You are on page 1of 14

Journal of Volcanology and Geothermal Research 388 (2019) 106686

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


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

Caldera or flank collapse in the Fogo volcano? What age?


Consequences for risk assessment in volcanic islands
F.O. Marques a,∗ , A. Hildenbrand b , S.S. Victória c , C. Cunha d , P. Dias a
a
Universidade de Lisboa, Lisboa, Portugal
b
GEOPS, Univ. Paris-Sud, CNRS, Université Paris-Saclay, 91405, Orsay, France
c
Universidade de Cabo Verde, Cape Verde
d
Instituto Politécnico de Coimbra, Escola Superior Agrária, Coimbra, Portugal

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

Article history: The target of this study is the Fogo Island in Cape Verde, where the hazards (caldera and/or flank collapse)
Received 26 March 2019 and the timing (120 ka or younger?) are still controversial. Using high-resolution DEM, field geological and
Received in revised form 8 October 2019 structural analysis, and high-precision K-Ar dating, we produced an age-calibrated volcanic stratigraphy
Accepted 8 October 2019
of Fogo’s summit (Chã das Caldeiras, for the last ca. 220 ka). From this, we infer the following evolution
Available online 2 November 2019
and associated processes: (1) the Fogo Volcano formed during seven stages of construction and partial
destruction; (2) three flank collapses can be recognised, the biggest of which occurred between 60 and
Keywords:
43 ka; (3) this collapse occurred retrogressively, producing at least two distinct collapse blocks; (4) the
Fogo Island
Cape Verde
innermost collapse was only partial, forming a flat step where a new volcano (Pico do Fogo) grew and
Geological hazard and risk formed the Chã das Caldeiras (literally Flat of the Calderas); (5) the removal of the buttressing eastern
Caldera collapse flank by the outermost collapse can be responsible for the seemingly “caldera” structure; (6) the growing
Flank collapse load of the young volcano can produce a new flank failure following the innermost fault; (7) the young
Large-scale mass wasting and closely spaced collapse ages here reported indicate a significant risk for the inhabitants of Fogo, in
Tsunami particular, and for the whole Cape Verde archipelago (and Atlantic seaboard) if a tsunami is produced by
a future catastrophic flank collapse.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction et al., 2001; Le Friant et al., 2003; Brunet et al., 2015). There was also
the idea that large-scale landslides were not expected to occur from
A geological hazard carries minimal risk for human society if the relatively small volcanic islands, but recent studies in the Azores
recurrence time is much longer than human life expectancy. Given have shown otherwise (Hildenbrand et al., 2012a, 2018; Costa et al.,
that risk increases greatly as the recurrence time approaches the 2014, 2015; Sibrant et al., 2014, 2015a, 2015b). This means that
lifetime of individuals, and also that it depends on the correct defi- every time a new landslide is discovered (see updated database for
nition of the associated hazard, we must correctly define the hazard the Atlantic in Blahut et al., 2018), the risk increases, especially if
and accurately estimate recurrence. Before the first recognition of the age of the landslide is estimated accurately and the recurrence
giant landslides in the Hawaiian Islands (Moore, 1964; Lipman et al., time decreases.
1988; Moore et al., 1989, 1994), there was no perception of this The Cape Verde archipelago ten inhabited intermediate-sized
risk because the hazard was unknown. After the first discoveries, islands (Fig. 1A) where many landslides have been recognized (Day
large-scale landslides have been found in oceanic islands world- et al., 1999 for Fogo Island; Ancochea et al., 2010 for S. Vicente
wide, namely in the Canaries (Holcomb and Searle, 1991; Krastel Island), mostly from offshore data (Le Bas et al., 2007; Masson et al.,
et al., 2001; Masson, 1996; Urgeles et al., 1997, 1999; Watts and 2008). Therefore, there is a great need for the characterization of
Masson, 1995), Reunion (Duffield et al., 1982; Labazuy, 1996), in the onshore structure of the landslides in Cape Verde and their age.
French Polynesia (Clouard et al., 2001; Clouard and Bonneville, The islands rise 5–8 km above the basement made up of Jurassic-
2004; Hildenbrand et al., 2004, 2006), and the Caribbean (Deplus Cretaceous aged seafloor. Fogo Island, located in SW Cape Verde,
records evidence of collapses, but the causes are not agreed upon:
caldera collapse, large-scale mass wasting, or both (Machado, 1965;
Brum da Silveira et al., 1997; Day et al., 1999; Madeira et al., 2008;
∗ Corresponding author.
Masson et al., 2008; Paris et al., 2011).
E-mail address: fomarques@fc.ul.pt (F.O. Marques).

https://doi.org/10.1016/j.jvolgeores.2019.106686
0377-0273/© 2019 Elsevier B.V. All rights reserved.
2 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Fig. 1. A – sketched Google Earth image to show geographic and tectonic settings of the Cape Verde archipelago. B – slope map to illustrate the morphology of Fogo Island.
Red, cyan and blue curves represent possible landslide faults discussed in this work, and the numbers the sequence of failure, all based on the interpretation of the current
structure and topography, and on field evidence collected in this study. Red and magenta straight lines represent dykes associated with Fogo and Monte Amarelo volcanic
centres. Note that the red straights converge to a point, which we infer to be the eruptive centre of the ancestral Fogo Volcano. C – topographic profiles with position of
collapse faults inferred in this study. D – 3-D Google image with location (red stars) of dated samples and respective approximate age (see Table 1 for details). Red circles –
main localities (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.).

Machado and Torre de Assunção (1965) published the first geo- (the Old Complex), unconformably overlain by a nephelinite com-
logical map of Fogo, and divided the outcropping rocks into pre- and plex of lava flows and dykes that makes up most of the island. This
post-collapse, which comprise three units: two pre-collapse (Old complex is cut by a composite listric fault forming a caldera, whose
Complex and Nephelinite Complex), and one post-collapse (Pico scarp is a prominent topographic relief that is locally called the Bor-
do Fogo Complex). Pre-collapse Fogo includes a basement complex deira (see Fig. 1D). Post-collapse lavas are (mostly) basanites that
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 3

