ArticlePDF Available

Turbulent particle-gas feedback exacerbates the hazard impacts of pyroclastic density currents

Authors:

Abstract and Figures

Causing one-third of all volcanic fatalities, pyroclastic density currents create destruction far beyond our current scientific explanation. Opportunities to interrogate the mechanisms behind this hazard have long been desired, but pyroclastic density currents persistently defy internal observation. Here we show, through direct measurements of destruction-causing dynamic pressure in large-scale experiments, that pressure maxima exceed theoretical values used in hazard assessments by more than one order of magnitude. These distinct pressure excursions occur through the clustering of high-momentum particles at the peripheries of coherent turbulence structures. Particle loading modifies these eddies and generates repeated high-pressure loading impacts at the frequency of the turbulence structures. Collisions of particle clusters against stationary objects generate even higher dynamic pressures that account for up to 75% of the local flow energy. To prevent severe underestimation of damage intensities, these multiphase feedback processes must be considered in hazard models that aim to mitigate volcanic risk globally.
Synthesizing pyroclastic density currents and dynamic pressure measurements in large-scale experiments a The time-variant flow structure of the experimental PDC during channel confined propagation along a proximal measurement profile, which is located at 1.8 m downstream from impact of the hot mixture of volcanic material and air with the channel. The height-time kymograph is obtained by the stacking of vertical columns of pixels recorded by a high-speed camera at the static observer location. Pixel brightness correlates positively with, both, bulk flow particle volume concentration and the abundance of highly reflective coarse-grained pumiceous glass shards and plagioclase crystals. At this location, the density current structure is fully developed comprising a leading head with a wake in its rear and a trailing density current body. The density current tail region is not shown. b Lateral view of the frontal c. 7 m of the advancing experimental PDC across the unconfined distal runout section approximately 1.5 m before buoyancy reversal. The red line shows the time-variant height of the upper flow boundary. c Snapshot of the front of the experimental PDC approaching one of the wing-shaped vertical profiles of dynamic pressure sensors. The shape of the wings is designed to reduce flow detachment and formation of a turbulent wake behind the profile. The sensing elements of the piezoelectric dynamic pressure sensors protrude, upstream from the wing, into the approaching density current.
… 
Two types of dynamic pressure and their probability density spectra All measurements of dynamic pressure shown in this figure are obtained at the location of 3.35 m from impact at a height 0.45 m above the flow base. a Variation of the entire dynamic pressure Pdyn as a function of time as recorded by piezoelectric sensor and the corresponding probability density function shown in b. The pressure signal of Pdyn shows abundant near-instantaneous high-pressure peaks of several hundreds to thousands of Pascal, which are characterized by pressure increases in less than one millisecond. The inset highlights an example of one of the near-instantaneous high-pressure signals, which are associated with impacts from individual particles with the pressure sensor. c Variation of the partial dynamic pressure signal Pimpact as a function of time recorded by the piezoelectric sensor, which shows only those parts of the entire pressure signal Pdyn that are associated with near-instantaneous pressure peaks. d The probability density function of Pimpact. e Variation of the partial dynamic pressure signal Pdusty gas as a function of time, which is the dynamic pressure of the continuous dusty gas phase computed as the difference between Pdyn(t) and Pimpact(t) via Eq. (2). The black line shows a 20-millisecond average of the timeseries Pdusty gas (depicted in blue) The orange line shows the corresponding timeseries of the dynamic pressure PBernoulli, computed independently from timeseries of flow velocity and flow density via Eq. (1). f probability density functions of Pdusty gas (blue) and PBernoulli (orange).
… 
This content is subject to copyright. Terms and conditions apply.
communications earth & environment Article
https://doi.org/10.1038/s43247-024-01305-x
Turbulent particle-gas feedback
exacerbates the hazard impacts of
pyroclastic density currents
Check for updates
Daniel H. Uhle 1,GertLube
1,EricC.P.Breard
2,3,EckartMeiburg 4, Josef Dufek3,JamesArdo 1,
Jim R. Jones 5,ErmannoBrosch
1, Lucas R. P. Corna1& Susanna F. Jenkins 6
Causing one-third of all volcanic fatalities, pyroclastic density currents create destruction far beyond
our current scientic explanation. Opportunities to interrogate the mechanisms behind this hazard
have long been desired, but pyroclastic density currents persistently defy internal observation. Here
we show, through direct measurements of destruction-causing dynamic pressure in large-scale
experiments, that pressure maxima exceed theoretical values used in hazard assessments by more
than one order of magnitude. These distinct pressure excursions occur through the clustering of high-
momentum particles at the peripheries of coherent turbulence structures. Particle loading modies
these eddies and generates repeated high-pressure loading impacts at the frequency of the
turbulence structures. Collisions of particle clusters against stationary objects generate even higher
dynamic pressures that account for up to 75% of the local ow energy. To prevent severe
underestimation of damage intensities, these multiphase feedback processes must be considered in
hazard models that aim to mitigate volcanic risk globally.
Dilute pyroclastic density currents (PDCs) are lethal and recurrent hazards
from volcanoes15. Over 200 million people are directly endangered by these
multiphase ows of hot volcanic particles and gas68. Over the past decade
alone, and despite strong advances in understanding volcanic hazards,
PDCs caused more than a thousand fatalities and signicant damage to
infrastructure globally913. So, how can we learn to forecast PDC hazards
more accurately? The key hazard agents of PDCs are the volcanic particles
carried inside them. As the main driver of PDC motion, particle con-
centration controls ow speed, destructive power and reach1422. Carrying
most of the thermal energy, particles are also important for PDC burn
hazards, while the readily respirable particles cause inhalation injury and
asphyxia23,24. To model and mitigate against PDC hazards we must learn to
better understand the behaviour of particles, more specically their motion
and sedimentation, inside ows1,35.
However, there is a fundamental roadblock hindering this endeavour:
the motion of particles within a PDC is modied through complex, long-
hypothesized but poorly understood coupling and feedback mechanisms
between particle and gas phases4,2528. Unlike a homogenous suspension of
particlesinauid, these interactions modify the ow and turbulence
structure in PDCs, leading to particle clustering, enhanced sedimentation,
high pore-pressure and uidization2935. Recent advances through large-
scale PDC experiments discovered that these interactions focus particles,
and thus mechanical and thermal energies, into large eddies, mesoscale
turbulence structures and internal gravity waves36. However, how the
feedback mechanisms between particles and gas control the runout and
hazardbehaviourofPDCsremainsunknown. The opportunity to explore
the multiphase physics of PDCs is long desired, but their ferocity, opacity
and unpredictability hampers direct observation and measurements of their
internal characteristics.
An important example of the large uncertainties for hazard planning
that result from this gap in understanding is forecasting of the destruction-
causing dynamic pressure of PDCs. To estimate local values of dynamic
pressure, researchers have systematically mapped the degree of damage to
infrastructure and vegetation in the aftermath of volcanic eruptions3740.
