ArticlePDF Available

Abstract and Figures

Five tephra layers named BRH1 to 5 were sampled in an ice cliff located on the north-eastern flank of Mount Melbourne (northern Victoria Land, Antarctica). The texture, componentry, mineralogy, and major and trace element compositions of glass shards have been used to characterize these layers. These properties suggest that they are primary fall deposits produced from discrete eruptions that experienced varying degrees of magma/water interaction. The major and trace element glass shard analyses on single glass shards indicate that Mount Melbourne Volcanic Field is the source of these tephra layers and the geochemical diversity highlights that the eruptions were fed by compositionally diverse melts that are interpreted to be from a complex magma system with a mafic melt remobilizing more evolved trachy-andesitic to trachytic magma pockets. Geochemical compositions, along with textural and mineralogical data, have allowed correlations between two of the englacial tephra and distal cryptotephra from Mount Melbourne, recovered within a marine sediment core in the Edisto Inlet (~ 280 km northeast of Mount Melbourne), and constrain the age of these englacial tephra layers to between the third and the fourth century CE. This work provides new evidence of the intense historical explosive activity of the Mount Melbourne Volcanic Field and better constrains the rates of volcanism in northern Victoria Land. These data grant new clues on the eruptive dynamics and tephra dispersal, and considerably expand the geochemical (major and trace elements) dataset available for the Mount Melbourne Volcanic Field. In the future, this will facilitate the precise identification of tephra layers from this volcanic source and will help define the temporal and spatial correlation between Antarctic records using tephra layers. Finally, this work also yields new valuable time-stratigraphic marker horizons for future dating, synchronization, and correlations of different palaeoenvironmental and palaeoclimatic records across large regions of Antarctica.
Content may be subject to copyright.
Vol.:(0123456789)
1 3
Bulletin of Volcanology (2023) 85:39
https://doi.org/10.1007/s00445-023-01651-2
RESEARCH ARTICLE
Historical explosive activity ofMount Melbourne Volcanic Field
(Antarctica) revealed byenglacial tephra deposits
PaolaDelCarlo1· AlessioDiRoberto1 · GiuseppeRe1· PaulG.Albert2· VictoriaC.Smith3· GaetanoGiudice4·
GrazianoLarocca4· BiancaScateni1,5· AndreaCannata4,6
Received: 28 February 2023 / Accepted: 13 May 2023
© The Author(s) 2023
Abstract
Five tephra layers named BRH1 to 5 were sampled in an ice cliff located on the north-eastern flank of Mount Melbourne
(northern Victoria Land, Antarctica). The texture, componentry, mineralogy, and major and trace element compositions
of glass shards have been used to characterize these layers. These properties suggest that they are primary fall deposits
produced from discrete eruptions that experienced varying degrees of magma/water interaction. The major and trace
element glass shard analyses on single glass shards indicate that Mount Melbourne Volcanic Field is the source of these
tephra layers and the geochemical diversity highlights that the eruptions were fed by compositionally diverse melts that
are interpreted to be from a complex magma system with a mafic melt remobilizing more evolved trachy-andesitic to
trachytic magma pockets. Geochemical compositions, along with textural and mineralogical data, have allowed correlations
between two of the englacial tephra and distal cryptotephra from Mount Melbourne, recovered within a marine sediment
core in the Edisto Inlet (~ 280km northeast of Mount Melbourne), and constrain the age of these englacial tephra layers to
between the third and the fourth century CE. This work provides new evidence of the intense historical explosive activity
of the Mount Melbourne Volcanic Field and better constrains the rates of volcanism in northern Victoria Land. These
data grant new clues on the eruptive dynamics and tephra dispersal, and considerably expand the geochemical (major and
trace elements) dataset available for the Mount Melbourne Volcanic Field. In the future, this will facilitate the precise
identification of tephra layers from this volcanic source and will help define the temporal and spatial correlation between
Antarctic records using tephra layers. Finally, this work also yields new valuable time-stratigraphic marker horizons for
future dating, synchronization, and correlations of different palaeoenvironmental and palaeoclimatic records across large
regions of Antarctica.
Keywords Antarctica· Mount Melbourne Volcanic Field· Explosive eruptions· Englacial tephra· Glass geochemistry·
Historical eruptions
Editorial responsibility: J.L. Smellie
* Alessio Di Roberto
alessio.diroberto@ingv.it
Paola Del Carlo
paola.delcarlo@ingv.it
Giuseppe Re
giuseppe.re@ingv.it
1 Istituto Nazionale di Geofisica e Vulcanologia, Sezione di
Pisa, Via C. Battisti 53, 56125Pisa, Italy
2 Department ofGeography, Swansea University, Singleton
Park, SwanseaSA28PP, UK
3 Research Laboratory forArchaeology andtheHistory ofArt,
University ofOxford, 1 South Parks Road, OxfordOX13TG,
UK
4 Istituto Nazionale di Geofisica e Vulcanologia, Osservatorio
Etneo, Piazza Roma 2, 95125Catania, Italy
5 Dipartimento di Scienze Della Terra, Università di Pisa,
56126Pisa, Italy
6 Dipartimento di Scienze Biologiche, Geologiche e
Ambientali, Università di Catania, Corso Italia 57,
Catania95125, Italy
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 2 of 18
Introduction
Tephra in Antarctic ice records (ice cores or blue ice) provide
important sources of data for evaluating the history and evo-
lution of Antarctic explosive volcanism. The typical absence
of pyroclastic deposits at proximal locations on the flanks of
the volcanoes, due to high rates of post-depositional erosion
and/or remobilization and also (mainly) the extensive cover
of snow and ice obscuring deposits on the volcanic centres,
limits the reconstruction of past volcanic eruptions, including
their timing, scale, and impact on the environment. This gap
in knowledge can be overcome by the study of tephra and
cryptotephra (not visible, highly disseminated, fine-grained
tephra) at medial to distal sites, which are preserved and
remain unaltered within the snow, transformed into glacial
ice upon burial (Smellie 1999; Harpel etal. 2008; Iverson
etal. 2017). In polar regions, and particularly in Antarctica,
tephrochronological studies in ice cores and blue ice records
have been well developed over recent decades (e.g. Narcisi
etal. 2005; Kurbatov etal. 2006; Curzio etal. 2008; Iver-
son etal. 2014; Wolff etal. 2010; Narcisi etal. 2012; Severi
etal. 2012; Kim etal. 2020; Nardin etal. 2021). Along with
tephra included in marine sediment sequences (Hillenbrand
etal. 2008; Del Carlo etal. 2015; Di Roberto etal. 2019;
2020; 2021a; b; 2023), tephra deposits in englacial sequences
can help improve near-vent eruption records, in terms of both
frequency/chronology of explosive volcanism, and magma
chemical evolution through time, whilst facilitating advance-
ments in dating, correlating, and synchronizing high-resolu-
tion climate and atmospheric composition records through the
Late Quaternary (Narcisi and Petit 2021). In addition, tephro-
chronological correlations between different archives (e.g. ice
and marine records), located at proximal and distal sites from
an eruptive source, are greatly enhanced once a numerical
age is obtained for tephra. Specifically, the correlation of two
tephra layers allows the age-transfer from one site to another
by the simple use of stratigraphy, heightening tephrochronol-
ogy to an age-equivalent dating method (Lowe 2011).
Mount Melbourne is an active volcano in northern
Victoria Land, Antarctica, and future activity on the volcano
would pose a significant tephra fall risk to nearby scientific
stations, e.g. Mario Zucchelli Station (Italian), Jang Bogo
(Korea), Gondwana (Germany), and the new China Antarctic
research base (Fig.1a), as well as to Austral hemisphere air
traffic. Therefore, a good knowledge of the recent volcanic
activity of this volcano is important to better evaluate the
possible volcanic hazards (Geyer 2021).
The earliest investigations on Mount Melbourne began at
the end of the 1960s whereas the first geophysical observa-
tions started in 1988 by Italian National Antarctic Research
Program (PNRA); since then no eruptive activity has been
observed at this volcano. The age of the most recent eruption
is proposed by Lyon (1986) who performed the stable isotope
analysis of two snow profiles sampled on Mount Melbourne
at ca. 2000m of altitude and on the Campbell Glacier obtain-
ing snow accumulation rates of 0.5–2.2m/a. Using these data,
the author derived the age of englacial ash layers outcropping
in an ice cliff at ca. 1200m of altitude on the western slope
of Mount Melbourne and photographed in 1965 by Adam-
son and Cavaney (1967). They roughly estimated the depth
of snow from the surface to the first major ash layer to be
between 29 and 36m and, considering an accumulation rate
of 0.5 ± 0.16m/a, the last major eruption should have occurred
between 1862 and 1922 AD. Although this method has several
limitations (e.g. the lack of direct measurement of the snow
thickness, or the detailed study of snow profile aimed at iden-
tifying possible hiatus or erosion surfaces), it does provide an
estimate of the age of the latest eruptive activity of Mount Mel-
bourne. Unfortunately, the englacial tephra layers described
in Lyon (1986) and Adamson and Cavaney (1967) were never
sampled nor characterized in their textural and geochemical
features because the outcrop did not allow safe sampling.
During the 2017 austral summer, the ICE-VOLC project
(PNRA) facilitated new geological surveys around Mount Mel-
bourne which revealed the occurrence of multiple englacial
volcanic ash layers in an ice cliff located on the northeast flank
of the volcano at ca. 800m of altitude, very close to Baker
Rock location (Gambino etal. 2021; Fig.1). The better acces-
sibility of the outcrop allowed the sampling of tephra and the
measurements of the thickness between the different ash layers.
In this paper, we present data on the texture, mineral
phases, and major and trace element geochemical data per-
formed on single glass shards analyzed via Electron Probe
Microanalysis (EMPA) and Laser Ablation Inductively
Coupled Plasma Mass Spectrometry (LA-ICP-MS) of five
englacial tephra layers sampled for the first time from the
north-eastern flank of Mount Melbourne. Geochemical data
indicate that the tephra layers are derived from Mount Mel-
bourne explosive activity. The texture of the particles and
their morphological features were important for clues to the
style of the eruptions and fragmentation mechanisms. In
addition, geochemical fingerprints of englacial tephra have
allowed their correlations and dating with cryptotephra layers
recently found in marine sediments in Edisto Inlet near Cape
Hallett (Di Roberto etal. 2023), highlighting that Mount
Melbourne has been very active during historical times.
Geological setting
Mount Melbourne stands as a 2732-m-high stratovolcano
with a basal diameter of about 21–24km on the coast of
northern Victoria Land, between Wood Bay and Terra
Nova Bay, in Antarctica (Fig.1b). The Mount Melbourne
Bulletin of Volcanology (2023) 85:39
1 3
Page 3 of 18 39
volcano also includes several parasitic cones and second-
ary vents located on the flank of the main edifice (Smellie
etal. 2023). The main volcano edifice shows a gentle shape
with undissected flanks, apart from a possible slump scar
on the east side (Giordano etal. 2012), and a well-formed
ice-filled crater ca. 700m in diameter that Armienti etal.
(1991) interpreted as a summit caldera. It is largely covered
by snow and ice except for the summit region where rock
outcrops extend downslope on the east side to ca. 1800m.
The Mount Melbourne Volcanic Field (MMVF), together
with The Pleiades, Mount Overlord, Mount Rittmann, and
the Malta Plateau volcanoes, is part of the Melbourne Vol-
canic Province within the McMurdo Volcanic Group (Smel-
lie and Rocchi 2021). A synthesis of MMVF volcanic his-
tory was reported by Giordano etal. (2012) and Smellie
etal. (2023) based on stratigraphic and volcanological stud-
ies, geochemical data, and age determinations. According
to recent updates based on radioisotopic ages, the MMVF
volcanism began in the Late Miocene (c. 12.5Ma for centres
in the north by Tinker Glacier) but developed mainly from
the Pliocene (c. 4Ma; Smellie etal. 2023) particularly after
c. 3 million years ago (Rocchi and Smellie 2021 and refer-
ences therein). The activity can be subdivided into different
evolution stages: (i) the older Cape Washington shield vol-
cano (Late Miocene-Late Pliocene), (ii) the Random Hills
Period (Lower-Middle Pleistocene), (iii) the Shield Nunatak
Period (Middle Pleistocene), and (iv) the Mount Melbourne
Period (Upper Pleistocene-present; Giordano etal. 2012;
Smellie etal. 2023).
The oldest deposit that can be linked directly to an
eruption of the Mount Melbourne stratovolcano, re-dated
as 115ka by Smellie etal. (2023) at Edmonson Point, is
a trachytic ignimbrite. This was emplaced during a large
Plinian eruption and indicates the formation of a crustal
magma chamber during the recent Mount Melbourne stage
(Giordano etal. 2012). Subsequently, the Adelie Penguin
Rookery lava field succession was produced compris-
ing alkali basaltic, hawaiitic, and subordinate benmore-
itic lavas, scoria cones, and spatter cones that are dated at
90.7 ± 19.0ka. (Giordano etal. 2012).
