0% found this document useful (0 votes)
29 views23 pages

JGR Solid Earth - 2021 - Heap - Alteration Induced Volcano Instability at La Soufri Re de Guadeloupe Eastern Caribbean

Uploaded by

aichaelbakri468
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
29 views23 pages

JGR Solid Earth - 2021 - Heap - Alteration Induced Volcano Instability at La Soufri Re de Guadeloupe Eastern Caribbean

Uploaded by

aichaelbakri468
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 23

RESEARCH ARTICLE Alteration-Induced Volcano Instability at La Soufrière de

10.1029/2021JB022514
Guadeloupe (Eastern Caribbean)
Key Points:
Michael J. Heap1,2 , Tobias S. Baumann3,4, Marina Rosas-Carbajal5 ,
• L aboratory experiments show that
Jean-Christophe Komorowski5 , H. Albert Gilg6 , Marlène Villeneuve7,
hydrothermal alteration reduces the
strength of volcanic rock from La Roberto Moretti5,8 , Patrick Baud1 , Lucille Carbillet1 , Claire Harnett9, and
Soufrière Thierry Reuschlé1
• Numerical modeling shows that
1
hydrothermal alteration significantly Université de Strasbourg, CNRS, Institut Terre et Environnement de Strasbourg, Strasbourg, France, 2Institut
increases volcano deformation and Universitaire de France (IUF), Paris, France, 3Institute of Geosciences, Johannes Gutenberg University Mainz, Mainz,
collapse volume Germany, 4Terrestrial Magmatic Systems (TeMaS) Research Platform, Mainz, Germany, 5Université de Paris, Institut de
• We provide a 3D strength map of
La Soufrière that exposes a low-
Physique du Globe de Paris, CNRS, Paris, France, 6Department of Civil, Geo and Environmental Engineering, Technical
strength zone coincident with the University of Munich, Munich, Germany, 7Department Mineral Resources Engineering, Montanuniversität Leoben,
hydrothermal system Leoben, Austria, 8Observatoire Volcanologique et Sismologique de Guadeloupe, Institut de Physique du Globe de Paris,
Gourbeyre, France, 9School of Earth Sciences, University College Dublin, Dublin, Ireland
Supporting Information:
Supporting Information may be found
in the online version of this article.
Abstract Volcanoes are unstable structures that deform laterally and frequently experience mass
wasting events. Hydrothermal alteration is often invoked as a mechanism that contributes significantly
to volcano instability. We present a study that combines laboratory experiments, geophysical data, and
Correspondence to:
large-scale numerical modeling to better understand the influence of alteration on volcano stability, using
M. J. Heap,
[email protected] La Soufrière de Guadeloupe (Eastern Caribbean) as a case study. Laboratory experiments on variably
altered (advanced argillic alteration) blocks show that uniaxial compressive strength, Young's modulus,
Citation: and cohesion decrease as a function of increasing alteration, but that the internal friction angle does
Heap, M. J., Baumann, T. S., Rosas- not change systematically. Simplified volcano cross sections were prepared (a homogenous volcano, a
Carbajal, M., Komorowski, J.-C., Gilg, volcano containing the alteration zone identified by a recent electrical survey, and a volcano with an
H. A., Villeneuve, M., et al. (2021).
artificially enlarged area of alteration) and mechanical properties were assigned to zones corresponding to
Alteration-induced volcano instability
at La Soufrière de Guadeloupe unaltered and altered rock. Numerical modeling performed on these cross sections, using a hydro-thermo-
(Eastern Caribbean). Journal of mechanical modeling code, show (a) the importance of using upscaled values in large-scale models and
Geophysical Research: Solid Earth,
(b) that alteration significantly increases volcano deformation and collapse volume. Finally, we combined
126, e2021JB022514. https://doi.
org/10.1029/2021JB022514 published muon tomography data with our laboratory data to create a 3D strength map, exposing a low-
strength zone beneath the southern flank of the volcano coincident with the hydrothermal system. We
Received 29 MAY 2021 conclude that hydrothermal alteration decreases volcano stability and thus expedites volcano spreading
Accepted 29 JUL 2021
and increases the likelihood of mass wasting events and associated volcanic hazards. Hydrothermal
alteration, and its evolution, should therefore be monitored at active volcanoes worldwide.

Plain Language Summary The rocks forming a volcanic edifice can be altered by circulating
hydrothermal fluids. This alteration can influence the physical and mechanical properties of these rocks,
which could jeopardize volcano stability. The stability of a volcanic edifice is an important consideration
in volcanic hazards and risk assessments due to the potentially dire consequences of partial volcanic flank
collapse. Using a combination of experimental data, geophysical data, and modeling, and La Soufrière de
Guadeloupe (Eastern Caribbean, France) as a case study, we find that hydrothermal alteration decreases
volcano stability and thus promotes volcano instability and associated volcanic hazards. As a result,
we conclude that hydrothermal alteration, and its evolution, should be monitored at active volcanoes
worldwide.

1. Introduction
Volcanoes are inherently unstable structures that are built haphazardly, in both space and time, from the
products of successive effusive and explosive eruptions and endogenous growth. These materials have
highly variable physical and mechanical properties (Heap & Violay, 2021) and often form oversteepened
© 2021. American Geophysical Union. and unstable slopes. As a result, volcano deformation (such as volcano spreading; Borgia et al., 2000) and
All Rights Reserved. mass wasting events (such as debris avalanches resulting from partial flank collapse, lahars, and rockfalls;

HEAP ET AL. 1 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Roverato et al., 2021) are commonplace at many volcanoes worldwide. Catastrophic collapse resulting from
volcano spreading (van Wyk de Vries & Francis, 1997) and mass wasting events present a significant volcan-
ic hazard, and can also trigger hazardous laterally directed explosions and devastating—both economically
and in terms of loss of life—pyroclastic density currents (Cole et al., 2015; Glicken, 1996; Sparks et al., 2002;
Voight et al., 1981). Indeed, partial flank collapses at about 200 volcanoes have resulted in at least 20,000
fatalities in the last 10,000 years (Siebert et al., 2010). Although about 52% of these flank collapses are asso-
ciated with magmatic eruptions, about 22% are associated with non-magmatic phreatic and/or hydrother-
mal eruptions (Siebert, 1984; Siebert et al., 1987, 2010). As a result, stability assessments at volcanoes are an
essential component of volcano monitoring and volcanic hazard mitigation.

Hydrothermal alteration, common to many volcanoes worldwide, is often invoked as a mechanism that
contributes significantly to volcano instability. Indeed, studies using geological evidence, geochemical or
geophysical data, numerical modeling, laboratory experiments, or a combination of these approaches have
highlighted that hydrothermal alteration has or can weaken a volcanic slope sufficiently to promote col-
lapse (Ball et al., 2013, 2015; 2018; Cecchi et al., 2004; del Potro & Hürlimann, 2009; Finn et al., 2018; John
et al., 2008; López & Williams, 1993; Opfergelt et al., 2006; Reid et al., 2001; Rosas-Carbajal et al., 2016;
Salaün et al., 2011; van Wyk de Vries et al., 2000; Voight et al., 2002; Watters et al., 2000). Fluid and pore
fluid pressure re-distributions caused by alteration are also thought to promote instability (Ball et al., 2018;
Day, 1996; Heap, Baumann, et al., 2021; Reid, 2004) and erratic explosive behavior (de Moor et al., 2019;
Heap et al., 2019; Mick et al., 2021).

Geological evidence is provided by the abundance of hydrothermally altered materials typically found in
debris avalanche deposits resulting from partial flank collapse. For example, in their analysis of the De-
cember 26, 1997 partial edifice collapse at Soufrière Hills volcano (Montserrat, Eastern Caribbean), Voight
et al. (2002) found varicolored, hydrothermally altered materials within the avalanche deposits, highlight-
ing that alteration could have contributed to the collapse of the southern sector of the volcano. Avalanche
deposits and material ejected during phreatic explosions at La Soufrière de Guadeloupe (Eastern Caribbean,
France) were also found to contain various parts of the active and ancient hydrothermal systems of the
volcano (Salaün et al., 2011). Debris avalanche deposits at Tutupaca volcano (Peru) also contain hydrother-
mally altered materials (Samaniego et al., 2015). Extensive fumarole activity and hydrothermal alteration
on the unstable flanks of Casita volcano (Nicaragua) was documented by van Wyk de Vries et al. (2000), a
volcano that experienced a partial flank collapse in 1998 with associated debris avalanches that killed about
2,500 people (Kerle & De Vries, 2001; van Wyk de Vries et al., 2000). Opfergelt et al. (2006) further suggested
that the presence of abundant smectite (up to 50 wt.%) in the hydrothermally altered core of Casita volcano
contributed to slope instability by acting as a barrier to water infiltration thereby promoting fluid circulation
along marked discontinuities, gradually decreasing the shear strength of the host-rock. Finally, more than
55 debris flows have originated from Mt Rainer (USA) in the Holocene, and hydrothermally derived clays
have been found in some of the most widespread lahar deposits (Reid et al., 2001).

Geophysical and geochemical evidence has been provided by a wide range of different techniques. For ex-
ample, electrical tomography data exposed the extent of the hydrothermally altered zone, the location and
geometry of low-strength detachment planes, and potential collapse volumes at La Soufrière de Guadeloupe
(Rosas-Carbajal et al., 2016). Hyperspectral remote sensing (Kereszturi et al., 2021) and helicopter electro-
magnetic and magnetic measurements (Finn et al., 2018) have been used to map hydrothermal alteration
zones to inform slope stability modeling at Mt Ruapehu (New Zealand) and Mount Baker (USA), respec-
tively, highlighting the most likely location of future collapses. Finally, geochemical data and modeling
provided by López and Williams (1993) suggested that hydrothermal alteration along faults can create weak
sliding planes that can facilitate slope failure.

Numerical and analog modeling has also provided key insights into how hydrothermal alteration can in-
fluence volcano stability. For example, using analogue experiments in which sand-plaster mixtures and
silicone were used to represent fresh and altered rock, respectively, van Wyk de Vries et al. (2000) showed
that a weak and ductile (hydrothermally altered) core can reproduce the deformation structures observed
at Casita volcano, suggesting that hydrothermal alteration destabilized the volcano. Motivated by intensive
alteration on the flanks and summit of Mt Rainer, Reid et al. (2001) examined the influence of alteration
on volcano stability using a geotechnical method of 3D column limit-equilibrium slope stability analysis.

HEAP ET AL. 2 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Using a best-estimate distribution of the alteration, and typical values of


cohesion, internal friction angle, and rock unit weight for fresh, lightly
altered, and highly altered rocks, Reid et al. (2001) concluded that steep
slopes and large volumes of altered and weak rock can promote potential-
ly large gravitational failures.

Although these aforementioned studies, and others, have improved our


understanding of alteration-induced volcano instability, the shape and
size of the alteration zones and/or the geomechanical properties used in
large-scale stability modeling are often poorly constrained or arbitrarily
chosen. Indeed, there is a paucity of mechanical data for variably altered
volcanic rocks for use in large-scale numerical models designed to better
understand flank stability. As noted by Ball et al. (2018), the scarcity of
rock physical and mechanical properties for altered volcanic rocks limits
the investigation of volcano stability using large-scale models. We pres-
ent here a multidisciplinary study in which we first provide physical and
mechanical properties for variably altered andesites from La Soufrière
de Guadeloupe. We then assign laboratory-scale and upscaled (i.e., mod-
ified to account for large-scale discontinuities that are not captured in
laboratory-scale samples) physical and mechanical properties to zones
identified by a recent electrical survey of the dome of La Soufrière de
Guadeloupe (Rosas-Carbajal et al., 2016) and perform large-scale numer-
ical modeling to showcase the importance of hydrothermal alteration
on volcano stability. Finally, we combine recent muon tomography data
(Rosas-Carbajal et al., 2017) with our laboratory data to create a 3D com-
pressive strength map of the volcano. Although we use La Soufrière de
Guadeloupe as a case study, we consider our salient conclusions relevant
for andesitic stratovolcanoes worldwide.

2. La Soufrière de Guadeloupe (Eastern Caribbean,


France)
La Soufrière de Guadeloupe (Figure 1a), henceforth called La Soufrière,
is located on the island of Guadeloupe in the Eastern Caribbean (Fig-
ure 1b) and is an archetype of the hazardous andesitic volcanoes that
occur in many subduction zones. The current lava dome formed in 1530
Figure 1. (a) Photograph of La Soufrière de Guadeloupe (Eastern
Caribbean, France). This photograph also shows the collapse scar CE and, since 1635 CE, there have been six non-magmatic phreatic or hy-
(indicated by the white arrow) triggered by the November 21, 2004 Les drothermal eruptions (Komorowski et al., 2005). The largest and most-re-
Saintes magnitude Mw 6.3 regional earthquake (Feuillet et al., 2011). (b) cent eruption occurred in 1976–1977. After months of seismic unrest, the
Map of La Soufrière de Guadeloupe (taken from Google Maps®) showing volcano exploded suddenly and without precursory signs in 1976–1977,
the sampling locations for the 17 blocks acquired for this study. This image
resulting in a volcanic crisis that led to the evacuation of about 70,000
also shows the location of the dome and the 2009 collapse scar (indicated
by white arrows). Inset shows a map of Guadeloupe in which the location people and serious socioeconomic consequences for the island (Feuillard
of La Soufrière de Guadeloupe is indicated by a red triangle. et al., 1983; Hincks et al., 2014; Komorowski et al., 2005, 2015).

