynamics: Outlook. In: Iceland Geodynamics:
Crustal Deformation and Divergent Plate Tectonics. 175–176.
Sigmundsson, F., Hooper, A., Parks, M., Spaans, K., Gudmundsson, G.B., Drouin,
V., Samsonov, S., White, R.S., Hensch, M., Pedersen, R., Bennett, R.A.,
Greenfield, T., Green, R.G., Sturkell, E., Bean, C.J., Mo, M., Femina, P.C. La,
Bjo, H., Pa, F., Braiden, A.K., Eibl, E.P.S., 2014. Segmented lateral dyke
growth in a rifting event at Barðarbunga volcanic system, Iceland. Nature.
Simmons, G. & Cooper, H.W., 1978. Three Igneous Rocks, 15, 145–148.
83
Bibliography
Siratovich, P. A., von Aulock, F.W., Lavallee, Y., Cole, J.W., Kennedy, B.M.,
Villeneuve, M.C. 2015. Thermoelastic properties of the Rotokawa Andesite: A
geothermal reservoir constraint. Journal of Volcanology and Geothermal
Research, 301, 1–13.
Stevenson, D.S. & Blake, S., 1998. Modelling the dynamics and thermodynamics of
volcanic degassing. Bulletin of Volcanology, 60, 307–317.
Sturkell, E., Einarsson, P., Sigmundsson, F., Geirsson, H., Ólafsson, H., Pedersen,
R., de Zeeuw-van Dalfsen, E., Linde, A.T., Sacks, S.I., Stefánsson, R., 2006a.
Volcano geodesy and magma dynamics in Iceland. Journal of Volcanol.
Geothermal Research 150, 14–34.
Sturkell, E. & Sigmundsson, F., 2000. Continuous deflation of the Askja caldera,
Iceland, during the 1983–1998 noneruptive period. Journal of Geophysical
Research, 105, 256-271.
Sturkell, E., Sigmundsson, F., Slunga, R., 2006b. 1983-2003 decaying rate of
deflation at Askja caldera: Pressure decrease in an extensive magma plumbing
system at a spreading plate boundary. Bulletin of Volcanology, 68, 727–735.
Thirlwall, M. F., Singer, B. S., & Marriner, G. F. 2000. 39 Ar–40 Ar ages and
geochemistry of the basaltic shield stage of Tenerife, Canary Islands, Spain.
Journal of Volcanology and Geothermal Research, 103, 247-297.
Trasatti, E., Giunchi, C., Agostinetti, N.P., 2008. Numerical inversion of
deformation caused by pressure sources: Application to Mount Etna (Italy).
Geophysical Journal International, 172, 873–884.
Turcotte, D. L., & Schubert, G. 2002. Geodynamics, Cambridge University Press,
New York 456 pp.
Vasseur, J., Wadsworth, F.B., Lavallée, Y., Hess, K.U., Dingwell, D.B., 2013.
Volcanic sintering: Timescales of viscous densification and strength recovery.
Geophysical Research Letters, 40, 5658–5664.
84
Bibliography
Vinciguerra, S. et al., 2005. Relating seismic velocities, thermal cracking and
permeability in Mt. Etna and Iceland basalts. International Journal of Rock
Mechanics and Mining Sciences, 42, 900–910.
Wadsworth, F.B., Vasseur, J., Aulock, F.W. Von, Hess, K.U., Scheu, B., Lavallée,
Y., Dingwell, D.B., 2014. Nonisothermal viscous sintering of volcanic ash.
Journal of Geophysical Research, 119, 8792-8804
Walker, G.P.L., 1960. Zeolite zones and dike distributions in relation to the structure
of the basalts of eastern iceland. Journal of Geology, 68, 515–528.
Walker GPL. 1975 Intrusive sheet swarms and the identity of Crustal Layer 3 in
Iceland. Journal of the Geological Society of London, 131, 143-161.
Wilson, C.J.N., Houghton, B.F., McWilliams, M.O., Lanphere, M. a, Weaver, S.D.,
Briggs, R.M., 1994. Volcanic and structural evolution of Taupo Volcanic Zone,
New Zealand: a review. Journal of Volcanology and Geothermal Research, 68,
1–28.
Yun, S., Segall, P., Zebker, H., 2006. Constraints on magma chamber geometry at
Sierra Negra Volcano, Galapagos Islands, based on InSAR observations.
Journal of Volcanology and Geothermal Research 150, 232–243
85
Appendices
Appendix A
Additional experimental measurements
A1
Differential scanning calorimetry conducted at University of Liverpool (a) and
Lancaster (b&c) University was used to provide insights into melting and
crystallisation regimes in all rock types tested. All DSC results require a degree of
interpretation (shown as annotation on each figure) that is highly subjective.
Therefore these results were used to compliment TMA data, rather than act as standalone evidence for meting and annealing. Clear endothermic peaks indicate partial
melting in AP and IB around 1100C, and a peak at approximately 750C in NKD
matches well with the expected Tg temperature.
84
Appendices
A2
Hot-stage microscopy results proved largely inconclusive for evidence of melting at
high temperature. AP displayed no observational changes. Evidence of melt
relaxation was observed in IB over a 10 minute period in the temperature range
1100C to 1200C. Here we see microscopic changes in the light passing through the
edge of a pyroxene crystal, indicating in melt delamination around the crystal
surface.
85
Appendices
A3
Acoustic emission hit energy as a function of temperature for three rock types
(Basalt, IB, Phonolite, AP, Dacite, NKD). Samples heated at a rate of 4˚C/min and
cooled naturally (black dashed line). Onset of high energy bursts occur at the onset
of cooling (blue lines) with most energy generated around 800˚C upon cooling.
Similarly AE hit energy is shown on different scales for the same reason.
Corresponding b values are calculated using Aki’s maximum likelihood method for
200 hits at 100 hit intervals shown as orange lines.
86
Appendices
A4
SEM images and micro-crack analysis of a non-heat treated Anaga Phonolite. a)
original SEM image x 250 zoom, c) x 500 zoom of the same area, b) crack map
drawn in Adobe Illustrator and then imported into MATLAB, d) rose diagram
showing the orientation and frequency of cracks.
87
Appendices
A5
SEM images and micro-crack analysis of a heat treated Anaga Phonolite. a) original
SEM image x 250 zoom, c) x 500 zoom of a different area, b) crack map drawn in
Adobe Illustrator and then imported into MATLAB, d) rose diagram showing the
orientation and frequency of cracks.
88
Appendices
Appendix B
Supplementary data of Browning et al., (in press) (Chapter 3).
B1
The mean lava flow volume (common eruption) at Nea Kameni was calculated by
averaging all known lava volumes. The original data was collected and published by
Nomikou et al., (2014), partly using bathymetric datasets and satellite techniques.
89
Appendices
Appendix C
Supplementary data of Browning and Gudmundsson (2015)
(Chapter 3)
C1
Numerical setup showing the continuation of an inclined sheet that has been captured
within the fault and propagated vertically up the fault.
90