Fig. 1 (a) Map of the volcanoes in northern Victoria Land. The
location of the studied outcrop is labelled with a red star whereas
the location of the marine sediment core TR17-08 in Edisto Inlet
(Di Roberto et al. 2023) and of the deep ice cores at Talos Dome
and Styx glacier are annotated with a green star and blue hexagons,
respectively. Green dots indicate the location of the permanent sci-
entific bases. The red box outlines the area of the topographic map in
(b) that illustrates Mount Melbourne and locations of the BRH sec-
tions. (c) Picture of the studied outcrop where the BRH tephra layers
are exposed, Mount Melbourne in the background
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 4 of 18
Evidence of more recent activity, based on tephrostrati-
graphic analysis and 40Ar-39Ar dating of proximal pyroclastic
sequence exposed on the Mount Melbourne summit, indi-
cates there were at least four Strombolian or Vulcanian style
to sub-Plinian/Plinian eruptions during the Late Pleistocene to
Holocene (Del Carlo etal. 2022). The most intense of these
eruptions was a sub-Plinian to Plinian eruption that yielded
an age < 17.8ka (13.5 ± 4.3ka). Furthermore, three trachytic
cryptotephra with glass compositions similar to Mount Mel-
bourne products were found intercalated in marine sediments of
Edisto Inlet, near Cape Hallett (Di Roberto etal. 2023). These
cryptotephra layers were interpreted as derived from historic
explosive eruptions of Mount Melbourne that occurred between
1615cal. years BP and 1677cal. years BP, i.e. between the third
and fourth century CE (Di Roberto etal. 2023).
Presently, the Mount Melbourne volcano is quiescent with
thermal anomalies, including noticeable steaming or fuma-
rolic activity both in the crater and on the north-western side
of the volcano. These emissions of steam and volcanic gas
have produced several ice towers and a complex network of
ice caves, as recently reported after geochemical surveys in
the frame of the ICE-VOLC project (Gambino etal. 2021).
Material andmethods
During the XXXII Italian Antarctic Expedition in 2017, the
area surrounding Mount Melbourne was aerially surveyed
by helicopter flights. A series of dark, sub-horizontal engla-
cial tephra layers were found exposed along a ca. 50-m-high
ice cliff located on the north-eastern flank of the volcano at
about 800m of altitude close to Baker Rock (74.24096 S,
164.72032 E; red star in Fig.1). Five sub-horizontal tephra
layers, named BRH1 to 5 (from the top to the bottom), were
distinguished and sampled by the Italian alpine climbing
guides (Fig.2). Tephra layers are between 10 and 20cm
thick and are each separated by 1.5 to 7.5m of ice (Fig.2).
Each sample is collected the “bulk” of the tephra layer and
represents the whole unit; unfortunately, due to the extreme
sampling conditions, it was not possible to sub-sample dif-
ferent stratigraphic heights within each unit. The tephra
samples were recovered still embedded in the ice, which
was then melted to recover the clastic fraction.
Sample preparation was carried out at laboratories of
Istituto Nazionale di Geofisica e Vulcanologia, Sezione di
Pisa (INGV-Pisa). Samples were washed with deionized
water in an ultrasonic bath to remove impurities, dried at
60°C, mounted with epoxy resin in 1-in. stubs, polished, and
prepared for textural and geochemical analyses. Textures,
components, and mineral assemblages of each tephra were
studied with an optical microscope and a scanning electron
microscope (SEM) Zeiss EVO MA and images of 3D silhou-
ettes and 2D cross-sections have been collected in secondary
(SE) and back-scattered (BSE) electrons mode, respectively.
The major and minor elements glass composition of samples
was determined using a JEOL 8600 wavelength-dispersive
electron microprobe equipped with four spectrometers at
the Research Laboratory for Archaeology and the History
of Art, the University of Oxford (operating conditions:
15-kV accelerating voltage, 6-nA beam current, and a beam
diameter of 10µm). The JEOL 8600 electron microprobe
was calibrated with a suite of appropriate mineral stand-
ards; peak count times were 30s for all elements except Mn
(40s), Na (12s), Cl (50s), and P (60s). The PAP absorption
correction method was used for quantification. Reference
glasses from the Max Planck Institute (MPI-DING suite;
Jochum etal. 2006) bracketing the possible chemistries
were also analyzed. These included felsic [ATHO-G (rhyo-
lite)], through intermediate [StHs6/80-G (andesite)] to mafic
[GOR132-G (komatiite)] glasses. All glass data have been
normalized to 100% for comparative purposes. Uncertainties
are typically < ± 0.8% RSD for Si; ~ ± 5% for most other
Fig. 2 Picture of the glacier
cliff where BRH tephra layers
are exposed, taken during the
sampling performed by the Ital-
ian alpine guides. On the picture
are annotated the five englacial
tephra and the ice thickness
between them
Bulletin of Volcanology (2023) 85:39
1 3
Page 5 of 18 39
major elements, except for the low abundance elements, for
instance Ti (~ ± 7%) and Mn (~ ± 30%).
Trace elements analysis of volcanic glass was per-
formed using an Agilent 8900 triple quadrupole ICP-MS
(ICP QQQ) coupled to a Resonetics 193nm ArF excimer
laser-ablation in the Department of Earth Sciences, Royal
Holloway, University of London. Full analytical procedures
used are reported in Tomlinson etal. (2010). Spot sizes 20
and 25mm were used depending on the vesicularity, crystal
content, and ultimately the size of available glass surfaces.
The repetition rate was 5Hz, with a count time of 40s on the
sample, and 40s on the gas blank to allow the subtraction of
the background signal. Blocks of eight or nine glass shards
and one MPI-DING reference glass were bracketed by the
NIST612 glass calibration standard (GeoREM 11/2006). In
addition, MPI-DING reference glasses were used to monitor
analytical accuracy (Jochum etal. 2005). The internal stand-
ard applied was 29Si (determined by the EMPA analysis).
Where individual shards were arranged into a grid formation
across an epoxy mount, they are given a unique row (A, B,
C) and shard number (1, 2, 3). In these instances, the grain-
specific 29Si content was applied as the internal standard.
Where tephra samples were mounted in epoxy mounts with-
out shard mapping, average 29Si contents for the appropriate
compositional groupings within the tephra deposit, deter-
mined based on EMPA analysis, were applied as the inter-
nal standard to the individual glass shards ablated. Internal
standard values applied to individual shard ablations are pro-
vided in the Supplementary Information. LA-ICP-MS data
reduction was performed in Microsoft Excel. Accuracies of
LA-ICP-MS analyses of MPI-DING glass standards ATHO-
G and StHs6/80-G were typically < 5%. Full glass datasets
and MPI-DING standard glass analyses are provided in Sup-
plemental Information.
Results
Texture andcomponents oftephra layers
BRH samples are similar in terms of components, compris-
ing a range of clasts with different textures ranging from
dense to highly vesicular and glassy to microlite-rich (Figs.3
and 4a–e). The morphology of juvenile clasts ranges from
equant with rounded vesicles (Fig.4f) to fluidal with elon-
gated tubular vesicles (Fig.4g, h). Dense poorly vesicular
juveniles display features such as stepped surfaces and
hackle lines, quenching cracks, and pitted surfaces (Fig.4l,
m), similar to the glassy shards that range from blocky to
platy (Fig.4i, n).
Overall, the grain size, scarcity of non-volcanic detrital
material, and unabraded pristine shapes of the clasts indicate
that these decimetre-thick layers are emplaced as primary
fallout deposits. Extremely fragile particles such as glass
fibres (Pele’s hair) survived, and the external morphology
of vesicular particles preserves pristine elements like fragile
glass fibres, glass tips, spiny glass edges (Figs.3 and 4), or
glass coatings around magmatic crystals. Particles do not
exhibit surficial alteration textures due to weathering, sur-
face abrasion, or rounding. Lithic or detrital fragments are
also scarce. All these features along with the geochemical
fingerprints (see next paragraph) indicate very minor or no
aeolian remobilization, re-sedimentation, or other transport
after the deposition from the volcanic plume. Moreover, the
layers occur close to the surface of the ice cover and well
above the glacier bed, hence are unlikely to be glacier bed-
load brought to the surface by shearing.
As mentioned in the “Material and methods” section, the
studied material comprises “bulk” samples of each tephra
layer. Possible internal stratifications resulting from varia-
tions in eruptive style, energy, dynamics, magma reflected
by changes in the glass composition, grain size, componen-
try, etc. were lost.
BRH1 tephra is the uppermost tephra exposed at the top
of the ice cliff. It is made of coarse to very fine ash (Fig.4a)
and comprises two discrete particle populations differing
in colour, shape, and texture. The first population, which is
the more abundant, consists of shiny, black to dark brown,
vesicular glassy clasts (Fig.3a and g) with spiny to fluidal
shapes (Fig.3d), volcanic glass fibres, and glass droplets
(Pele’s hair and tears; Fig.3c) and minor amounts of grey,
dense, and blocky clasts (Fig.3b). Often, particles with flu-
idal shapes are bounded by sharp-planar breakage surfaces
(Fig.3g). Fluidal clasts (Fig.3a) are usually glassy aphy-
ric to microlite-poor, poorly to moderately vesicular with
spherical vesicles. Blocky clasts (Fig.3b) are almost non-
vesicular to very poorly vesicular with abundant microlites
and microphenocrysts of andesine plagioclase (up to 60μm),
Fe-augite clinopyroxene (up to 30μm), olivine (~ Fo45; up
to 10μm), Fe-Ti oxides (up to 10μm), and rare apatite (up
to 10μm). Also, crystals (up to ca. 250μm) of plagioclase,
olivine, and clinopyroxene occur coated in a dark glass.
The second particle population represents approximately
30% of the sample and has a finer grain size (fine ash). Par-
ticles comprise white to pale yellow, highly vesicular pum-
ice fragments, and nearly transparent glass shards (Fig.4a).
Pumice fragments often show elongated to tubular vesicles
(Fig.3f), and bear within an aphyric groundmass abundant
labradorite plagioclase (up to 80μm) and olivine (~ Fo75;
up to 30μm) with rare Mg-augite clinopyroxene (up to
20μm) and Mg-chromite spinel (up to 10μm). Lithic clasts
are scarce in BRH1 tephra and consist of reddish or altered
volcanic rocks (Fig.3e).
BRH2 tephra is a poorly sorted lapilli (up to 4mm) and
ash (Fig.4b) layer. It is predominantly composed of black/
dark brown, shiny, vesicular glass fragments with spiny and
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 6 of 18
fluidal shapes again often bounded by sharp-planar break-
age surfaces (Fig.3d), Pele’s hair and tears, highly vesicu-
lar fragments, and grey dense, blocky clasts. Fluidal clasts
(Fig.4b) have glassy groundmass with spherical to elliptical
vesicles and contain abundant microlites of andesine to oli-
goclase plagioclase (up to 50μm), Fe-rich olivine (~ Fo35; up
to 15μm), Fe-augite clinopyroxene (up to 30μm), and Fe-Ti
oxides (~ 10mm). Grey dense and blocky clasts (Fig.4b) are
porphyritic with microphenocrysts of oligoclase to anortho-
clase feldspar, Fe-augite, and Fe-Ti oxides within a glassy to
microcrystalline groundmass. Tubular pumice fragments and
bubble wall glass shards are abundant in the fine-grained,
fine ash-sized portion of the deposit. Pumices have a glassy
groundmass sometimes with acicular microlites of anorthitic
feldspar. Scarce lithic clasts consist of reddish, altered vol-
canic rocks and intrusive rock fragments. Loose crystals of
anorthoclase, clinopyroxene, and fayalite are abundant.
BRH3 tephra (Fig.4c) is fine ash, with scarce scoriaceous
fine lapilli, made of black to dark brown, shiny, poorly to
moderately vesicular, fluidal glass particles (Fig.4g) and
grey, blocky fragments (Fig.4n). Dark glass-coated crystals
of plagioclase, clinopyroxene, and olivine are also abundant.
Fluidal clasts are moderately vesicular, mainly with rounded
tubular vesicles (Fig.4g), and have glassy groundmass with
phenocrysts of oligoclase plagioclase (up to 250μm) and Fe-
augite clinopyroxene (up to 350μm); accessory minerals are
anorthoclase feldspar (~ 90μm), olivine (~ Fo20; ~ 25μm),
Fe-Ti oxides (~ 15μm), and apatite (~ 45μm). Blocky frag-
ments are poorly vesicular with abundant microlites of
plagioclase, clinopyroxene, olivine, and Fe-Ti oxides. Pale
yellow to white fibrous pumiceous clasts are also abundant
in BRH3 tephra. Pumice fragments are mostly aphyric,
with rare microphenocrysts of plagioclase and Fe-Ti oxides
within a glassy highly vesicular groundmass with elongated
Fig. 3 Optical stereo-microscope images of the BRH4 ash. Images in the small boxes represent the different types of components: (A) vesicular;
(B) blocky; (C) Pele’s tear; (D) spiny fluidal; (E) lithic clast; (F) pumice; (G) vesicular fragment with sharp-planar breakage surface
Bulletin of Volcanology (2023) 85:39
1 3
Page 7 of 18 39
vesicles. Lithic clasts consist of red, oxidized lava fragments
and altered tuffaceous rock fragments.