Another approach uses relationships between sediment transport and
resulting deposit characteristics in turbulent uid ow to estimate local
time-averaged values of ow velocity and density, which together dene
dynamic pressure26,4143. Recently, direct measurements in large-scale
1Volcanic Risk Solutions, Massey University, Palmerston North, New Zealand. 2School of Geosciences, The University of Edinb urgh, Edinburgh, UK. 3Department
of Earth Science, University of Oregon, Eugene, OR, USA. 4Department of Mechanical Engineering, University of California at Santa Barbara, Santa Barbara, CA,
USA. 5School of Food and Advanced Technology, Massey University, Palmerston North, New Zealand. 6Earth Observatory of Singapore, Asian School of
Environment, Nanyang Technological University, Singapore, Singapore. e-mail: d.uhle@massey.ac.nz
Communications Earth & Environment | (2024) 5:245 1
1234567890():,;
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
experiments, natural PDCs and snowavalanchesdemonstratedthat
dynamic pressure calculated from turbulent uctuations in velocity and
density exceed eld- and theory-derived pressure estimates36.Thisnding
demonstrates the importance of turbulent multiphase ow in volcanic
hazard assessments and raises caution regarding the applicability of current
approaches, which use (time) average ow properties.
Here we present the rst high-resolution direct measurements of
destruction-causing dynamic pressure in large-scale PDC experiments to
interrogate how dynamic pressures form and evolve in PDCs. These mea-
surements reveal that dynamic pressure reaches pronounced maxima in the
peripheries of turbulence structures and generate regular pulses of high
dynamic pressure at the characteristic frequency of these structures. This
occurs through turbulent gas-particle feedbacks as particles, whose ratio of
characteristic response time to uid motion to the characteristic timescale of
uidmotionislessthanorequaltoone,clusteratthemarginsofeddies.We
demonstrate that this process results in two different types of dynamic
pressure effects with distinct hazard impacts to life and infrastructure.
Importantly, the resulting energy spectra of dynamic pressure are char-
acterized by peak pressure values that largely exceed pressure estimates
based on bulk ow characteristics, which are currently used for hazard
assessments.
Results
Synthesizing pyroclastic density currents and measuring
dynamic pressure in large-scale experiments
Over the last decade, the development of large-scale experimental facilities
to synthesize scaled analogues of PDCs16,4446 has enabled a novel approach
to study their internal ow characteristics under safe conditions. The
experimental apparatuses in Italy45 and New Zealand44, which use heated
natural volcanic material, lend themselves to the interrogation of the scaled
uid mechanic and thermodynamic processes inside fully turbulent PDC
analogues. Here we report the results of direct measurements of dynamic
pressure inside large-scale experimental PDCs conducted at the New
Zealand facility PELE (the Pyroclastic ow Eruption Large-scale Experi-
ment). At PELE, experimental PDCs are generated by the controlled
gravitational collapse of a suspension of hot volcanic particles and air from
an elevated hopper into an instrumented runout section (Supplementary
Fig. 1; Supplementary Movie). For the experiments reported here, we used a
0.7 m3hopper in which a 124 kg mixture of natural volcanic particles was
heated to 120 °C (the ambient temperature was 15 °C) over a period of 72 h
to allow for thermal equilibration and evaporation of residual moisture
inside the pre-dried mixture.
The volcanic material comprises a mixture of two well-characterised
deposits of PDCs of the 232 CE Taupo eruption in New Zealand47.Themain
components of the mixture are highly vesicular pumice, glass shards, free
crystals and rare lithic particles. The mixture has a weakly bimodal grainsize
distribution ranging from 2 µm to 16 mm, with a main mode at 250 µm and
a minor mode at 11 µm. The particle density distribution ranges from c.
3002,600 kg m3and is skewed towards low densities. Further details of the
material characteristics are provided in Supplementary Fig. 1 and in the
Methods section.
The hopper is lifted to a vertical drop height of 7 m. It is mounted onto
four load cells recording its mass discharge, which in this case lasts c. 4.6 s at
an average discharge rate of c. 27 kg s1. On impact into a 0.5 m-wide and 12
m-long umewithaslopeof6°,theaeratedparticle-airmixtureformsa
channelized dilute gravity current with an initial front velocity of c. 5.5 m s1
and, on average, particle volume concentration of c. 0.25 vol.%. The ow is
characterized by a leading, c. 1.12 m thick gravity current head, which is
trailed by a gravity current wake that overlies a c. 0.91.6 m thick gravity
current body, and followed by a c. 1.83.5 m thick gravity current tail region
(Fig. 1a, b). Downstream of the inclined ume section, the experimental
PDC propagates to 17 m along a horizontal, 0.5 m wide channel-conned
section before spreading across an unconned horizontal concrete pad.
Around 8 s after impact, the vertically density-stratied current has pro-
pagated23.5m,slowedtoc.0.5ms
1and has emplaced approximately
99.7% of its total deposit volume. At 24 m, the depth-and time-integrated
Fig. 1 | Synthesizing pyroclastic density currents and dynamic pressure mea-
surements in large-scale experiments. a The time-variant ow structure of the
experimental PDC during channel conned propagation along a proximal measure-
ment prole,whichislocatedat1.8mdownstreamfromimpactofthehotmixtureof
volcanic material and air with the channel. The height-time kymograph is obtained by
the stacking of vertical columns of pixels recorded by a high-speed camera at the static
observer location. Pixel brightness correlates positively with, both, bulk ow particle
volume concentration and the abundance of highly reective coarse-grained pumiceous
glass shards and plagioclase crystals. At this location, the density current structure is
fully developed comprising a leading head with a wake in its rear and a trailing density
current body. The density current tail region is not shown. bLateral view of the frontal
c. 7 m of the advancing experimental PDC across the unconned distal runout section
approximately 1.5 m before buoyancy reversal. The red line shows the time-variant
height of the upper ow boundary. cSnapshot of the front of the experimental PDC
approaching one of the wing-shaped vertical proles of dynamic pressure sensors. The
shapeofthewingsisdesignedtoreduceow detachment and formation of a turbulent
wake behind the prole. The sensing elements of the piezoelectric dynamic pressure
sensors protrude, upstream from the wing, into the approaching density current.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
particle volume concentration has decreased to values smaller than
0.01 vol.% due to both deposition and entrainment of ambient air; the
current becomes positively buoyant and nally ascends as a series of ther-
mals along the ow length (Supplementary Movie).
To investigate the origin and evolution of dynamic pressure inside the
experimental PDCs, seven vertical, airfoil-shaped proles with up to six
piezo-electric dynamic pressure sensors, which comprise circular, 15 mm
diameter frontal steel diaphragms, were placed into the ow centerline at
runout distances of 1.8, 3.35, 5.4, 9.6, 12, 16 and 20 m (Fig. 1c). Measurements
of time (t)-variant dynamic pressure P
dyn
(t, z) are recorded at a frequency of
1 kHz, and are complemented by measurements of time series, in vertical
proles, of ow velocity u(t,z), particle volume concentration Cs(t, z), ow
density ρC(t, z),ow temperature and grain-size distribution, where zis the
height above the ow base in the direction perpendicular to slope (see
methods for details). Flow velocity components u(t,z)are measured using
high-speed video at 0.5 kHz through the umes tempered glass sides. The
sidewalls introduce boundary effects that are not present in unconned real-
world ows and in the ow centerline of the experimental ows where
dynamic pressure is measured. We minimize these boundary effects through
the use of hydraulically smooth sidewalls (i.e., thickness of laminar layer/wall
roughness >1) while the owsReynoldsnumber,whichisinverselyrelatedto
the thickness of the viscous boundary layer, is high (Re =1.5×106). Velocity
data together with measurements of time-variant and height-variant grain-
size distribution, ow density and temperatures allow for an independent
measurement of dynamic pressure, dened as:
Pdyn Bernoulli ¼1
2ρCu
jj
2;ð1Þ
where u
jj
is the magnitude of the local time-variant velocity.