erupted from the young Pico do Fogo and associated vents, and from collapse. Finally, we collected a sample (FO17 N) from outside of the
Strombolian cones scattered on the slopes of Fogo Island. collapse structure, at the base of a small paleocliff in northern Fogo,
Day et al. (1999) followed the previously established volcanic close to Fajãzinha, to constrain the oldest age of the Intermediate
stratigraphy, but gave the units local names and interpreted the col- Volcanic Complex.
lapse mechanism differently. Instead of caldera collapse, Day et al. Microscopic examination of thin sections was carried out on all
(1999) recognized an eastward major flank collapse, which they samples to identify their textural and petrographic characteristics,
called the Monte Amarelo collapse. Martínez-Moreno et al. (2018) and ensure that they had not suffered significant alteration. From
also followed the stratigraphy defined by Machado and Torre de the general abundance of olivine and/or pyroxenes and/or plagio-
Assunção (1965), but called Lower, Middle and Upper Volcanic units clase phenocrysts in thin section, all our new samples from Fogo are
the Old, the Nephelinite and the Pico do Fogo complexes, respec- basic to slightly evolved lavas; therefore, groundmass was selected
tively. for dating. The samples were crushed and sieved to the chosen size
The topography of the Bordeira scarp in the central-eastern part fraction (typically 63–125 ␮m or 125–250 ␮m). After ultrasonic
of Fogo Island has been diversely interpreted: scarp of a caldera col- cleaning in de-ionized water and drying at moderate temperature
lapse (Machado, 1965), the eastern half of which was subsequently (< 65 ◦ C), heavy liquids were used to separate groundmass from
removed by a normal fault or by erosion; scarp of two calderas sepa- phenocrysts (e.g. plagioclase, pyroxene and olivine), which are K-
rated by the Monte Amarelo spur, the missing eastern half of which poor and may carry potentially unsuitable inherited excess 40 Ar.
was subsequently removed by a flank collapse (Brum da Silveira The samples were then dated with the K-Ar Cassignol-Gillot tech-
et al., 1997); headwall scarp of a giant flank collapse (Day et al., nique (Cassignol and Gillot, 1982; Gillot and Cornette, 1986; Gillot
1999; Masson et al., 2008; Paris et al., 2011). et al., 2006), which has been shown to be well-suited to date Qua-
Analysis of the literature and the slope map shown in Fig. 1B ternary volcanic products with high precision, even low-K basalts
leads to the recognition of one main (composite) problem: is the and andesites (e.g. Germa et al., 2010, 2011; Boulesteix et al., 2012,
current structure and topography of eastern Fogo Island the prod- 2013; Ricci et al., 2015, 2017; Marques et al., 2018). This technique
uct of caldera collapse, flank failure, or both? At what age? What has been used extensively to constrain ages of flank instabilities
are the consequences for risk assessment in volcanic islands? Were on volcanic islands distributed worldwide (e.g., Gillot et al., 1994;
there more collapses in Fogo? This main question encompasses the Hildenbrand et al., 2003, 2004, 2008; 2012a, 2012b, 2014, 2018;
following problems/questions: (1) Where is/are the faults/scarps Quidelleur et al., 2008; Samper et al., 2008; Boulesteix et al., 2012,
produced by the large-scale landslide inferred for Fogo? Lines 1 2013; Costa et al., 2014, 2015; Sibrant et al., 2014, 2015a, 2015b,
(red), 2 (cyan) and 3 (blue) are three possibilities put forward in 2016).
this work; (2) was it a composite slide made of faults whose traces Potassium was measured by flame absorption spectrometry and
are represented by the red, cyan and blue curves? Or a slide along a compared with standards MDO-G (Gillot et al., 1992) and BCR-2
single fault?; (3) if composite, was failure along all faults complete (Rackzek et al., 2001), which were both prepared and analyzed
or only partial?; (4) What is the volume of the flank collapse? following the same protocol as the samples. At least two rounds
We analysed aerial imagery and DEMs for geomorphological and of independent acid attacks (i.e. the dissolution and leaching of
geological interpretation, and we carried out fieldwork in target acid-susceptible constituents) and K measurements were achieved
areas to recognize the major unconformities and volcanic units, to for each sample (see Gillot et al., 2006 for more information); the
unravel the geometry and position of main volcanoes, and to find K-value (Table 1) is considered final when the relative deviation
the critical collapse markers. Based on this information, we col- to the average is better than 1%. Argon was measured by a 180◦
lected critical samples for high-precision K-Ar dating to calibrate multi-collector sector mass spectrometer with the unspiked K-Ar
the volcanic stratigraphy, and to constrain the age of the major Cassignol-Gillot technique (Cassignol and Gillot, 1982; Gillot and
collapse episodes. Given that the volcanic and collapse structures Cornette, 1986). Decay constants and isotopic ratios of Steiger and
extend to the sea, we include high-resolution bathymetric data Jäger (1977) were used. At least two independent age determina-
(collected on the UK vessel RRS Charles Darwin during cruise CD168 tions were realized for each sample (Table 1). The uncertainty on
in February 2005) to discuss the link between the various structures each determination is obtained as the quadratic sum of all inde-
and associated offshore deposits. pendent sources of uncertainty, i.e. uncertainty on the K content
(1%), uncertainty on the calibration of the mass spectrometer (1%),
and the uncertainty on the amount of radiogenic argon (see details
2. Sampling strategy and K-Ar dating in previous works developing and applying the method, e.g., Gillot
et al., 2006; Germa et al., 2010; Boulesteix et al., 2013; Hildenbrand
Deposits of the pre-collapse unit(s) and the lava flows that et al., 2018). The final age (Table 1, bold font) is obtained as the
unconformably overlie (cascading, meaning downward flow as mean of individual age determinations (Table 1, normal font) when
deduced in the field) the inferred collapse scarps were sampled the latter are mutually consistent at the 1␴ level. Additional argon
(Fig. 1D). The main caldera wall (Bordeira) provided access to the analyses have been realized on samples FO17 J and FO18B, in order
base of the older volcanic succession (sample FO17A), which con- to obtain a more robust age determination for these young lavas
strains the earlier exposed stages of Fogo’s volcanic evolution. The despite their relatively low radiogenic yield (Table 1).
topmost part of the volcanic succession, cut by the Bordeira, is com- The new K-Ar ages obtained in this study range between
posed generally of lava flows with pronounced external dips (up to 212 ± 20 ka and 20 ± 4 ka (cf. Table 1 and Figs. 1D and 2), and are
30◦ ). To constrain the maximum age of the collapse, three samples fully consistent with the ages reported in Cornu (2017). These new
(FO17 G, FO18C and FO18D) were collected along the Bordeira’s rim ages allow better calibration of the volcanic stratigraphy, and better
(topmost lava flows), one of which is associated with a sectioned age constraints of the inferred collapses and their mechanisms.
volcanic cone. A lava flow from the (seemingly) same succession
was collected further east, at the Espigão scarp (sample FO17 L). To
constrain the minimum age of the collapse, a sample was collected 3. Field data
at the base of a lava pile that cascaded over the base of the Bordeira
(FO18B). Two lava flows (FO17 J and FO17 K) that cascaded over the We concentrated field work on Chã das Caldeiras and the
southern portion of the Espigão scarp (between Cova Figueira and eastern flank of Pico do Fogo. The Bordeira (local name of the
Tinteira were sampled to constrain the minimum age of the (outer) caldera/landslide scarp) exposes up to 1000 m of steeply dipping
4 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Fig. 2. Panorama photograph of Chã das Caldeiras (ca. 1700 m altitude) and Bordeira (scarp on the background – peak at 2700 m) as seen from the top of Pico do Fogo (peak
at ca. 2829 m). The blue dashed line marks the unconformity separating two main events of the evolution of Fogo (see Fig. 3 for volcanic stratigraphy, and Fig. 4 for details
of this unconformity). The green dashed line at the base of the Bordeira marks the youngest unconformity. Red stars – location of samples dated by K-Ar in this work. Red
circles – main localities (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.).

Fig. 3. Photographs with drawn volcanic stratigraphy based on major unconformities and calibrated by K-Ar geochronology. A – main stratigraphy of Chã das Caldeiras,
mostly made of two thick volcanic complexes on the Bordeira. B – detail of the lowermost units observed at Boca Fonte. BC – Basement Complex (Old Complex of Machado
and Torre de Assunção, 1965); LSU – Lower Sedimentary Unit; LVC – Lower Volcanic Complex; IVC – Intermediate Volcanic Complex (these three units correspond to the
Nephelinite Complex of Machado and Torre de Assunção, 1965); UVSC – Upper Volcano-sedimentary Complex (Pico do Fogo Complex of Machado and Torre de Assunção,
1965).
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 5

Table 1
Summary of the new K-Ar ages on separated fresh groundmass. The mean age is obtained by weighing by the amount of radiogenic argon (40 Ar*), except in the case of
samples FO18B and FO17 J, for which more than 2 independent argon analyses have been achieved; for the latter, the mean age (*) is calculated by weighing by the inverse
of variance. Measured ages in normal font, and mean ages in bold font. Uncertainties quoted at the 1␴ level.

Sample Latitude (N) Longitude (W) Altitude (m) K% 40Ar* (%) 40Ar* (1011 at./g) Age (ka) Uncertainty (ka)

FO17B 14.96884 24.37923 1792 1.856 1.1 4.3103 222 20


1.0 3.8766 200 19
mean 212 20
FO17N 15.04199 24.34609 58 2.347 2.1 2.5455 104 5
2.6 2.5107 102 4
mean 103 5
FO17L 14.91457 24.29687 426 2.975 3.3 2.4792 80 3
3.9 2.4503 79 2
mean 79 2
FO17G 14.91895 24.35125 2119 2.335 1.5 1.5132 62 4
1.6 1.4193 58 4
mean 60 4
FO18C 14.96019 24.39255 2666 2.626 2.4 1.5843 58 3
2.6 1.6312 59 2
mean 59 2
FO18D 14.93763 24.39353 2396 3.823 3.7 2.3124 58 2
3.3 2.3699 59 2
mean 59 2
FO18B 14.95965 24.38538 1770 2.312 0.2 1.0242 42 17
0.2 0.96058 40 16
0.3 1.0401 43 14
0.3 1.3268 55 17
0.3 0.91213 38 13
mean* 43 7
FO17J 14.90942 24.29416 544 2.274 0.4 0.49267 21 6
0.3 0.37005 16 5
0.5 0.58425 25 5
mean* 20 3
FO17K 14.91123 24.29390 524 2.312 0.6 0.49188 20 4
0.6 0.48917 20 4
mean 20 4