BRH4 tephra (Fig.4d) is a very poorly sorted lapilli (up
to 2mm) to very fine ash layer. The sample is almost com-
pletely formed by black, shiny, spiny glass fragments and
Pele’s hair and tears, grey, dense and blocky clasts, and glass
fragments (Fig.4i). Golden-coloured, moderately vesicular
glassy fragments are also present. Fluidal clasts (Fig.4h) and
golden vesicular fragments have porphyritic texture, with
microphenocrysts of andesine plagioclase (~ 100μm), olivine
(~ Fo50; ~ 20μm), Fe-Ti oxides (~ 10μm), within a glassy
groundmass with rounded vesicles. The grey blocky parti-
cles (Fig.4d) are highly porphyritic with abundant micro-
phenocrysts of oligoclase plagioclase (~ 80μm), Fe-Ti oxides
Fig. 4 Back-scattered electron (BSE) images of the studied tephra:
(A) BRH1; (B) BRH2; (C) BRH3; (D) BRH4; (E) BRH5. Letters
annotated within these five pictures represent the different types of
components: B—blocky; F—fluidal; L—lithic clast; V—vesicular
pumice fragment. Secondary electron images (F–N) of 3D silhouettes
of selected clasts illustrating the morphology of magmatic (FH)
and phreatomagmatic (IN) juvenile clasts. Yellow arrows indicate
stepped features and hackle marks, blue arrows point at branching
quenching cracks, and the red arrow highlights pitted surfaces
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 8 of 18
(~ 45μm), and olivine (~ Fo45; ~ 40μm), and subordinate
anorthoclase/sanidine feldspar (~ 30μm), Fe-augite clino-
pyroxene (~ 40μm), and apatite (~ 25μm), within a microlite-
rich glassy groundmass. Many clasts show truncated shapes.
Loose crystals of plagioclase, clinopyroxene, and olivine are
abundant. Lithic clasts comprise lava and tuff fragments.
A minor quantity of whitish, spongy pumice clasts is also
present in the fine ash-sized fraction of the deposit. Pumice
particles are aphyric and often present altered groundmass.
BRH5 tephra (Fig.4e) is a very poorly sorted fine lapilli
(up to 2mm) to a fine ash layer. It comprises dark brown
to black and shiny, vesicular glass fragments with spiny
(Fig.4i), fluidal, and blocky shapes, glass shards, and a few
grey dense blocky clasts. Dense clasts are poorly vesicu-
lar, with spaced rounded to amoeboid vesicles, and por-
phyritic, with abundant microphenocrysts of labradorite
plagioclase (up to 140μm), olivine (~ Fo60; ~ 50μm), Fe-Ti
oxides (~ 30μm), within a glassy groundmass with abun-
dant microlites of plagioclase, Fe-rich olivine, Fe-augite
clinopyroxene (up to 15μm), and rare apatite (~ 10μm).
Clear to light grey pumice fragments are also abundant (ca.
20%). Pumices are highly vesicular with rounded and non-
collapsed to tubular and stretched vesicles (Fig.4e), bearing
rare microphenocrysts of acicular labradorite to andesine
plagioclase (up to 200mm) and Fe-rich olivine (~ 25μm)
within a glassy groundmass with rare microlites of plagio-
clase, clinopyroxene olivine, and Fe-Ti oxides. Some pum-
ice fragments, which have coarse amoeboid and collapsed
vesicles, have domains with glassy groundmass and domains
with microcrystalline ones (Fig.4e). Loose crystals of pla-
gioclase, clinopyroxene, and olivine are also present. Lithic
clasts include reddish to pinkish, oxidized pumices and rare
fragments of granitoid rocks.
Major, minor, andtrace elements glass
geochemistry
Representative major, minor, and trace element glass chem-
istry from the BRH 1 to 5 tephra layers are provided in
Table1, whilst the full datasets are provided in the Supple-
mentary Information.
The tephra layers analyzed (BRH1 to 5) display more
than one compositional component, consistent with their
variable textures and componentry; consequently, composi-
tions range from basalts, through basaltic trachy-andesites
and trachy-andesites to more evolved trachytes (Fig.5a).
Glasses are all characterized by high analytical totals sug-
gesting a very minor amount of post-depositional alteration
and/or reworking.
BRH1 to 5 volcanic glasses predominantly straddle the
high-K calc-alkaline and Shoshonitic boundary, with the
more evolved trachytic end-members residing more clearly
within the shoshonitic field (Fig. 5b). The absence of
negative anomalies at Nb and Ta in the mantle-normalized
trace element profiles of the BRH1 to 5 volcanic glasses is
clearly consistent with Antarctic alkaline regional volcanism
within the West Antarctic Rift System (Panter 2021), and the
absence of significant crustal involvement in magma genesis
(Fig.6). As is to be expected, overall levels of incompat-
ible trace element enrichment are significantly greater in
the trachytic end-member glasses of the tephra units relative
to their more mafic components (Fig.7). Light Rare Earth
Elements (LREE) are enriched relative to the Heavy Rare
Earth Elements (HREE), and the relative level of enrichment
remains fairly constant spanning from the basalts through
to the trachytes (La/YbN ~ 13–14). Trachytic end-member
glasses display depletions in Ba, Sr, and Eu consistent with
feldspar fractionation, with these anomalies absent in the
more primitive (basaltic) glasses. Owing to the significant
heterogeneity of the BRH tephra layers, we describe the
chemical signature of dominant chemical clusters or compo-
nents, and therefore rare outlying analyses are not described
in detail.
BRH1 (n = 18/24), BRH2 (n = 13/33), and BRH5
(n = 15/30) tephra layers all contain a significant mafic glass
component associated with the black glassy and fluidal par-
ticles, which display a relatively narrow compositional range
within the basalt compositional field (44.5 to 46.0 wt% SiO2,
and 3.9 to 5.3 Na2O + K2O), with minor overlap into the
basanite field (Fig.5a). More noticeable compositional vari-
ability in these basalts is observed in other major oxides, for
instance with 7.3–4.7 wt% MgO, 14–12 wt% FeOt, 12.1–9.7
wt% CaO, and 5.0–3.3 wt% TiO2, where this variability is
consistent between the basalts of all three layers (Fig.5).
Tephra layer BRH3 also contains dark glassy material entirely
consistent with the basaltic glasses of the BRH1, 2, and 5
tephra layers; however, these compositions are not as well
represented in the chemical analysis of the tephra (n = 10/74),
which are dominated by a trachytic end-member (Fig.5a).
The incompatible trace elements content of these basaltic
glasses, predominantly those successfully analyzed from the
BRH1 and BRH5 layers, reveal significant variability, for
instance 128–236ppm Zr, 15.3–28.0ppm Y, and 2.6–4.7ppm
Th (Fig.7). Ratios of High Field Strength Elements (HFSE)
vs. Th in these basaltic glasses show some variability (e.g. Zr/
Th = 46.3–54.3; Nb/Th = 14.2–15.5; Y/Th = 5.6–6.6). These
basaltic glasses display high LREE relative to the HREE where
La/YbN = 13.0 ± 0.9 (1 s.d). Despite the overall variability
within the basaltic glasses analyzed, they are chemically indis-
tinguishable through the successive tephra layers, suggesting a
common magmatic source.
BRH1 glasses, consistent with the observed compo-
nentry (see above), are chemically bimodal; in addition
to the dominant dark glassy basaltic component already
described, the tephra layer also contains a homogeneous
secondary trachytic component (SiO2 = 65.2 ± 0.1 wt% and
Bulletin of Volcanology (2023) 85:39
1 3
Page 9 of 18 39
Table 1 Representative major, minor, and trace elements glass data for the BRH1 to 5 tephra deposits investigated. Bas, basalt; Bas Tra-And, basaltic trachy-andesite; Tra-And, trachy-andesite;
Tra, trachyte. Full geochemical data sets are available in the Supplementary material file
BRH 1 BRH 2 BRH 3 BRH 4 BRH 5
TAS class Bas Tra Bas Bas Tra-And Tra-And Basalt Bas. Tra-And Tra Bas Tra-And Tra-And Tra Bas Tra Tra
Norm. wt%
SiO245.52 65.11 45.27 55.28 61.09 45.28 54.70 66.67 54.45 57.63 64.13 45.17 64.85 66.77
TiO24.83 0.50 4.64 2.03 0.99 3.95 2.21 0.35 1.97 1.69 0.84 4.11 0.46 0.40
Al2O314.31 15.87 14.90 15.15 15.08 14.70 14.39 15.54 15.43 14.94 14.73 14.92 15.85 15.33
FeOt 13.96 5.30 13.61 11.46 8.88 12.98 11.66 4.76 11.21 10.44 6.99 12.87 5.68 4.85
MnO 0.15 0.11 0.21 0.25 0.19 0.17 0.25 0.12 0.19 0.21 0.21 0.15 0.16 0.17
MgO 5.12 0.16 5.24 2.04 0.81 6.12 2.07 0.16 1.91 1.50 0.60 5.94 0.21 0.19
CaO 10.21 1.85 10.45 5.97 3.65 11.61 5.95 1.78 6.05 4.60 2.37 11.70 1.92 1.80
Na2O 3.61 5.91 3.67 4.37 5.33 3.34 5.19 5.47 4.82 4.73 5.10 3.31 5.62 5.41
K2O 1.26 4.98 1.20 2.56 3.55 1.21 2.71 4.91 2.90 3.56 4.66 1.12 4.99 4.85
P2O50.94 0.03 0.74 0.78 0.31 0.56 0.77 0.04 0.96 0.55 0.20 0.64 0.04 0.07
Cl 0.08 0.19 0.06 0.11 0.12 0.07 0.10 0.19 0.13 0.14 0.17 0.06 0.22 0.18
Ana. Total 99.16 99.54 99.19 99.62 99.04 98.36 99.69 99.13 98.75 98.19 99.51 98.69 99.02 99.75
Na2O + K2O 4.87 10.89 4.87 6.93 8.88 4.56 7.90 10.39 7.71 8.29 9.76 4.43 10.62 10.26
ppm
Rb 32.0 167.8 19.6 - 115.5 26.4 74.4 154.9 83.1 94.7 139.2 26.9 160.2 157.9
Sr 643 142 561 - 408 660 612 148 454 437 218 625 135 157
Y 27.0 54.4 15.3 - 47.5 23.8 42.9 51.8 48.8 44.1 46.7 23.0 56.8 54.0
Zr 222 687 128 - 504 197 373 685 449 469 605 189 709 711
Nb 68 165 37 - 127 59 102 152 118 120 141 54 165 157
Ba 371 1097 231 - 1052 328 831 1117 843 922 977 301 1044 1188
La 40.6 101.8 23.0 - 85.1 35.2 73.1 96.8 82.3 77.5 88.7 32.1 103.9 100.6
Ce 84.8 197.9 48.1 - 169.4 75.0 150.7 190.1 169.1 157.4 176.3 67.6 202.4 193.5
Pr 9.6 20.5 5.5 - 18.7 8.5 16.8 19.7 18.7 17.0 18.4 7.9 21.3 20.3
Nd 41.5 77.5 23.3 - 71.9 36.2 68.9 74.4 75.9 69.4 69.2 33.6 83.3 73.3
Sm 9.0 14.6 4.7 - 13.8 7.6 12.8 13.6 15.0 13.3 13.0 7.3 15.0 14.4
Eu 2.9 2.9 1.6 - 3.8 2.6 4.3 2.7 4.4 3.9 2.9 2.3 3.0 2.7
Gd 7.4 11.3 4.4 - 11.4 6.6 10.9 10.9 12.6 10.8 10.7 6.4 12.0 11.5
Dy 5.9 10.5 3.5 - 9.6 5.2 9.1 10.3 10.5 9.4 9.3 5.0 11.7 11.0
Er 2.7 5.7 1.6 - 4.7 2.5 4.3 5.4 5.0 4.5 4.7 2.5 6.1 5.8
Yb 2.3 5.4 1.2 - 4.3 1.8 3.6 5.5 4.2 4.3 4.7 2.0 5.6 5.4
Lu 0.3 0.7 0.2 - 0.6 0.3 0.5 0.7 0.6 0.6 0.6 0.3 0.8 0.8
Hf 5.3 15.1 3.1 - 10.8 4.7 8.6 15.4 10.3 10.5 13.2 4.5 16.2 16.1
Ta 4.1 9.2 2.2 - 6.7 3.6 5.9 8.6 6.6 6.8 7.7 3.2 9.1 8.9
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 10 of 18
Na2O + K2O = 10.78 ± 0.1 wt%; n = 5/23). The incompat-
ible trace element contents of these trachytic glasses (n = 4)
are relatively homogeneous, for instance 676 ± 40 ppm
Zr, 54 ± 4ppm Y, and 21 ± 2ppm Th, whilst these levels
of enrichment far exceed those of the dominant basaltic
end-member (Fig.7). Ratios of HFSE vs. Th become sig-
nificantly lower (Zr/Th = 33.0 ± 1.3; Nb/Th = 8.1 ± 0.5; Y/
Th = 2.6 ± 0.1 [1s.d]) than those of the basaltic component
(Fig.7). These evolved trachytic glasses display particularly
strong negative Sr anomalies (Sr/PrN = 0.09 ± 0.003), but
also less pronounced anomalies in Ba and Eu (Fig.6).