The experimental PDCs generated under these conditions scale well to
natural dilute fully turbulent PDCs (i.e., pyroclastic surges or blasts; see a
comparison of scaling parameters for natural and experimental PDCs in
Supplementary Table 1). Amongst the non-dimensional products of char-
acteristic length-, time and temperature-scales we highlight the Reynolds
number (comparing inertial and viscous forces) reaching values of 1.5 × 106,
Stokes number (comparing particle time-scale to turbulent ow time-scale)
of 103100, Stability number (comparing time-scales of particle settling and
turbulent uid motion) of 102101, Richardson number (characterizing the
stability of stratication in turbulent ows) of 102101,andthermal
Richardson number (assessing the ratio of forced and buoyant convection)
of 0.024.5. The overlap of ranges in Reynolds, Stokes and Stability numbers
in experimental and natural PDCs ensures that the complete range of
particle-gas feedback mechanisms and turbulent particle transport is
reproduced.
Two types of dynamic pressure
Figure 2a and b depict a typical example of the time series of dynamic
pressure inside the experimental PDCs (here for the vertical prole 2 at a
runout distance of 3.35 m at approximately mid-ow height of 0.45 m
from the ow base). The local time-averaged dynamic pressure takes a
value of 45 Pa, while maximum values reach several kilopascals. The wide
discrepancy between average and maximum pressures, as well as the
absolute values of recorded maximum pressure are surprising. Recent
analysis of dynamic pressure in experimental PDCs using measurements
of ow velocity and ow density and Eq. 1showed that turbulent uc-
tuations in velocity and density generate dynamic pressures that exceed
average values by a factor of 3536. This coincides with our measurements
of dynamic pressure through velocity and density time-series (that is
P
dyn_Bernoulli
) and Eq. 1giving maximum and average values of 79 and
Fig. 2 | Two types of dynamic pressure and their probability density spectra. All
measurements of dynamic pressure shown in this gure are obtained at the location
of 3.35 m from impact at a height 0.45 m above the ow base. aVariation of the entire
dynamic pressure P
dyn
as a function of time as recorded by piezoelectric sensor and
the corresponding probability density function shown in b. The pressure signal of
P
dyn
shows abundant near-instantaneous high-pressure peaks of several hundreds to
thousands of Pascal, which are characterized by pressure increases in less than one
millisecond. The inset highlights an example of one of the near-instantaneous high-
pressure signals, which are associated with impacts from individual particl es with the
pressure sensor. cVariation of the partial dynamic pressure signal P
impact
as a
function of time recorded by the piezoelectric sensor, which shows only those parts
of the entire pressure signal P
dyn
that are associated with near-instantaneous pres-
sure peaks. dThe probability density function of P
impact
.eVariation of the partial
dynamic pressure signal P
dusty gas
as a function of time, which is the dynamic pressure
of the continuous dusty gas phase computed as the difference between P
dyn
(t) and
P
impact
(t) via Eq. (2). The black line shows a 20-millisecond average of the timeseries
P
dusty gas
(depicted in blue) The orange line shows the corresponding timeseries of
the dynamic pressure P
Bernoulli
, computed independently from timeseries of ow
velocity and ow density via Eq. (1). fprobability density functions of P
dusty gas
(blue)
and P
Bernoulli
(orange).
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
24 Pa (i.e., a ratio of c. 3.3), respectively. In stark contrast, time-averaged
(45 Pa) and maximum dynamic pressures (4830 Pa) recorded by the
dynamic pressure sensors (Fig. 2a) differ by two orders of magnitude.
They can, therefore, not be explained by turbulent uctuations in velocity
and density of the multiphase ow. Thus, from the perspective of
understanding and mitigating hazard impacts of PDCs, an important
question arises regarding the origin and nature of these unexpectedly
large pressures.
An important clue is given by the form of those pressure signals that
exceed the theoretical values of dynamic pressure calculated through
turbulent uctuations of ow velocity and density through Eq. 1.These
high-pressure signals are characterized by an abrupt pressure increase to
a maximum in less than one millisecond (inset of Fig. 2a; the sensor
response time (12 μs) « sampling rate (1 ms)). Furthermore, we deter-
mined that in all occurrences of these near-instantaneous high-pressure
signals, the rate of pressure change with time was always larger than
c. 220 Pa ms1. These characteristics in the form of pressure signals are
consistent with impacts of individual particles on the pressure sensor.
Considering particle diameters and size-averaged particle densities
determined from ow samples, and local velocities measured in high-
speed video, theoretical values of dynamic pressure caused by impacts
of individual particles can be estimated (Supplementary note 1 and
Supplementary Fig. 3). At time-averaged local velocities, impacts of
particles with diameters larger than c. 100 µm generate dynamic pres-
sures larger than 500 Pa, and thus, considerably larger than theoretical
values of dynamic pressure calculated through turbulent uctuations of
ow velocity and density and Eq. 1. Maximum measured values of
dynamic pressure (e.g., 4.8 kPa at prole 2) can be explained by impacts
of, either, low-density pumiceous particles with particle diameters larger
than c. 2 millimeters or abundant high-density plagioclase particles with
particle diameters larger than c. 600µm. Hence, the near-instantaneous
high-pressure signals are explained by individual particle impacts.
This nding allows a sub-division of the total dynamic pressure signal
P
dyn
into the dynamic pressure associated with impacts from individual
particles P
impact
(Fig. 2c and d) and the dynamic pressure of the (continuous)
dusty gas phase P
dusty gas
(Fig. 2e) where:
Pdyn ¼Pdustygas þPimpact :ð2Þ
Probability density functions of P
dusty gas
and P
dyn_Bernoulli
coincide
markedly well (Fig. 2f). This justies the subdivision of the total dynamic
pressure into its continuous dusty gas, P
dusty gas
, and particle impact, P
impact
,
components. Furthermore, the piezoelectric signal of the dynamic pressure
sensor has a superior temporal resolution (1 kHz) in comparison to the
timeseries of dynamic pressure P
dyn_Bernoulli
computed, via Eq. 1,from
timeseries of ow density data (with a sampling rate of 20 Hz) and velocity
(sampled at 0.5 kHz). Thus, the piezoelectric signal shows a much more
detailed record of the dynamic pressure of the continuous dusty gas phase in
the high-pressure end of its spectrum.
Despite the strong differences in pressure ranges between the dynamic
pressures of the dusty gas phase (up to 390 Pa at prole 2) and particle
impacts (up to 4.8 kPa at prole 2), the probability density functions and the
time-series of both signals, and hence that of the total dynamic pressure,
share two characteristics: rst, the pressure distributions are strongly skewed
towards large pressures (Fig. 2b, d and f). Second, largest pressures and
highest occurrences of large pressures occur during passages of the head
(c. 01sinFig.2a, c and e) and mid-body regions (c. 2.43sinFig.2a, c and
e) of the experimental PDC. For the dynamic pressure of particle impacts
P
impact
, time-series of the rate of particle impacts, R, provide further insights
(Fig. 3eh). Along ow runout, the rate of particle impacts, at approxi-
mately mid-ow height (i.e., at z= 0.45 m), decays strongly from more than
275 Hz at a distance of 1.8 m, to 95 Hz at 5.4 m and no recorded impacts at
9.6 m and beyond. This suggests that dynamic pressure signals associated
with impacts from individual particles are limited to particles with a critical
Fig. 3 | Spatiotemporal variation of the frequency of particle impacts. The time-
variant ow structure of the experimental PDC at four different distances from
impact with the channel at 1.8 m (a), 3.35 m (b), 5.4 m (c) and 9.6 m (d). As in Fig. 1a,
the height-time kymograph plots are generated by the stacking of vertical columns of
pixels recorded by high-speed cameras at the static observer locations. ehTime-
series of the frequency of particle impacts f
impact
at the four different distances. The
frequency of particle impacts is computed through convolution of time series of
binary impacts using a box function with a duration of 200 ms. Time series of binary
impacts, shown in Fig. 5, were generated by assigning times of occurrences of particle
impacts a value of 1 and times without impacts a value of zero. Due to the sampling
rate of 1 kHz, impact rates of up to 500 Hz can be resolved. The frequency strongly
declines with distance. ilTimeseries of the depth-integrated ow density
ρCtðÞat
the four different distances. The regular low-frequency oscillations in ow density
are positively correlated in time with the oscillations of particle impacts.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 4
Content courtesy of Springer Nature, terms of use apply. Rights reserved
minimum momentum. We note that the pattern of decay of the number of
particle impacts with runout distance is similar to the decay of particles with
diameters of 125500 µm in ow samples (Fig. 4). This is the grain-size
range that constitutes the coarse tail of ow grain-size distributions in
proximal to distal runout reaches.