Volcanic unrest has been increasing since the volcano reawakened in


1992, marked by the restart of fumarolic and seismic activity (Komorowski
et al., 2005) and, in April 2018, the largest felt volcano-tectonic earthquake since the 1976–1977 crisis (ML
4.1 or MW 3.7; Moretti et al., 2020). The 2018 episode of accelerated unrest was interpreted as a failed phreat-
ic eruption and raised serious concerns for the stability of the southwest flank of the dome, which shows the
largest displacements (up to 9 mm/year toward the southwest) over the last 20 years (Moretti et al., 2020).
The main features of the ongoing unrest include the continuous expansion of the outgassing area on top of
the lava dome and the appearance of new steam-dominated fumaroles in July 2014 and February 2016 that
are characterized by a deep magmatic gas component (Brombach et al., 2000; Moretti et al., 2020; Villemant
et al., 2014). Such expansion was accompanied by an increase in the heat output from the entire dome,
which evolved from ~1.2 MW in 2010 to ~7.6 MW in 2020 (Jessop et al., 2021). This increase in heat output

HEAP ET AL. 3 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

is of great concern for La Soufrière as it has been recently demonstrated that volcanoes with a high heat flux
are more likely to experience magmatic or phreatic eruptions (Girona et al., 2021). High-flow rate thermal
acid sulfate-chloride springs, associated with degassing of the underlying magmatic body, are observed
on the slopes and at the base of the dome and supply hot acid fluids and CO2, but also SO2, H2S, HCl, and
HF, to the above hydrothermal system (Villemant et al., 2005, 2014). These acidic fluids are (a) transferred
through the dome and feed the steam-rich summit fumaroles and (b) circulate through the dome structure
along listric, concave-upward structural discontinuities that drain to the southwest (Brombach et al., 2000;
Rosas-Carbajal et al., 2016; Salaün et al., 2011; Villemant et al., 2014). Evidence for the preferential circu-
lation of acidic thermal fluids along these structures is provided by the 3D geometry of thermal springs
(Moretti et al., 2020; Villemant et al., 2005) and geophysical imaging (Bouligand et al., 2016; Brothelande
et al., 2014; Nicollin et al., 2006; Rosas-Carbajal et al., 2016). The shallow hydrothermal system is consid-
ered responsible for the recent observed shallow deformation and seismic activity (at depths at or above
sea level, with magnitudes less than one) (Moretti et al., 2020). Indeed, the monthly bulletins from the ob-
servatory (Observatoire Volcanologique et Sismologique de Guadeloupe, Institut de Physique du Globe de
Paris [OVSG-IPGP], 1999–2021; available at http://www.ipgp.fr/fr/ovsg/bulletins-mensuels-de-lovsg) docu-
ment hundreds of shallow, low-magnitude (M < 1) earthquakes each month. Deeper and higher magnitude
off-axis seismic activity are thought to be related to increases in pore pressure in response to the pulsatory
arrival of hot magmatic fluids along the larger faults that cut the dome (Moretti et al., 2020).

The recent unrest, suggesting that La Soufrière is undergoing a renewed period of structural and chemical
weakening, raises concern for the stability of the volcano. Indeed, the most-recent unrest and instability
at La Soufrière was associated with hydrothermal activity. For example, the materials ejected during the
explosive hydrothermal eruption of 1976–1977 were largely hydrothermally altered (Feuillard et al., 1983)
and geological studies have also shown that La Soufrière has an exceptional record of partial edifice collapse
(at least eight collapses in the past 9,150 years) that have produced extensive debris avalanche deposits
dominantly composed of hydrothermally altered materials (Boudon et al., 1987; Komorowski et al., 2005; Le
Friant et al., 2006; Peruzzetto et al., 2019; Rosas-Carbajal et al., 2016; Salaün et al., 2011). The simulations of
Le Friant et al. (2006) and Peruzzetto et al. (2019) suggest that the northern and eastern parts of the town of
Saint Claude could be impacted by small-volume partial dome collapses and that, in the worst-case scenario
of a major dome collapse, a large part of Basse-Terre could be affected. As a result, not only is La Soufrière
an ideal natural laboratory to study the influence of hydrothermal alteration on volcano stability, but such
a study is also timely because of the ongoing, and increasing, unrest at the volcano.

3. Materials and Methods


A total of 17 blocks were collected from La Soufrière during a field campaign in 2019 (sampling locations
shown as red stars in Figure 1b). Our aim was to sample a range of materials from different locations on the
volcano that best represent the variability in porosity and hydrothermal alteration of the rocks forming the
edifice. A large proportion of the blocks (eight out of 17) were collected from the collapse scar of the No-
vember 19, 2009 landslide, triggered by extreme rainfall (Moretti et al., 2020; Peruzzetto et al., 2019), which
was seen as an opportunity to sample blocks that are more representative of the interior of the volcano.

Eight blocks were taken from the collapse scar of the 2009 landslide (H2A, H2B, H3, H4A, H5A, H6, H25,
and H29), one block was taken from the collapse scar of the landslide triggered by the November 21, 2004
Les Saintes magnitude Mw 6.3 regional earthquake (Feuillet et al., 2011, Figure 1b) (WP1285), two blocks
were taken from the West wall of the fault “Faille 30 août” (H14 and H15), and one block, a volcanic bomb
from the 1976–1977 eruption, was taken from the roof of a small disused thermal bathhouse to the South
of the dome (WP1317). The remaining samples were taken from the dome: one block was taken from the
“Lacroix Supérieur” outgassing fracture (H18), and four blocks were taken from the lava spines of the 1530
CE dome: two blocks from “Cratère Sud Central” (H19 and H20) and two blocks from an adjacent site (H21
and H22). Site descriptions and Global Positioning System (GPS) coordinates are available for each block in
the Supporting Information.

Cylindrical samples were prepared from each of the blocks to a diameter of 20 mm and then cut and preci-
sion-ground to a nominal length of 40 mm. The samples were washed and then dried in a vacuum-oven at

HEAP ET AL. 4 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

40°C for at least 48 h. The connected porosity of each sample was calculated using the bulk sample volume
and the skeletal (solid) sample volume measured by a helium pycnometer. Dry and wet bulk sample den-
sities were calculated using the bulk sample volume and the dry and wet mass, respectively. Polished thin
sections were prepared from offcuts of the samples for microstructural analysis, performed using a Tescan
Vega 2 XMU scanning electron microscope (SEM). The mineral content of the 17 blocks was analyzed and
quantified using X-ray powder diffraction (XRPD) on powdered offcuts of the samples, assisted by Raman
spectroscopy on the prepared thin sections (see Supporting Information S1 for more details).

Dry uniaxial compressive strength was measured on oven-dry samples from all 17 blocks in a uniaxial
load frame. Samples were deformed under ambient laboratory pressure and temperature at a constant axial
strain rate of 10−5 s−1 until macroscopic failure. The static Young's modulus was determined from the elastic
portion of the uniaxial stress-strain curves (Heap, Villeneuve, et al., 2020). Triaxial compression experi-
ments were performed on blocks selected to best represent the variability in observed alteration (samples
H2B, H3, and H18). These samples were vacuum-saturated in deionized water, inserted into a rubber jack-
et, placed inside a pressure vessel, and then taken to the target confining (from 10.5 to 25 MPa) and pore
fluid pressures (10 MPa for all experiments) using servo-controlled pumps. Samples were deformed under
ambient laboratory temperature at a constant axial strain rate of 10−5 s−1 until either macroscopic failure
(for brittle experiments) or until 3% axial strain (for ductile experiments). The triaxial apparatus was also
used to measure the water-saturated uniaxial compressive strength of samples H2B, H3, and H18. For both
uniaxial and triaxial experiments, axial displacement and axial load were measured using a linear variable
differential transducer and a load cell, respectively. Axial displacement (minus the displacement accumu-
lated within the load chain) and axial load were converted to axial strain and axial stress using the sample
dimensions. Here, we define the effective pressure, Peff , as the confining pressure, Pc, minus the pore fluid
pressure, Pp. We adopt the convention that compressive stresses and strains are positive.

4. Results
4.1. Microstructure, Mineralogy, and Alteration

The mineral contents for each of the 17 blocks are given in Table 1, and Figure 2 shows backscattered SEM
images of three blocks selected to represent the range of alteration intensity (defined by their percentage
of secondary minerals) observed in the rocks collected for this study (relatively unaltered, H18; moderately
altered, H3; and highly altered, H2B). SEM images of all the blocks are available in the Supporting Informa-
tion. All of the andesite blocks are characterized by a porphyritic texture comprising phenocrysts of domi-
nantly plagioclase and pyroxene (and high-density oxides) within a crystallized groundmass (Figure 2). All
of the samples also contain secondary minerals, including silica polymorphs (quartz, cristobalite, tridymite,
and opal-A), hematite, pyrite, alunite or natroalunite, gypsum, kaolinite, and talc (Table 1). Plagioclase phe-
nocrysts are often pervasively altered and partially replaced by kaolinite (e.g., Figure 2c) or opal-A. Second-
ary minerals (natroalunite, alunite, cristobalite, tridymite, pyrite, and kaolinite) are also found precipitated
within pores and microcracks (see Supplementary Information S1 for additional SEM images).

4.2. Mechanical Data

Representative uniaxial stress-strain curves for selected samples are shown in Figure 3a. Stress is first a
non-linearly increasing function of strain, often attributed to the closure of pre-existing microcracks. This
stage is followed by a quasi-linear elastic stage. In the next stage, stress is a nonlinearly decreasing func-
tion of strain, a result of the initiation, propagation, and coalescence of microcracks. Finally, a stress drop
signals the formation of a macroscopic shear fracture. Uniaxial compressive strength and Young's modulus
are plotted as a function of connected porosity in Figures 3b and 3c, respectively (data available in Table 2).

strength and Young's modulus of our sample suite varies from ~3.5 to ~150 MPa and from ~1 to ∼43 GPa, re-
We first note that the connected porosity of our samples varies from 0.12 to 0.43. The uniaxial compressive

spectively (Figure 3). Our data show that uniaxial compressive strength and Young's modulus both decrease
as a function of increasing porosity (Figure 3).

Samples from three blocks were selected for triaxial compression experiments: a relatively unaltered
block (H18), a moderately altered block (H3), and a very altered block (H2B). The microstructure of these

HEAP ET AL. 5 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 1
Mineral Contents, Measured by X-Ray Powder Diffraction and Refined Using Raman Spectroscopy and Optical Microscopy, of the 17 Blocks Collected for This
Study (Sampling Locations Shown in Figure 1b)
Mineral H2A H2B H3 H4A H5A H6 H14 H15 H18 H19 H20 H21 H22 H25 H29 WP1285 WP1317
Plagioclase 56.7 12.3 46.6 23.3 41.3 30.0 60.7 22.5 61.2 22.0 28.7 24.2 59.5 38.7 62.4 64.7 61.6
Clinopyroxene 8.7 3.4 5.6 4.9 5.2 6.4 6.3 7.3 8.4 5.0 8.9 12.4 8.9 5.3 7.8 5.2 5.9
Orthopyroxene 10.8 9.5 11.8 11.8 11.1 10.8 8.6 9.2 12.2 10.2 15.0 19.3 13.6 10.2 11.2 13.2 15.6
(Ti-) Magnetite 0.7 - 0.8 - - - 0.8 - 2.9 - 2.4 3.1 0.8 - 2.7 3.5 0.7
a
Quartz 1.0 0.5 0.6 0.6 0.5 0.5 1.7 0.7 0.7 1.7 0.3 0.2 0.6 0.3 0.4 0.2 0.7
Cristobalitea 11.3 12.8 10.6 11.8 13.0 11.1 13.5 10.2 11.7 9.5 11.4 11.7 10.6 9.8 12.4 - -
a
Tridymite - - - - - - - 0.7 - - - - - - - 13.2 13.2
Hematitea - - - - - - 3.4 - 2.8 2.4 - - - - 3.1 - -
a
Pyrite 3.5 - 3.8 2.3 - - - - - - - 0.4 3.1 0.6 - - -
Alunitea - - - - - - - - - - - - - - - - 2.4
Na-Alunitea 1.4 1.6 2.8 1.3 5.4 5.1 5.1 15.0 - 14.2 0.5 0.5 - 9.8 - - -
a
Gypsum - - - 0.7 - - - - - - 0.8 1.2 - - - - -
Kaolinitea 6 59.7 17.4 43.3 23.5 36.0 <1 34.3 - 2.0 2.0 2.0 <1 25.3 - - -
Talca - - - - - - - - - - - - 2.9 - - - -
Opal-Aa - - - - - - - - - 33.0 30.0 25.0 - - 10.0 - -
Note. Values in wt.%.
a
A secondary/alteration mineral.

samples is provided in Figure 2 and their mineral componentry in Table 1. The percentages of secondary
minerals (determined using the XRPD data; Table 1) for these rocks are 15.2%, 35.2%, and 74.6%, respective-
ly. Triaxial experiments were performed on H18 samples at effective pressures of 0, 2.5, 5, 10, and 15 MPa
(Figure 4a), on H3 samples at effective pressures of 0, 2.5, 5, and 7.5 MPa (Figure 4b), and on H2B samples at
effective pressures of 0, 0.5, 1, and 2 MPa (Figure 4c) (data available in Table 3). Different effective pressures
were chosen for the different samples in an attempt to ensure the brittle behavior required to calculate the
cohesion and angle of internal friction. Blocks H18 (Figure 4a) and H3 (Figure 4b) were brittle over the cho-
sen range of effective pressures. The compressive strength of blocks H18 and H3 increased as a function of
increasing effective pressure: H18 increased from 65.4 to 143.8 MPa and H3 increased from 31.5 to 54.8 MPa
as effective pressure was increased from 0 to 15 and 7.5 MPa, respectively (Figure 4). The stress-strain curves
for block H2B are very different to those for blocks H18 and H3 (Figure 4). Subtle strain softening associated
with brittle behavior was observed for the experiments performed at effective pressures of 0 and 0.5 MPa,
but no strain softening was seen at 1 and 2 MPa, suggestive of ductile behavior (Figure 4c).