Overall, BRH2 glasses cover a wide compositional range,
consistent with the mixed componentry (see the “Texture
and components of tephra layers” section). As with BRH1,
the basaltic glass forms a significant component (1) of
BRH2. In addition, at least two other compositional compo-
nents or clusters are observed in BRH2, which plot along an
evolutionary trend that extends towards the highly evolved
trachytes, similar to the overlying BRH1 tephra (~ 65 wt%
SiO2). BRH2 glass component 2 straddles the boundary
between the basaltic trachy-andesite to trachy-andesite com-
positional fields with between 54.5 and 56.5 wt% SiO2 and
where Na2O + K2O = 6.9–8.4 wt% (Fig.5a). BRH2 compo-
nent 3 glasses are more variable and evolved than compo-
nent 2 predominantly clustering at the boundary between
the trachy-andesite and trachytic field, albeit with a small
number of analyses extending towards the more evolved tra-
chytic endmember of BRH1 (Fig.5a).
Trace element analysis of BRH2 component 2 (basaltic
trachy-andesite/trachy-andesite) glasses was unsuccessful
due to the presence of microlites, but trace element analyses
were acquired for glass component 3. The trachy-andesites
through to trachytes are equally heterogeneous at the trace
element level, with 388–720ppm Zr and 10.0–21.9ppm
Th. The glasses show depletions in Sr (Sr/PrN = 0.30 ± 0.1
[1s.d]) and Eu; however, these are not as pronounced as
those observed in the more evolved trachytes of BRH1
(Fig.6). HFSE ratios to Th in these glasses remain constant
despite the variability in absolute concentrations (Zr/Th
33.6 ± 1.6; Nb/Th = 8.4 ± 0.9; Y/Th = 3.1 ± 0.2 [1s.d]). The
Zr/Th and Nb/Th ratios are broadly consistent with the more
evolved (trachytic) glasses of the overlying BRH1, but again
are significantly lower than those of the basaltic glasses in
this tephra (and BRH1; Fig.7).
In addition to a basaltic component, the BRH3 tephra
has a minor population of intermediate glasses which strad-
dle the basaltic trachy-andesite to trachy-andesite composi-
tional fields, similar to component 2 of the overlying tephra
BRH2 (Fig.5a). A significant component of BRH3 analyses
(n = 51/73) lies firmly in the trachytic compositional field,
with 62.1–67.2 wt% SiO2 and 8.8–10.6 wt% Na2O + K2O
and relate to the pumiceous material within this tephra layer.
Within these variable trachytic compositions, a dominant
Table 1 (continued)
BRH 1 BRH 2 BRH 3 BRH 4 BRH 5
TAS class Bas Tra Bas Bas Tra-And Tra-And Basalt Bas. Tra-And Tra Bas Tra-And Tra-And Tra Bas Tra Tra
Pb 2.3 23.8 1.4 - 14.8 1.9 8.7 18.7 9.2 11.3 16.1 3.0 19.3 19.2
Th 4.4 21.0 2.6 - 14.6 3.9 10.8 20.8 12.2 13.0 17.2 3.7 22.0 22.1
U 1.4 5.0 0.7 - 3.4 1.2 2.5 4.4 2.9 3.1 4.1 1.3 5.0 4.6
Zr/Th 50.4 32.6 49.1 - 34.4 50.9 34.4 32.9 36.9 36.1 35.1 51.5 32.2 32.2
Nb/Th 15.4 7.8 14.3 - 8.7 15.2 9.4 7.3 9.7 9.2 8.2 14.7 7.5 7.1
Y/Th 6.1 2.6 5.9 - 3.2 6.1 4.0 2.5 4.0 3.4 2.7 6.3 2.6 2.4
Bulletin of Volcanology (2023) 85:39
1 3
Page 11 of 18 39
Fig. 5 Bivariate major elements plots depicting the significant vol-
canic glass heterogeneity of the BRH1 to 5 tephra deposits which
range from basalts through to trachytes. (A) Total alkalis vs. silica
classification diagram (TAS) following LeBas etal. (1986); (B) K2O
vs. SiO2 following Peccerillo and Taylor (1976). The trachytic glasses
of the BRH 1 to 5 tephra layers plot along the same evolution trend
as those previously analyzed from Mount Melbourne (Del Carlo etal.
2022), and are offset from glasses erupted at nearby volcanic centres,
such as Mount Rittmann (Di Roberto etal. 2019), the Pleiades (Lee
et al. 2019), and Mount Erebus (Harpel et al. 2008). Black dashed
line encircling the grey compositional field represents the TR17-08
tephra from Edisto Inlet and the black symbols indicate tephras from
Brimstone Peak (+ BIT121, × BIT122; Dunbar 2003)
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 12 of 18
cluster of shoshonitic glasses is clearly recognized at ~ 66
wt% SiO2, and ~ 4.9 K2O (Fig.5b), where Na2O is > K2O
(K2O/Na2O = 0.9). These most evolved trachytes (n = 16)
are extremely enriched in incompatible trace elements,
for instance 613–728ppm Zr and 18.3–22.1 ppm Th.
HFSE ratios vs. Th in these trachytic glasses remain con-
stant where Zr/Th = 33.0 ± 0.8, Nb/Th = 7.5 ± 0.3, and Y/
Th = 2.5 ± 0.1 (Fig.7). They also display depletions in Ba,
Sr (Sr/PrN = 0.10 ± 0.01), and Eu (Fig.6).
The glass compositions of BRH4 are heterogeneous,
largely bimodal, with some glasses extending between the
two dominant end-members. The least evolved glass compo-
nent plots in the basaltic trachy-andesite compositional field
(SiO2 = 53.8–55.7 wt% SiO2, Na2O + K2O = 7.0–8.3), whilst
straddling the boundary with trachy-andesites (Fig.5a).
The more evolved end-member is characterized by slightly
more variable trachytic glass compositions with 62.6–66.5
wt% SiO2 and 9.1–10.5 wt% Na2O + K2O (Fig. 5a) and
these relate to the pumice component of the deposit. The
basaltic trachy-andesitic glasses (n = 5) have some incom-
patible trace element variability, for instance 400–449ppm
Zr and 11.3–12.3ppm Th (Fig.7). The HFSE ratios in
Fig. 6 Primitive mantle normal-
ized (McDonough and Sun
1995) trace elements con-
centrations of volcanic glass
(component averages) for the
BRH1 to 5 tephra layers. The
data presented clearly illustrate
a chemical link between the
trachytic end-member glasses
of the BRH tephra layers and
those found at the summit of
Mount Melbourne (Del Carlo
etal. 2022). Bas—basalt; Tra-
Bas—trachy-basalt; Bas Tra-
And—basaltic trachy-andesite;
Tra-And—trachy-andesite;
Tra—trachyte
Fig. 7 Bivariate trace elements
plots, specifically (A) Th vs Zr,
(B) Th vs Nb, (C) Th vs Y, and
(D) Th vs Hf, depicting the signifi-
cant heterogeneity of the BRH1
to 5 tephra layers consistent with
their major elements variability
ranging from basalts to trachytes.
The plots illustrate the significant
chemical overlap of these succes-
sive tephra deposits, which clearly
lie upon an evolutionary trend
indistinguishable from the prod-
ucts of Mount Melbourne (mainly
trachytic), and offset from those of
the nearby Mount Rittmann (after
Del Carlo et al. 2022). The grey
compositional field is for Edisto
Inlet tephra TR17-08, whereas the
black symbols indicate the aver-
age composition of Brimstone
Peak tephras (+ BIT121, × BIT122;
Dunbar 2003)
Bulletin of Volcanology (2023) 85:39
1 3
Page 13 of 18 39
these glasses are constant (36.7 ± 0.2 Zr/Th, 9.6 ± 0.1 Nb/
Th, and 4.0 ± 0.02 Y/Th [1s.d]), whilst they show a small
anomaly in Sr (Sr/PrN = 0.4 ± 0.1 [1s.d]). The trachytic end-
member glasses (n = 5) are more enriched in incompatible
elements than the basaltic trachy-andesites (Fig.7), and
show considerable variability, for instance 472–604ppm
Zr and 13.8–17.2ppm Th. These trachytic glasses display
constant HFSE ratios (e.g. 34.7 ± 1.0 Zr/Th, 8.2 ± 0.4 Nb/
Th, and 2.8 ± 0.3 Y/Th), but are slightly lower than those
of the basaltic trachy-andesites. The trachytic glasses dis-
play a greater depletion in Sr (Sr/PrN = 0.3 ± 0.1), along with
additional depletions in Ba and Eu in contrast to the basaltic
trachy-andesites (Fig.6). However, the removal of Sr from
the melt composition of BRH4 is not as significant as for
some of the more evolved trachytic glasses observed in the
overlying tephra deposits (e.g. BRH1, BRH3; Fig.6).
Compositionally, the BRH5 tephra layer is largely bi-
modal, with some very minor evidence of intermediate
glass compositions spanning an evolutionary trend between
the two end-members, which are characterized by basal-
tic and trachytic glass populations (Fig.5a). The basaltic
component is entirely consistent with those observed and
described (above) from the BRH1, BRH2, and BRH3 tephra
layers (Figs.5, 6, and 7). Whilst the trachytic end-mem-
ber glasses are variable, they remain consistent with both
dominant trachytic components of BRH1 and BRH3, where
SiO2 contents of ~ 64.9 wt% and ~ 66.8 wt% respectively are
observed (Fig.5a). The incompatible trace element con-
tents of the BRH5 trachytic endmember (n = 5) glasses are
homogeneous with 737 ± 30ppm Zr and 23 ± 1ppm Th.
The levels of enrichment in some of these trachytic glasses
extend to the highest observed throughout this englacial suc-
cession (Fig.7). Furthermore, strong depletions in Ba, Sr
(Sr/PrN = 0.09 ± 0.01 [1s.d]), and Eu are observed in these
glasses (Fig.6). HFSE ratios to Th are constant (32.0 ± 0.3
Zr/Th, 7.3 ± 0.2 Nb/Th, 2.5 ± 0.1 Y/Th [1s.d]) and are
entirely consistent with the most evolved trachytic products
seen in the BRH1 and BRH3 layers.
Discussion
Volcanic source andproximal‑distal correlation
Major and trace elements glass composition data show a
strong geochemical affinity between the BRH tephra and
the products of MMVF, plotting along the same evolution-
ary trend. This compositional similarity and the proximity
of these tephra layers to MMVF suggest they were erupted
from its vents (Figs.5 and 7). BRH major and trace elements
compositions plot along the evolutionary trend defined by
volcanic glasses of MMVF (Figs.5 and 7), which is distinct
from other volcanoes in the region. However, there is no
direct match between the composition of BRH tephra and
that of late Pleistocene to Holocene tephra layers exposed
in the summit of Mount Melbourne, recently characterized
compositionally by Del Carlo etal. (2022) are character-
ized by generally higher alkali and lower silica contents. In
addition, Mount Melbourne summit proximal deposits lack
the basaltic end-member compositions which instead are
common in the BRH products except for the BRH4 tephra.
Thus, the BRH tephra layers are likely to result from erup-
tions at separate vent areas, possibly from parasitic cones
located on the flank of the volcano or at the periphery of the
Mount Melbourne volcanic complex. On the eastern flank
of Mount Melbourne, a few kilometres NNW of Edmon-
son Point ropey basaltic lava and a hawaiite scoria cone are
observed, known respectively as the ROL and SCC outcrops
of Giordano etal. (2012). These sources are close enough
and their deposits are compatible with the characteristics and
the eruption styles of BRH tephra. Also, the parasitic cones
NNW of the summit of Mount Melbourne, which have tra-
chytic compositions, and possibly scoria cones on the south
side of Random Hills further to the north (Smellie etal.
2023) could be considered possible sources. Unfortunately,
these only the eruptive centres close to Edmonson Point are
dated, and not precisely. They are only constrained to the last
90ka (Giordano etal. 2012; Smellie etal. 2023), hampering
direct correlations.
We also compared the composition of the BRH tephra
with visible tephra and cryptotephra layers identified in the
marine record of the Ross Sea and in the glacial records of
northern Victoria Land.
Interestingly, the composition of trachytic particle popu-
lation in BRH3 and BRH5, the deepest and middle tephra
respectively among those studied, correlates very well with
the composition of three cryptotephra recently identified
in the TR17-08 piston sediment core recovered in Edisto
Inlet, near Cape Hallett, more than 280km from Melbourne
volcano (Di Roberto etal. 2023; Fig.1). These cryptote-
phra layers, namely TR17-08-512, -518, and -524, have a
trachytic composition (Figs.5 and 7) and comprise col-
ourless to light-green glass shards and pumice fragments,
characterized by pristine textures among blocky, y-shaped,
and bubble wall morphology whilst preserving fragile glass
tips. Within core TR17-08, tephra age was defined using
radiocarbon dating of carbonate material, intercalated with
tephra, and was constrained between 1615cal. years BP and
1677cal. years BP, i.e. between the third and fourth century
AD (Di Roberto etal. 2023). These TR17-08 tephra layers
have been interpreted as produced by a series of explosive
eruptions from the Mount Melbourne volcanic complex that
occurred closely spaced in time, with a maximum interval of
c. 60years between events (Di Roberto etal. 2023).