(up to ca. 30◦ outwards) volcanic (mostly) and sedimentary (minor) pyroclasts cut and fed by a dyke swarm. This complex is deeply
rock (see Figs. 2 and 3), which represents a great fraction of the metasomatized by the numerous dykes intruding it, giving the rock
volcanic sequence of the pre-collapse Fogo Volcano. Therefore, a green colour typical of minerals like chlorite (see Supp. Fig. 1). The
the volcanic stratigraphy based on observations realized on the rocks of the overlying Intermediate Volcanic Complex do not show
Bordeira and proposed here represents an important part of the this kind of alteration. The number of dykes cutting and feeding
evolution of the Fogo volcano. the Lower Volcanic Complex increases towards the Monte Amarelo
spur (Fig. 2), where a large number of dykes with different ori-
entations cross-cut, and number of dykes decreases towards the
3.1. Volcanic stratigraphy of Chã das Caldeiras
north. The Lower Volcanic Complex corresponds to the lower part
of what was called the pre-collapse complex by Machado and Torre
3.1.1. Lower sedimentary unit
de Assunção (1965), and Monte Amarelo Group or Volcano by Day
The oldest rocks we recognised at Chã das Caldeiras occur in
et al. (1999). The oldest age reported by Cornu et al. (2017) on lavas
one of the few canyons that cut the base of the Bordeira, locally
sampled at the base of the Bordeira is ca. 158 ka. From this age
known as Boca Fonte (Fig. 3B), just south of Monte Amarelo. Here,
and the oldest age measured in the present work (ca. 212 ka), we
the lowermost deposits make up the Lower Sedimentary Unit, in
infer that the ca. 158 ka corresponds probably to the oldest age of
which three conglomerates could be distinguished by clast size and
the overlying Intermediate Volcanic Complex. Therefore, the age of
composition, separated by sharp contacts (Fig. 3B). The base of the
this unit is here constrained between ca. 160 and 212 ka.
lowermost conglomerate is not exposed. The lower conglomerate
is polygenetic and consists mostly of large blocks (meters in size),
including greenish yellow blocks of volcanic breccia that were not
3.1.3. Intermediate volcanic complex
observed in the overlying conglomerates. The intermediate con-
The Intermediate Volcanic Complex lies unconformably on the
glomerate consists of (mostly) monogenetic boulders (< 50 cm). The
Lower Volcanic Complex, and is comprised of thick (up to 20 m)
uppermost one is made of metre-sized blocks. These conglomerates
lava flows intercalated with thick (up to 20 m) pyroclastic strom-
are cut by dykes feeding the younger units. The Lower Sedimen-
bolian deposits. Most importantly, this unit includes volcanic cones
tary Unit seems to have filled a major unconformity between an
at the Bordeira’s rim and sectioned by the scarp (see Fig. 3A). The
underlying volcanic complex (Old Complex of Machado and Torre
unconformity, here called the Monte Amarelo Unconformity, can
de Assunção, 1965) and the overlying Lower Volcanic Complex. The
be recognised by the different colours of rock below (light green)
age of the Lower Sedimentary Unit is constrained here at > 212 ka
and above (dark brown and red) (see Figs. 2 and 4), by the geometry
(cf. Fig. 1D and Table 1), because this is the oldest age we measured
of flows and dykes, by dykes that only exist in the Lower Volcanic
in the overlying Lower Volcanic Complex.
Complex (NNE-SSW trend), and by a small gap in age.
The Monte Amarelo Unconformity does not appear to be erosive
3.1.2. Lower volcanic complex in origin (no erosion-related sediments found to date). It could rep-
The Lower Volcanic Complex unconformably overlies the con- resent a shift in the eruptive centre towards the south, as indicated
glomerates, and is composed of intercalated lava flows and by dyke distribution and trend (cf. section 3.2 below), with conse-
6 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Fig. 4. Photographs of the unconformity between the Lower and Intermediate Volcanic Complexes, here called the Monte Amarelo unconformity. Successive zooming in
from A to D. C – note the great number of NNE-SSW dykes in the Lower Volcanic Complex (marked with white arrows) that are absent in the Intermediate Volcanic Complex.
D – note the great difference in colour between the two complexes, mostly greenish in the Lower Volcanic Complex due to metasomatization induced by the numerous
dykes.

quent overlap of younger on older lavas. The Intermediate Volcanic centre of the southern Chã das Caldeiras (dashed green circle in
Complex corresponds to the upper part of what was called the pre- Fig. 5). From this we infer the position of the main emission cen-
collapse complex by Machado and Torre de Assunção (1965), and tre and crater of the ancestral Fogo volcano (160-60 ka). We note
Monte Amarelo Group or Volcano by Day et al. (1999). We keep that the horizontal distance between the craters of Fogo inferred
the name Monte Amarelo Volcano, but we use the name Fogo Vol- here and Pico do Fogo is very similar to that between the Bordeira
cano for younger and more voluminous lavas. Previous work has rim and the flat bottom of Chã das Caldeiras (considering that the
not recognised the Monte Amarelo unconformity or distinguished original flat has been filled by recent volcanism). Coincidently, the
the two volcanic complexes described here. The age of the Inter- direction of this displacement is approximately the same expected
mediate Volcanic Complex is here constrained between ca. 59 and for the collapse, which can be constrained by the geometry of the
ca. 158 ka (this work and Cornu et al., 2017). scarp. From the similarity of lengths and directions, we infer that
the feeder centre moved together with the partially collapsed block,
from the edge of which the new volcano Pico do Fogo grew. Dyke
3.1.4. Upper volcano-sedimentary complex
trends represented by the yellow dashed lines in Fig. 5 converge at
The structures attributed to the Upper Volcano-sedimentary
Monte Amarelo, which was active between ca. 212 and 160 ka.
Complex lie unconformably on the older volcanic complexes, both
Sills are also abundant on the Bordeira, and can be distinguished
inside and outside Chã das Caldeiras, comprising Pico do Fogo vol-
from lava flows by their different structure (typical sills do not have
cano, many Strombolian cones inside and outside Chã das Caldeiras,
clinker at the contact with the host rock), by the thermal effect on
and thin lava flows intercalated in recent talus deposits at the base
host rock in both contacts (lava flows only cook the underlying
of the Bordeira (Supp. Fig. 2). Most importantly, this unit includes
rocks), and by the absence of weathering at the top.
lava flows that flowed over the collapse scarps, and thus post-date
collapse. The age of this unit is constrained between ca. 43 and 0
ka.
3.3. Evidence of collapse inside Chã das Caldeiras

3.2. Dykes at Chã das Caldeiras The Bordeira scarp, as a whole, appears to be the scarp of a nor-
mal fault: ca. 60 ka rim in the footwall, and younger top in the
We measured 273 dykes at Chã das Caldeiras (Fig. 5 and Supp. hanging wall separated by hundreds of meters. However, normal
Fig. 3), and two main groups can be distinguished. The first group faults typically exhibit dip-slip striations, which are not obvious at
is restricted to the Lower Volcanic Complex (yellow dashed lines in Chã das Caldeiras, most likely because what we observe is a scarp,
Fig. 5). The other group consists of dykes that cut all units (orange and not the fault itself, which can be as high as 1000 m. Locally at
dashed lines in Fig. 5). To the north and south of the spur, the mean the rim of the Bordeira scarp we observe sectioned volcanic cones
trend is NNE-SSW, but within the spur many dykes cross-cut one with an age of 59 ± 2 ka, (sample FO18C) (Supp. Figs. 4 and 5),
another. Dykes between the spur and the southern end of the Bor- hinting at the maximum age of the collapse. Locally, at the base of
deira show a radial trend that converges towards a point in the the Bordeira, we can observe lava flows cascading over the scarp,
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 7

Fig. 5. Google image centred at Chã das Caldeiras to show the mean trend of the measured 273 dykes (by sectors for clarity, but all dykes plotted as short red lines along the
base of the scarp) and the inferred geological meaning. Orange dashed lines represent mean trend of dykes feeding the Fogo Volcano (numbering as in Supp. Fig. 3); yellow
dashed lines represent the dykes feeding the Monte Amarelo (literally Yellow Hill) volcano; and blue dashed line represents the mean strike of the measured 273 dykes. Note
that this mean strike is, within error, parallel to the elongation of Chã das Caldeiras. Green arrows added to compare displacement due to collapse and displacement of the
volcanic emission centre (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.).