The cohesion and angle of internal friction calculated using these triaxial data (excluding the two ductile
H2B experiments), necessary for the large-scale numerical modeling, were 14.8 MPa and 46.7° for H18,
7.1 MPa and 37.7° for H3, and 0.4 MPa and 42.5° for H2B, respectively (Figure 4). The method for determin-
ing these values is explained in the Supporting Information.

5. Discussion
5.1. Hydrothermal Alteration and Mineral Precipitation

Volcanic andesitic systems discharging acid-chloride-sulfate hydrothermal fluids (Browne, 1978; Ellis &
Mahon, 1977; Taran & Kalacheva, 2020) are typically characterized by either propylitic alteration, result-
ing from the rapid ascension of acidic fluids and the lack of time available for neutralization (Taran &
Kalacheva, 2020), or intermediate to advanced argillic alteration in zones where fluids can circulate. The
samples measured for this study were sourced from (a) zones proximal to summit fumarolic vents that

HEAP ET AL. 6 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Figure 2. Backscattered scanning electron microscope images of three andesites from La Soufrière de Guadeloupe
(Eastern Caribbean, France) selected to represent the range of observed alteration: (a) relatively unaltered, H18,
(b) moderately altered, H3, and (c) highly altered, H2B. Black, porosity; grayscale, rock groundmass. Insets show
photographs of the 20 mm-diameter cylindrical samples.

HEAP ET AL. 7 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Figure 3. (a) Representative stress-strain curves for selected andesite samples from La Soufrière de Guadeloupe
(Eastern Caribbean, France). Inset shows an isolated curve indicating from where the Young's modulus (the slope of
the stress-strain curve), E , was determined (the quasi-linear elastic stage is indicated by the gray zone). Also labeled
is the uniaxial compressive strength,  p , which is the maximum axial stress the sample attained prior to macroscopic
failure (signaled by the stress drop). Uniaxial compressive strength (b) and Young's modulus (c) as a function of
connected porosity (data available in Table 2).

HEAP ET AL. 8 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 2 tap the central zone characterized by the rapid ascent of steam separat-
Connected Porosity, Uniaxial Compressive Strength, Young's Modulus, ed from a boiling hydrothermal aquifer (Brombach et al., 2000; Moretti
and the Percentage of Secondary Minerals (i.e., Alteration) for the Samples et al., 2020) (e.g., H18–22; Table 1) and (b) from zones far from currently
Prepared for This Study active fumarolic zones, where the shallow groundwaters are heated via
Uniaxial Young's Percentage conduction and mixed with Na-Cl liquids from the hydrothermal aqui-
Connected compressive modulus of secondary fer (Brombach et al., 2000; Villemant et al., 2014) (e.g., those sampled
Sample porosity strength (MPa) (GPa) minerals from the collapse scar of the 2009 landslide: H2A, H2B, H3, H4A, H5A,
H2A_2 0.18 63.7 31.9 23.2 H6, H25, and H29; Table 1) (Figure 1b).
H2A_5 0.19 80.5 25.5 23.2
The secondary mineral assemblage in our samples (Table 1) is essential-
H2A_6 0.19 65.4 25.8 23.2 ly the same as that found in the debris avalanche deposits to the south-
H2A_11 0.19 59.3 24.4 23.2 west of La Soufrière (Salaün et al., 2011). However, the debris-avalanche
H2A_12 0.18 70.4 28.0 23.2 deposits investigated by Salaün et al. (2011) contain abundant smectite,
H2B_3 0.42 6.6 1.8 74.6
whereas the samples collected for this study contain kaolinite and no
smectite (Table 1). The presence of kaolinite, a key mineral in the argillic
H2B_10 0.42 6.4 1.9 74.6

fluid temperatures (<150–200°C) and a pH of ∼4.5–6 (Fulignati, 2020;


alteration facies, typically indicates fluid-rock interaction at relatively low
H2B_11 0.41 8.7 2.3 74.6
H2B_12 0.41 8.1 2.3 74.6 Inoue, 1995), but can also occur in the advanced argillic alteration facies
H2B_15 0.43 4.6 1.6 74.6 (pH < 3; temperatures up to about 300°C). The altered samples analyzed
H3_3 0.16 70.3 27.3 35.2 in this study therefore represent a more advanced stage of hydrolysis re-
actions than the smectite-bearing products from the Holocene debris-av-
H3_7 0.16 69.9 24.9 35.2
alanche deposits described in Salaün et al. (2011), indicating that the
H3_8 0.16 64.8 27.5 35.2
rocks analyzed here were altered by hydrothermal fluids with a higher
H3_11 0.16 59.7 24.0 35.2 H+ activity that led to complete alkali leaching.
H3_13 0.16 71.2 26.2 35.2
The most abundant secondary minerals are cristobalite, amorphous sili-
H4A_2 0.23 40.3 10.4 60.0
ca, natroalunite, and kaolinite (Table 1). Cristobalite is a silica polymorph
H4A_4 0.23 42.0 10.6 60.0 stable at low-pressure and high-temperature in its cubic β-form (Schipper
H4A_6 0.23 43.9 8.1 60.0 et al., 2020). Because it is kinetically favored, metastable cristobalite is
H4A_8 0.23 37.2 9.3 60.0 commonly found in preference to quartz in lava domes, where it precip-
H4A_9 0.22 40.4 9.9 60.0 itates from a vapor-phase enriched in magmatic components or results
from groundmass devitrification (Horwell et al., 2013; Martel et al., 2021;
H5A_2 0.16 88.3 26.4 42.4
Schipper et al., 2020). Our samples contain tetragonal α-cristobalite, sug-
H5A_3 0.16 86.6 22.8 42.4
gesting a final temperature lower than 240°C, the nominal temperature
H5A_5 0.16 83.9 22.5 42.4 for the transition from β- to α-cristobalite (Damby et al., 2014). Schipper
H5A_8 0.18 68.8 21.2 42.4 et al. (2020) identified a lower and broader range for the β-α transition in
H5A_10 0.17 74.5 23.0 42.4 volcanic cristobalites (∼150–200°C), which overlaps with the tempera-
H6_6 0.18 55.1 13.5 52.7 ture interval inferred by the presence of kaolinite. The cristobalite in the
studied samples precipitated onto pore walls (see Figure 2 and the Sup-
H6_9 0.18 68.0 18.0 52.7
porting Information S1) and so we consider that it predominately formed
H6_9 0.18 53.0 12.5 52.7
by vapor deposition. Amorphous silica (opal-A) partially to completely
H6_12 0.18 64.2 17.6 52.7 replaced the plagioclase phenocrysts (see Figure 2 and the Supporting
H6_13 0.18 63.0 18.3 52.7 Information S1), as often observed in acid-sulfate alteration zones (Mc-
H14_2 0.18 65.4 26.1 23.7 Collom et al., 2013).
H14_3 0.21 21.1 11.6 23.7 Alunite is a characteristic family of sulfate minerals produced by hy-
H14_5 0.21 26.8 13.1 23.7 pogene magmatic-hydrothermal and steam-heated alteration or super-
H14_6 0.19 29.8 12.1 23.7 gene processes (Rye et al., 1992). As such it is commonly found in the
H14_10 0.21 16.1 6.6 23.7 alteration products of rhyolitic to andesitic rock compositions, including
surficial solfataric or acid hot spring environments. Alunite and natroal-
H15_3 0.28 24.6 10.0 60.9
unite are the two most common minerals of the alunite group (Stoffre-
H15_4 0.28 27.1 11.0 60.9
gen et al., 2000). Only the volcanic bomb from the 1976–1977 eruption
H15_5 0.29 22.8 9.4 60.9 (sample WP1317) contains alunite, whereas natroalunite is commonly
H18_3 0.13 101.3 37.9 15.2 observed (Table 1). Natroalunite is the high temperature end-mem-
H18_4 0.13 89.8 39.1 15.2 ber of the alunite group and typically forms at 200–350°C, but is also

HEAP ET AL. 9 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 2 stable at 380–450°C (Henley & Berger, 2011; Meyer & Hemley, 1967).
Continued Natroalunite precipitates from an acidic (pH < 4) fluid with a high sul-
fate content and high [Na]/[K] ratio (Hemley et al., 1969; Stoffregen &
Uniaxial Young's Percentage
Connected compressive modulus of secondary Cygan, 1990). Natroalunite precipitation requires fluids with [Na]/[K]
Sample porosity strength (MPa) (GPa) minerals about 100 (at pH = 3; Deyell & Dipple, 2005), much higher than meas-
H18_5 0.13 98.8 41.0 15.2
ured in circulating fluids and in whole rock and groundmass composi-
tions at La Soufrière (Boudon et al., 2008; Metcalfe et al., 2021; Ville-
H18_6 0.13 93.2 43.1 15.2
mant et al., 2005), suggesting that equilibrium was not obtained or that
H18_7 0.13 113.7 35.1 15.2 the fluids responsible are different from those discharged at the surface
H19_3 0.21 39.7 12.7 62.8 (Deyell & Dipple, 2005). If the fluids are different, acidic Na-Cl-SO4 flu-
H19_8 0.22 47.4 15.8 62.8 ids would be required, which form H2S and H2SO4 following the interac-
H19_9 0.22 34.1 12.7 62.8 tion between Na-Cl waters and elemental sulfur at depth (Taran & Kal-
acheva, 2020). This hypothesis is compatible with the observation that
H19_12 0.20 41.2 17.4 62.8
highly acidic chlorine-sulfate rich fluids (pH < 1) have been rising and
H19_15 0.22 39.6 13.2 62.8
circulating within the dome as of early 1998, forming acid ponds with-
H20_2 0.37 4.8 1.2 45.0 in summit craters (Komorowski et al., 2005; Rosas-Carbajal et al., 2016;
H20_3 0.37 4.4 1.3 45.0 Villemant et al., 2014). Alternative hypotheses include the interaction of
H20_7 0.37 4.6 1.4 45.0 Na-Cl neutral water (water/rock volume ratio >> 1) with rocks already
H20_8 0.37 6.1 1.8 45.0 showing advanced argillic alteration, and that the rocks collected repre-
sent old and buried solfataric fields (Marini et al., 2003) resulting from
H20_10 0.37 3.4 1.1 45.0
previous volcanic cycles.
H21_3 0.17 90.2 26.6 41.0
H21_8 0.16 65.2 17.4 41.0 We conclude that the observed secondary mineral assemblages (Table 1)
are the result of intense fluid-rock interactions. Advanced argillic alter-
H21_11 0.16 86.9 25.8 41.0
ation of the shallow portion of the edifice was promoted by the efficient
H21_12 0.16 73.8 19.2 41.0
circulation of cooled (below 350°C, and down to 150°C–200°C), acid-
H21_13 0.16 87.2 20.1 41.0 ic (pH < 4) hydrothermal fluids, possibly mixed with meteoric waters.
H22_2 0.12 145.7 29.7 17.2 The alteration assemblages observed at La Soufrière are similar to those
H22_3 0.12 144.3 30.1 17.2 commonly identified at domes and craters of active volcanoes worldwide
H22_4 0.12 139.1 30.5 17.2 (Heap et al., 2019; Yilmaz et al., 2021; Zimbelman et al., 2005), although
the rocks from Mt Unzen volcano (Japan) were overprinted by propyli-
H22_5 0.12 149.7 28.9 17.2
tic alteration (Yilmaz et al., 2021), not observed here. Multiple reaction
H22_6 0.12 133.6 31.1 17.2
paths and variable alteration intensity reflect a variable degree of reaction
H25_2 0.14 99.9 29.3 45.8 progress. For a given alteration rate, reaction progress will increase as
H25_4 0.14 97.7 29.3 45.8 exposure to altering fluids and pathway length increase, or as water flow
H25_5 0.14 94.2 26.2 45.8 and water/rock volume ratio decrease. Variations in these environmental
H25_9 0.14 87.1 29.9 45.8 parameters simply reflect the heterogeneity and structural complexity of
the dome and thus the complex architecture of the hydrothermal system,
H25_12 0.15 83.1 24.0 45.8
as revealed by geological and geophysical data (Brothelande et al., 2014;
H29_2 0.22 53.9 25.4 25.9
Jessop et al., 2021; Moretti et al., 2020; Nicollin et al., 2006; Rosas-Car-
H29_8 0.23 59.1 24.5 25.9 bajal et al., 2016; Tamburello et al., 2019). In particular, the complex 3D
H29_9 0.22 75.2 27.9 25.9 geometry of large faults and fractures in the dome allow for the efficient
H29_12 0.25 53.4 23.5 25.9 drainage of fluids and the compartmentalization of supergene and hypo-
H29_16 0.24 50.7 21.3 25.9
gene alteration. In such an environment: (a) intermediate to advanced
argillic alteration will develop preferentially along discontinuities favor-
WP1285_2 0.13 78.5 22.8 13.4
ing fluid circulation, leading to the formation of low-strength layers dom-
WP1285_8 0.13 63.0 25.7 13.4 inated by clay and sulfate minerals; (b) partial edifice collapse or erosion
WP1285_10 0.13 78.3 21.4 13.4 favored in these weakened zones will expose pyrite-rich zones of the hy-
WP1285_11 0.12 87.2 21.1 13.4 drothermal system below the water table and promote the deepening of
WP1285_15 0.12 84.6 25.6 13.4 supergene oxidation and the alteration aureole (as seen in the rain-trig-
gered landslide in 2009; Rosas-Carbajal et al., 2016); and (c) H2S will be
WP1317_2 0.16 67.3 20.4 16.3
easily released by deeper, boiling fluids or by the disproportionation of
WP1317_6 0.16 84.5 24.7 16.3
magmatic SO2 to H2S and aqueous sulfate during the condensation of
WP1317_7 0.15 91.5 22.6 16.3 magmatic vapor plume at intermediate depths.