The geochemical correlation defined between BRH proxi-
mal tephra and distal marine tephra allows the former to be
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 14 of 18
indirectly dated and consequently defines the onset (and an
overall age) of this eruptive period of Mount Melbourne
between the third and fourth century AD. The results con-
firm that Mount Melbourne has been very active in historical
times, with more explosive eruptive events than previously
thought.
For the sake of completeness, it should be said that some
differences exist between the BRH and TR17-08 cryptote-
phra. The most notable regards the lack of basaltic glass
populations in TR17-08 samples. The lower viscosity of
the denser melts typically results in less fragmentation and
lower plume heights of the basaltic phases that limit the
tephra dispersal. This is consistent with the significantly
larger grain size characterizing the basaltic ash (and lapilli)
in the studied tephra. The higher abundance of fine ash in the
trachytic end-member testifies to more efficient magmatic
fragmentation, and these finer particles are transported fur-
ther, especially those associated with the intense phases of
the eruption with high plumes. Moreover, whilst three cryp-
totephra have been identified within the TR17-08 marine
core, only two tephra with similar trachytic compositions
were recovered at the BRH site. Another layer may be pre-
served deeper in the ice sequence, below the BRH5 tephra,
or was eroded at the BRH site.
Furthermore, we explored correlations with tephra from
the ice core records of Styx Glacier (SG), Talos Dome (TD),
and the blue ice areas of Brimstone Peak (75.888S 158.55E;
BIT Dunbar etal. 2003), located at ca. 35km, 230km, and
250km ESE from the sampling site, respectively. Unfortu-
nately, most of the geochemical data for SG, TD, and BIT
tephra layers comprise only major element glass composi-
tion for which the accuracy is not clear, and lack trace ele-
ment glass data, which hamper reliable correlations with the
BRH tephra layers.
In spite of this, we found a major similarity between the
trachytic compositions of BRH4 tephra and the TD85 tephra
layer in the Talos Dome record dated at c. 670 yrs BP (1280
CE; Narcisi etal. 2012; Severi etal. 2012), but it is not a via-
ble correlation based on other tephrostratigraphic evidence.
In the Talos Dome record, the TD87a tephra correlates to the
1254 CE eruption of Mount Rittmann, and is located c. 2m
below the TD85 (Severi etal. 2012). Given that the 1254
CE tephra is ubiquitous in the marine and glacial records
of the Ross Sea and northern Victoria Land, a significant
and widespread tephra marker (Di Roberto etal. 2019 and
references therein), and it is not between BRH4 and BRH5
in the proximal exposure, we conclude that BRH4 does not
correlate to TD85.
A tephrostratigraphy similar to that of Talos Dome has
been developed for the Styx glacier ice core (which is the
ice record closer to the studied BRH site), where the 1254
CE tephra was detected at a depth of 99.18m and is only
overlaid by another tephra at 97.01m (Han etal. 2015),
instead of five ash layers.
Vice versa, our data confirm that TD85 likely derives
from Mount Melbourne volcanic activity, as previously
suggested by Narcisi etal. (2012), based on the major ele-
ment data. If this correlation would be verified using trace
element glass compositions, it would affirm that Mount
Melbourne was active in even more recent times, less than
700years ago. This further shows that incomplete geochemi-
cal databases strongly limit correlation efficacy, and high-
quality geochemical data are necessary to make reliable
correlations.
A general geochemical affinity also exists between BRH1,
BRH2, and BRH5 tephra and the composition of BIT vol-
canic ash (Dunbar 2003). In detail, there is a strong simi-
larity in major element compositions between the BRH1,
BRH2, and BRH5 basalt to basanite glass population and
the BIT121 and BIT122 ash layers (Fig.5). Trace element
compositions of BIT are not entirely determined, and the
same similarity can be only documented in the Th vs Hf
plot (Fig.7d). BIT121 and BIT122 layers are respectively
a thick dark brown unit made of clear, aphyric, and dense
blocky shards (BIT121) and a grey diffuse layer with clear
to olive green glass, blocky shards with spherical vesicles,
plus droplets and glass hairs. In particular, the latter has
strikingly similar textural characteristics to the BRH tephra.
Unfortunately, BIT volcanic ash layers lack chronological
and appropriate trace element information, and thus cannot
be definitively correlated with BRH.
Eruption dynamics
Componentry, textures, and geochemical fingerprints
of the studied tephra indicate that these layers have been
emplaced as primary fallout deposits. As mentioned above,
the external morphology of particles that preserves pristine
features like fragile glass tips, Pele’s hair, or spiny glass
edges (Figs.3 and 4) and glass coatings around magmatic
crystals indicate minor or no aeolian remobilization, re-
sedimentation, or other transport after deposition from the
volcanic plume.
The volcanic glass composition of each sample, which
is strongly bimodal with other analyses along geochemical
trends, always lies on the same evolutionary compositional
lineage indicating the same volcanic source (see Del Carlo
etal. 2018). The proximity of these decimetre-thick units to
the MMVF, and their shared geochemical affinity, is consist-
ently being from Mt. Melbourne or other vents around the
summit. The absence of large amounts of detrital material,
including glass particles from other volcanic sources, is evi-
dent and indicates negligible reworking, re-sedimentation,
and redistribution of pyroclastic products.
Bulletin of Volcanology (2023) 85:39
1 3
Page 15 of 18 39
Glass particles forming the studied deposits comprise a
variety of fragments differing in external shape, vesicular-
ity, mineral content, and composition. Internal and external
textural inhomogeneity is widely documented in the prod-
ucts of explosive eruptions from many volcanoes and has
been ascribed to changes in the ascent and/or flow conditions
during the eruption and to diverse mechanisms of particle
fragmentation (i.e. hydromagmatic vs magmatic), transport,
and the eruptive environment (see, e.g. D’Oriano etal. 2022
and references therein).
The almost ubiquitous presence of fluidal fragments
including Pele’s hair and tears (i.e. fibres and droplets of
volcanic glass; Figs.3 and 4) indicates specific eruptive
dynamics and magma properties typical of magmatic erup-
tions. Fluidal fragments and Pele’s hairs and tears formed
by the stretching, deformation, and final magma breakup
of very hot, low-viscosity magma due to interfacial shear
between gas and melt (see Lin and Reitz 1998; Eggers and
Villermaux 2008). Pele’s hairs and tears form during suba-
erial Hawaiian-style eruptions fed by low-viscosity basaltic
melts (Heiken and Wohletz 1985) and are also common, but
minor, in basaltic eruptions occurring in sub-aqueous envi-
ronments (Clague etal. 2000, 2003). Similarly, the presence
of abundant moderately to highly vesicular basaltic parti-
cles with cuspate to spongy external surfaces is indicative of
pure magmatic fragmentation of gas-rich melts dominated
by exsolution and expansion of magmatic gas, for example
in a Strombolian column.
More equant and blocky particles have irregular con-
tours, are bounded by sharp planar edges, and are made
of dense to poorly vesicular glass, bearing features like
stepped surfaces and quenching cracks, as well as hackle
lines and surface pitting (Fig.4i–l), which are distinctive
of phreatomagmatic fragmentation (Fisher and Schmincke
1984; Heiken and Wohletz 1985; Morrissey etal. 2000;
Dürig etal. 2012; Zimanowski etal. 2015; Ross etal.
2022). Stepped features are a consequence of the extreme
and intensive brittle fracturing of the melt (Zimanowski
etal. 2015). The water, in direct contact with the hot melt,
expands and exerts a massive hydraulic pressure onto the
melt itself, which behaves in a brittle way, so that cracks are
occurring, driven by the applied stress; still, liquid water
is pushed into the propagating crack, thus increasing the
interface area between magma and water and therefore
increasing the heat flux from magma to water, generating
an accelerating thermohydraulic feedback loop (Dürig and
Zimanowski 2012; Dürig etal. 2012). Hackle lines indicate
the direction of propagation of a crack and, since these can
form during slow or fast cracking, are not strictly diagnos-
tic of melt-water interaction, although some fragmentation
experiments with water produce increased proportions of
particles with hackle lines relative to others without water
(Ross etal. 2022). Quenching cracks form immediately
after fragmentation due to the sudden quenching and conse-
quent contraction of still-hot particles due to the fast contact
with liquid water (Zimanowski etal. 2015). Pitted surfaces
indicate incipient alteration resulting from the interaction
of glass with hydrothermal fluids (or excess water) in the
eruptive column (Zimanowski etal. 2015; Ross etal. 2022).
The simultaneous presence of particle morphologies
with features indicative of magmatic and hydromagmatic
fragmentation in BRH tephra layers is evidence that vari-
ous degrees of magma-water interaction occurred dur-
ing the forming eruptions. For example, texturally mixed
deposits could derive from the transformation of an initially
subglacial, hydromagmatic eruption fuelled by the interac-
tion between magma and meltwater, into a relatively dry
magmatic subaerial Hawaiian- to Strombolian-style erup-
tion, once the water source was exhausted and drained or
the magma no longer came into contact with the meltwater
(vent completely subaerial). As mentioned above, the major
and trace element glass compositions of the heterogeneous
BRH1 to 5 tephra deposits are largely overlapping, and
clearly reside on the same evolutionary trends indicating
they all likely derive from the same magmatic system. The
overall chemical variability (ranging from basalts through
to trachytes) and in particular the observed clustering of
the erupted volcanic glass compositions along the overall
evolutionary trend might indicate that the successive erup-
tions were fed by a complex and vertically extensive magma
storage region beneath the source volcano (Cashman etal.
2017; Giordano and Caricchi 2022). This is quite common
as reported for several eruptions (see Shane and Hoverd
2002;Shane etal.2007, 2008).
Interestingly, apart from BRH4, all the remaining BRH
tephra deposits comprise a basaltic glass component, and
the interaction of the mafic melt with more evolved tra-
chy-andesitic to trachytic magma pockets, or mush, is the
likely trigger of these eruptions. In this regard, it should be
highlighted that at least during some of the eruptive phases
emplacing BRH tephra, a trachytic magma was fragmented
efficiently enough to disperse the ash several hundred kil-
ometres away. This is possible only during high-intensity
eruptions able to produce eruptive columns of significant
heights, as hypothesized by Di Roberto etal. (2023).
Conclusions
BRH1 to 5 tephra are primary fallout layers derived from
five explosive eruptions that occurred in a short time from
within the Mount Melbourne Volcanic Field. The nature and
texture of particles forming the tephra layers indicate that
the eruptions were mostly Hawaiian to Strombolian in style
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 16 of 18
and characterized by magmatic and possibly hydromagmatic
fragmentation occurring over the time of the eruption. The
distal dispersal of trachytic ashes to some hundreds of kilo-
metres from the source (c. 280km) also demonstrates that
the BRH eruptions were at some time characterized by a
strong explosivity and very efficient magma fragmenta-
tion, and possibly produced several kilometre-high eruptive
columns.
The magma compositions feeding the eruptions range
from basalts to trachytes and the clustering of the erupted
product compositions indicates that the eruptions were fed
by a complex and vertically extensive magma system region
beneath the source volcano with the mafic melt remobilizing
more evolved trachy-andesitic to trachytic magma pockets.
Although the geochemical data confirm that the volcanic
source of BRH1 to 5 tephra is the Mount Melbourne vol-
cano, there is no direct match between the composition of
tephra studied here and Pleistocene to Holocene tephra lay-
ers exposed in the summit part of Mount Melbourne char-
acterized by Del Carlo etal. (2022; MEL samples). This
potentially indicates that BRH tephra originated from vents
located away from the summit area.
The geochemical correlations assessed between the BRH
tephra deposits and three marine cryptotephra intercalated
in sediments of the Edisto Inlet (TR17-08-512, -518, and
-524), which were dated by the radiocarbon method between
1615cal. years BP and 1677cal. years BP, allowed to indi-
rectly constrain the ages of BRH eruptions between the third
and fourth century CE. This finding improves our knowledge
of the eruptive history of Mount Melbourne volcano, which
has produced at least five eruptions during historical time,
increasing the awareness of potential hazards to the several
permanent scientific bases close to Mount Melbourne.
Moreover, the attribution of a numerical age to the BRH
tephra elevates them to new regional isochron markers that
will possibly enhance the correlation and synchronization
of climatic records of the northern Victoria Land and Ross
Sea areas.
This study demonstrates the robustness of the tephro-
chronological method in terms of correlation and dating of
glacial, terrestrial, and marine records, if carried out with a
modern approach and up-to-date techniques. The need for
high-quality textural, mineralogical, and compositional data
(major and trace element glass compositions) on tephra is
even more evident for Antarctica, where outcrops are scarce.
Only with an exhaustive fingerprinting of tephra layers is
possible to establish effective correlations between differ-
ent records enhancing synchronization and correlations for
palaeoenvironmental and palaeoclimatic reconstructions.