The low numbers of particle impacts at medial to distal prole locations
(i.e., > 10 m) limit further analysis. However, at the proximal proles, impact
numbers are sufciently large to see that the rate of impacts obeys a regular
low-frequency oscillation (Fig. 3e and f), which is positively correlated in
time with a low-frequency oscillation in ow density (Fig. 3iandj;i.e,the
number of impacts is large when ow density is high).
Figure 5depicts timeseries of particle impacts in binary form (i.e., times
during which an impact occurs are assigned a value of 1, and times where no
impacts occur a value of zero). This shows that if a particle impact occurs, it
is typically followed by multiple further impacts. By contrast, isolated single
particle impacts are rare. The mean of the number of successive, in other
words, clustered impacts ranges from 214 impacts with an average number
of clustered impacts of c. 4.
Energy spectra of dynamic pressure from particle impacts and
dusty gas
Recently, low-frequency oscillations in ow density, velocity and hence
dynamic pressure have been linked to the occurrence of large-scale coherent
turbulence structures inside experimental and real-world PDCs and snow
avalanches36. The frequency fof the most energetic coherent structure in a
turbulent ow is given by the Strouhal number Str as
Str ¼fL
Uð3Þ
where Land Uare the owscharacteristic length- and velocity-scales48.For
highly turbulent ow conditions with Reynolds numbers Re >10
5,the
Strouhal number approaches a constant value of approximately 0.336,4951.In
our high-Reynolds number experimental PDCs (Re =1.5×10
6), the
measured main frequencies of ow density oscillations of c. 1.26 Hz (i.e.,
a period of 794 ms) and particle impact rates Rof c. 1.33 Hz (a period of
752 ms) coincide closely (Supplementary Fig. 4a and b), and correspond
with the period of the visually observed occurrences of the largest ow
structures (Fig. 3ad; Supplementary Fig. 4c). Substituting the experimen-
tally determined frequency f~ 1.3 Hz, the time-averaged height of the
gravity current body region L~ 1.15 m and the time-averaged ow velocity
U~4.76ms
1into Eq. (3), gives a Strouhal number Str = 0.31 close to the
commonly observed value.
The energy peak associated with the most energetic coherent struc-
tures, at a frequency of f~ 1.3 Hz, is clearly visible in energy spectra of the
total dynamic pressure P
dyn
, the dynamic pressure of particle impacts
P
impacts
and the dynamic pressure of the continuum dusty gas phase P
dusty gas
at the top of their respective energy cascades (red bars in Fig. 6ac). The
energy of P
dusty gas
decreases strongly with increasing frequency consistent
with the transfer of turbulent kinetic energy (per unit volume) to smaller
scales of coherent structures (Fig. 6c). By contrast, the energies of P
impacts
(Fig. 6b) and P
dyn
(Fig. 6a) show a markedly atter decay of pressure energy
with increasing frequency (and decreasing length-scale of coherent turbu-
lence structure). Importantly, the maximum energy values of both, P
impacts
and P
dyn
, which also reach the energy values of the most energetic coherent
structures at f=1.3Hz,occuracrossadened band of frequencies fof c.
716Hz(greybarsinFig.6a and b). The wavenumbers associated with this
frequency band correspond to length-scales of c. 0.250.6 m (Fig. 6d and e).
The multiphase origin of dynamic pressure uctuations
We hypothesize that the origin of the unusually high and temporally clus-
tered dynamic pressures is rooted in the concentration of particles at the
peripheries of coherent turbulence structures. PDCs comprise a wide
spectrum of particle-gas feedback mechanisms from simple one-way cou-
pling where the motion of particles is affected by the motion of the uid
phase but not vice versa, over two-way coupled particle-gas systems where
particle-gas interactions can modify the ow and turbulence structure, to
four-way coupling that strongly alters the ow structure4.Thedegreeof
Fig. 4 | Spatial evolution of particle impacts and ow grain-size distribution. The
number of particle impacts with the piezoelectric dynamic pressure sensors at mid-
ow height (0.45 m from the ow base) as a function of ow distance (white dia-
mond symbols). In comparison, the time-integrated cumulative grain-size dis-
tribution of the ow captured in ow samplers at heights of 0.45 m is shown as a
function of ow distance. The decrease of particle impacts with distance follows a
similar trend as the decay of particle sizes 125500 µm in ow samples. Particles with
diameters larger than 2 mm (not shown) completely sediment from the ow by c.
1.8 m; particles with diameters >1 mm and >500 µm sediment from the ow at
distances of c. 5 m and c. 7.5 m, respectively.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Fig. 6 | Energy spectra of dynamic pressure in Fourier space. Energy spectra of
dynamic pressure against frequency for P
dyn
(a), P
impact
(b) and P
dusty gas
(c). Energy
peaks at a frequency f~ 1.3 Hz associated with the largest coherent structures are
highlighted by a red bar. This frequency coincides with the measured frequency of
the largest coherent structures f1:3Hz ¼U Str
L, where Uand Lare the time-
averaged velocity and thickness of the density current body region and Str is the
Strouhal number. The highest-pressure energies occur in P
dyn
(a) and P
impact
(b)
across a narrow band of frequencies f
max
from c.716 Hz (where the frequency range
f
max
was visually handpicked) highlighted by a grey bar. Energy spectra of dynamic
pressure against wavenumber kfor P
dyn
(d), P
impact
(e) and P
dusty gas
(f). For P
dyn
and
P
impact
, the high-energy band associated with f
max
is highlighted by grey bars. This
band corresponds to wavenumber of c.1.
64 1/m or length-scales of coherent
structures 1/kof c. 0.250.6 m. The orange line with slope of 5/3 shows the Kol-
mogorov scaling of the inertial range.
Fig. 5 | Clustered distribution of particle impacts. Timeseries of particle impacts at
0.45 m above the ow base at four different distances from impact: 1.8 m (a), 3.35 m
(b), 5.4 m (c) and 9.6 m (d). Occurrences of particle impacts are shown in binary
form. That is times of impacts are assigned a value of 1 and times without impacts are
assigned a value of zero. The number of particle impacts is strongly decreasing with
ow distance and no particle impacts are recorded at the position at 9.6 m. When
particle impacts occur, they tend to occur in clusters of several successive impacts.