HEAP ET AL. 10 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 2 5.2. The Influence of Porosity and Alteration on Uniaxial


Continued Compressive Strength
Uniaxial Young's Percentage New mechanical data show that the uniaxial compressive strength and
Connected compressive modulus of secondary Young's modulus of variably altered andesites from La Soufrière both de-
Sample porosity strength (MPa) (GPa) minerals
crease as a function of increasing porosity (Figure 3), in accordance with
WP1317_8 0.15 95.8 25.9 16.3 previous studies on volcanic rocks (Heap, Xu, et al., 2014; Heap, Ville-
WP1317_12 0.15 104.0 24.7 16.3 neuve, et al., 2020; Heap & Violay, 2021; Mordensky et al., 2018; Schaefer
Note. Sample locations are provided in Figure 1b (site descriptions and
et al., 2015; Zhu et al., 2016). We compare our new data with those previ-
GPS coordinates are available in the Supporting Information). ously published for andesites in Figures 5a and 5b, which show that the
data for the rocks from La Soufrière are in broad agreement with those
previously published. This comparison shows that the andesites form-
ing La Soufrière are generally more porous, and therefore weaker (Fig-
ure 5a) and softer (Figure 5b), than andesites from other stratovolcanoes (the block with the lowest porosity,
H22, has a porosity of 0.12). Indeed, muon tomography data have shown that the material forming the
dome at La Soufrière is typically low-density (typically ≤1,500 kg.m−3; Lesparre et al., 2012; Rosas-Carbajal
et al., 2017) and, therefore, likely comprises high-porosity rocks. However, we note that the relative abun-
dance of high-porosity rocks with a low strength and a low Young's modulus from La Soufrière may be a
consequence of sampling only 17 blocks. In other words, low-porosity rocks may exist at La Soufrière and,
if many more blocks were sampled, maybe the data would encompass a similar range to the compiled data
set for other volcanoes.

Of interest to the goal of this study is whether hydrothermal alteration has influenced the compressive
strength and Young's modulus of the materials collected. Our triaxial data highlight that alteration (per-
cent of secondary minerals) reduces rock cohesion, but does not systematically influence the angle of in-
ternal friction (Figure 4). Experimental studies have shown that alteration can influence the strength of
volcanic rock (del Potro & Hürlimann, 2009; Farquharson et al., 2019; Frolova et al., 2014; Heap, Gravley,
et al., 2020; Heap, Wadsworth, et al., 2021, Mordensky et al., 2018, 2019; Wyering et al., 2014). The type of
alteration, porosity increasing dissolution and alteration to clay (del Potro & Hürlimann, 2009; Farquhar-
son et al., 2019; Opfergelt et al., 2006; Watters & Delahaut, 1995) or porosity decreasing mineral precipi-
tation (Heap, Gravley, et al., 2020; Heap, Baumann, et al., 2021; Heap, Wadsworth, et al., 2021), dictates
whether the alteration decreases or increases the strength and Young's modulus of the rock. Figures 5c
and 5d show plots of uniaxial compressive strength and Young's modulus, respectively, as a function of the
percentage of secondary minerals for the rocks from La Soufrière. The symbols in Figures 5c and 5d are
color-coded to show their porosity: red colors indicate low porosity and yellow colors indicate high porosity.
Although there are two notable outliers, Figures 5c and 5d, respectively, show that uniaxial compressive

pressive strength and Young's modulus are reduced from ∼60–140 to <10 MPa and from 20–45 to <5 GPa,
strength and Young's modulus decrease as a function of increasing hydrothermal alteration. Uniaxial com-

respectively, as alteration increases from ∼16–18 to ∼75% (Figures 5c and 5d). The low uniaxial compres-
sive strengths and Young's moduli of the two relatively unaltered blocks that form the most obvious outliers
(H14 and H20) can be explained by meso-scale fractures found within sample H14 (sample photographs are
provided in the Supporting Information) and the high porosity of sample H20, respectively. Although the
data in Figures 5c and 5d suggest that an increase in alteration results in a decrease in strength and Young's
modulus, it is difficult to untangle the influence of porosity. In other words, are the weak samples weak
because of alteration or are the more porous samples simply more altered because of their higher surface
area available for rock-fluid interactions? Indeed, our data do show that, in general, alteration increases
as a function of increasing porosity. Although laboratory-controlled hydrothermal alteration experiments
(Farquharson et al., 2019) are required to definitively answer this question, we contend here that the ob-
served alteration served to reduce the strength and Young's modulus of these andesites from La Soufrière,
regardless of their porosity, due to the presence of abundant clay minerals, phases considered to reduce the
strength of volcanic rocks (del Potro & Hürlimann, 2009; Nicolas et al., 2020; Opfergelt et al., 2006; Watters
& Delahaut, 1995).

HEAP ET AL. 11 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Figure 4. Stress-strain curves for andesites from La Soufrière de Guadeloupe (Eastern Caribbean, France) deformed
under different effective pressures (given by the number adjacent to each curve). (a) Relatively unaltered sample H18.
(b) Moderately altered sample H3. (c) Very altered sample H2B. Differential stress is the axial stress minus the confining
pressure. Insets show photographs of the 20 mm-diameter cylindrical samples. The values of cohesion and internal
friction angle, calculated using these data, are also provided for each of the three rocks.

HEAP ET AL. 12 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 3
Experimental Summary for the Triaxial Deformation Experiments (and Wet Uniaxial Experiments) Performed
on Relatively Unaltered (H18), Moderately Altered (H3), and Very Altered (H2B) Andesites From La Soufrière de
Guadeloupe (Eastern Caribbean, France)
Connected Percentage of Pore fluid Confining Effective Peak differential
Sample porosity secondary minerals pressure (MPa) pressure (MPa) pressure (MPa) stress (MPa)
H18_11 0.13 15.2 0 (wet) 0 0 65.7
H18_10 0.14 15.2 10 12.5 2.5 97.7
H18_2 0.12 15.2 10 15 5 104.0
H18_8 0.12 15.2 10 20 10 135.2
H3_14 0.16 35.2 0 (wet) 0 0 31.5
H3_12 0.15 35.2 10 12.5 2.5 31.7
H3_4 0.15 35.2 10 15 5 39.4
H3_5 0.16 35.2 10 17.5 7.5 54.9
H2B_7 0.44 74.6 0 (wet) 0 0 2.0
H2B_8 0.44 74.6 10 10.5 0.5 3.8
H2B_6 0.39 74.6 10 11 1 -
H2B_5 0.38 74.6 10 12 2 -
Note. Sample locations are provided in Figure 1b (site descriptions and GPS coordinates are available in the Supporting
Information) and microstructural images are provided in Figure 2.

5.3. Influence of Alteration on the Stability of La Soufrière de Guadeloupe: Large-Scale


Numerical Modeling

We used the open-source hydro-thermo-mechanical modeling code Lithosphere and Mantle Evolution
Model (LaMEM; https://bitbucket.org/bkaus/lamem/src/master/; Kaus et al., 2016), which models the
non-linear, visco-elastoplastic deformation of rocks, to investigate the influence of upscaling and alteration
on the stability of La Soufrière. This model has recently been used to model large-scale volcano deforma-
tion (Heap, Baumann, et al., 2021). Mathematically, the model solves a coupled system of conservation
equations, the conservation equations of momentum and mass. The various deformation mechanisms are
connected in an additive constitutive relationship, where the total strain rate is the sum of the individual
strain rates of elastic, viscous, and plastic deformation. LaMEM uses the Drucker-Prager failure criterion
to determine the magnitude of the plastic strain rate. Numerically, LaMEM uses a staggered-grid finite dif-
ference method for discretization and a marker-and-cell method to assign and track rock properties within
an Eulerian advection framework (Harlow & Welch, 1965). This approach enables the model to attain the
large deformations required in, for example, salt tectonics, continental collision, and magmatic systems
(Baumann et al., 2014; Pusok & Kaus, 2015; Reuber et al., 2018). The ability to model large deformations
is seen here as an advantage of LaMEM over commonly used commercial packages such as FLAC (by Itas-
ca Consulting Group), RS2 (by Rocscience), and PLAXIS 2D (by Virtuosity). We refer the reader to Kaus
et al. (2016) and the Supporting Information for a more detailed description of LaMEM.

Models were run on simplified 2D North-South cross-sections of the volcano (643 km Easting), built using
the geomIO tool (Bauville & Baumann, 2019). The model domain was first separated into deformable and
non-deformable domains, guided by geological observations and the recent electrical survey of Rosas-Car-
bajal et al. (2016). The models have a dimension of 2,000 × 800 m, a model resolution of 4 × 3 m, free-slip
boundary conditions at the sides, and a free surface at the top. All our models assumed an entirely liquid
saturated volcano. The average topographic height was used as the reference surface for determining the
hydrostatic pressure, ph , using p f ph  i  plith  ph , where p f and plith are the pore fluid and lithostatic
pressure, respectively, and i is the pore fluid pressure ratio. Following Heap, Baumann, et al. (2021), we
assumed a value of 0.2 for i (Table 4). To account for the uncertainty in i, additional models using i = 0.4
are provided in the Supplementary Information. We highlight that, although the assumption of complete
volcano saturation is an oversimplification (Ball et al., 2015, 2018; Hurwitz et al., 2003), the main goal of

HEAP ET AL. 13 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Figure 5. (a) Uniaxial compressive strength and (b) Young's modulus as a function of porosity for andesites from La Soufrière de Guadeloupe (Eastern
Caribbean, France) and from volcanoes worldwide (data from: Harnett et al., 2019; Heap, Lavallée, et al., 2014; Heap, Kennedy, et al., 2015; Heap & Violay, 2021;
Karaman & Kesimal, 2015; Kennedy et al., 2020; Mordensky et al., 2018; Siratovich et al., 2014; Wyering et al., 2014). Uniaxial compressive strength (c) and
Young's modulus (d) as a function of alteration (the percentage of secondary minerals). Symbol color indicates the connected porosity (red colors indicate low
porosity and yellow colors indicate high porosity). Outliers—blocks H14 and H20—are labeled on panels (c) and (d).

our modeling was to understand, all else being equal, whether hydrothermal alteration is a contributing
factor to the deformation observed at La Soufrière. Further, the distribution of liquid-saturated and dry
zones within the volcano is currently poorly constrained. However, using the average topographic height as
the reference surface resulted in a “dry” volcano summit with a pore fluid pressure of zero (see Supporting
Information S1 for further details).

By modifying the segmentation of the deformable domain, we prepared three cross-sections: (a) a homo-
geneous volcano, (b) a volcano containing alteration zones identified by the electrical conductivity data
of Rosas-Carbajal et al. (2016), and (c) a volcano in which we artificially enlarged the highly altered zone
identified by Rosas-Carbajal et al. (2016). Zones within the volcano cross-sections were designated as either
slightly altered (V3), moderately altered (V2), or highly altered (V1) (Figures 6a–6d). Zones V3, V2, and
V1 were assigned the mechanical properties of samples H18, H6, and H2B, respectively, the three blocks
chosen to represent relatively unaltered, moderately altered, and highly altered rock (Figure 2; Table 1). The
laboratory data used in the modeling were: water-saturated bulk density, Young’ modulus, cohesion, and
the angle of internal friction. Poisson's ratio, not measured here, was assumed to be 0.2 (as recommended
by Heap, Villeneuve, et al., 2020). The model parameters for V3, V2, and V1 are provided in Table 4. As

HEAP ET AL. 14 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Table 4 the models are dynamic, we must decide when the models reach a qua-
Laboratory-Scale and Upscaled Input Parameters for the Numerical si-steady state and can be compared (see Baumann & Kaus, 2015). Here,
Modeling Performed Using Hydro-Thermo-Mechanical Modeling Code we stop the models after an initial deformation phase that is dominated
Lithosphere and Mantle Evolution Model by an elastic response. For each of the three modeled scenarios, we ran
V2 V3 one model in which we used the laboratory data and one model in which
V1 (highly (moderately (relatively we used upscaled values of the laboratory data (Table 4).
altered) altered) unaltered)
The laboratory-scale Young's modulus and Poisson's ratio were upscaled
Laboratory-scale input parameters
using empirical relationships for Young's modulus (Hoek & Dieder-
Bulk density (kg/m³) 1,900 2,400 2,500
ichs, 2006) and Poisson's ratio (Vásárhelyi, 2009) that account for rock-
Young's modulus (GPa) 0.4 9.6 20.3 mass structure using the Geological Strength Index (GSI), as described in
Poisson's ratio 0.2 0.2 0.2 Heap, Villeneuve, et al. (2020). Following Heap, Villeneuve, et al. (2020),
Cohesion (MPa) 0.4 7.1 14.8 GSI was assumed to be 55 and the damage parameter in the Hoek-Died-
Internal friction angle (°) 42.5 37.7 46.7
erichs equation was considered zero. The cohesion and angle of internal
friction were upscaled by transforming the generalized form of the Hoek-
Pore fluid pressure ratio 0.2 0.2 0.2
Brown failure criterion to the Mohr-Coulomb failure criterion (as in Hoek
Upscaled input parameters et al., 2002). We used the mean uniaxial compressive strength of samples
Bulk density (kg/m³) 1,900 2,400 2,500 H18, H6, and H2B for layers V3, V2, and V1, respectively. The empirical
Young's modulus (GPa) 0.2 3.9 8.3 fitting parameter mi required for the generalized Hoek-Brown failure cri-
Poisson's ratio 0.3 0.3 0.3 terion was determined from the triaxial data (Figure 4) using RocData
(Rocscience; https://www.rocscience.com). To do so, we again assumed
Cohesion (MPa) 0.67 1.09 2.02
a GSI of 55 and a damage parameter of zero. Since a depth is required to
Internal friction angle (°) 30.1 33.5 46.4
upscale the cohesion and angle of internal friction, we selected a depth
Pore fluid pressure ratio 0.2 0.2 0.2 half of that between the summit of the volcano and the top of the un-
deformable zone in the model (275 m). The upscaled model parameters
for V3, V2, and V1 are provided in Table 4. A more detailed description
of the upscaling methods can be found in the Supporting Information.