Supplementary information The online version contains supplemen-
tary material available at https:// doi. org/ 10. 1007/ s00445- 023- 01651-2.
Acknowledgements We acknowledge ENEA for providing field logis-
tics at Mario Zucchelli Station. We are grateful to the pilots J. Henery
and B. McElhinney for the helicopter surveys and the Italian alpine
guides M. Bussani and D. De Podestà for sampling the ice cliff and
their assistance in the fieldwork. Dr C. Manning is also acknowledged
for her assistance with the LA-ICP-MS analysis. We also thank the
editors J.L. Smellie and G. Giordano, and an anonymous reviewer for
their revisions that greatly improved the paper. This paper is dedicated
in memory of Antonio De Sio (Tony), an Antarctic friend of ours.
Author contribution P.D.C., A.C., G.G., and G.L. carried out the field-
work and sampling on the flanks of Mt. Melbourne. P.D.C. and A.D.R.
conceived the research. P.D.C., A.D.R., G.R., and B.S. carried out the
textural analysis of the volcanic glass and the petrographic analysis,
and interpreted geochemical and petrological data. P.G.A. and V.C.S.
carried out the major and trace element geochemical analyses, respec-
tively. All authors contributed to data interpretation, the writing of the
manuscript, and the preparation of the figures.
Funding Open access funding provided by Istituto Nazionale di Geofi-
sica e Vulcanologia within the CRUI-CARE Agreement. This work
was funded by the Projects: ICE-VOLC (multiparametrIC Experiment
at Antarctica VOLCanoes: data from volcano and cryosphere-ocean-
atmosphere dynamics, www. icevo lc- proje ct. com/; PNRA 14_00011),
TRACERS (TephRochronology and mArker events for the CorrElation
of natural archives in the Ross Sea, Antarctica; PNRA2016—Linea
A3/00055), and CHIMERA (CryptotepHra In Marine sEquences of
the Ross Sea, Antarctica: implications and potential applications;
PNRA18_00158-A). We acknowledge PNRA, the Italian Programma
Nazionale di Ricerche in Antartide, for funding the projects. This paper
is sponsored by the SCAR Expert Group, AntVolc.
Open Access This article is licensed under a Creative Commons Attri-
bution 4.0 International License, which permits use, sharing, adapta-
tion, distribution and reproduction in any medium or format, as long
as you give appropriate credit to the original author(s) and the source,
provide a link to the Creative Commons licence, and indicate if changes
were made. The images or other third party material in this article are
included in the article's Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in
the article's Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
References
Adamson RG, Cavaney RJ (1967) Volcanic debris-layers near Mount
Melbourne, Northern Victoria Land, Antarctica. New Zeal J Geol
Geophys 10:418–421. https:// doi. org/ 10. 1080/ 00288 306. 1967.
10426 745
Armienti P, Civetta L, Innocenti F, Manetti P, Tripodo S, Villari L,
Vita G (1991) New petrological and geochemical data on Mt.
Melbourne Volcanic Field, Northern Victoria Land, Antarctica.
(II Italian Antarctic Expedition). Mem Soc Geol It 46:397–424
Cashman KV, Sparks RSJ, Blundy JD (2017) Vertically extensive and
unstable magmatic systems: a unified view of igneous processes.
Science 355:eaag3055-11. https:// doi. org/ 10. 1126/ scien ce. aag30 55
Clague DA, Moore JG, Reynolds JR (2000) Formation of submarine
flat-topped volcanic cones in Hawai’i. Bull Volcanol 62:214–233.
https:// doi. org/ 10. 1007/ s0044 50000 088
Bulletin of Volcanology (2023) 85:39
1 3
Page 17 of 18 39
Clague DA, Batiza R, Head JW, Davis AS (2003) Pyroclastic and
hydroclastic deposits on Loihi Seamount, Hawaii. Explosive Sub-
aqueous Volcanism. AGU Geophys. Monogr 140:73–95. https://
doi. org/ 10. 1029/ 140GM 05
Curzio P, Folco L, Ada Laurenzi M etal (2008) A tephra chronostrati-
graphic framework for the Frontier Mountain blue-ice field (north-
ern Victoria Land, Antarctica). Quat Sci Rev 27:602–620. https://
doi. org/ 10. 1016/j. quasc irev. 2007. 11. 017
D’Oriano C, Del Carlo P, Andronico D, Cioni R, Gabellini P, Cristaldi
A, Pompilio M (2022) Syn-eruptive processes during the Janu-
ary–February 2019 ash-rich emissions cycle at Mt. Etna (Italy):
implications for petrological monitoring of volcanic ash. Front
Earth Sci 10:824872. https:// doi. org/ 10. 3389/ feart. 2022. 824872
Del Carlo P, Di Roberto A, Di Vincenzo G, Bertagnini A, Landi P,
Pompilio M, Colizza E, Giordano G (2015) Late Pleistocene-
Holocene volcanic activity in northern Victoria Land recorded
in Ross Sea (Antarctica) marine sediments. Bull Volcanol 77:36.
https:// doi. org/ 10. 1007/ s00445- 015- 0924-0
Del Carlo P, Di Roberto A, D’Orazio M, Petrelli M, Angioletti A,
Zanchetta G, Maggi V, Daga R, Nazzari M, Rocchi S (2018) Late
Glacial-Holocene tephra from southern Patagonia and Tierra del
Fuego (Argentina, Chile): a complete textural and geochemical
fingerprinting for distal correlations in the Southern Hemisphere.
Quatern Sci Rev 195:153–170. https:// doi. org/ 10. 1016/j. quasc irev.
2018. 07. 028
Del Carlo P, Di Roberto A, Di Vincenzo G, Re G, Albert PG, Nazzari
M, Smith VC, Cannata A (2022) Tephrostratigraphy of proximal
pyroclastic sequences at Mount Melbourne (northern Victoria
Land, Antarctica): insights into the volcanic activity since the
last glacial period. J Volcanol Geotherm Res 422:107457. https://
doi. org/ 10. 1016/j. jvolg eores. 2021. 107457
Di Roberto A, Colizza E, Del Carlo P, Petrelli M, Finocchiaro F, Kuhn
G (2019) First marine cryptotephra in Antarctica found in sedi-
ments of the western Ross Sea correlates with englacial tephras
and climate records. Sci Rep 9:10628. https:// doi. org/ 10. 1038/
s41598- 019- 47188-3
Di Roberto A, Albert PG, Colizza E, Del Carlo P, Di Vincenzo G, Gal-
lerani A, Giglio F, Kuhn G, Macrì P, Manning CJ, Melis R, Mis-
erocchi S, Scateni B, Smith VC, Torricella F, Winkler A (2020)
Evidence for a large-magnitude Holocene eruption of Mount
Rittmann (Antarctica): a volcanological reconstruction using the
marine tephra record. Quat Sci Rev 250:106629. https:// doi. org/
10. 1016/j. quasc irev. 2020. 106629
Di Roberto A, Scateni B, Di Vincenzo G, Petrelli M, Fisauli G,
Barker SJ, Del Carlo P, Colleoni F, Kulhanek DK, McKay R, De
Santis L (2021) Tephrochronology and provenance of an Early
Pleistocene (Calabrian) tephra from IODP Expedition 374 Site
U1524, Ross Sea (Antarctica). Geochem Geophys Geosystems
22:e2021GC009739. https:// doi. org/ 10. 1029/ 2021G C0097 39
Di Roberto A, Del Carlo P, Pompilio M (2021b) Chapter6.1 Marine
record of Antarctic volcanism from drill cores. Geol Soc London
Mem 55:631 LP – 647. https:// doi. org/ 10. 1144/ M55- 2018- 49
Di Roberto A, Re G, Scateni B, Petrelli M, Tesi T, Capotondi L, Morigi
C, Galli G, Colizza E, Melis R, Torricella F, Giordano P, Giglio
F, Gallerani A, Gariboldi K (2023) Cryptotephras in the marine
sediment record of the Edisto Inlet, Ross Sea: implications for
the volcanology and tephrochronology of northern Victoria Land,
Antarctica. Quat Sci Adv 100079. https:// doi. org/ 10. 1016/j. qsa.
2023. 100079
Dunbar NW, Zielinski GA, Voisins DT (2003) Tephra layers in the
Siple Dome and Taylor Dome ice cores, Antarctica: sources and
correlations. J Geophys Res Solid Earth 108:2374. https:// doi. org/
10. 1029/ 2002J B0020 56
Dunbar N (2003) Blue Ice Tephra II - Brimstone Peak. U.S. Antarctic
Program (USAP) Data Center. https:// doi. org/ 10. 7265/ N5MG7
MDK
Dürig T, Zimanowski B (2012) “Breaking news” on the formation
of volcanic ash: fracture dynamics in silicate glass. Earth Planet
Sci Lett 335–336:1–8. https:// doi. org/ 10. 1016/j. epsl. 2012. 05. 001
Dürig T, Mele D, Dellino P, Zimanowski B (2012) Comparative analy-
ses of glass fragments from brittle fracture experiments and vol-
canic ash particles. Bull Volcanol 74:691–704. https:// doi. org/ 10.
1007/ s00445- 011- 0562-0
Eggers J, Villermaux E (2008) Physics of liquid jets. Reports Prog
Phys 71:36601. https:// doi. org/ 10. 1088/ 0034- 4885/ 71/3/ 036601
Fisher RV, Schmincke J-U (1984) Pyroclastic rocks. Springer-Verlag,
Berlin. https:// doi. org/ 10. 1007/ 978-3- 642- 74864-6
Gambino S, Armienti P, Cannata A, Del Carlo P, Giudice G, Giuffrida
G, Liuzzo M, Pompilio M (2021) Chapter7.3 Mount Melbourne
and Mount Rittmann. Geol Soc London, Mem 55:M55–2018–43.
https:// doi. org/ 10. 1144/ M55- 2018- 43
Geyer A (2021) Chapter1.4 Antarctic volcanism: active volcanism
overview. Geol Soc London, Mem 55:M55–2020–12. https:// doi.
org/ 10. 1144/ M55- 2020- 12
Giordano G, Caricchi L (2022) Determining the state of activity of
transcrustal magmatic systems and their volcanoes. Annu Rev
Earth Planet Sci 50(2022):231–259. https:// doi. org/ 10. 1146/ annur
ev- earth- 032320- 084733
Giordano G, Lucci F, Phillips D, Cozzupoli D, Runci V (2012) Stra-
tigraphy, geochronology and evolution of the Mt. Melbourne
volcanic field (North Victoria Land, Antarctica). Bull Volcanol
74:1985–2005. https:// doi. org/ 10. 1007/ s00445- 012- 0643-8
Han Y, Jun SJ, Miyahara M, Lee H, Ahn J, Chung JW, Hur SD, Hong
SB (2015) Shallow ice-core drilling on Styx glacier, northern
Victoria Land, Antarctica in the 2014–2015 summer TT. J Geol
Soc Korea 51(3):343–355. https:// doi. org/ 10. 14770/ jgsk. 2015.