Non-clustered particle impacts are rare. The largest number of particle impacts
occur in the head region (c. 01 s) and in the mid-body region (c. 2.43 s) of the
experimental PDC.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 6
Content courtesy of Springer Nature, terms of use apply. Rights reserved
coupling of particles to turbulence structures can be approximated through
the particle Stokes number St dened as
St ¼τp
τε
;ð4Þ
where τpis the characteristic timescale of particles and τεis the turbulence
timescale. Particles with low Stokes numbers St 1 are well coupled with the
uidphaseandtendtofollowtheuid motion; particles with St 1accu-
mulate at the margins of coherent turbulence structures, while high Stokes
number particles St 1 are decoupled from uid turbulence and move
independent of the turbulence motion25,27,30,48,52,53. For the condition St =1of
particles concentrating at the peripheries of coherent structures it follows:
τp¼τε$τ1
p¼fε;and ð5Þ
$fε¼τ1
p¼18ρFνFF
ρPd2;ð6Þ
where fεis the frequency associated with this condition, ρFand ρPare uid
and particle densities, respectively, νFis the kinematic viscosity of the uid,
Fis the Stokes drag factor as dened by ref. 54.anddthe particle diameter.
Equation (6) and our experimental data of the particle size-averaged particle
density ρPðdÞcan be combined to calculate the critical frequency fεdðÞof
particle clustering at margins of coherent structures as a function of particle
diameter (Fig. 7a). For the polydisperse volcanic mixture used in our
experiments (Supplementary Fig. 1), these critical frequencies range over
four orders of magnitude from 3×1014×103Hz. However, the particle
sizes associated with the measured energy maximum in P
impacts
and P
dyn
,at
fεmax ~716 Hz (Fig. 5a and b and grey horizontal bar in Fig. 7a),
constitute a narrow band of particle diameters of 117229 µm (grey vertical
barinFig.7a). In ow samples, this particle size range occupies the coarse
tail of grainsize distributions (Fig. 7b). Due to sedimentation, the proportion
of particles larger than 117 µm strongly decreases from c. 70 wt.% in the
initial mixture to <5 wt.% at a ow distance of 20 m (Fig. 4). Similarly, the
dynamic pressure energy associated with particle impacts P
impacts
,which,at
a runout distance of 1.8 m, accounts forc.75%ofthetotaldynamicpressure
P
dyn
, decreases strongly downstream approximately following a power law
decay (Fig. 7c).
The largest particle size supported by turbulence in the experimental
ows, which is limited by the characteristic length- and velocity-scales of the
most energetic coherent turbulence structure (Eq. (3)) and associated with a
measured frequency f~ 1.3 Hz, corresponds to c. 2 millimetres (red vertical
barinFig.7a and b). Therefore, the experimental PDC strongly depletes in
particles larger than 2 millimetres even before a proper gravity current
structure has formed at around 1.8 m after impact. Particle sizes smaller
than 2 mm (associated with the frequency of the largest coherent structure),
but larger than 117229 µm (associated with the frequencies at the energy
maximum of dynamic pressure fεmax of c. 716 Hz) do not form notable
peaks in dynamic pressure energy (Fig. 6). This is explained by the rare
abundance of particles of this size range, which occupy the coarse tail of ow
grainsize distributions (Fig. 7b).
To visualize the process of clustering of particles with critical Stokes
numbers at the margins of coherent turbulence structures we conducted a
numerical simulation of the experimental PDC using a large eddy resolving,
Eulerian-Eulerian approach with particle-uid four-way coupling. In this
multiphase simulation, ve different particle sizes are modelled (2 mm,
0.5 mm, 0.125 mm, 0.032 mm and 0.008 mm) whose relative proportions
and particle size-dependent densities are the same as in the physical
experiment (see Methods for details). Figure 8depicts contour plots of
Fig. 7 | Particle sizes of critical Stokes numbers and their sequential sedi-
mentation. a The frequency fεassociated with the condition of particle Stokes
number St = 1 as a function of grainsize. The grey horizontal bar depicts the
frequency band in fεof c.717 Hz of highest-pressure energies. The grey
vertical bar delimit the experimentally determined corresponding critical
grainsizes for St =1atfεof c. 127224 µm. The red horizontal bar and the red
vertical bar mark the frequency associated with the most energetic coherent
ow structure at a frequency f~ 1.3 and the corresponding grainsize of c.
2 mm, respectively. bHistograms of time-integrated ow grainsize
distributions captured in ow samplers at ow heights of 0.45 m above the ow
base for various owdistances.Asina, the grey vertical bar highlights the
grain-size range of c. 127224 µm associated with fε,whiletheredverticalbar
marks the critical grainsize of c. 2 mm supported by the most energetic
coherent turbulence structures at f~ 1.3. Note the downstream depletion of the
ow in critical Stokes number particles associated with fε(c. 127224 µm).
cThe fractional time-integrated energy of particle impacts P
impact
relative to
theentirepressureenergyP
dyn
as a function of ow distance. The black line is a
best-t power law to the data.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
particle volume concentration of the evolving current at approximately half
of the total runout length for each particle size. High Stokes number particles
(2 mm in diameter; Fig. 8a) do not highlight any coherent turbulence
structures and have sedimented into the basal ow. Low Stokes number
particles (0.032 mm and 0.008 mm in diameter; Fig. 8d and e) remain
homogeneously suspended throughout the ow or tend to concentrate
slightly in the central regions of coherent structures. By contrast, particles
with diameters of 0.5 and 0.125 mm concentrate at the peripheries of eddies
highlighting the coherent turbulence structure of the ow (Fig. 8b and c).
Similar to our physical experiment, the highest particle concentration in
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 8
Content courtesy of Springer Nature, terms of use apply. Rights reserved
eddy peripheries occurs for particles with diameters of 0.125 mm. The
average size of the large coherent structures (highlighted by concentrated
margins) in the numerical simulation accounts to approximately 0.5 m
(Fig. 8c). This corresponds to the length-scales of c. 0.250.6 m that are
associated with the narrow band of maxima in dynamic pressure as seen in
the energy spectra of dynamic pressure as a function of wavenumber (where
thelength-scaleistheinverseofthewavenumberk;Fig.6dande).
Spatiotemporal evolution of dynamic pressure uctuations
To assess the destruction potential of PDCs, volcanologists routinely use
estimates of either bulk ow or local time-averaged ow velocity and density
values to evaluate average dynamic pressures26,4143.Ournding of the
concentration of particles with critical Stokes numbers (i.e., St ¼Oð1Þ)at
the peripheries of coherent turbulence structures imply strong uctuations
of dynamic pressure to occur at the characteristic frequency fεmax of these
structures. To prevent underestimation of the destruction potential of
PDCs, it is important to understand how turbulent uctuations of dynamic
pressure around eld-estimated average values evolve during ow
propagation.
Figure 9a compares the time-averaged dynamic pressure P
dusty
gas_ave
, the maximum dynamic pressure of the continuum phase P
dusty
gas_max
and the maximum dynamic pressure associated with particle
impacts P
impact max
as a function of ow distance. P
dusty gas_ave
(i.e, the
equivalent of the time-averaged loading pressure estimated in hazard
assessments to assess damage to infrastructure) steadily decreases with
ow distance Dfollowing a powerlaw of the form Pdusty gas ave /1=D.By
contrast, and up to 9.6 m, the maximum loading pressure P
dusty gas_max
exceeds the time-averaged pressure by an order of magnitude. In distal
reaches, D>9.6m, P
dusty gas_max
starts to decline more strongly. Max-
imum dynamic pressures from particle impacts P
impact max
exceed
P
dusty gas_ave
by two orders of magnitude, strongly decline with distance
and are restricted to the proximal ow reaches of D5.4 m.