which we dated at ca. 43 ka and indicate a minimum age of the that a proto-Bordeira scarp existed prior to ca. 212 ka and would
collapse. We did not find deposits typical of caldera collapse (e.g. have been the source of the conglomerates, which could be a proto-
ignimbrites), which characterize most calderas around the world caldera or simply a deep crater.
(e.g. Druitt and Sparks, 1984; Gooday et al., 2018; Jordan et al., 2018; Our data indicate that the volcanic feeding centre shifted from
Rosi et al., 1996). We also found no evidence for caldera resurgence the earlier Monte Amarelo Volcano in the north to the younger
(e.g. Tibaldi and Vezzoli, 2004, for Ischia, Italy) or for caldera forma- Fogo Volcano in the south. This can explain why the Bordeira is
tion by magma chamber growth (Gudmundsson, 1988, for Iceland). much taller south of the Monte Amarelo spur than to the north. We
Therefore, the evidence for a collapse fault seems clear, but not the suggest, therefore, that the pre-collapse Fogo edifice did not have
nature and age of the collapse itself, which we discuss below. a single main eruptive centre, but two – Monte Amarelo and Fogo.
The stresses that produce shallow tabular intrusions are reason-
ably well understood: the least compressive stress (␴3 ) is typically
3.4. Evidence of collapse outside Chã das Caldeiras
perpendicular to the tabular body, and ␴1 is parallel to the tabular
intrusion (e.g., Daniels and Menand, 2015, and references therein).
Outside of Chã das Caldeiras, but onshore, there is good evi-
Both dykes and sills occur pervasively on the Bordeira, and most
dence of collapse on the scarp following the southern Bordeira end
cut one another with no specific chronological order. This indicates
to the east, the so-called Espigão scarp (Figs. 6 and 7). This scarp is
that the stress field was sometimes dominated by horizontal ␴3 to
partly buried by post-collapse lava flows (Fig. 7) but is still conspic-
form vertical dykes, and sometimes by steeply dipping ␴3 to form
uous. The scarp following the north end of the Bordeira to the east
the sills sub-parallel to lava flows dipping ca. 30◦ outwards from
towards Mosteiros is not evident, because it has been concealed
the volcanic centre. We infer from the overall radial geometry of
by post-collapse lavas. We can envisage two ways for linking these
dykes and the pervasive intrusion of sills that the stress field was
two scarps: (1) the most obvious is through the Bordeira scarp,
produced by a shallow magma reservoir inside the volcanic edi-
as proposed by Day et al. (1999); (2) through the slope break at
fice. Inflation can be responsible for radial dyke intrusion, whereas
the eastern edge of Chã das Caldeiras (blue dashed line in Fig. 6),
deflation can cause sill emplacement (e.g. Polteau et al., 2008).
which is still well visible despite blanketing by Pico do Fogo and
partial burial by post-collapse volcanism, as proposed here. Given
the existence of well-preserved and vigorous scarps closer to the
4.1. Collapse structure and nature
coast, and towards the north (Mosteiros) and south (below Cova
Figueira) of the collapse delimited by the blue dashed line, we pro-
There are two contrasting views on the nature of the Chã das
pose the existence of another collapse fault, the headwall scarp of
Caldeiras (caldera or flank collapse) and on the position of the
which is represented by the red dashed line in Fig. 6.
main fault along which the flank collapse occurred: on the Bor-
deira scarp (Day et al., 1999, and followers), or at the eastern edge
4. Discussion of the Chã das Caldeiras flat (Brum da Silveira et al., 1997, and fol-
lowers). Regarding the nature of the Chã das Caldeiras, we found
Conglomerates with large blocks occur under specific condi- no evidence for extensive ignimbrite deposits that are common to
tions, and typically in settings of very high potential energy related most calderas. We also found no evidence of caldera formed by the
to high and steep topography. At Chã das Caldeiras, at ca. 1700 m growth of a shallow magma chamber (Gudmundsson, 1988) We
altitude, there are not many options to explain the existence of also do not favour the existence of two calderas (Brum da Silveira
the Lower Sedimentary Unit conglomerates. Our interpretation is et al., 1997), one to the north and the other to the south of the Monte
8 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Fig. 6. 3D Google image of Fogo viewed from the east to show the surface trace of the collapse faults proposed in this study. The white dashed line highlights the slope breaks
that help us constrain the position of the scarps produced by the flank collapse. Surface traces of the scarps are represented by the red, cyan and blue dashed lines. Red stars
– ages reported here. Red circles – main localities (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.).

Fig. 7. 3-D Google image with geometry of lava flows and unconformities, and ages of critical samples dated by K-Ar in this study (red stars). From flows cascading on a
previous steep topography, and from the ages above and below the unconformable contact, we infer the position of the fault trace, which should be responsible for the flank
collapse and submarine deposit reported in Day et al. (1999). Red circles – localities (For interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.).

Amarelo spur, because the Monte Amarelo spur is a volcanic feeding significant flat made by the parasitic cone. The volcanic products
centre strengthened by a great amount of dykes, which therefore either cover the flank as a small cone of pozzolan (volcanic ash) or
could have resisted collapse. In summary, we found no evidence flow downhill (lava flows). Therefore, we discard this hypothesis.
for caldera collapse, and thus we do not favour the caldera collapse Here, we propose a different scenario, which is partial failure
hypothesis. Regarding the position of the collapse fault, the hypoth- of the flank collapse fault most upstream (west) in a composite
esis that the flank collapse occurred on the Bordeira has a problem, collapse structure (e.g. Hildenbrand et al., 2012a, 2012b; Hunt et al.,
as illustrated in Supp. Fig. 6: in this case (Supp. Fig. 6A), all surfaces 2013; Costa et al., 2015) with one (or two) partial failure(s) and one
are non-cylindrical (collapse fault is listric and concave to the east, complete failure (Figs. 6 and 7, and Supp. Fig. 6). When the eastern
and the new volcano Pico do Fogo is a cone), which makes it vir- flank failed along the outermost fault (red dashed line in Figs. 1B
tually impossible to produce a horizontal surface like the Chã das and 6), the edifice was no longer buttressed on the east, and so the
Caldeiras. This situation is similar to parasitic cones on the flanks summit collapsed in that direction. This summit collapse was not
of major conical volcanic edifices, in which there is no record of a total, resulting in the Chã das Caldeiras as discussed above (Supp.
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 9