The model results, the velocity magnitude at the surface and the second invariant of the deviatoric strain
rate tensor, which we refer to as the strain rate, within the deformable domain, are presented in Figure 6.
The strain rate allows us to understand whether the deformation is widespread or localized, and the veloc-
ity at the surface can be compared to geophysical data. Figures 6e–6l show model runs using the labora-
tory-scale and upscaled values, respectively. We first highlight that our results emphasize the importance
of using upscaled values for large-scale modeling. Not only are the strain rates and velocities much lower
when using the laboratory-scale values, but the deformation also manifests slightly differently (Figure 6).

To discuss the influence of alteration, we will restrict our discussion to the upscaled models (Figures 6i–6l).
We first note that strain rates and velocities are low in the homogenously altered volcanoes, that deforma-
tion is largely restricted to the surface and the base of southern slope, and that there is little difference in
strain rates and velocities when the alteration is increased from slightly to moderately altered (Figures 6i
and 6j). When the highly altered zone guided by the recent electrical survey of Rosas-Carbajal et al. (2016)

<0.01 to ∼0.05 mm/yr and a large listric, concave-upward sliding surface emerges, which stretches from
is included, the strain rates and velocities increase significantly (Figure 6k). Surface velocities increase from

the top of the dome to near the base of the southern flank (Figure 6k). A localized slip surface at the base
of the deformable domain and two smaller slip surfaces also emerge near the base of the southern flank
(Figure 6k). Therefore, the inclusion of the highly altered zone, the consequence of pervasive alteration
resulting from the circulation of acid hydrothermal fluids (Rosas-Carbajal et al., 2016), has significant-
ly increased volcano deformation rates and therefore decreased the stability of the volcano. As outlined
above, recent manifestations of increased hydrothermal activity, such as fumaroles and hot springs (Jessop
et al., 2021; Moretti et al., 2020), also provide evidence that the measured ground deformation is, in part,

System (GNSS) measurements over ∼20 years show that the dome is sliding at a rate of 0.3–0.7 cm/yr
due to the alteration of the rock forming the dome. Continuous and campaign Global Navigation Satellite

(Moretti et al., 2020). Therefore, although our modeling captures the dynamics of the deformation of the
southern flank, our simplified 2D model setup (Figure 6) underestimates the deformation measured at
the volcano. This underestimation is likely the result of the presence of large-scale faults, discontinuities,

HEAP ET AL. 15 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Figure 6. Results of the large-scale 2D numerical modeling designed to better understand the influence of alteration on the stability of La Soufrière de
Guadeloupe Eastern (Caribbean, France). Panels (a–d) show the model setup, in which white, light gray, gray, and dark gray indicate non-deformable, slightly
altered (V3), moderately altered (V2), and highly altered (V1) zones. Panels (e–h) and (i–l) show the model results for the models run using laboratory-scale and
upscaled mechanical properties, respectively. The color of the line representing the surface of the volcano indicates the surface velocity (mm/yr) and the color
inside the volcano indicates the strain rate (s−1).

extremely weak clay-rich layers, and zones of high pore pressure, not included in our modeling. Indeed, the
strain rate in the southern flank increases when we increase the pore fluid pressure ratio, i, from 0.2 to 0.4
(see additional models presented in the Supporting Information S1). Future models will aim to incorporate
these features.

In a final step, and to further explore how alteration can influence volcano stability, we modeled a scenario
in which the highly altered zone identified by Rosas-Carbajal et al. (2016) was artificially enlarged to incor-

velocities from ∼0.05 to ∼0.1 mm/yr and significantly increases the strain rate within the flank (Figure 6l).
porate the entire southern flank (Figure 6d). Increasing the size of the alteration zone increases the surface

The slip surface that at the base of the deformable domain is now larger and has a higher strain rate, and
the entire southern flank is now characterized by higher strain rates (Figure 6l). In addition, we also ob-
serve that expanding the alteration zone likely increases the potential collapse volume (Figure 6l). Based on
these models, increasing the size of the highly altered zone (i.e., to a size larger than the present-day altered
zone) will significantly increase the risk posed by the volcano. Deformation snapshots from additional mod-
els, presented in the Supplementary Information, show the progression of deformation for the present-day
and increased-alteration scenarios from 5 to 100 kyrs (using upscaled physical and mechanical properties).

HEAP ET AL. 16 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

These additional models also show that increasing the extent of the alter-
ation increases the rate of volcano spreading.

5.4. Three-Dimensional Strength Model for La Soufrière de


Guadeloupe

The results of the previous section highlight the importance of combin-


ing geophysical surveys, laboratory data, and large-scale modeling to
evaluate the influence of alteration on volcanic flank instability. Our 2D
model geometry is based on qualitative interpretations from an electrical
conductivity model to delimit different regions based on their degree of
alteration. A quantitative approach would require a petrophysical rela-
tion to convert the conductivity values to rock strength and cohesion, but
such relations are highly nonunique as electrical conductivity depends
on many factors including clay content, fluid temperature, and compo-
sition. Instead, rock strength is controlled to a first order by porosity, as
shown in Figure 3b, and thus to the bulk rock density.

Here, we combine a 3D model of the bulk density of the present-day lava


dome (Rosas-Carbajal et al., 2017) with an empirical power-law function
relating wet bulk density and uniaxial compressive strength,  p, where
 p  1.326 25  w 9.67 ( p in Pa and w in kg/m3) to create a 3D strength
map of the volcano (Figure 7). The empirical function was specifically
fitted to our new data for La Soufrière (Table 2; Figure 7a). Because the
distribution of liquid-saturated and dry zones within the volcano is cur-
rently poorly constrained, we chose to use the bulk density of our sam-
ples saturated with liquid water, a scenario we consider better represents
the natural state of the volcano than a completely dry volcano. The 3D
bulk density model of Rosas-Carbajal et al. (2017) was obtained by jointly
inverting gravity data and muon data acquired using three simultaneous
muon detectors that scanned the lava dome from different points of view.
Muon tomography is a novel technique, which uses the attenuation ex-
perienced by cosmic muons as they pass through the volcano to estimate
its density (Tanaka et al., 2010). Whereas a single detector can provide 2D
profiles of the average density of the volcano in the different directions
sampled, multiple simultaneous detectors from different points of view
can be used to create a 3D model of the volcano (Bonechi et al., 2020).

Figure 7 highlights that large portions of the dome are characterized by


very low values of strength (<5 MPa). In particular, the southern flank
of the volcano hosts a significant weak zone (Figures 7b and 7c). This
Figure 7. (a) Uniaxial compressive strength as a function of wet bulk low-strength zone is coincident with the location of the main hydrother-
sample density for andesites from La Soufrière de Guadeloupe (Eastern mal reservoir at La Soufrière, as exposed by electrical conductivity meas-
Caribbean, France). Curve is the best-fit empirical power-law function. urements (Lesparre et al., 2012; Rosas-Carbajal et al., 2016). Based on
(b) 3D uniaxial compressive strength map of La Soufrière de Guadeloupe
(a) the influence of hydrothermally altered rock on edifice stability and
(Eastern Caribbean, France). The black line, which runs North-South,
indicates the position of the cross-section in panel (c). (c) North-South spreading rates at La Soufrière revealed by the numerical modeling in the
cross-section through La Soufrière showing the distribution of uniaxial previous section, and (b) the large volume of low-strength rock in the La
compressive strength. Color scale is the Roma (seismic tomography) color Soufrière dome highlighted by our 3D strength model, we conclude that
scale of Fabio Crameri (Crameri et al., 2020). UCS, uniaxial compressive the stability of La Soufrière is compromised by its high degree of rock al-
strength.
teration, and that the rate of volcano spreading and the probability of an
eventual partial edifice collapse may increase solely due to the ongoing
alteration caused by circulating hydrothermal fluids, and may not neces-
sarily require intensified hydrothermal and/or magmatic activity.

HEAP ET AL. 17 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Based on our 3D strength model, we can provide estimates for the volumes of rock within the lava dome
below a certain strength threshold. For example, we estimate that rock with a strength lower than 1 and
10 MPa comprises minimum dome volumes of 2.7 × 106 and 36 × 106 m3, respectively, corresponding to at
least 4% and 55% of the dome above 1,200 m. In other words, more than half of the La Soufrière dome could
be considered to have a low strength. We highlight that these volume and percentage estimates should be
considered conservative due to the positioning of the three telescopes at La Soufrière, which provides better
coverage of the southern flank of the volcano (Rosas-Carbajal et al., 2017). It is likely, therefore, that the
volumes and percentages of weak rock within the dome at La Soufrière are higher than estimated here.

Although the strength map shown in Figure 7 is currently simplistic (for example, it is well known that
strength increases with depth in the brittle field; Heap, Farquharson, et al., 2015, Figure 4), this 3D model
permits the quick identification of zones within the volcano that could promote instability and more exper-
imental data, when available, will allow us to account for depth and other factors that can influence rock
mechanical behavior, such as temperature and saturation state (Heap & Violay, 2021).

6. Conclusions
Hydrothermal alteration is often invoked as a mechanism that greatly contributes to volcano instability
(Reid et al., 2001; van Wyk de Vries et al., 2000). However, owing to the paucity of experimental data and
the difficulty in accurately imaging subsurface alteration zones, the rock properties and the shape and size
of the alteration zones required for large-scale modeling are often poorly constrained or even assumed.
Here we use a combination of laboratory and geophysical data to numerically model the influence of hy-
drothermal alteration on volcano stability, using La Soufrière de Guadeloupe as a case study. Our numerical
modeling suggests that hydrothermal alteration can significantly increase the deformation rate within the
flanks of the volcano, and also increase the potential collapse volume.

The volcano stability modeling and 3D strength map presented herein, firsts for La Soufriere, help to val-
idate the hypothesis forwarded by the numerous geophysical studies performed at the volcano: that the
identified hydrothermal reservoir characterized by altered materials promotes volcano instability that could
result in collapse (Bouligand et al., 2016; Brothelande et al., 2014; Lesparre et al., 2012; Nicollin et al., 2006;
Rosas-Carbajal et al., 2016, 2017). As discussed above, eventual partial edifice collapse at La Soufrière may
not necessarily require intensified hydrothermal and/or magmatic activity. Not only would the debris ava-
lanche resulting from partial or major dome collapse threaten the inhabited areas surrounding the volcano
(Le Friant et al., 2006; Peruzzetto et al., 2019), but collapse could also depressurize the hydrothermal system
leading to steam-driven eruptions, dangerous laterally directed explosions, and associated high energy py-
roclastic density currents, similar to the activity during the 1976–1977 volcanic crisis (Feuillard et al., 1983;
Komorowski et al., 2005). We also highlight that the fragmentation threshold (the overpressure required for
complete fragmentation upon depressurization), the proportion of fines produced as a result of fragmenta-
tion, and the ejection speed are also influenced by hydrothermal alteration (Mayer et al., 2015, 2016, 2017;
Montanaro et al., 2016), reinforcing the need for a multidisciplinary approach to unravel how hydrothermal
alteration affects volcanic hazards (see also de Moor et al., 2019; Mick et al., 2021).

Our modeling and 3D strength map therefore highlight the importance of monitoring the extent and evo-
lution of hydrothermal alteration, and the formation of low-strength altered layers, at La Soufrière and
at other active volcanoes worldwide. Such monitoring can be achieved, for example, using geophysical
methods such as electrical tomography (Ahmed et al., 2018; Byrdina et al., 2017; Ghorbani et al., 2018; Ro-
sas-Carbajal et al., 2016) and muon tomography (Lesparre et al., 2012; Rosas-Carbajal et al., 2017), near-sur-
face seismic imaging (Amoroso et al., 2018), thermal and gas monitoring (de Moor et al., 2019; Edmonds
et al., 2003; Jessop et al., 2021; Moretti et al., 2020; Tamburello et al., 2019), geological mapping (van Wyk de
Vries et al., 2000), deformation monitoring (Moretti et al., 2020), magnetic methods (Finn et al., 2007), and
remote and/or ground-based optical and spectroscopic methods (Crowley & Zimbelman, 1997; Darmawan
et al., 2018; John et al., 2008; Kereszturi et al., 2020; Mueller et al., 2021). Further, although we document
a reduction in strength as a function of alteration (Figure 5c), alteration can also increase strength and
potentially promote volcanic instability by creating zones of high pore fluid pressure (Heap, Baumann,
et al., 2021; Reid, 2004). As noted in Heap, Baumann, et al. (2021), it is important not only to map the extent

HEAP ET AL. 18 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

and evolution of hydrothermal alteration, but also the locations in the volcano prone to porosity increasing
and porosity decreasing alteration.

We have also shown that microtextural and mineralogical analysis of host-rock material erupted during
phreatic explosive eruptions, as well as material from landslides and debris-avalanche deposits, can provide
highly valuable insights into the nature, extent, and dynamics of alteration processes and fluid circulation
in the pre-collapse edifice (see also Boudon et al., 1998; Komorowski et al., 2010; Salaün et al., 2011). Hence,
we highlight that it is also important to quantify the mineral content of new material exposed by landslides
and collapses, as well as their deposits, and to continuously monitor fluid compositions and water table
fluctuations (Hurwitz et al., 2003) with respect to the geometry of weakened low-strength structures that
are characteristic of the internal framework volcanic edifices hosting hydrothermal systems. A better un-
derstanding of the spatio-temporal evolution of hydrothermal alteration, and how it can influence volcano
stability using large-scale modeling, will help improve the mitigation of volcanic hazards at active volcanoes
worldwide.

Data Availability Statement


All the mechanical data collected for this study can be downloaded here: https://doi.org/10.6084/m9.
figshare.14701137.v1.