Figure 9bshowsthepressureratioofP
dusty gas_max
over P
dusty gas_ave
against ow distance. This ratio assesses by how much current eld-based
estimates of dynamic pressure underestimate actual, turbulence-enforced
maximum loading pressures and can be written as:
Pdusty gas max
Pdusty gas ave
¼ρmaxUmax 2
ρaveU2
ave
:ð7Þ
The pressure ratio is independent of scale, and experimental estimates
can be applied to natural ow scales. However, the spatiotemporal evolution
of the dynamic pressure ratio is limited by the abundance of large Stokes
number particles inside ows. Following the decoupling of high Stokes
number particles from the margins of coherent structures, their motion is
independent of uid turbulence and results in their rapid sedimentation
from the ow.
Immediately following formation of a gravity current structure at c.
1.8 m and up to 5.4 m, the pressure ratio continuously increases from initial
values of c. 11.8 to maxima of c. 13 in 0.74 s (Fig. 9b). This duration coincid es
with the characteristic eddy timescale of the largest coherent structure of
τε=1/f~ 1/1.3 = 0.77 s. This suggests that the characteristic timescale of the
motion of high Stokes number particles from a homogeneous suspension
into eddy peripheries can be estimated by the overturn time of the largest
coherent structure.
After the time τε, the particle volume fraction θVof particles with
St ¼Oð1Þdeclines below 104. This appears to correspond to the limiting
mass loading for maximum particle clustering to occur and, downstream of
5.4 m, the pressure ratio decreases strongly. At 20m, where the volume
fraction of particles with St ¼Oð1Þhas decreased to values of
θV5×106, the pressure ratio approaches a value of 3.84. In the case of
exhaustion of St ~ 1 particles carried inside ows, Eq. (7) reduces to:
Pdusty gas max
Pdusty gas ave
Umax
Uave

4
ð8Þ
(see Supplementary note 2). Our velocity measurement of the max-
imum and average velocity at 20 m give a value of Umax=Uave

4of 3:9in
Eq. 8, which closely corresponds with the measured pressure ratio of 3.84.
Discussion
Current approaches to assess the destruction potential of pyroclastic density
currents envisage the loading force of a continuum multiphase uid of
characteristic velocity and density across structural surfaces opposing ow
direction.e.g. 3740 In the absence of detailed measurements inside ows,
volcanologists estimate local time-averaged values of ow velocity and
density from deposit or damage characteristics to guide assessments of
potential hazard impact15,4143. Our simulations of pyroclastic density cur-
rents through large-scale physical experiments and numerical modelling
demonstrate a fundamental shortcoming of this approach: the modication
of the ow and turbulence structure due to coupled feedbacks between
particle and gas phases. This feedback leads to the preferential clustering of
large, high-momentum particles at the margins of coherent turbulence
structures (Fig. 10a) and generate two different types of distinct hazard
impacts (Fig. 10bc).
The rst type, the dynamic pressure of the continuous dusty gas phase
P
dusty gas
,signicantly exceeds the time-integrated dynamic pressure by up
to one order of magnitude (Fig. 10d). The cause of these strong turbulent
pressure uctuations is the modication of the ow into coherent structures
with high-density margins (Fig. 10a), spatially separating the ow into high
density (eddy peripheries) and lower density domains (central regions of
eddies). The size of the coherent structures, and hence the characteristic
frequency of high-density/high dynamic pressure ow pulsing fεmax ,is
determined by the abundance of particles whose characteristic response
time to uid motion is equal to or smaller than the characteristic time-scale
of changes in the uid motion (that is, particle Stokes numbers St 1).
Depending on the mass loading of particles with this critical Stokes number
condition, the ratio of the clustering-enforced maximum dynamic pressure
Pdusty gas and the eld-estimated average pressure can range between c.413.
To prevent an underestimation of the intensity of hazard impact, we
strongly suggest that these factors are applied to conventional estimates of
dynamic pressure that apply bulk ow values of ow velocity and
density15,4143. For instance, for two- to three-story, reinforced brick, stone
and concrete buildings, a mean dynamic pressure of 5 kPa results in failure
of doors and partial damage to door and window frames. By comparison,
413 times larger maximum clustering-enforced dynamic pressures of
2065 kPa lead to failure of peaked roofs and front-facing exterior walls (c.
20 kPa) and complete failure of all building elements (c. 65 kPa)15.Oscil-
lation in dynamic pressure at the frequency fεmax further intensies the
destructiveness of ows due to the progressive structural weakening of
Fig. 8 | Clustering of critical Stokes number particles at the peripheries of
coherent structures. aeSnapshots from a numerical multiphase Eulerian-Eulerian
uid particle four-way coupled simulation of the physical experiment at approxi-
mately mid-ow runout distance. The contour plots show the spatial variation across
the simulated PDC in the particle concentration of ve modelled grainsizes of 2 mm
(a), 500 µm (b), 125 µm (c), 32 µm (d) and 8 µm (e). Technical aspects of the
numerical simulation are detailed in the Methods section. Particles with Stokes
numbers St >1(a), which are only supported by turbulence at the s cales of the largest
coherent structures, have sedimented into the basal ow and do highlight any
coherent structures at mid-ow runout. Particles with critical Stokes numbers St =1
(band c) concentrate at the peripheries of coherent structures. The effect is stronger
for particle sizes of 125 µm than for particle sizes of 500 µm because of the sig-
nicantly higher abundance of the 125 µm particles over 500 µm particles. Particles
with low Stokes numbers St <1(dand e) tend to be homogeneously suspended or
slightly concentrate in the central parts of coherent structures.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
materials amid repeated high-pressure loading. The damage inicted
depends on the maximum dynamic pressure associated with the high-
density margins of coherent turbulence structures and the characteristic
integral length scale of these structures relative the characteristic length scale
of an infrastructure. Our experimental method does not allow us to deter-
mine the range of three-dimensional geometries of coherent turbulence
structures and future physical and numerical experiments are needed to
address this problem. However, our experimental results can give some
guidance. The characteristic length scale of the most energetic turbulence
structures with particle clusters of c. 0.250.6 m is of the order of the height
of the lower ow boundary to roughly half of the thickness of the body the
density current. This length scale (i.e, half the thickness of the PDC) is
considerably larger than the height of the common built environment at the
Earths surface. Limited by the mass loading of particles with particle Stokes
numbers of order unity in a PDC, this indicates that clustering-enforced
destructiveness will be effective in typical PDC/infrastructure interactions.
The modication of the ow and ow turbulence structure through particle-
gas feedback also implies associated modications of the local concentration
of readily respirable ne-ash particles and local ow temperature, with
effects on the respiratory and burns hazards of PDCs.
The second type of hazard impact, the force of particle collisi ons, which
is associated with (and measured in our experiments as) the dynamic
pressure of particle impacts P
impact
, can exceed average dynamic pressures,
even stronger than P
dusty gas
, by two orders of magnitude (Fig. 10d). Unlike a
loading continuum pressure, this pressure corresponding to the force of
particle impacts results when particles with Stokes numbers St 1uncouple
from the ow that engulfs a structure and directly impinge onto it. The
resulting piercing particle impacts can be seen as pockmarks on trees and
infrastructure55 (Fig. 10c) and are illustrated for the case of the Merapi 2010
eruption in Supplementary Fig. 5. Piercing particle impacts from PDCs are
also likely contributing to the failure of brittle structures such as glass
windows38,55. Importantly, the magnitude of particle impact forces, because
they are distributed over relatively small impact areas of the order of the
cross-sectional area of a particle, must be considered as a potential con-
tributor to the high fatality and injury rates in PDC-forming eruptions
(Fig. 10d).