51.3. 343
Harpel CJ, Kyle PR, Dunbar NW (2008) Englacial tephrostratigraphy
of Erebus volcano, Antarctica. J Volcanol Geotherm Res 177:549–
568. https:// doi. org/ 10. 1016/j. jvolg eores. 2008. 06. 001
Heiken G, Wohletz K (1985) Volcanic ash. University of California
Press, Berkeley
Hillenbrand C, Moreton SG, Caburlotto A, Pudsey CJ, Lucchi RG,
Smellie JL, Benetti S, Grobe H, Hunt JB, Larter RD (2008) Vol-
canic time-markers for Marine Isotopic Stages 6 and 5 in South-
ern Ocean sediments and Antarctic ice cores: implications for
tephra correlations between palaeoclimatic records. Quatern Sci
Rev 27:518–540. https:// doi. org/ 10. 1016/j. quasc irev. 2007. 11. 009
Iverson NA, Kyle PR, Dunbar NW, McIntosh WC, Pearce NJG (2014)
Eruptive history and magmatic stability of Erebus volcano, Ant-
arctica: insights from englacial tephra. Geochem Geophys Geo-
systems 15:4180–4202. https:// doi. org/ 10. 1002/ 2014G C0054 35
Iverson NA, Kalteyer D, Dunbar NW, Kurbatov A, Yates M (2017)
Advancements and best practices for analysis and correlation of
tephra and cryptotephra in ice. Quat Geochronol 40:45–55. https://
doi. org/ 10. 1016/j. quageo. 2016. 09. 008
Jochum KP, Pfänder J, Woodhead JD, Willbold M, Stoll B, Herwig
K, Amini M, Abouchami W, Hofmann AW (2005) MPI-DING
glasses: new geological reference materials for insitu Pb isotope
analysis. Geochem Geophys Geosystems 6(10). https:// doi. org/
10. 1029/ 2005G C0009 95
Jochum KP, Stoll B, Herwig K, Willbold M, Hofmann AW, Amini M,
Aarburg S, Abouchami W, Hellebrand E, Mocek B, Raczek I,
Stracke A, Alard O, Bouman C, Becker S, Dücking M, Brätz H,
Klemd R, de Bruin D, Canil D, Cornell D, de Hoog C, Dalpé C,
Danyushevsky L, Eisenhauer A, Gao Y, Snow JE, Groschopf N,
Günther D, Latkoczy C, Guillong M, Hauri E, Höfer HE, Lahaye
Bulletin of Volcanology (2023) 85:39
1 3
39 Page 18 of 18
Y, Horz K, Jacob DE, Kasemann SA, Kent AJR, Ludwig T, Zack
T, Mason PRD, Meixner A, Rosner M, Misawa K, Nash BP,
Pfänder J, Premo WR, Sun WD, Tiepolo M, Vannucci R, Venne-
mann T, Wayne D, Woodhead JD (2006) MPI-DING reference
glasses for insitu microanalysis: new reference values for element
concentrations and isotope ratios. Geochem Geophys Geosystems
7(2). https:// doi. org/ 10. 1029/ 2005G C0010 60
Kim D, Prior DJ, Han Y, Qi C, Han H, Ju HT (2020) Microstructures
and fabric transitions of natural ice from the Styx Glacier, North-
ern Victoria Land. Antarct Miner 10:892. https:// doi. org/ 10. 3390/
min10 100892
Kurbatov AV, Zielinski GA, Dunbar NW, Mayewski PA, Meyerson
EA, Sneed SB, Taylor KC (2006) A 12,000 year record of explo-
sive volcanism in the Siple Dome Ice Core, West Antarctica. J
Geophys Res Atmos 111:D12307. https:// doi. org/ 10. 1029/ 2005J
D0060 72
LeBas MJL, Maitre RWL, Streckeisen A, Zanettin B, IUGS Subcom-
mission on the Systematics of Igneous Rocks (1986) A chemical
classification of volcanic rocks based on the total alkali-silica
diagram. J Petrol 27:745–750. https:// doi. org/ 10. 1093/ petro logy/
27.3. 745
Lee MJ, Kyle PR, Iverson NA, Lee JI, Han Y (2019) Rittmann volcano,
Antarctica as the source of a widespread 1252 ± 2 CE tephra layer
in Antarctica ice. Earth Planet Sci Lett 521:169–176. https:// doi.
org/ 10. 1016/j. epsl. 2019. 06. 002
Lin SP, Reitz RD (1998) Drop and spray formation from a liquid jet.
Ann Rev Fluid Mech 30:85–105. https:// doi. org/ 10. 1146/ annur
ev. fluid. 30.1. 85
Lowe DJ (2011) Tephrochronology and its application: a review. Quat
Geochronol 6:107–153. https:// doi. org/ 10. 1016/j. quageo. 2010. 08. 003
Lyon GL (1986) Stable isotope stratigraphy of ice cores and the age
of the last eruption at Mount Melbourne, Antarctica. New Zeal
J Geol Geophys 29:135–138. https:// doi. org/ 10. 1080/ 00288 306.
1986. 10427 528
McDonough WF, Sun S-S (1995) The composition of the Earth.
Chem Geol 120:223–253. https:// doi. org/ 10. 1016/ 0009- 2541(94)
00140-4
Morrissey M, Zimanowski B, Wohletz K, Buettner R (2000) Phreato-
magmatic fragmentation. In: Sigurdsson H, Houghton B, McNutt
SR, Rymer H, Stix J (eds) Encyclopedia of volcanoes. Academic,
London, pp 431–445
Narcisi B, Petit JR, Delmonte B, Basile-Doelsch I, Maggi V (2005)
Characteristics and sources of tephra layers in the EPICA-Dome
C ice record (East Antarctica): implications for past atmospheric
circulation and ice core stratigraphic correlations. Earth Planet
Sci Lett 239:253–265. https:// doi. org/ 10. 1016/j. epsl. 2005. 09. 005
Narcisi B, Petit JR, Delmonte B, Scarchilli C, Stenni B (2012) A
16,000-yr tephra framework for the Antarctic ice sheet: a contri-
bution from the new Talos Dome core. Quat Sci Rev 49:52–63.
https:// doi. org/ 10. 1016/j. quasc irev. 2012. 06. 011
Narcisi B, Petit JR (2021) Chapter6.2 Englacial tephras of East Ant-
arctica. Geol Soc London, Mem 55:649 LP-664. https:// doi. org/
10. 1144/ M55- 2018- 86
Nardin R, Severi M, Amore A, Becagli S, Burgay F, Caiazzo L, Ciar-
dini V, Dreossi G, Frezzotti M, Hong S-B, Khan I, Narcisi BM,
Proposito M, Scarchilli C, Selmo E, Spolaor A, Stenni B, Traversi
R (2021) Dating of the GV7 East Antarctic ice core by high-reso-
lution chemical records and focus on the accumulation rate vari-
ability in the last millennium. Clim past 17:2073–2089. https://
doi. org/ 10. 5194/ cp- 17- 2073- 2021
Panter KS (2021) Antarctic volcanism: petrology and tectonomagmatic
overview. Geol Soc Lond Mem 55:43–53. https:// doi. org/ 10. 1144/
M55- 2020- 10
Peccerillo A, Taylor SR (1976) Geochemistry of eocene calc-alkaline
volcanic rocks from the Kastamonu area, Northern Turkey. Con-
trib Miner Petrol 58:63–81. https:// doi. org/ 10. 1007/ BF003 84745
Rocchi S, Smellie JL (2021) Chapter5.1b Northern Victoria Land:
petrology. Geol Soc London Mem 55:383 LP – 413. https:// doi.
org/ 10. 1144/ M55- 2019- 19
Ross P-S, Dürig T, Comida PP, Lefebvre N, White JDL, Andronico
D, Thivet S, Eychenne J, Gurioli L (2022) Standardized analysis
of juvenile pyroclasts in comparative studies of primary magma
fragmentation; 1. Overview and Workflow. Bull Volcanol 84:13.
https:// doi. org/ 10. 1007/ s00445- 021- 01516-6
Severi M, Udisti R, Becagli S, Stenni B, Traversi R (2012) Volcanic
synchronisation of the EPICA-DC and TALDICE ice cores for
the last 42 kyr BP. Clim past 8:509–517. https:// doi. org/ 10. 5194/
cp-8- 509- 2012
Shane P, Hoverd J (2002) Distal record of multi-sourced tephra in One-
poto Basin, Auckland, New Zealand: implications for volcanic
chronology, frequency and hazards. Bull Volcanol 64:441–454.
https:// doi. org/ 10. 1007/ s00445- 002- 0217-2
Shane P, Martin SB, Smith VC, Beggs KF, Darragh MB, Cole JW,
Nairn IA (2007) Multiple rhyolite magmas and basalt injection in
the 17.7 ka Rerewhakaaitu eruption episode from Tarawera vol-
canic complex. New Zealand J Volcanol Geotherm Res 164:1–26.
https:// doi. org/ 10. 1016/j. jvolg eores. 2007. 04. 003
Shane P, Nairn IA, Smith VC, Darragh M, Beggs K, Cole JW (2008) Silicic
recharge of multiple rhyolite magmas by basaltic intrusion during the
22.6 ka Okareka Eruption Episode, New Zealand. LITHOS 103:527–
549. https:// doi. org/ 10. 1016/j. lithos. 2007. 11. 002
Smellie JL (1999) The upper Cenozoic tephra record in the south polar
region: a review. Glob Planet Change 21:51–70. https:// doi. org/
10. 1016/ S0921- 8181(99) 00007-7
Smellie JL, Rocchi S, Di Vincenzo G (2023) Controlling influence
of water and ice on eruptive style and edifice construction in the
Mount Melbourne Volcanic Field (northern Victoria Land, Ant-
arctica). Front Earth Sci - Volcanol 10:1061515. https:// doi. org/
10. 3389/ feart. 2022. 10615 15
Smellie JL, Rocchi S (2021) Northern Victoria Land: volcanology.
Geological Society, London, Memoirs, 55. https:// doi. org/ 10.
1144/ M55- 2018- 60
Tomlinson EL, Thordarson T, Müller W, Thirlwall M, Menzies MA
(2010) Microanalysis of tephra by LA-ICP-MS - strategies, advan-
tages and limitations assessed using the Thorsmörk ignimbrite
(Southern Iceland). Chem Geol 279:73–89. https:// doi. org/ 10.
1016/j. chemg eo. 2010. 09. 013
Wolff EW, Barbante C, Becagli S, Bigler M, Boutron CF, Castellano
E, de Angelis M, Federer U, Fischer H, Fundel F, Hansson M,
Hutterli M, Jonsell U, Karlin T, Kaufmann P, Lambert F, Littot
GC, Mulvaney R, Röthlisberger R, Ruth U, Severi M, Siggaard-
Andersen ML, Sime LC, Steffensen JP, Stocker TF, Traversi R,
Twarloh B, Udisti R, Wagenbach D, Wegner A (2010) Changes in
environment over the last 800,000 years from chemical analysis
of the EPICA Dome C ice core. Quat Sci Rev 29(1–2):285–295.
https:// doi. org/ 10. 1016/j. quasc irev. 2009. 06. 013
Zimanowski B, Büttner R, Delino P, White JDL, Wohletz K (2015)
Magma-water interaction and phreatomagmatic fragmentation. In:
Sigurdsson H, Houghton B, McNutt SR, Rymer H, Stix J (eds)
Encyclopedia of Volcanoes, 2nd edn. Academic Press, London,
pp 473–484
Article
Monitoring active volcanoes is necessary to analyze their current status to pose a mitigation hazard. Mount Melbourne is an active volcano that has erupted in the past, and future eruptions are possible. This condition could threaten future eruptions, particularly near scientific bases. Jang Bogo, a South Korean research station, is located only 30 km from the summit and could be affected by significant ash fallout in case of an explosive eruption. This condition leads to the necessity of observing Mount Melbourne’s activity frequently. This study used Sentinel-1 SAR data acquired from 2017 to 2024 to monitor the volcanic activity of Mount Melbourne by utilizing InSAR multitemporal time-series analysis implementing the improved combined scatterers interferometry with optimized point scatterers (ICOPS) method. The ICOPS method combined persistent scatterer (PS) and distributed scatterer (DS) with measurement point (MP) optimization based on convolutional neural network (CNN) and optimized hot spot analysis (OHSA). The ICOPS measurement results maintain reliable MP along the Mount Melbourne summit and around Jang Bogo station. The absence of GPS stations around these two areas makes it difficult to validate the result with the ground truth measurement, so the comparison with another method, small baseline (SBAS) measurement, is made to evaluate the reliability of the ICOPS measurement points. The comparison between the MP from ICOPS and the SBAS methods shows a good correlation with R2 of about 0.8134 in the Melbourne area and 0.8678 in the Jang Bogo area. The selected time-series plot around the summit of Mount Melbourne and the Jang Bogo area shows a stable trend of surface deformation. Thus, a total accumulated deformation of around 0.82 cm and an average deformation of about 0.10 cm/year was found around Mount Melbourne. Meanwhile, the Jang Bogo area exhibits a total deformation of about 0.15 cm with an average deformation of about 0.02. Overall, this research is a preliminary study of the ability of the ICOPS algorithm to monitor volcanic activity in snow-covered areas.
Article
Full-text available
The Mount Melbourne Volcanic Field (MMVF) is part of the West Antarctic Rift System, one of Earth’s largest intra-continental rift zones. It contains numerous small, compositionally diverse (alkali basalt–benmoreite) flank and satellite vents of Late Miocene–Pliocene age (≤12.50 Ma; mainly less than 2.5 Ma). They demonstrate a wide range of morphologies and eruptive mechanisms despite overlapping compositions and elevations, and they occur in a relatively small area surrounding the active Mount Melbourne stratovolcano. The volcanic outcrops fall into several main categories based on eruptive style: scoria cones, tuff cones, megapillow complexes, and shield volcanoes. Using the analysis of lithofacies and appraisal of the internal architectures of the outcrops, we have interpreted the likely eruptive setting for each center and examined the links between the environmental conditions and the resulting volcanic edifice types. Previous investigations assumed a glacial setting for most of the centers but without giving supporting evidence. We demonstrate that the local contemporary environmental conditions exerted a dominant control on the resulting volcanic edifices (i.e., the presence or absence of water, including ice or snow). The scoria cones erupted under dry subaerial conditions. Products of highly explosive hydrovolcanic eruptions are represented by tuff cones. The water involved was mainly glacial (meltwater) but may have been marine in a few examples, based on a comparison of the contrasting internal architectures of tuff cones erupted in confined (glacial) and unconfined (marine, lacustrine) settings. One of the glaciovolcanic tuff cones ceased activity shortly after it began transitioning to a tuya. The megapillow complexes are highly distinctive and have not been previously recognized in glaciovolcanic successions. They are subglacial effusive sequences emplaced as interconnected megapillows, lobes, and thick simple sheet lavas. They are believed to have erupted at moderately high discharge and reduced cooling rates in partially drained englacial vaults under ice, probably several hundred meters in thickness. Finally, several overlapping small shield volcanoes crop out mainly in the Cape Washington peninsula area. They are constructed of previously unrecognized multiple ‘a‘ā lava-fed deltas, erupted in association with a thin draping ice cover c. 50–145 m thick. Our study highlights how effectively water in all its forms (e.g., snow, ice, and any meltwater) or its absence exerts a fundamental control on eruption dynamics and volcano construction. When linked to published ages and ⁴⁰Ar/³⁹Ar dates produced by this study, the new environmental information indicates that the Late Pliocene–Pleistocene landscape was mainly an icefield rather than a persistent topography-drowning ice sheet. Ice thicknesses also generally increased toward the present.