Acknowledgments References
This work was supported by the TelluS
Program of INSU-CNRS (“Assessing Ahmed, A. S., Revil, A., Byrdina, S., Coperey, A., Gailler, L., Grobbe, N., et al. (2018). 3D electrical conductivity tomography of volcanoes.
the role of hydrothermal alteration Journal of Volcanology and Geothermal Research, 356, 243–263. https://doi.org/10.1016/j.jvolgeores.2018.03.017
on volcanic hazards”) and ANR grant Amoroso, O., Festa, G., Bruno, P. P., D'Auria, L., De Landro, G., Di Fiore, V., et al. (2018). Integrated tomographic methods for seismic
MYGALE (“Modeling the phYsical and imaging and monitoring of volcanic caldera structures and geothermal areas. Journal of Applied Geophysics, 156, 16–30. https://doi.
chemical Gradients of hydrothermal org/10.1016/j.jappgeo.2017.11.012
ALteration for warning systems of Ball, J. L., Calder, E. S., Hubbard, B. E., & Bernstein, M. L. (2013). An assessment of hydrothermal alteration in the Santiaguito lava
flank collapse at Explosive volcanoes”), dome complex, Guatemala: Implications for dome collapse hazards. Bulletin of Volcanology, 75(1), 1–18. https://doi.org/10.1007/
awarded to the first author. M. Heap s00445-012-0676-z
also acknowledges support from the Ball, J. L., Stauffer, P. H., Calder, E. S., & Valentine, G. A. (2015). The hydrothermal alteration of cooling lava domes. Bulletin of Volcanology,
Institut Universitaire de France (IUF). 77(12), 1–16. https://doi.org/10.1007/s00445-015-0986-z
We thank Tomaso Esposti Ongaro, Ball, J. L., Taron, J., Reid, M. E., Hurwitz, S., Finn, C., & Bedrosian, P. (2018). Combining multiphase groundwater flow and slope stability
Gilles Morvan, Christophe Nevado, models to assess stratovolcano flank collapse in the Cascade Range. Journal of Geophysical Research: Solid Earth, 123(4), 2787–2805.
Marie Cherrier, Océane Rocher, and https://doi.org/10.1002/2017jb015156
Lisa Tomasetto. We thank the IPGP for Baumann, T. S., & Kaus, B. J. (2015). Geodynamic inversion to constrain the non-linear rheology of the lithosphere. Geophysical Journal
general funding for the Observatoires International, 202(2), 1289–1316. https://doi.org/10.1093/gji/ggv201
Volcanologiques et Sismologiques Baumann, T. S., Kaus, B. J., & Popov, A. A. (2014). Constraining effective rheology through parallel joint geodynamic inversion. Tectono-
(OVS), INSU-CNRS for the funding physics, 631, 197–211. https://doi.org/10.1016/j.tecto.2014.04.037
provided by the Service National Bauville, A., & Baumann, T. S. (2019). geomIO: An open-source MATLAB toolbox to create the initial configuration of 2-D/3-D ther-
d'Observation en Volcanologie (SNOV), mo-mechanical simulations from 2-D vector drawings. Geochemistry, Geophysics, Geosystems, 20(3), 1665–1675. https://doi.
and the Ministère pour la Transition org/10.1029/2018gc008057
Ecologique (MTE) for financial support Bonechi, L., D'Alessandro, R., & Giammanco, A. (2020). Atmospheric muons as an imaging tool. Reviews in Physics, 5, 100038. https://doi.
for the monitoring of the instable org/10.1016/j.revip.2020.100038
flank of La Soufrière de Guadeloupe. Borgia, A., Delaney, P. T., & Denlinger, R. P. (2000). Spreading volcanoes. Annual Review of Earth and Planetary Sciences, 28(1), 539–570.
This study contributes to the IdEx https://doi.org/10.1146/annurev.earth.28.1.539
Université de Paris ANR-18-IDEX-0001. Boudon, G., Komorowski, J.-C., Villemant, B., & Semet, M. P. (2008). A new scenario for the last magmatic eruption of La Soufrière of
Olivier Roche, Jessica Ball, and Gabor Guadeloupe (Lesser Antilles) in 1530 AD Evidence from stratigraphy radiocarbon dating and magmatic evolution of erupted products.
Kereszturi are thanked for providing Journal of Volcanology and Geothermal Research, 178(3), 474–490. https://doi.org/10.1016/j.jvolgeores.2008.03.006
comments that helped improve this Boudon, G., Semet, M. P., & Vincent, P. M. (1987). Magma and hydrothermally driven sector collapses: The 3100 and 11,500 y. BP eruptions
manuscript. of la Grande Decouverte (la Soufriere) volcano, Guadeloupe, French West Indies. Journal of Volcanology and Geothermal Research,
33(4), 317–323. https://doi.org/10.1016/0377-0273(87)90021-7
Boudon, G., Villemant, B., Komorowski, J.-C., Ildefonse, P., & Semet, M. P. (1998). The hydrothermal system at Soufriere Hills volcano,
Montserrat (West Indies): Characterization and role in the ongoing eruption. Geophysical Research Letters, 25, 3693–3696. https://doi.
org/10.1029/98gl00985
Bouligand, C., Coutant, O., & Glen, J. M. (2016). Sub-surface structure of La Soufrière of Guadeloupe lava dome deduced from a ground-
based magnetic survey. Journal of Volcanology and Geothermal Research, 321, 171–181. https://doi.org/10.1016/j.jvolgeores.2016.04.037
Brombach, T., Marini, L., & Hunziker, J. C. (2000). Geochemistry of the thermal springs and fumaroles of Basse-Terre Island, Guadeloupe,
Lesser Antilles. Bulletin of Volcanology, 61(7), 477–490. https://doi.org/10.1007/pl00008913
Brothelande, E., Finizola, A., Peltier, A., Delcher, E., Komorowski, J. C., Di Gangi, F., et al. (2014). Fluid circulation pattern inside La Sou-
frière volcano (Guadeloupe) inferred from combined electrical resistivity tomography, self-potential, soil temperature and diffuse de-
gassing measurements. Journal of Volcanology and Geothermal Research, 288, 105–122. https://doi.org/10.1016/j.jvolgeores.2014.10.007
Browne, P. R. L. (1978). Hydrothermal alteration in active geothermal fields. Annual Review of Earth and Planetary Sciences, 6, 229–250.
https://doi.org/10.1146/annurev.ea.06.050178.001305

HEAP ET AL. 19 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Byrdina, S., Friedel, S., Vandemeulebrouck, J., Budi-Santoso, A., Suryanto, W., Rizal, M. H., & Winata, E. (2017). Geophysical image of
the hydrothermal system of Merapi volcano. Journal of Volcanology and Geothermal Research, 329, 30–40. https://doi.org/10.1016/j.
jvolgeores.2016.11.011
Cecchi, E., van Wyk de Vries, B., & Lavest, J. M. (2004). Flank spreading and collapse of weak-cored volcanoes. Bulletin of Volcanology,
67(1), 72–91. https://doi.org/10.1007/s00445-004-0369-3
Cole, P. D., Neri, A., & Baxter, P. J. (2015). Hazards from pyroclastic density currents. In The encyclopedia of volcanoes (pp. 943–956). Aca-
demic Press. https://doi.org/10.1016/b978-0-12-385938-9.00054-7
Crameri, F., Shephard, G. E., & Heron, P. J. (2020). The misuse of colour in science communication. Nature Communications, 11(1), 1–10.
https://doi.org/10.1038/s41467-020-19160-7
Crowley, J. K., & Zimbelman, D. R. (1997). Mapping hydrothermally altered rocks on Mount Rainier, Washington, with airborne visible/infra-
red imaging spectrometer (AVIRIS) data. Geology, 25(6), 559–562. https://doi.org/10.1130/0091-7613(1997)025<0559:mharom>2.3.co;2
Damby, D. E., Llewellin, E. W., Horwell, C. J., Williamson, B. J., Najorka, J., Cressey, G., & Carpenter, M. (2014). The α–β phase transition
in volcanic cristobalite. Journal of Applied Crystallography, 47(4), 1205–1215. https://doi.org/10.1107/s160057671401070x
Darmawan, H., Walter, T. R., Brotopuspito, K. S., & Nandaka, I. G. M. A. (2018). Morphological and structural changes at the Merapi lava
dome monitored in 2012–15 using unmanned aerial vehicles (UAVs). Journal of Volcanology and Geothermal Research, 349, 256–267.
https://doi.org/10.1016/j.jvolgeores.2017.11.006
Day, S. J. (1996). Hydrothermal pore fluid pressure and the stability of porous, permeable volcanoes. Geological Society, London, Special
Publications, 110(1), 77–93. https://doi.org/10.1144/gsl.sp.1996.110.01.06
de Moor, J. M., Stix, J., Avard, G., Muller, C., Corrales, E., Diaz, J. A., et al. (2019). Insights on hydrothermal-magmatic interactions and
eruptive processes at Poás Volcano (Costa Rica) from high-frequency gas monitoring and drone measurements. Geophysical Research
Letters, 46(3), 1293–1302. https://doi.org/10.1029/2018gl080301
del Potro, R., & Hürlimann, M. (2009). The decrease in the shear strength of volcanic materials with argillic hydrothermal alteration,
insights from the summit region of Teide stratovolcano, Tenerife. Engineering Geology, 104(1–2), 135–143. https://doi.org/10.1016/j.
enggeo.2008.09.005
Deyell, C. L., & Dipple, G. M. (2005). Equilibrium mineral–fluid calculations and their application to the solid solution between alunite
and natroalunite in the El Indio–Pascua belt of Chile and Argentina. Chemical Geology, 215(1–4), 219–234. https://doi.org/10.1016/j.
chemgeo.2004.06.039
Edmonds, M., Oppenheimer, C., Pyle, D. M., Herd, R. A., & Thompson, G. (2003). SO2 emissions from Soufrière Hills Volcano and their
relationship to conduit permeability, hydrothermal interaction and degassing regime. Journal of Volcanology and Geothermal Research,
124(1–2), 23–43. https://doi.org/10.1016/s0377-0273(03)00041-6
Ellis, A. J., & Mahon, W. A. J. (1977). Chemistry and geothermal systems. (No. 553.79 E4).
Farquharson, J. I., Wild, B., Kushnir, A. R., Heap, M. J., Baud, P., & Kennedy, B. (2019). Acid-induced dissolution of andesite: Evolution of
permeability and strength. Journal of Geophysical Research: Solid Earth, 124(1), 257–273. https://doi.org/10.1029/2018jb016130
Feuillard, M., Allegre, C. J., Brandeis, G., Gaulon, R., Le Mouel, J. L., Mercier, J. C., et al. (1983). The 1975–1977 crisis of La Soufrière de
Guadeloupe (FWI): A still-born magmatic eruption. Journal of Volcanology and Geothermal Research, 16(3–4), 317–334. https://doi.
org/10.1016/0377-0273(83)90036-7
Feuillet, N., Beauducel, F., Jacques, E., Tapponnier, P., Delouis, B., Bazin, S., et al. (2011). The Mw = 6.3, November 21, 2004, Les Saintes
earthquake (Guadeloupe): Tectonic setting, slip model and static stress changes. Journal of Geophysical Research, 116, B10301. https://
doi.org/10.1029/2011JB008310
Finn, C. A., Deszcz-Pan, M., Anderson, E. D., & John, D. A. (2007). Three-dimensional geophysical mapping of rock alteration and
water content at Mount Adams, Washington: Implications for lahar hazards. Journal of Geophysical Research, 112(B10). https://doi.
org/10.1029/2006jb004783
Finn, C. A., Deszcz-Pan, M., Ball, J. L., Bloss, B. J., & Minsley, B. J. (2018). Three-dimensional geophysical mapping of shallow water satu-
rated altered rocks at Mount Baker, Washington: Implications for slope stability. Journal of Volcanology and Geothermal Research, 357,
261–275. https://doi.org/10.1016/j.jvolgeores.2018.04.013
Frolova, J., Ladygin, V., Rychagov, S., & Zukhubaya, D. (2014). Effects of hydrothermal alterations on physical and mechanical properties
of rocks in the Kuril–Kamchatka island arc. Engineering Geology, 183, 80–95.
Fulignati, P. (2020). Clay minerals in hydrothermal systems. Minerals, 10(10), 919.
Ghorbani, A., Revil, A., Coperey, A., Ahmed, A. S., Roque, S., Heap, M. J., et al. (2018). Complex conductivity of volcanic rocks and the ge-
ophysical mapping of alteration in volcanoes. Journal of Volcanology and Geothermal Research, 357, 106–127. https://doi.org/10.1016/j.
jvolgeores.2018.04.014
Girona, T., Realmuto, V., & Lundgren, P. (2021). Large-scale thermal unrest of volcanoes for years prior to eruption. Nature Geoscience,
14(4), 238–241. https://doi.org/10.1038/s41561-021-00705-4
Glicken, H. (1996). Rockslide-debris avalanche of May 18, 1980, Mount St. Helens volcano. (No. 96-677). US Geological Survey.
Harlow, F. H., & Welch, J. E. (1965). Numerical calculation of time-dependent viscous incompressible flow of fluid with free surface. The
Physics of Fluids, 8(12), 2182–2189. https://doi.org/10.1063/1.1761178
Harnett, C. E., Kendrick, J. E., Lamur, A., Thomas, M. E., Stinton, A., Wallace, P. A., et al. (2019). Evolution of mechanical properties
of lava dome rocks across the 1995–2010 eruption of Soufrière Hills volcano, Montserrat. Frontiers of Earth Science, 7, 7. https://doi.
org/10.3389/feart.2019.00007
Heap, M. J., Baumann, T., Gilg, H. A., Kolzenburg, S., Ryan, A., Villeneuve, M., et al. (2021). Hydrothermal alteration can result in pore
pressurization and volcano instability. Geology, 49. https://doi.org/10.1130/G49063.1
Heap, M. J., Farquharson, J. I., Baud, P., Lavallée, Y., & Reuschlé, T. (2015). Fracture and compaction of andesite in a volcanic edifice.
Bulletin of Volcanology, 77(6), 55. https://doi.org/10.1007/s00445-015-0938-7
Heap, M. J., Gravley, D. M., Kennedy, B. M., Gilg, H. A., Bertolett, E., & Barker, S. L. (2020). Quantifying the role of hydrothermal alteration
in creating geothermal and epithermal mineral resources: The Ohakuri ignimbrite (Taupō Volcanic Zone, New Zealand). Journal of
Volcanology and Geothermal Research, 390, 106703. https://doi.org/10.1016/j.jvolgeores.2019.106703
Heap, M. J., Kennedy, B. M., Pernin, N., Jacquemard, L., Baud, P., Farquharson, J. I., et al. (2015). Mechanical behaviour and failure modes
in the Whakaari (White Island volcano) hydrothermal system, New Zealand. Journal of Volcanology and Geothermal Research, 295,
26–42. https://doi.org/10.1016/j.jvolgeores.2015.02.012
Heap, M. J., Lavallée, Y., Petrakova, L., Baud, P., Reuschlé, T., Varley, N. R., & Dingwell, D. B. (2014). Microstructural controls on the phys-
ical and mechanical properties of edifice-forming andesites at Volcán de Colima, Mexico. Journal of Geophysical Research: Solid Earth,
119(4), 2925–2963. https://doi.org/10.1002/2013jb010521