Our experiments demonstrate the role of turbulent particle-gas feed-
back in modifying the ow and turbulence structure and exacerbating the
intensities of hazard impacts. These ndings are also applicable to other
types of high-Reynolds number, polydisperse particle-gas ows, such as
powder snow avalanches, dust storms, and the debris-laden basal ow
regions of tornados and explosions, and should be considered in the
assessments of their potential hazard impacts28,5659.
Methods
Large-scale experiments
The Pyroclastic ow Eruption Large-scale Experiment (PELE), which is
fully described in ref. 44., is an international test facility to synthesize, view
and measure inside the interior of pyroclastic density currents. Experi-
mental PDCs of up to six tonnes of volcanic material and gas move at
velocities of 732 m s1,are24.5 m thick and propagate to runout length of
>35 m. We conducted a series of three large-scale experiments to conrm
that the experimental results reported are typical. The currents are generated
by the controlled gravitational collapse of variably diluted suspensions of
pyroclastic particles and gas from an elevated hopper into an instrumented
runout section. PELE is operated inside a 16 m high, 25 m long and 18 m
wide disused boiler house. The apparatusiscomposedoffourmainstruc-
tural components: (i) Tower. A 13 m-high construct that lifts either a 4.2 m3
Fig. 9 | Spatiotemporal evolution of dynamic pressure during ow propagation.
aDynamic pressure as a function of ow distance D. The red circles show time-
averaged values of dynamic pressure Pdusty gas ave. The red line is a best-t powerlaw
through the data that yields Pdusty gas ave ¼136D1:019 . The blue square symbols
represent the maximum dusty gas pressure Pdusty gas max. The black circles show
measurements of the maximum particle impact pressure Pimpact max .bThe pressure
ratio Pdusty gas max =Pdusty gas ave as a function of ow distance (black diamonds). The
contour plot shows the particle volume concentration of particles with diameters
>125 μm, associated with the condition St O1ðÞ. The vertical red dotted lines
demark particle concentrations of, from left to right, 103,10
4, and 105.
The secondary, non-linear horizontal time axis depicts the times of ow front arrival
corresponding to these distances. After formation of a gravity current structure at c.
1.8 m and up until 5.4 m, the pressure ratio increases slightly to maximum values of
around 13. The ow duration associated with this increase coincides with the eddy
time scale tεhighlighted on the secondary x-axi s. Downstream from 5.4 m, the
pressure ratio decreases in association with a reduction in the particle concentration
of critical Stokes number particles larger than 125 µm. The horizontal dashed line
corresponds to the condition predicted by Eq. (8) where the pressure ratio takes a
critical value of c. 3.9, when the ow is depleted in particles with St 1.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved
hopper (for moderate to high discharge rates of 3001,500 kg s1)ora0.7m
3
hopper (for discharge rates of 30200 kg s1) to the desired discharge height.
The hoppers are equipped with internal hopper heating units to heat the
pyroclastic material to target temperatures of up to 400 °C. The heating
process is measured by thermocouples and the hopper is mounted onto four
load cells to capture the time-variant mass discharge. (ii) Column. A 9m
high shroud of heat-resistant cloth through which the discharge particle-air
mixtures accelerate under gravity. (iii) Chute. A 17 m long and 0.5 m wide
multi-instrumented channel section with 0.61.8 m high sides of
temperature-resistant glass. The rst 12 m are variable adjustable to
slope angle between and 25° while last 5 m of the channel is horizontal.
(iv) Outow. A 25 m-long at instrumented and unconned runout section
that extends outside the building. The physical properties of the particle-gas
suspensions prior to impact with the channel (velocity, mass ux, volume
ux, particle concentration), the characteristics of the solid components
(grainsize distribution, particle density distribution, particle temperature),
and boundary conditions (substrate roughness, chute slope and channel
width)can be modied to generate a wide range of reproducible natural ow
conditions44. For the experiments reported in this study, we used the small
hopper of 0.7 m3to generate fully turbulent expe rimental PDCs with a basal
bedload region, but without a dense basal underow, which would form at
the high discharge rates in the large hopper setup condition.
The use of pyroclastic solids material and air is an important pre-
requisite to generate natural stress coupling between the uid and solid
Fig. 10 | The effects of gas-particle feedbacks on pyroclastic density current
hazards. a Schematic presentation of the processes of turbulent gas-particle feed-
back in PDCs at the scale of an idealized eddy at three different times. For the
polydisperse multiphase ow, the spectrum of concurrent behaviours of particles
with different degrees of coupling to turbulent uid motion is illustrated for particles
with three different particle Stokes numbers St.Left eddy: initial stage where all
particles are homogenously dispersed. Middle eddy: intermediate stage, after the
overturn time of the largest coherent turbulence structure. St ~ 1 particles pre-
ferentially cluster on eddy margins, while St « 1 particles remain homogenously
dispersed following turbulence motion and St » 1 particles move unhindered by
turbulence. Right eddy: later stage, when clusters of dominantly St ~ 1 particles
decouple from eddy margins and sediment as mesoscale turbulence clusters.
bExample of the damage effects of clustering-enforced dynamic loading pressure
P
dusty gas
as repeated high-pressure impacts at the frequency fεmax .cExample of the
piercing damage effects of the direct impact of high St particles or particle clusters
with structures generating pockmarks on the upstream face of a concrete pole.
dSchematic probability density function of dynamic pressure P
dyn
relative to the
average dynamic pressure
Pdyn concurrently considered in hazard assessments. The
dynamic pressure of the continuum dusty gas phase P
dusty gas
exceeds the average
pressure by one order of magnitude. The dynamic pressure associated with particle
impacts P
impact
exceeds the average dynamic pressure by two orders of magnitude.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
phases. The pyroclastic material, containing particle sizes from 2 µm to
16 mm, consists of a mixture of two well-characterized ignimbrite deposits
F1 and F2 from the 232 CE Taupo eruption47.Therst component (F1) is a
proximal medium-ash-dominated ignimbrite deposit with a unimodal
grain-size distribution, a median diameter of 366 µm, and 4.5 wt.% of
extremely ne ash (<63 µm). The second component (F2) is a ne ash rich
facies from the base of the proximal Taupo ignimbrite deposit with a
polymodal grain-size distribution, a median diameter of 103 µm, and 36.5
wt.% of extremely ne ash. The experiments reported in this study involve a
material blend with F1 = 60 wt.% and F2 = 40 wt.% (the grain-size dis-
tribution is shown in Supplementary Fig. 2a) yielding a mixture with 20
wt.%ofparticlessmallerthan6m.Themainsolidscomponentsofthe
pyroclastic material are highly vesicular pumice, glass shards, free crystals
and lithic particles. The relative proportion of these components vary with
grain-size. The average particle density as a function of particle diameter is
showninSupplementaryFig.2b.
The resulting experimental pyroclastic density currents are fully tur-
bulent with Reynolds numbers in the lower 106(and up to 107in proximal
regions). Dimensionless products quantifying the scaling similarity of nat-
ural and experimental currents for the bulk ow are depicted in Supple-
mentary Table 1. Further details of the experimental protocol,propertiesthe
volcanic material, and measurement techniques are reported in ref. 44,but
the measurements and analytical methods specic to the results presented
here are detailed below.