Article
Full-text available
Low-intensity emission of volcanic ash represents the most frequent eruptive activity worldwide, spanning the whole range of magma compositions, from basalts to rhyolites. The associated ash component is typically characterized by heterogeneous texture and chemical composition, leading to misinterpretation of the role of syn-eruptive processes, such as cooling and degassing during magma ascent or even magma fragmentation. Despite their low intensity, the ash emission eruptions can be continuous for enough time to create problems to health and life networks of the communities all around the volcano. The lack of geophysical and/or geochemical precursor signals makes the petrological monitoring of the emitted ash the only instrument we have to understand the leading mechanisms and their evolution. Formation of low-level plumes related to ash-rich emissions has increasingly become a common eruptive scenario at Mt. Etna (Italy). In January–February 2019, an eruptive cycle of ash-rich emissions started. The onset of this activity was preceded on 24 December 2018 by a powerful Strombolian-like eruption from a fissure opened at the base of the New Southeast Crater. A lava flow from the same fissure and an ash-rich plume, 8–9 km high a.s.l., from the crater Bocca Nuova occurred concurrently. After about 4 weeks of intra-crater strombolian-like activity and strong vent degassing at summit craters, starting from 23 January 2019, at least four episodes of ash-rich emissions were recorded, mainly issued from the Northeast Crater. The episodes were spaced in time every 4–13 days, each lasting about 3–4 days, with the most intense phases of few hours. They formed weak plumes, up to 1 km high above the crater, that were rapidly dispersed toward different directions by dominant winds and recorded up to a distance of 30 km from the vent. By combining observations on the deposits with data on textural and chemical features of the ash components, we were able to discriminate between clasts originated from different crater sources and suggest an interpretive model for syn-eruptive processes and their evolution. Data indicate the occurrence of scarce (<10 vol.%) fresh juvenile material, including at least four groups of clasts with marked differences in microlite content and number density, and matrix glasses and minerals composition. Moreover, a large amount of non-juvenile clasts has been recognized, particularly abundant at the beginning of each episode. We propose that the low amount of juvenile ash results from episodic fast ascent of small magma batches from shallow reservoirs, traveling within a slow rising magma column subjected to cooling, degassing, and crystallization. The large number of non-juvenile clasts deriving from the thick crater infill of variably sealed or thermally altered material at the top of the magma column is suggested to contribute to the ash generation. The presence of a massive, granular crater infilling accumulating in the vent area may contribute to buffer the different geophysical signals associated with the active magma fragmentation process during the low-energy ash eruptions, as already evidenced at other volcanoes.
Article
Full-text available
Juvenile pyroclasts, especially in the ash size range, provide important information on primary fragmentation processes, i.e., initial explosive magma fragmentation, and on the state of the magma both prior to and at the point of fragmentation and quenching. There exists an extensive body of literature focusing on the quantification of juvenile particle morphology (shape), internal textures, and surface features spanning several decades; however, a standardized method has yet to emerge for comparative studies. No community-wide consensus currently exists (i) regarding the most representative size fraction(s) to be examined, (ii) on sample preparation procedures (such as whether to use whole-particle silhouettes or 2D cross-sections), (iii) on imaging techniques and image acquisition parameters, or (iv) on the optimal morphometric parameters to measure. Lack of a standardized method precludes robust comparison between different studies and laboratories. We propose here a preliminary “best practices” and workflow for characterization of juvenile pyroclasts, for comparative studies of primary fragmentation. If the community follows such a standardized method, it will become possible to accumulate a large volume of consistent data on juvenile pyroclasts from a range of eruption styles, fragmentation mechanisms, and magma compositions. This will ultimately allow deeper insights into the full panoply of magma-to-pyroclast processes that drive particle-producing volcanic eruptions. One or more “fragmentation diagrams” may eventually be developed to allow different types of magmatic and phreatomagmatic explosive eruptions to be distinguished based on their products.
Article
Full-text available
Ice core dating is the first step for a correct interpretation of climatic and environmental changes. In this work, we release the dating of the uppermost 197 m of the 250 m deep GV7(B) ice core (drill site, 70∘41′ S, 158∘52′ E; 1950 m a.s.l. in Oates Land, East Antarctica) with a sub-annual resolution. Chemical records of NO3-, MSA (methanesulfonic acid), non-sea-salt SO42- (nssSO42-), sea-salt ions and water stable isotopes (δ18O) were studied as candidates for dating due to their seasonal pattern. Different procedures were tested but the nssSO42- record proved to be the most reliable on the short- and long-term scales, so it was chosen for annual layer counting along the whole ice core. The dating was constrained by using volcanic signatures from historically known events as tie points, thus providing an accurate age–depth relationship for the period 1179–2009 CE. The achievement of the complete age scale allowed us to calculate the annual mean accumulation rate throughout the analyzed 197 m of the core, yielding an annually resolved history of the snow accumulation on site in the last millennium. A small yet consistent rise in accumulation rate (Tr = 1.6, p<0.001) was found for the last 830 years starting around mid-18th century.
Article
Full-text available
We present a full characterization of a 20 cm‐thick tephra layer found intercalated in the marine sediments recovered at Site U1524 during International Ocean Discovery Program (IODP) Expedition 374, in the Ross Sea, Antarctica. Tephra bedforms, mineral paragenesis, and major‐ and trace‐element composition on individual glass shards were investigated and the tephra age was constrained by ⁴⁰Ar‐³⁹Ar on sanidine crystals. The ⁴⁰Ar‐³⁹Ar data indicate that sanidine grains are variably contaminated by excess Ar, with the best age estimate of 1.282 ± 0.012 Ma, based on both single‐grain total fusion analyses and step‐heating experiments on multi‐grain aliquots. The tephra is characterized by a very homogeneous rhyolitic composition and a peculiar mineral assemblage, dominated by sanidine, quartz, and minor aenigmatite and arfvedsonite‐riebeckite amphiboles. The tephra from Site U1524 compositionally matches with a ca. 1.3 Ma, rhyolitic pumice fall deposit on the rim of the Chang Peak volcano summit caldera, in the Marie Byrd Land, located ca. 1,300 km from Site U1524. This contribution offers important volcanological data on the eruptive history of Chang Peak volcano and adds a new tephrochronologic marker for the dating, correlation, and synchronization of marine and continental early Pleistocene records of West Antarctica.
Article
We present the results o the tephrochronology study o a 14.49 m long marine sediment core (TR 17–08) collected in the Edisto Inlet, Ross Sea (Antarctica). The core contains our cryptotephra layers at 55–56, 512–513, 517–518, and 524–525 cm o depth, which have been characterised by a detailed description o the texture, mineral assemblage, and single glass shards major and trace element geochemistry. The age model o the investigated sedimentary sequence, based on radiocarbon dating, indicates that the topmost cryptotephra correlates with the widespread 1254 CE tephra erupted by a historical eruption (696 ± 2 cal yrs BP) o Mount Rittmann, in northern Victoria Land. Deeper cryptotephra layers were derived rom previously unknown explosive eruptions o Mount Melbourne volcano and were emplaced between 1615 cal yrs BP and 1677 cal yrs BP, e.g. between the 3rd and 4th centuries CE. This discovery demonstrates that the Mount Melbourne volcanic complex has been highly active in historical times allowing signicant progress in the current understanding o regional eruptive history. Moreover, rom a teph�rochronological point o view, the detected cryptotephra provide new regional isochron markers to acilitate high-precision correlations and help stratigraphically constrain changes in environmental and climatic conditions that are identied by multidisciplinary studies.
Article
Polygenetic volcanoes and calderas produce eruptions of a wide variety of magnitudes, chemistries, and recurrence times. Understanding the interplay between long- and short-term and deep and shallow processes associated with accumulation and transfer of eruptible magma is essential for assessing the potential for future eruptions to occur and estimating their magnitude, which remains one of the foremost challenges in the Earth sciences. We review literature and use existing data for emblematic volcanic systems to identify the essential data sets required to define the state of activity of volcanoes and their plumbing systems. We explore global eruptive records in combination with heat flux and other geological and geophysical data to determine the evolutionary stage of plumbing systems. We define a Volcanic Activity Index applicable to any volcano that provides an estimate of the potential of a system to erupt in the future, which is especially important for long-quiescent volcanoes. ▪ Magmatic plumbing systems that feed volcanic activity extend across Earth's crust and are long-lived at depth and ephemeral in their shallowest portions. ▪ We revise and update the definitions of active, quiescent, and extinct volcanoes based on physical proxies for the architecture, longevity, amount, and distribution of eruptible magma in the crust. ▪ We propose a Volcanic Activity Index, which provides a relative measure of the state of activity of a volcano with respect to all other volcanoes in the world. ▪ New imaging and monitoring strategies are required to improve our ability to detect lower and middle crust magmatic processes and forecast eruptions and their potential size. Expected final online publication date for the Annual Review of Earth and Planetary Sciences, Volume 50 is May 2022. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
We report on the characterization of a thick sequence of pyroclastic deposits exposed on the summit area and flanks of Mount Melbourne volcano, in northern Victoria Land, Antarctica related to eruptions during the Late Glacial period. We provide a complete characterization of tephra deposits including mineralogy, single shard major- and trace-element glass compositions, and an ⁴⁰Ar³⁹Ar age of feldspar crystals extracted from the deposit. The pyroclastic deposits are trachybasaltic to trachytic in composition and are interpreted to have resulted from four Strombolian/Vulcanian to sub-Plinian/Plinian eruptions. The younger and more intense sub-Plinian/Plinian eruption (our eruption 2) yielded an ⁴⁰Ar³⁹Ar age of 13.5 ± 4.3 ka (±2σ). The study of Mount Melbourne proximal deposits provides significant new data for the reconstruction of the volcano eruptive history and a better assessment of the volcanic risk connected to a possible future eruption. We also explore geochemical correlations between Mount Melbourne proximal deposits and distal tephra layers recognized in ice cores and blue ice fields of East Antarctica. A good geochemical match exists between the composition of products from the trachytic sub-Plinian/Plinian eruption 2 and some tephra layers from Talos Dome and shards in Siple Dome which is also compatible in age (c. 9.3 ka) with our ⁴⁰Ar³⁹Ar age determination. Our new insights into the volcanic history of Mount Melbourne and the new high-quality electron microprobe and trace element composition data on its proximal products will help improve future correlations and synchronization of tephra archives in the region.
Article
We review here data and information on Antarctic volcanism resulting from recent tephrostratigraphic investigations on marine cores. Records include deep drill cores recovered during oceanographic expeditions: DSDP, ODP and IODP drill cores recovered during ice-based and land-based international cooperative drilling programmes DVDP 15, MSSTS-1, CIROS-1 and CIROS-2, DVDP 15, CRP-1, CRP-2/2A and CRP-3, ANDRILL-MIS and ANDRILL-SMS, and shallow gravity and piston cores recovered in the Antarctic and sub-Antarctic oceans. We report on the identification of visible volcaniclastic horizons and, in particular, of primary tephra within the marine sequences. Where available, the results of analyses carried out on these products are presented. The volcanic material identified differs in its nature, composition and emplacement mechanisms. It was derived from different sources on the Antarctic continent and was emplaced over a wide time span. Marine sediments contain a more complete record of the explosive activity from Antarctic volcanoes and are complementary to those obtained by land-based studies. This record provides important information for volcanological reconstructions including approximate intensities and magnitudes of eruptions, and their duration, age and recurrence, as well as their eruptive dynamics. In addition, characterized tephra layers represent an invaluable chronological tool essential in establishing correlations between different archives and in synchronizing climate records.
Article
Cenozoic magmatic rocks related to the West Antarctic Rift System crop out right across Antarctica, in Victoria Land, Marie Byrd Land and into Ellsworth Land. Northern Victoria Land, located at the northwestern tip of the western rift shoulder, is unique in hosting the longest record of the rift-related igneous activity: plutonic rocks and cogenetic dyke swarms cover the time span from c. 50 to 20 Ma, and volcanic rocks are recorded from 15 Ma to the present. The origin of the entire igneous suite is debated; nevertheless, the combination of geochemical and isotopic data with the regional tectonic history supports a model with no role for a mantle plume. Amagmatic extension during the Cretaceous generated an autometasomatized mantle source that, during Eocene–present activity, produced magma by small degrees of melting induced by the transtensional activity of translithospheric fault systems. The emplacement of Eocene–Oligocene plutons and dyke swarms was focused along these fault systems. Conversely, the location of the mid-Miocene–present volcanoes is governed by lithospheric necking along the Ross Sea coast for the largest volcanic edifices; while inland, smaller central volcanoes and scoria cones are related to the establishment of magma chambers in thicker crust.