HEAP ET AL. 20 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Heap, M. J., Troll, V. R., Kushnir, A. R., Gilg, H. A., Collinson, A. S., Deegan, F. M., et al. (2019). Hydrothermal alteration of andesitic
lava domes can lead to explosive volcanic behaviour. Nature Communications, 10(1), 1–10. https://doi.org/10.1038/s41467-019-13102-8
Heap, M. J., Villeneuve, M., Albino, F., Farquharson, J. I., Brothelande, E., Amelung, F., et al. (2020). Towards more realistic values
of elastic moduli for volcano modelling. Journal of Volcanology and Geothermal Research, 390, 106684. https://doi.org/10.1016/j.
jvolgeores.2019.106684
Heap, M. J., & Violay, M. E. (2021). The mechanical behaviour and failure modes of volcanic rocks: A review. Bulletin of Volcanology, 83(5),
1–47. https://doi.org/10.1007/s00445-021-01447-2
Heap, M. J., Wadsworth, F. B., Heng, Z., Xu, T., Griffiths, L., Velasco, A. A., et al. (2021). The tensile strength of volcanic rocks: Experiments
and models. Journal of Volcanology and Geothermal Research, 418, 107348. https://doi.org/10.1016/j.jvolgeores.2021.107348
Heap, M. J., Xu, T., & Chen, C.-F. (2014). The influence of porosity and vesicle size on the brittle strength of volcanic rocks and magma.
Bulletin of Volcanology, 76(9), 856. https://doi.org/10.1007/s00445-014-0856-0
Hemley, J. J., Hostetler, P. B., Gude, A. J., & Mountjoy, W. T. (1969). Some stability relations of alunite. Economic Geology, 64(6), 599–612.
https://doi.org/10.2113/gsecongeo.64.6.599
Henley, R. W., & Berger, B. R. (2011). Magmatic-vapor expansion and the formation of high-sulfidation gold deposits: Chemical controls on
alteration and mineralization. Ore Geology Reviews, 39(1–2), 63–74. https://doi.org/10.1016/j.oregeorev.2010.11.003
Hincks, T. K., Komorowski, J. C., Sparks, S. R., & Aspinall, W. P. (2014). Retrospective analysis of uncertain eruption precursors at La
Soufrière volcano, Guadeloupe, 1975–77: Volcanic hazard assessment using a Bayesian Belief Network approach. Journal of Applied
Volcanology, 3(1), 3. https://doi.org/10.1186/2191-5040-3-3
Hoek, E., & Diederichs, M. S. (2006). Empirical estimation of rock mass modulus. International Journal of Rock Mechanics and Mining
Sciences, 43(2), 203–215. https://doi.org/10.1016/j.ijrmms.2005.06.005
Horwell, C. J., Williamson, B. J., Llewellin, E. W., Damby, D. E., & Le Blond, J. S. (2013). The nature and formation of cristobalite at the
Soufrière Hills volcano, Montserrat: Implications for the petrology and stability of silicic lava domes. Bulletin of Volcanology, 75(3), 696.
https://doi.org/10.1007/s00445-013-0696-3
Hurwitz, S., Kipp, K. L., Ingebritsen, S. E., & Reid, M. E. (2003). Groundwater flow, heat transport, and water table position within
volcanic edifices: Implications for volcanic processes in the Cascade Range. Journal of Geophysical Research, 108(B12). https://doi.
org/10.1029/2003jb002565
Inoue, A. (1995). Formation of clay minerals in hydrothermal environments. In Origin and mineralogy of clays (pp. 268–329). Springer.
https://doi.org/10.1007/978-3-662-12648-6_7
Jessop, D. E., Moune, S., Moretti, R., Gibert, D., Komorowski, J. C., Robert, V., et al. (2021). A multi-decadal view of the heat and mass
budget of a volcano in unrest: La Soufrière de Guadeloupe (French West Indies). Bulletin of Volcanology, 83(3), 1–19. https://doi.
org/10.1007/s00445-021-01439-2
John, D. A., Sisson, T. W., Breit, G. N., Rye, R. O., & Vallance, J. W. (2008). Characteristics, extent and origin of hydrothermal alteration
at Mount Rainier Volcano, Cascades Arc, USA: Implications for debris-flow hazards and mineral deposits. Journal of Volcanology and
Geothermal Research, 175(3), 289–314. https://doi.org/10.1016/j.jvolgeores.2008.04.004
Karaman, K., & Kesimal, A. (2015). Evaluation of the influence of porosity on the engineering properties of volcanic rocks from the East-
ern Black Sea Region: NE Turkey. Arabian Journal of Geosciences, 8(1), 557–564. https://doi.org/10.1007/s12517-013-1217-6
Kaus, B., Popov, A. A., Baumann, T., Pusok, A., Bauville, A., Fernandez, N., & Collignon, M. (2016). Forward and inverse modelling of
lithospheric deformation on geological timescales. In K. Binder, M. Müller, M. Kremer, & A. Schnurpfeil (Eds.), Proceedings of NIC
Symposium (Vol. 48). John von Neumann Institute for Computing (NIC), NIC Series.
Kennedy, B. M., Farquhar, A., Hilderman, R., Villeneuve, M. C., Heap, M. J., Mordensky, S., et al. (2020). Pressure controlled permeability
in a conduit filled with fractured hydrothermal breccia reconstructed from ballistics from Whakaari (White Island), New Zealand. Geo-
sciences, 10(4), 138. https://doi.org/10.3390/geosciences10040138
Kereszturi, G., Schaefer, L., Mead, S., Miller, C., Procter, J., & Kennedy, B. (2021). Synthesis of hydrothermal alteration, rock mechanics and
geophysical mapping to constrain failure and debris avalanche hazards at Mt. Ruapehu (New Zealand). New Zealand Journal of Geology
and Geophysics, 64, 1–22. https://doi.org/10.1080/00288306.2021.1885048
Kereszturi, G., Schaefer, L. N., Miller, C., & Mead, S. (2020). Hydrothermal alteration on composite volcanoes: Mineralogy, hyperspectral
imaging, and aeromagnetic study of Mt Ruapehu, New Zealand. Geochemistry, Geophysics, Geosystems, 21(9), e2020GC009270. https://
doi.org/10.1029/2020gc009270
Kerle, N., & De Vries, B. V. W. (2001). The 1998 debris avalanche at Casita volcano, Nicaragua—Investigation of structural deformation
as the cause of slope instability using remote sensing. Journal of Volcanology and Geothermal Research, 105(1–2), 49–63. https://doi.
org/10.1016/s0377-0273(00)00244-4
Komorowski, J.-C., Boudon, G., Semet, M., Beauducel, F., Anténor-Habazac, C., Bazin, S., & Hammmouya, G. (2005). Guadeloupe. In J.
Lindsay, R. Robertson, J. Shepherd, & S. Ali (Eds.), Volcanic Atlas of the lesser Antilles (pp. 65–102). University of the Fnrech West Indies,
Seismic Research Unit.
Komorowski, J.-C., Hincks, T., Sparks, R., & Aspinall, W., & CASAVA ANR Project Consortium. (2015). Improving crisis decision-making
at times of uncertain volcanic unrest (Guadeloupe, 1976). In S. C. Loughlin, R. S. J. Sparks, S. K. Brown, S. F. Jenkins, & C. Vye-Brown
(Eds.), Global volcanic hazards and risk (pp. 255–261). Cambridge University Press.
Komorowski, J.-C., Legendre, Y., Christopher, T., Bernstein, L., Stewart, R., Joseph, E., et al. (2010). Insights into processes and deposits
of hazardous vulcanian explosions at Soufrière Hills Volcano during 2008 and 2009 (Montserrat, West Indies). Geophysical Research
Letters, 37, L00E19. https://doi.org/10.1029/2010GL042558
Le Friant, A., Boudon, G., Komorowski, J. C., Heinrich, P., & Semet, M. P. (2006). Potential flank-collapse of Soufriere Volcano, Guade-
loupe, lesser Antilles? Numerical simulation and hazards. Natural Hazards, 39(3), 381–393. https://doi.org/10.1007/s11069-005-6128-8
Lesparre, N., Gibert, D., Marteau, J., Komorowski, J. C., Nicollin, F., & Coutant, O. (2012). Density muon radiography of La Soufriere
of Guadeloupe volcano: Comparison with geological, electrical resistivity and gravity data. Geophysical Journal International, 190(2),
1008–1019. https://doi.org/10.1111/j.1365-246x.2012.05546.x
López, D. L., & Williams, S. N. (1993). Catastrophic volcanic collapse: Relation to hydrothermal processes. Science, 260(5115), 1794–1796.
https://doi.org/10.1126/science.260.5115.1794
Marini, L., Zuccolini, M. V., & Saldi, G. (2003). The bimodal pH distribution of volcanic lake waters. Journal of Volcanology and Geothermal
Research, 121(1–2), 83–98. https://doi.org/10.1016/s0377-0273(02)00413-4
Martel, C., Pichavant, M., Di Carlo, I., Champallier, R., Wille, G., Castro, J. M., et al. (2021). Experimental constraints on the crystallization
of Silica Phases in Silicic Magmas. Journal of Petrology, 62(1), egab004. https://doi.org/10.1093/petrology/egab004