Fig. 6). It also seems to have been formed by one (or two) partial At Chã das Caldeiras and eastern flank of Fogo, there is good
collapses as inferred from slope breaks east of Chã das Caldeiras evidence for a younger age of the flank collapse. Lava flows cascade
(Fig. 6). The Pico do Fogo therefore sits on the eastern edge of the over the base of the Bordeira scarp, one of which we dated at ca. 43
Chã das Caldeiras, and is apparently fed by the same conduit as the ka (Supp. Fig. 7). We could find no evidence for the source of these
ancestral Fogo (see Fig. 5). flows at the top of the Bordeira scarp, and therefore we infer that the
The young Pico do Fogo volcano has been growing very fast at the source of lava is somewhere on the scarp, i.e. a vent from which lava
edge of Chã das Caldeiras (eruptions every 20 years on average since fountained (Supp. Fig. 7). Also at Chã das Caldeiras, at the Bordeira’s
island discovery, and currently ca. 1000 m in height), thus loading top rim, we dated lava flows at ca. 59 ka that are unambiguously
the block (Chã das Caldeiras) underlain by the partially failed col- cut by the scarp (samples FO17 G and FO18D), and could observe a
lapse fault (Bordeira fault). The load of the new volcano and of the sectioned volcanic cone that is also ca. 59 ka old. This cone either
lavas filling the Chã das Caldeiras can therefore reactivate the weak- predates flank collapse, and the section is a product of the collapse,
ened Bordeira fault and produce a (near?) future massive landslide or it post-dates collapse, and its destruction is simply a result of
of the whole Chã das Caldeiras block. subsequent erosion. The problem with the latter solution is that
the dyke that fed eruption of the cone outcrops and is ca. 3 m thick
as far as the westernmost extension of the historical lava flows (cf.
4.2. Volume of the eastern Fogo flank collapse
Supp. Fig. 4), where the trace of the collapse fault should be. Given
that dykes cannot flow through air, and that the sectioned cone
Here we estimate the volume of the flank collapse for two end-
is well preserved, we infer that the scarp is very young, and that
member cases. (1) Listric collapse flattening out at sea bottom
the missing portion of the original volcano (towards the observer
(Fig. 8B), and (2) listric collapse above ca. −500 m, as proposed by
in Supp. Fig. 4) was continuous at the time of the collapse with
Martínez-Moreno et al. (2018; their Fig. 8) (Fig. 8C). If the vertex of
the structures observed currently (dykes flowing across the earlier
the conical island were coincident with the Bordeira (Fig. 8A), there
main volcano and feeding the original unsectioned cone). From our
would be no collapse. A shallow collapse above −500 m (Fig. 8C)
new age data of the topmost lava flows and cones, and the age
cannot explain the missing flank at -1000 m. Therefore, the solution
of lava flows cascading on the partial collapse scarp, we therefore
that best fits the current data is the one represented in Fig. 8B.
propose that the age of the collapse lies between ca. 59 ± 2 ka and
To calculate volumes for the two cases, we approximate the
43 ± 7 ka.
shape of the collapse fault to a spherical cap, because the collapse
On the eastern flank of Fogo, one collapse scar is prominent
fault is curved both along dip and along strike; i.e., it is a combina-
(the southern Espigão scarp – Fig. 7) and another (north, towards
tion of half an ellipse in vertical cross-section (cf. Fig. 8) with a circle
Mosteiros) has been mostly concealed by young and historic vol-
in plan view (cf. circumference drawn in Fig. 9A). Then we calculate
canic products. The ages obtained here on the lavas from the
the volume of a partial sphere assuming values of the radius of the
Espigão scarp (ca. 79 ka) and on cascading lava flows (ca. 20 ka)
approximated circumference (6 and 10 km for shallow and deep
also point to a young age (< ca. 60 ka) for the flank collapse. As both
collapses, respectively, measured horizontally as shown with the
scarps formed within the last 60 ka, they may be contemporaneous
circle in Fig. 9) and the height of the spherical cap (1 and 2 km for
and result from one single main phase of instability. In this case,
shallow and deep collapses, respectively, with greatest thickness
we propose that flank collapse along the Espigão-Mosteiros fault
measured perpendicular to the collapse fault in Fig. 8). We obtain
removed part of the eastern flank of Fogo, eliminating the buttress
ca. 20 and 120 km3 for the shallow and deep collapses, respec-
of the summit from the east and triggering partial collapse along the
tively. A volume of 20 km3 is significantly less than the ca. 110 km3
Bordeira fault. We note that the post-collapse lavas dated between
estimated by Martínez-Moreno et al. (2018), and ca. 120 km3 is
ca. 20 and 43 ka represent a significant phase of volcanism drap-
comparable with the estimates in Masson et al. (2008 – 130 to
ing/concealing the collapse structures. By analogy to what has been
160 km3 ), Martínez-Moreno et al. (2018), and Paris et al. (2011 –
proposed for many other volcanic islands where flank collapses are
115 km3 ), but significantly less than estimated by Day et al. (1999
soon followed by vigorous volcanism (e.g. Hildenbrand et al., 2003,
– 150 to 300 km3 ). Note that our estimates do not consider mate-
2004; Quidelleur et al., 2008; Boulesteix et al., 2012, 2013; Hunt
rial eventually incorporated from submarine failure of the slopes,
et al., 2018), we infer that the composite collapse occurred between
nor entrainment of material and subsequent failures of the seafloor
60 and 43 ka ago.
from loading of the landslide debris. Such incorporation may result
in significant difference between volume removed by the collapse
4.4. Flank collapses of Fogo from topographic data
and the volume of the deposits (e.g., Hildenbrand et al., 2006).
Analysis of the topography of Fogo Island and the adjacent sea
4.3. Age of the eastern Fogo flank collapse bottom (Fig. 9A and Supp. Fig. 8) hints at the existence of more than
one collapse of the volcanic edifice. If we draw three circles enclos-
There is no consensus regarding the age of the flank collapse, ing and tangent to the 0, -1000 and -2,000 m isobaths (cyan, orange
which has been proposed to be between ca. 59 and ca. 124 ka and green dashed circles, respectively, in Fig. 9), it becomes appar-
(Foeken et al., 2009; Paris et al., 2011); Ramalho et al., 2015; ent that most of the edifice at those depths is coincident with the
Cornu, 2017). Using cosmogenic 3 He, Foeken et al. (2009) dated circles, which means that it was originally conical and centred at
the collapse at 62–123 ka. Paris et al. (2011) bracketed the age Chã das Caldeiras. However, three sectors of the edifice are straight
of the eastern Fogo collapse at 86–124 ka, based on a U–Th age (east) or concave towards the sea (northwest and southwest), from
(123.6 ± 3.9 ka) of a coral branch in a tsunami deposit found on which we can infer that, similarly to the east Fogo collapse, parts
Santiago Island, and a K-Ar age (86 ± 3 ka) of a “post-collapse” lava of Fogo’s edifice are missing (magenta shaded areas in Fig. 9A).
along the Bordeira in Fogo. Assuming that large blocks in north San- The eastern Fogo collapse is associated with a prominent debris
tiago Island represent a tsunami deposit produced by the eastern deposit (outlined by solid red line in Fig. 9A), but the inferred col-
Fogo collapse, Ramalho et al. (2015) used exposure ages to propose lapse to the NW also shows a debris deposit offshore, although not
an age between 65 and 97 ka for the collapse. Cornu (2017) con- as prominent as in the east. In Fig. 9A, we can also distinguish three
strained the age of the collapse between ca. 68 (youngest lava flow kinds of deposits: (1) proximal collapse-related debris deposit with
cut by the collapse scar) and ca. 59 ka (oldest lava flow interpreted mixed large blocks and finer matrix (delimited by the red line); (2)
to have cascaded over the scarp). distal fine-grained, collapse-related debris deposit, in clear conti-
10 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Fig. 8. A – topographic profile along the white dashed line in Fig. 9A. Red line reproduces the topography of the current Fogo’s western slope. B and C – two possible
configurations of the East Fogo Flank Collapse: B – listric collapse flattening out at sea bottom; B – listric collapse above ca. −500 m, as proposed by Martínez-Moreno et al.
(2018; their Fig. 8). Bo – Bordeira; CC – Chã das Caldeiras; PF – Pico do Fogo (For interpretation of the references to colour in this figure legend, the reader is referred to the
web version of this article.).

nuity with the coarser deposit upstream (less energy away from the In this study, we not only show that Fogo’s eastern flank collapse is
source = smaller grain size); (3) recent Pico do Fogo/erosion-related younger than previously inferred, but also found strong evidence
volcanic/sedimentary deposits. In Fig. 9B we draw the displaced for two new and recent collapses in northwest and southwest of the
Fogo’s crater at 1000 m altitude based on two arguments: (1) Pico island, which, if confirmed, would mean that the risk is significantly
do Fogo could not have made the flat of Chã das Caldeiras, as dis- higher than previously thought.
cussed above (cf. section 4.1); (2) Fig. 9B in Martínez-Moreno et al. A similar situation happened in the Azores; before 2012 the risk
(2018) shows a flat low-resistivity layer at ca. 1000 m depth below related to potential large-scale island flank collapse was underesti-
Chã das Caldeiras, inconsistent with the curved shape of the col- mated, as very few old events of catastrophic lateral destabilization
lapse fault, but consistent with the hypothesis of partial collapse of had been reported (Woodhall, 1974; Mitchell, 2003). Currently the
a flat original crater floor of Fogo Volcano. risk is much higher, because many flank collapses have been recog-
nised in the course of two research projects, MEGAHazards 1 and
4.5. Hazard and risk 2 (Hildenbrand et al., 2012a, 2018; Costa et al., 2014, 2015; Sibrant
et al., 2014, 2015a, 2015b). The same may be true for many other
The hazard discussed here, large-scale flank collapse(s), com- oceanic volcanic islands, where landslides have continuously been
prises two kinds of risk: one inherent to the collapse itself (for recognised, so increasing the number of collapses and therefore the
example villages can suddenly be carried into the sea in the case of a risk.
catastrophic collapse), and the other is related to the tsunamigenic
potential of the collapse, i.e., depending on the collapse velocity and 4.6. Evolution of Fogo
internal structure, a tsunami can be generated by the collapse and
have devastating consequences on the island itself and surround- Having discussed all available and new data, we now propose
ing coast lines. In both cases, the hazard is not easy to translate into the following Fogo’s evolution (see Fig. 10, and also Fig. 3 for the
risk, because there are many unknowns to consider. How many col- stratigraphy defined and calibrated in this study). The evolution
lapses occurred on Fogo and what is the recurrence time? Before of the island began with a seamount stage according to Day et al.
identifying all individual flank collapses on Fogo, and date them, we (1999) and Foeken et al. (2009) (Fig. 10), possibly older than 4.5
cannot estimate recurrence time, and therefore we cannot evaluate Ma (Foeken et al., 2009). The seamount emerged from the sea,
the risk. Did the eastern Fogo flank collapse occur catastrophically, thus giving birth to a new island during the first subaerial build-
or slowly as a slump? Did it occur as one main solid block, or as ing stage (Figs. 2–10). The early seamount stage must have been
shattered mass? Without this knowledge, we cannot evaluate the followed by an island stage, because all rocks observed on Fogo
tsunamigenic potential of the collapse. We can have an idea of the are sub-aerial; no submarine lavas (e.g. pillow lavas) have been
tsunamigenic potential if we find a coeval tsunami-related deposit reported on Fogo. This volcanic complex forms the unreachable
onshore, because we can estimate the amplitude of the wave from core of Fogo, which must be older than ca. 212 ka, the oldest lavas
the altitude at which we find the highest tsunami deposit. How- we found so far on Fogo. The growing island had steep topography
ever, such a deposit has not been identified onshore Fogo. Blocky or became gravitationally unstable, thus producing the first sub-
deposits interpreted as tsunami deposits generated by the Fogo aerial erosional stage (Figs. 3–10). This stage is deduced from the
eastern flank collapse have been described on the neighbouring major unconformity with conglomerates of the Lower Sedimentary
island of Santiago (Paris et al., 2011; Ramalho et al., 2015), but our Unit. Given the large dimensions and diversity of the blocks, we
new age data for the collapse is significantly younger than the mean infer that the conglomerates formed due to high topographic relief
age estimated for those deposits, rendering the link between the (simply a relatively deep crater? Caldera collapse? Flank collapse?).
two processes unlikely. Yet, no unambiguous tsunami deposits less This erosional stage must be older than ca. 212 ka, because this is
than ca 59 ka in age has been reported elsewhere in the archipelago. the oldest age we found in the unconformably overlying unit (LVC).
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 11