HEAP ET AL. 21 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Mayer, K., Scheu, B., Gilg, H. A., Heap, M. J., Kennedy, B. M., Lavallée, Y., et al. (2015). Experimental constraints on phreatic eruption
processes at Whakaari (White Island volcano). Journal of Volcanology and Geothermal Research, 302, 150–162. https://doi.org/10.1016/j.
jvolgeores.2015.06.014
Mayer, K., Scheu, B., Montanaro, C., Yilmaz, T. I., Isaia, R., Aßbichler, D., & Dingwell, D. B. (2016). Hydrothermal alteration of surficial
rocks at Solfatara (Campi Flegrei): Petrophysical properties and implications for phreatic eruption processes. Journal of Volcanology and
Geothermal Research, 320, 128–143. https://doi.org/10.1016/j.jvolgeores.2016.04.020
Mayer, K., Scheu, B., Yilmaz, T. I., Montanaro, C., Gilg, H. A., Rott, S., et al. (2017). Phreatic activity and hydrothermal alteration in the
Valley of Desolation, Dominica, Lesser Antilles. Bulletin of Volcanology, 79(12), 82. https://doi.org/10.1007/s00445-017-1166-0
McCollom, T. M., Hynek, B. M., Rogers, K., Moskowitz, B., & Berquó, T. S. (2013). Chemical and mineralogical trends during acid-sulfate
alteration of pyroclastic basalt at Cerro Negro volcano and implications for early Mars. Journal of Geophysical Research: Planets, 118,
1719–1751. https://doi.org/10.1002/jgre.20114
Metcalfe, A., Moune, S., Komorowski, J. C., Kilgour, G., Jessop, D. E., Moretti, R., & Legendre, Y. (2021). Magmatic Processes at La Soufrière
de Guadeloupe: Insights from crystal studies and diffusion timescales for eruption onset. Frontiers of Earth Science, 9, 78. https://doi.
org/10.3389/feart.2021.617294
Meyer, C., & Hemley, J. J. (1967). Wall rock alteration. In H. L. Barnes (Ed.), Geochemistry of hydrothermal ore deposits. Holt, Reinhart and
Winston. Inc.
Mick, E., Stix, J., de Moor, J. M., & Avard, G. (2021). Hydrothermal alteration and sealing at Turrialba volcano, Costa Rica, as a mech-
anism for phreatic eruption triggering. Journal of Volcanology and Geothermal Research, 416, 107297. https://doi.org/10.1016/j.
jvolgeores.2021.107297
Montanaro, C., Scheu, B., Mayer, K., Orsi, G., Moretti, R., Isaia, R., & Dingwell, D. B. (2016). Experimental investigations on the explosivity
of steam-driven eruptions: A case study of Solfatara volcano (Campi Flegrei). Journal of Geophysical Research: Solid Earth, 121(11),
7996–8014. https://doi.org/10.1002/2016jb013273
Mordensky, S., Heap, M., Kennedy, B., Gilg, A., Villeneuve, M., Farquharson, J., & Gravley, D. (2019). Influence of alteration on the me-
chanical behaviour and failure mode of andesite: Implications for shallow seismicity and volcano monitoring. Bulletin of Volcanology,
81, 44. https://doi.org/10.1007/s00445-019-1306-9
Mordensky, S. P., Villeneuve, M. C., Kennedy, B. M., Heap, M. J., Gravley, D. M., Farquharson, J. I., & Reuschlé, T. (2018). Physical and
mechanical property relationships of a shallow intrusion and volcanic host rock, Pinnacle Ridge, Mt. Ruapehu, New Zealand. Journal
of Volcanology and Geothermal Research, 359, 1–20. https://doi.org/10.1016/j.jvolgeores.2018.05.020
Moretti, R., Komorowski, J. C., Ucciani, G., Moune, S., Jessop, D., de Chabalier, J. B., et al. (2020). The 2018 unrest phase at La Soufrière of
Guadeloupe (French West Indies) andesitic volcano: Scrutiny of a failed but prodromal phreatic eruption. Journal of Volcanology and
Geothermal Research, 393, 106769. https://doi.org/10.1016/j.jvolgeores.2020.106769
Mueller, D., Bredemeyer, S., Zorn, E., De Paolo, E., & Walter, T. R. (2021). Surveying fumarole sites and hydrothermal alteration by unoccu-
pied aircraft systems (UAS) at the La Fossa cone, Vulcano Island (Italy). Journal of Volcanology and Geothermal Research, 413, 107208.
Nicolas, A., Lévy, L., Sissmann, O., Li, Z., Fortin, J., Gibert, B., & Sigmundsson, F. (2020). Influence of hydrothermal alteration on the
elastic behaviour and failure of heat-treated andesite from Guadeloupe. Geophysical Journal International, 223(3), 2038–2053. https://
doi.org/10.1093/gji/ggaa437
Nicollin, F., Gibert, D., Beauducel, F., Boudon, G., & Komorowski, J. C. (2006). Electrical tomography of La Soufrière of Guadeloupe Vol-
cano: Field experiments, 1D inversion and qualitative interpretation. Earth and Planetary Science Letters, 244(3–4), 709–724. https://
doi.org/10.1016/j.epsl.2006.02.020
Opfergelt, S., Delmelle, P., Boivin, P., & Delvaux, B. (2006). The 1998 debris avalanche at Casita volcano, Nicaragua: Investigation of the
role of hydrothermal smectite in promoting slope instability. Geophysical Research Letters, 33(15). https://doi.org/10.1029/2006gl026661
Peruzzetto, M., Komorowski, J.-C., Le Friant, A., Rosas-Carbajal, M., Mangeney, A., & Legendre, Y. (2019). Modeling of partial dome col-
lapse of La Soufrière of Guadeloupe volcano: Implications for hazard assessment and monitoring. Scientific Reports, 9(1), 1–15. https://
doi.org/10.1038/s41598-019-49507-0
Pusok, A. E., & Kaus, B. J. (2015). Development of topography in 3-D continental-collision models. Geochemistry, Geophysics, Geosystems,
16(5), 1378–1400. https://doi.org/10.1002/2015gc005732
Reid, M. E. (2004). Massive collapse of volcano edifices triggered by hydrothermal pressurization. Geology, 32(5), 373–376. https://doi.
org/10.1130/g20300.1
Reid, M. E., Sisson, T. W., & Brien, D. L. (2001). Volcano collapse promoted by hydrothermal alteration and edifice shape, Mount Rainier,
Washington. Geology, 29(9), 779–782. https://doi.org/10.1130/0091-7613(2001)029<0779:vcpbha>2.0.co;2
Reuber, G. S., Kaus, B. J., Popov, A. A., & Baumann, T. S. (2018). Unraveling the physics of the Yellowstone magmatic system using geody-
namic simulations. Frontiers of Earth Science, 6, 117. https://doi.org/10.3389/feart.2018.00117
Rosas-Carbajal, M., Jourde, K., Marteau, J., Deroussi, S., Komorowski, J. C., & Gibert, D. (2017). Three-dimensional density structure of
La Soufrière de Guadeloupe lava dome from simultaneous muon radiographies and gravity data. Geophysical Research Letters, 44(13),
6743–6751.
Rosas-Carbajal, M., Komorowski, J. C., Nicollin, F., & Gibert, D. (2016). Volcano electrical tomography unveils edifice collapse hazard
linked to hydrothermal system structure and dynamics. Scientific Reports, 6, 29899. https://doi.org/10.1038/srep29899
Roverato, M., Di Traglia, F., Procter, J., Paguican, E., & Dufresne, A. (2021). Factors contributing to volcano lateral collapse. In Volcanic
debris avalanches (pp. 91–119). Springer. https://doi.org/10.1007/978-3-030-57411-6_5
Rye, R. O., Bethke, P. M., & Wasserman, M. D. (1992). The stable isotope geochemistry of acid sulfate alteration. Economic Geology, 87(2),
225–262. https://doi.org/10.2113/gsecongeo.87.2.225
Salaün, A., Villemant, B., Gérard, M., Komorowski, J. C., & Michel, A. (2011). Hydrothermal alteration in andesitic volcanoes: Trace ele-
ment redistribution in active and ancient hydrothermal systems of Guadeloupe (Lesser Antilles). Journal of Geochemical Exploration,
111(3), 59–83. https://doi.org/10.1016/j.gexplo.2011.06.004
Samaniego, P., Valderrama, P., Mariño, J., van Wyk de Vries, B., Roche, O., Manrique, N., et al. (2015). The historical (218±14 aBP) ex-
plosive eruption of Tutupaca volcano (Southern Peru). Bulletin of Volcanology, 77(6), 1–18. https://doi.org/10.1007/s00445-015-0937-8
Schaefer, L. N., Kendrick, J. E., Oommen, T., Lavallée, Y., & Chigna, G. (2015). Geomechanical rock properties of a basaltic volcano. Fron-
tiers of Earth Science, 3, 29. https://doi.org/10.3389/feart.2015.00029
Schipper, C. I., Rickard, W. D., Reddy, S. M., Saxey, D. W., Castro, J. M., Fougerouse, D., et al. (2020). Volcanic SiO2-cristobalite: A nat-
ural product of chemical vapor deposition. American Mineralogist: Journal of Earth and Planetary Materials, 105(4), 510–524.
https://doi.org/10.2138/am-2020-7236

HEAP ET AL. 22 of 23
21699356, 2021, 8, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021JB022514 by Morocco Hinari NPL, Wiley Online Library on [05/12/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Solid Earth 10.1029/2021JB022514

Siebert, L. (1984). Large volcanic debris avalanches: Characteristics of source areas, deposits, and associated eruptions. Journal of Volcan-
ology and Geothermal Research, 22, 163–197. https://doi.org/10.1016/0377-0273(84)90002-7
Siebert, L., Glicken, H., & Ui, T. (1987). Volcanic hazards from Bezymianny- and Bandai-type eruptions. Bulletin of Volcanology, 49, 435–
459. https://doi.org/10.1007/bf01046635
Siebert, L., Simkin, T., & Kimberly, P. (2010). Volcanoes of the world. University of California Press.
Siratovich, P. A., Heap, M. J., Villenueve, M. C., Cole, J. W., & Reuschlé, T. (2014). Physical property relationships of the Rotokawa An-
desite, a significant geothermal reservoir rock in the Taupo Volcanic Zone, New Zealand. Geothermal Energy, 2(1), 1–31. https://doi.
org/10.1186/s40517-014-0010-4
Sparks, R. S. J., Barclay, J., Calder, E. S., Herd, R. A., Komorowski, J.-C., Norton, G. E., et al. (2002). Generation of a debris avalanche and
violent pyroclastic density current: The Boxing Day eruption of 26 December 1997 at the Soufrière Hills Volcano, Montserrat. In T. H.
Druitt, & B. P. Kokelaar (Eds.), The eruption of Soufrière Hills volcano, Montserrat, from 1995 to 1999 (Vol. 21, pp. 409–434). Geological
Society. https://doi.org/10.1144/gsl.mem.2002.021.01.18
Stoffregen, R. E., Alpers, C. N., & Jambor, J. L. (2000). Alunite-jarosite crystallography, thermodynamics, and geochronology. Reviews in
Mineralogy and Geochemistry, 40(1), 453–479. https://doi.org/10.2138/rmg.2000.40.9
Stoffregen, R. E., & Cygan, G. L. (1990). An experimental study of Na-K exchange between alunite and aqueous sulfate solutions. American
Mineralogist, 75(1–2), 209–220.
Tamburello, G., Moune, S., Allard, P., Venugopal, S., Robert, V., Rosas-Carbajal, M., et al. (2019). Spatio-temporal relationships between
fumarolic activity, hydrothermal fluid circulation and geophysical signals at an arc volcano in degassing unrest: La Soufrière of Guade-
loupe (French West Indies). Geosciences, 9(11), 480. https://doi.org/10.3390/geosciences9110480
Tanaka, H. K., Taira, H., Uchida, T., Tanaka, M., Takeo, M., Ohminato, T., et al. (2010). Three-dimensional computational axial tomography
scan of a volcano with cosmic ray muon radiography. Journal of Geophysical Research, 115(B12). https://doi.org/10.1029/2010jb007677
Taran, Y., & Kalacheva, E. (2020). Acid sulfate-chloride volcanic waters; Formation and potential for monitoring of volcanic activity. Jour-
nal of Volcanology and Geothermal Research, 405, 107036. https://doi.org/10.1016/j.jvolgeores.2020.107036
van Wyk de Vries, B., & Francis, P. W. (1997). Catastrophic collapse at stratovolcanoes induced by gradual volcano spreading. Nature,
387(6631), 387–390. https://doi.org/10.1038/387387a0
van Wyk de Vries, B., Kerle, N., & Petley, D. (2000). Sector collapse forming at Casita volcano, Nicaragua. Geology, 28(2), 167–170. https://
doi.org/10.1130/0091-7613(2000)028<0167:scfacv>2.3.co;2
Vásárhelyi, B. (2009). A possible method for estimating the Poisson's rate values of the rock masses. Acta Geodaetica et Geophysica Hunga-
rica, 44(3), 313–322. https://doi.org/10.1556/ageod.44.2009.3.4
Villemant, B., Hammouya, G., Michel, A., Semet, M., Komorowski, J.-C., Boudon, G., & Cheminée, J.-L. (2005). The memory of volcanic
waters: Shallow magma degassing revealed by halogen monitoring in thermal springs of La Soufrière volcano (Guadeloupe, Lesser
Antilles). Earth and Planetary Science Letters, 237, 710–728. https://doi.org/10.1016/j.epsl.2005.05.013
Villemant, B., Komorowski, J. C., Dessert, C., Michel, A., Crispi, O., Hammouya, G., et al. (2014). Evidence for a new shallow magma in-
trusion at La Soufrière of Guadeloupe (Lesser Antilles): Insights from long-term geochemical monitoring of halogen-rich hydrothermal
fluids. Journal of Volcanology and Geothermal Research, 285, 247–277. https://doi.org/10.1016/j.jvolgeores.2014.08.002
Voight, B., Glicken, H., Janda, R. J., & Douglass, P. M. (1981). Catastrophic rockslide avalanche of May 18. In P. W. Lipman, & D. R.
Mullineaux (Eds.), The 1980 eruptions of Mount St Helens, Washington (1250, pp. 347–378). US Geological Survey, Professional Papers.
Voight, B., Komorowski, J.-C., Norton, G. E., Belousov, A. B., Belousova, M., Boudon, G., et al. (2002). The 1997 Boxing Day sector col-
lapse and debris avalanche, Soufriere Hills Volcano, Montserrat, W. I. In T. H. Druitt, & B. P. Kokelaar (Eds.), The eruption of Sou-
frière Hills volcano, Montserrat, from 1995 to 1999 (21, pp. 363–407). Geological Society, London Memoirs. https://doi.org/10.1144/gsl.
mem.2002.021.01.17
Watters, R. J., & Delahaut, W. D. (1995). Effect of argillic alteration on rock mass stability. Clay and Shale Slope Instability, 10, 139.
Watters, R. J., Zimbelman, D. R., Bowman, S. D., & Crowley, J. K. (2000). Rock mass strength assessment and significance to edifice
stability, Mount Rainier and Mount Hood, Cascade Range volcanoes. Pure and Applied Geophysics, 157(6–8), 957–976. https://doi.
org/10.1007/s000240050012
Wyering, L. D., Villeneuve, M. C., Wallis, I. C., Siratovich, P. A., Kennedy, B. M., Gravley, D. M., & Cant, J. L. (2014). Mechanical and phys-
ical properties of hydrothermally altered rocks, Taupo Volcanic Zone, New Zealand. Journal of Volcanology and Geothermal Research,
288, 76–93. https://doi.org/10.1016/j.jvolgeores.2014.10.008
Yilmaz, T. I., Wadsworth, F. B., Gilg, H. A., Hess, K. U., Kendrick, J. E., Wallace, P. A., et al. (2021). Rapid alteration of fractured volcanic
conduits beneath Mt Unzen. Bulletin of Volcanology, 83(5), 1–14. https://doi.org/10.1007/s00445-021-01450-7
Zhu, W., Baud, P., Vinciguerra, S., & Wong, T. F. (2016). Micromechanics of brittle faulting and cataclastic flow in Mount Etna basalt.
Journal of Geophysical Research: Solid Earth, 121(6), 4268–4289. https://doi.org/10.1002/2016jb012826
Zimbelman, D. R., Rye, R. O., & Breit, G. N. (2005). Origin of secondary sulfate minerals on active andesitic stratovolcanoes. Chemical
Geology, 215(1–4), 37–60. https://doi.org/10.1016/j.chemgeo.2004.06.056

Reference From the Supporting Information


Hoek, E., Carranza-Torres, C. T., & Corkum, B. (2002). Hoek–Brown failure criterion—2002 edition. In R. Hammah, W. Bawden, J. Cur-
ran, & M. Telesnicki (Eds.), (Eds), Proceedings of the Fifth North American Rock Mechanics Symposium (NARMS-TAC), (pp. 267–273).
University of Toronto Press.

HEAP ET AL. 23 of 23

You might also like