Fig. 9. A – image with slope map onshore and bathymetry offshore Fogo Island with interpretation of different flank collapses and associated debris deposits. The red line
delimits the coarse debris deposit of the Eastern Fogo collapse, as defined by Masson et al. (2008). B – Profile showing the position of possible partial (blue) and full (cyan)
collapse faults. White dashed line – position of the topographic profile shown in Fig. 8. Black circumference – represents the geometry used to calculate the collapse volume.
Bo – Bordeira; CC – Chã das Caldeiras; PF – Pico do Fogo (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.).

The rocks composing this conglomerate should give us a rough idea tion stage took place in the new Fogo Volcano (Figs. 5–10). We
of what the ancestral Fogo looked like, both in terms of composi- attribute this stage to a shift of the eruptive centre from the Monte
tion and age (work in progress). The erosional stage was followed Amarelo Volcano to the south, the Fogo Volcano, thus producing the
by a second subaerial building stage, the Monte Amarelo Volcano Monte Amarelo unconformity. This stage corresponds to the Inter-
stage (Figs. 4–10). This stage comprises the rocks here attributed mediate Volcanic Complex, which we dated younger than ca. 60
to the Lower Volcanic Complex, which from our study is younger ka. The growing island became gravitationally unstable, thus pro-
than ca. 212 ka. Volcanic construction continued, but growth of ducing the second subaerial erosional stage (Figs. 6–10). This stage
the Monte Amarelo Volcano ceased and a third subaerial construc- corresponds to the eastern Fogo flank collapse, which we dated here
12 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

dated between ca. 124 and 86 ka by Paris et al. (2011), and between
ca. 65 and 97 ka by Ramalho et al. (2015). Given the stratigraphy
presented here, and possible multiple collapses around Fogo, we do
not favour the name Monte Amarelo Collapse proposed by Day et al.
(1999) for the main flank collapse in Fogo Island, and we propose
that it be called East Fogo Flank Collapse.

Declaration of Competing Interest

We do not have any conflicts of interest.

Acknowledgment

This is a contribution to Project MEGAHazards2 (PTDC/GEO-


GEO/0946/2014), funded by FCT, Portugal. We are very grateful to
Tim Le Bas for having made available to us the bathymetric data col-
lected on the UK vessel RRS Charles Darwin during cruise CD168 in
February 2005. Thanks are also due to Léa Ostorero her help in the
laboratory. Thorough and constructive reviews by James Hunt and
James Gardner helped to improve the quality of this manuscript.
We also thank the careful editorial work by Editor James Gardner.

Appendix A. Supplementary data

Supplementary material related to this article can be found,


in the online version, at doi:https://doi.org/10.1016/j.jvolgeores.
2019.106686.

Fig. 10. Cartoon with Fogo’s evolution inferred from field data and calibrated vol- References
canic stratigraphy. Red dashed lines in 3 are hypothetical flank collapses that can
justify the observed erosion (but are by no means mandatory). The red dashed lines
Ancochea, E.M., Huertas, J., Hernán, F., Brändle, J.L., 2010. Volcanic evolution of São
in 5 mark the location of the faults that led to the later flank collapse illustrated Vicente, Cape Verde Islands: the Praia Grande landslide. J. Volcanol. Geotherm.
in stage 6 (For interpretation of the references to colour in this figure legend, the Res. 198, 143–157.
reader is referred to the web version of this article.). Blahut, J., Klimeš, J., Rowberry, M., Kusák, M., 2018. Database of giant landslides on
volcanic islands - first results from the Atlantic Ocean. Landslides, http://dx.doi.
org/10.1007/s10346-018-0967-3.
between ca. 59 and 43 ka, i.e. around 50 ka. The flank collapse left a Brum da Silveira, A., Madeira, J., Serralheiro, A., 1997. A estrutura da Ilha do Fogo,
block only partially collapsed, the Chã das Caldeiras block, which is Cabo Verde. A Erupção Vulcânica de 1995 na Ilha do Fogo, Cabo Verde. Publ. IICT,
underlain by the Bordeira Fault. The collapse was soon followed by Lisboa, pp. 63–78.
Boulesteix, T., Hildenbrand, A., Gillot, P.Y., Soler, V., 2012. Eruptive response of
a fourth subaerial building stage (Figs. 7–10), which comprises the oceanic islands to giant landslides: new insights from the geomorphologic
Upper Volcano-sedimentary Complex and a new impressive vol- evolution of the Teide–pico Viejo volcanic complex (Tenerife, Canary). Geomor-
cano, the Pico do Fogo Volcano. Pico do Fogo grew on the eastern phology 138 (1), 61–73.
Boulesteix, T., Hildenbrand, A., Soler, V., Quidelleur, X., Gillot, P.Y., 2013. Coeval giant
edge of the partially collapsed Chã das Caldeiras and is still very landslides in the Canary Islands: implications for global, regional and local trig-
active (eruptions every 20 years on average since island discov- gers of giant flank collapses on oceanic volcanoes. J. Volcanol. Geotherm. Res.
ery), thus loading the unstable Chã das Caldeiras block and being 257, 90–98.
Brunet, M., Le Friant, A., Boudon, G., Lafuerza, S., Talling, P., Hornbach, M., Ishizuka,
capable of producing a (near?) future flank collapse. The oldest age O., Lebas, E., Guyard, H., IODP Expedition 340 Science Party, 2015. Com-
of this stage is here constrained around 43 ka. position, geometry, and emplacement dynamics of a large volcanic island
landslide offshore Martinique: from volcano flankcollapse to seafloor sediment
failure? Geochem. Geophys. Geosyst. 17, 699–724, http://dx.doi.org/10.1002/
5. Conclusions 2015GC006034.
Cassignol, C., Gillot, P.Y., 1982. Range and effectiveness of unspiked potassium-
Based on new and available data, we draw the following con- argon dating: experimental groundwork and applications. Numerical dating in
stratigraphy 1, 159–179, John Wiley, New York.
clusions: (1) there was no caldera collapse; (2) one major event of Clouard, V., Bonneville, A., Gillot, P.Y., 2001. A giant landslide on the southern flank
composite (two or three faults) flank collapse occurred between of Tahiti Island, French Polynesia. Geophys. Res. Lett. 28 (11), 2253–2256.
ca. 60 and 40 ka; (3) the fault most upstream (west) produced only Clouard, V., Bonneville, A., 2004. Submarine landslides in French Polynesia. In:
Hekinian, R., Stoffers, P., Cheminé, J.-L. (Eds.), Oceanic Hotspots: Intraplate Sub-
partial collapse, leaving a flat platform where renewed volcanism marine Magmatism and Tectonism. Springer-Verlag, pp. 209–238.
(including Pico do Fogo) occurred since at least ca. 20 ka; (4) the load Cornu, M.-N., 2017. Evolution magmatique d’un volcan bouclier océanique avant
of the new volcano and of the lavas filling the flat platforms (Chã et après une déstabilisation massive de ses flancs: Fogo, Cap Vert et Tenerife,
Canaries. Sciences de la Terre. PhD thesis. Université Clermont Auvergne, 295
das Caldeiras) can reactivate the two previous upstream faults and
pp. (in French).
produce a future massive landslide; (5) the on and offshore topog- Cornu, M.-N., Paris, R., Doucelance, R., Bachelery, P., Guillou, H., 2017. Exploring the
raphy hints at the existence of other collapses besides the eastern links between volcano flank collapse and magma evolution: fogo oceanic shield
volcano, Cape Verde. Geophysical Research Abstracts 19, EGU2017-4875-2.
Fogo collapse; (6) the increased number of collapses, the possibil-
Costa, A.C.G., Marques, F.O., Hildenbrand, A., Sibrant, A.L.R., Catita, C.M.S., 2014.
ity of a (near?) future collapse of Chã das Caldeiras, and the young Large-scale catastrophic flank collapses in a steep volcanic ridge: the Pico-Faial
and closely spaced ages here reported indicate a significant haz- Ridge, Azores Triple Junction. J. Volcanol. Geotherm. Res. 272, 111–125.
ard/risk for the inhabitants of Cape Verde and neighbouring coastal Costa, A.C.G., Hildenbrand, A., Marques, F.O., Sibrant, A.L.R., Santos de Campos, A.,
2015. Catastrophic flank collapses and slumping in Pico Island during the last
areas; (7) Fogo’s flank collapse (ca. 59 to 43 ka old) is unlikely to 130 kyr (Pico-Faial ridge, Azores Triple Junction). J. Volcanol. Geotherm. Res.
be responsible for the tsunami deposits described in Santiago and 302, 33–46.
F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686 13

Daniels, K.A., Menand, T., 2015. An experimental investigation of dyke injection (Eds.), Volcano Instability on the Earth and Other Planets. Special Publication.
under regional extensional stress. J. Geophys. Res. Solid Earth JB011627, http:// Geological Society of London, London, pp. 295–306.
dx.doi.org/10.1002/2014jb011627. Le Bas, T.P., Masson, D.G., Holtom, R.T., Grevemeyer, I., 2007. Slope failures of the
Day, S.J., Heleno Da Silva, S.I.N., Fonseca, J.F.B.D., 1999. A past giant lateral col- flanks of the southern Cape Verde Islands. In: Lykousis, V., Sakellariou, D., Locat,
lapse and present-day flank instability of Fogo, Cape Verde Islands. J. Volcanol. J. (Eds.), Submarine Mass Movements and Their Consequences: 3rd International
Geotherm. Res. 94, 191–218. Symposium. Springer, Netherlands, Dordrecht, pp. 337–345.
Deplus, C., Le Friant, A., Boudon, G., Komorowski, J.C., Villemenant, B., Hrford, C., Le Friant, A., Boudon, G., Deplus, C., Villemant, B., 2003. Large-scale flank collapse
Ségoufin, J., Cheminée, J.L., 2001. Submarine evidence for large-scale debris events during the activity of Montagne Pelée, Martinique, Lesser Antilles. J. Geo-
avalanches in the Lesser Antilles Arc. Earth Planet. Sci. Lett. 192, 145–157. phys. Res. 108 (B1), 2055.
Duffield, W.A., Stieljes, L., Varet, J., 1982. Huge landslide blocks in the growth of Piton Lipman, P.W., Normark, W.R., Moore, J.G., Wilson, J.B., Gutmacher, C.E., 1988. The
de la Fournaise, La Réunion and Kilauea volcano, Hawaii. J. Volcanol. Geotherm. giant Alika debris slide, Mauna Loa, Hawaii. J. Geophys. Res. 93, 4279–4299.
Res. 12, 147–160. Machado, F., 1965. Mechanism of Fogo volcano, Cape Verde Islands. Garcia de Orta
Druitt, T.H., Sparks, R.S.J., 1984. On the formation of calderas during ignimbrite Lisboa. 13, 51–56.
eruptions. Nature 310, 679–681. Machado, F., Torre de Assunção, C.F., 1965. Carta geológica de Cabo Verde na escala
Foeken, J.P.T., Day, S., Stuart, F.M., 2009. Cosmogenic 3He exposure dating of the Qua- de 1:100,000. Notícia explicativa da folha da ilha do Fogo – estudos petrográficos.
ternary basalts from Fogo, Cape Verdes: Implications for rift zone and magmatic Garcia de Orta (Lisboa) 13, 597–604.
reorganisation. Quat. Geochronol. 4 (1), 37–49. Madeira, J., Brum da Silveira, A., Mata, J., Mourão, C., Martins, S., 2008. The role of mass
Germa, A., Quidelleur, X., Gillot, P.Y., Tchilinguirian, P., 2010. Volcanic evolution of movements on the geomorphologic evolution of island volcanoes: examples
the back-arc Pleistocene Payun Matru volcanic field (Argentina). J. South Am. from Fogo and Brava in the Cape Verde archipelago. Comunicações Geológicas
Earth Sci. 29, 717–730. 95, 93–106.
Germa, A., Quidelleur, X., Lahitte, P., Labanieh, S., Chauvel, C., 2011. The K–Ar Marques, F.O., Hildenbrand, A., Hübscher, C., 2018. Evolution of a volcanic island on
Cassignol–gillot technique applied to western Martinique lavas: a record the shoulder of an oceanic rift and geodynamic implications: S. Jorge Island on
of Lesser Antilles arc activity from 2 Ma to Mount Pelée volcanism. Quat. the Terceira Rift, Azores Triple Junction. Tectonophysics 738, 41–50.
Geochronol. 6, 341–355. Martínez-Moreno, F.J., Monteiro Santos, F.A., Madeira, J., Pous, J., Bernardo, I.,
Gillot, P.Y., Cornette, Y., 1986. The Cassignol technique for potassium–argon dating, Soares, A., Esteves, M., Adão, F., Ribeiro, J., Mata, J., Brum da Silveira, A., 2018.
precision and accuracy – examples from the Late Pleistocene to recent volcanics Investigating collapse structures in oceanic islands using magnetotelluric sur-
from southern Italy. Chem. Geol. 59, 205–222. veys: the case of Fogo Island in Cape Verde. J. Volcanol. Geotherm. Res. 357,
Gillot, P.Y., Cornette, Y., Max, N., Floris, B., 1992. Two reference materials, trachytes 152–162.
MDO-G and ISH-G, for argon dating (K–Ar and Ar-40/Ar-39) of Pleistocene and Masson, D.G., 1996. Catastrophic collapse of the volcanic island of Hierro 15 ka ago
Holocene rocks. Geostandards Newsletter 16, 55–60. and the history of landslides in the Canary Islands. Geology 24, 231–234.
Gillot, P.Y., Lefèvre, J.C., Nativel, P.E., 1994. Model for the structural evolution of the Masson, D.G., Le Bas, T.P., Grevemeyer, I., Weinrebe, W., 2008. Flank collapse and
volcanoes of Rèunion Island. Earth Planet. Sci. Lett. 122, 291–302. largescale landsliding in the Cape Verde Islands, off West Africa. Geochem. Geo-
Gillot, P.Y., Hildenbrand, A., Lefevre, J.C., Albore-Livadie, C., 2006. The K-Ar dating phys. Geosystems 9 (7), Q07015.
method: principle, analytical techniques and application to Holocene volcanic Mitchell, N.C., 2003. Susceptibility of mid-ocean ridge volcanic islands and
eruptions in Southern Italy. Acta Vulcanol. 18, 55–66. seamounts to large-scale landsliding. J. Geophys. Res. 108, 2397, http://dx.doi.
Gooday, R.J., Brown, D.J., Goodenough, K.M., Kerr, A.C., 2018. A proximal record of org/10.1029/2002JB001997.
caldera-forming eruptions: the stratigraphy, eruptive history and collapse of the Moore, J.G., 1964. Giant submarine landslides on the hawaiian Ridge. In: Geological
Palaeogene Arran caldera, western Scotland. Bull. Volcanol. 80, 70, http://dx.doi. Survey Professional Paper 501-D., pp. D95–D98.
org/10.1007/s00445-018-1243-z. Moore, J.G., et al., 1989. Prodigious submarine landslides on the Hawaiian Ridge. J.
Gudmundsson, A., 1988. Formation of collapse calderas. Geology 16, 808–810. Geophys. Res. 94, 17465–17484.
Hildenbrand, A., Gillot, P.Y., 2006. Evidence for a differentiated ignimbritic activ- Paris, R., Giachetti, T., Chevalier, J., et al., 2011. Tsunami deposits in Santiago island
ity ending the building-stage of Tahiti-Nui (French Polynesia). Comptes Rendus (Cape Verde archipelago) as possible evidence of a massive flank failure of Fogo
Geosci. 338, 280–287. volcano. Sediment. Geol. 239, 129–145.
Hildenbrand, A., Gillot, P.-Y., Bonneville, A., 2006. Off shore evidence for a huge land- Polteau, S., Ferré, E.C., Planke, S., Neumann, E.-R., Chevallier, L., 2008. How are saucer-
slide of the northern flank of Tahiti-Nui (French Polynesia). Geochem. Geophys. shaped sills emplaced? Constraints from the Golden Valley Sill, South Africa. J.
Geosyst. 7, Q03006, http://dx.doi.org/10.1029/2005GC001003. Geophys. Res. 113, B12104, http://dx.doi.org/10.1029/2008JB005620.
Hildenbrand, A., Gillot, P.Y., Le Roy, I., 2004. Volcano-tectonic and geochemical evo- Quidelleur, X., Hildenbrand, A., Samper, A., 2008. Causal link between Quaternary
lution of an oceanic intra-plate volcano: tahiti-Nui (French Polynesia). Earth paleoclimatic changes and volcanic islands evolution. Geophys. Res. Lett. 35,
Planet. Sci. Lett. 217, 349–365. L02303.
Hildenbrand, A., Gillot, P.Y., Soler, V., Lahitte, P., 2003. Evidence for a persistent uplift- Rackzek, I., Stoll, B., Hofmann, A.W., Jochum, K.P., 2001. High-precision trace element
ing of La Palma (Canary Islands) inferred from morphological and radiometric data for the USGS reference materials BCR-1, BCR-2, BHVO-1, BHVO-2, AGV-1,
data. Earth Planet. Sci. Lett. 210, 277–289. AGV-2, DTS-1, DTS-2, GSP-1 and GSP-2 by ID-TIMS and MIC-SSMS. Geostandards
Hildenbrand, A., Madureira, P., Marques, F.O., Cruz, I., Henry, B., Silva, P., 2008. Multi- Newsletter 25, 77–86.
stage evolution of a sub-aerial volcanic ridge over the last 1.3 Myr: S. Jorge Island, Ramalho, R.S., Winckler, G., Madeira, J., Helffrich, G.R., Hipólito, A.R., Quartau, R.,
Azores Triple Junction. Earth Planet. Sci. Lett. 273, 289–298. Adena, K., Schaefer, J.M., 2015. Hazard potential of volcanic flank collapses raised
Hildenbrand, A., Marques, F.O., Catalão, J., 2018. Large-scale mass wasting on small by new megatsunami evidence. Sci. Adv. 1 (2015), e1500456.
volcanic islands revealed by the study of Flores Island (Azores). Sci. Rep. 8, 13898. Ricci, J., Quidelleur, X., Lahitte, P., 2015. Volcanic evolution of central Basse-Terre
Hildenbrand, A., Marques, F.O., Catalão, J., Catita, C.M.S., Costa, A.C.G., 2012a. Large- Island revisited on the basis of new geochronology and geomorphology data.
scale active slump of the southeastern flank of Pico Island. Azores. Geology 40, Bull. Volcanol. 77, 84, http://dx.doi.org/10.1007/s00445-015-0970-7.
939–942. Ricci, J., Quidelleur, X., Pallares, C., Lahitte, P., 2017. High-resolution K-Ar dating of
Hildenbrand, A., Marques, F.O., Costa, A.C.G., Sibrant, A.L.R., Silva, P.F., Henry, B., a complex magmatic system: the example of Basse-Terre Island (French West
Miranda, J.M., Madureira, P., 2012b. Reconstructing the architectural evolution Indies). J. Volcanol. Geotherm. Res. 345, 142–160.
of volcanic islands from combined K/Ar, morphologic, tectonic, and magnetic Rosi, M., Vezzoli, L., Aleotti, P., De Censi, M., 1996. Interaction between caldera
data: the Faial Island example (Azores). J. Volcanol. Geotherm. Res. 241-242, collapse and eruptive dynamics during the Campanian Ignimbrite eruption,
39–48. Phlegraean Fields, Italy. Bull. Volcanol. 57, 541–554.
Hildenbrand, A., Weis, D., Madureira, P., Marques, F.O., 2014. Recent plate Samper, A., Quidelleur, X., Boudon, G., Le Friant, A., Komorowski, J.C., 2008. Radio-
re-organization at the Azores Triple Junction: evidence from combined geo- metric dating of three large volume flank collapses in the Lesser Antilles arc. J.
chemical and geochronological data on Faial, S. Jorge and Terceira volcanic Volcanol. Geotherm. Res. 176, 485–492.
islands. Lithos 210-211, 27–39. Sibrant, A.L.R., Hildenbrand, A., Marques, F.O., Costa, A.C.G., 2015a. Volcano-tectonic
Holcomb, R.T., Searle, R.C., 1991. Large landslides from oceanic volcanoes. Mar. evolution of the Santa Maria Island (Azores): implications for paleostress evo-
Geotechnol. 10, 19–32. lution at the western Eurasia–nubia plate boundary. J. Volcanol. Geotherm. Res.
Hunt, J., Cassidy, M., Talling, P.J., 2018. Multi-stage volcanic island flank collapses 291, 49–62.
with coeval explosive caldera-forming eruptions. Sci. Rep. 8, 1146. Sibrant, A.L.R., Hildenbrand, A., Marques, F.O., Weiss, B., Boulesteix, T., Hübscher, C.,
Hunt, J.E., Wynn, R.B., Talling, P.J., Masson, D.G., 2013. Multistage collapse of eight Lüdmann, T., Costa, A.C.G., Catalão, J.C., 2015b. Morpho-structural evolution of a
western Canary Island landslides in the last 1.5 Ma: sedimentological and volcanic island developed inside an active oceanic rift: S. Miguel Island (Terceira
geochemical evidence from subunits in submarine flow deposits. Geochem. Rift, Azores). J. Volcanol. Geotherm. Res. 301, 90–106.
Geophys. Geosyst. 14, 2159–2181. Sibrant, A., Marques, F.O., Hildenbrand, A., 2014. Construction and destruction of a
Jordan, N.J., Rotolo, S.G., Williams, R., Speranza, F., McIntosh, W.C., Branney, M.J., volcanic island developed inside an oceanic rift: graciosa Island, Terceira Rift,
Scaillet, S., 2018. Explosive eruptive history of Pantelleria, Italy: repeated caldera Azores. J. Volcanol. Geotherm. Res. 284, 32–45.
collapse and ignimbrite emplacement at a peralkaline volcano. J. Volcanol. Steiger, R.H., Jager, E., 1977. Subcommission on geochronology: convention on the
Geotherm. Res. 349, 47–73. use of decay constants in geo and cosmochronology. Earth Planet. Sci. Lett. 36,
Krastel, S., Schmincke, H.-U., Jacobs, C.L., Rihm, R., Le Bas, T.P., Alibes, B., 2001. Sub- 359–362.
marine landslides around the Canary Islands. J. Geophys. Res. 106, 3977–3997. Tibaldi, A., Vezzoli, L., 2004. A new type of volcano flank failure: The resurgent caldera
Labazuy, P., 1996. Recurrent landslide events on the submarine flank of piton de sector collapse, Ischia, Italy. Geophys. Res. Lett. 31, L14605, http://dx.doi.org/10.
la fournaise volcano Reunion Island. In: McGuire, W.J., Jones, A.P., Neuberg, J. 1029/2004GL020419.
14 F.O. Marques, A. Hildenbrand, S.S. Victória et al. / Journal of Volcanology and Geothermal Research 388 (2019) 106686

Urgeles, R., Canals, M., Baraza, J., Alonso, B., Masson, D.G., 1997. The most recent Watts, A.B., Masson, D.G., 1995. A giant landslide on the north flank of Tenerife,
megaslides on the Canary Islands: the el golfo debris avalanche and the canary Canary Islands. J. Geophys. Res. 100, 24487–24498.
debris flow, west El Hierro Island. J. Geophys. Res. 102, 20305–20323. Woodhall, D., 1974. Geology and volcanic history of Pico Island Volcano, Azores.
Urgeles, R., Masson, D.G., Canals, M., Watts, A.B., Le Bas, T., 1999. Recurrent large- Nature 248, 663–665, http://dx.doi.org/10.1038/248663a0.
scale landsliding on the west flank of La Palma, Canary Islands. J. Geophys. Res.
104, 25331–25348.

You might also like