Sensors and analytical methods
Twenty fast cameras (60120 frames per second) and three normal-speed
cameras (2430 frames per second) positioned at different distances and
viewing angles, recorded the downstream propagation of the experimental
pyroclastic density currents. At ow distances of 1.8 m, 3.35 m, 5.4 m, and
9.6 m, four highspeed camera proles consisting of six highspeed cameras
recorded vertical proles of the passing ows at 500 frames per second. LED
oodlight arrays were used to achieve sufcient and even illumination,
which allowed for a detailed analysis of the gas-particle transport and
sedimentation with particle image analysis (PIV; using the algorithm
PIVlab60). Two-dimensional velocity elds were derived with PIV from the
highspeed videos at time intervals of 2 ms.
Timeseries of dynamic pressure P
dyn
(t) were measured with piezo-
electric pressure sensors (PCB Piezotronics 106B51 and signal conditioners
PCB 483 C) at the ow centerline, at owdistancesof1.8m,3.35m,5.4m,
9.6 m, 16 m, and 20 m from impact and were recorded at a sampling rate of
1 kHz. The sensors consist of a quartz crystal encapsulated in a solid steel
casing. The deformation of the steel casing due to pressure variations is
transferred to the quartz crystal, which induces a voltage that is directly
recorded as calibrated dynamic pressure signal. In our experiments, we
exposed only the frontal face of the sensor to the ow. This frontal face is
covered by a circular, 15.7 mm diameter, frontal steel diaphragm. The
sensors were mounted into 0.1 m-long bullet-shaped sensor mounts to
reduce ow separation at the upstream-directed sensor head. The encased
pressure sensors protruded out of 1.8 m-heigh wing-shaped proles
designed to reduce ow separation behind the proles and to guide cabling.
In this study, we report the results of dynamic pressure measurements
obtained at approximately the mid-depth of the density current body region
at a slope-normal height z=0.45m above the ow base. Supplementary
Note 1 and Supplementary Fig. 3 show a comparison of the measured
dynamic pressure signals due to the forces of impacts of individual particles
with the sensor with theoretical estimates of impact pressures for the cases of
purely elastic and purely inelastic collisions. For the case of inelastic particle
collisions, the particle collision time is given by the ratio of the diameter of
the particle colliding with the sensor and the velocity of impact. Due to the
fast response time of the dynamic pressure sensors of 12 μsand taking a
typical ow velocity of 8 ms1as reference, collisions of particles larger than
c. 100 μmcan be detected in our experiments.
To test the sensor response to particle impacts, we conducted addi-
tional laboratory-scale experiments using a particle gun and spherical glass
beads with three different ranges in particle diameter. The particle gun
consists of a 1 m long, 0.01 m diameter steel pipe connec ted to a compressed
air line. The air velocity was regulated through an inlet air-pressure valve
and the gas velocity was measured with a hotwire anemometer. A funnel
system in the particle gun allows the feeding of individual particles and
mixtures of particles into the airstream. The dynamic pressure sensor was
positioned 2 cm in front of the particle gun. Collision experiments were
performed with individual, spherical glass bead particles with particle dia-
meters of 170180 µm, 300355 µm and 400 µm and the pressure signal was
recordedatthesamesamplingrateof1kHzasinthelarge-scaleexperi-
ments. Particles were fed into the particle gun when the transient signal of
the piezoelectric dynamic pressure sensor due to the dynamic pressure of the
airow had decayed to a zero Pascal baseline. This allowed an analysis of the
recorded dynamic pressure signal associated with the impact force of a
particle colliding with the sensor without removing the component of
dynamic pressure corresponding to the air ow alone.
As detailed in Supplementary Note 1 and Supplementary Fig. 3, the
signals of dynamic pressure recorded by the dynamic pressure sensors due
to the force of a particle colliding with the sensor diaphragm is the impact
force distributed over the surface area of the particle impact (i.e, the cross-
sectional area of the particle). This is different from a piezoelectric force
sensor where the force is distributed over the entire surface of the sensor
head. In Supplementary Note 1 and Supplementary Fig. 3 it is also shown
that the particle collisions with the dynamic pressure sensor are close to and
well approximated as purely inelastic collisions.
At ow distances of 0.5 m, 1.8 m, 3.35 m, 5.4 m, 9.6 m, 12 m, 16 m, and
20 m from impact vertical arrays of transparent sediment samplers collected
the owing mixture. During the experiment, the sequential lling the ow
samplers was lmed with highspeed cameras. These samplers are open on
the upstream side allowing the ow to enter through the 1.69 cm2cross-
sectional area while on the downstream side, a 16 microns mesh allows only
the gas-phase of the ow to exit, leading to accumulation of the transported
particles inside the sampler. From this we measure continuous data of ow
sediment passing a position as a function of time. Downslope velocity
components u(t) of the ow at a position 5 cm upstream of each ow
sampler were obtained through PIV. The weight and density of the material
deposited inside the ow samplers was measured at selected time intervals to
calculate the time-variant porosity of the captured sediment, as well as the
sediment grain-size distribution. Particle solids-concentration C
S
are
dened as
Csz;tðÞ¼
Vdð1εÞ
uA
otð9Þ
where V
d
is the time-variant sediment accumulated volume inside the ow
sampler, uis the time-variant downslope velocity obtained through PIV at
the entrance of the ow sampler, A
0
is the cross-sectional area of the ow
sampler, tthe selected time intervals, εthe time-variant sediment porosity,
and zthe height in slope-perpendicular direction.
At each of the ow sampler locations, time-variant and height-variant
temperature T(z, t) is measured in vertical arrays of fast thermocouples (410-
345 Type K) with a sampling rate of 70 Hz. The response time of the
thermocouples is approximately 1 millisecond and hence faster than the
sampling rate. Together with the time-series of time-variant and height-
variant ow velocity u(z, t), grain-size distributions, volumetric particle
concentrations C
S
(z, t), the temperature time-series allow for the calculation
of dynamic pressure P
dyn_Bernoulli
dened by Eq. (1)wherebulkow density
ρCz;tðÞis given by
ρC¼CSρPþPa
γgRgT1CS
 ð10Þ
where ρPistheparticledensity,P
A
is the ambient pressure, γgis the mass
fraction of the gas components (including moisture) and R
g
is their gas
constants.
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 12
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Energy spectra
For the computation of discrete Fourier spectra, the NumPy FFT function of
ref. 61. is used. From the complex output provided by the FFT functions only
the real part is used. Due to the symmetry of the Fourier transform and
physical meaning only positive frequencies are considered.
For the calculation of wave numbers, the time axis (t) of the data is
transferred into advected distances (x) by multiplication with time averaged
velocities (x¼
ut). Using the data series as function of advected distance,
Fourier transforms are calculated using the NumPy implementation. Also,
for these spectra only the real and positive part of the output is used.
Multiphase ow simulation of the PELE experiments
Numerical simulations using the Eulerian-Eulerian approach , also known as
the two-uid method (TFM), were performed to reproduce the experi-
mental currents. We employed the open-source Multiphase Flow with
Interphase eXchanges (MFIX) solver, developed by the National Energy
Technology Laboratory (NETL) within the US Department of Energy. This
technique enables us to solve mass, momentum, and energy equations for
both uid and solid phases, capturing solid-uid 4-way coupling. A com-
prehensive description of all equations can be found in ref. 62.
We executed the 3D ow simulation on the ARCHER2 cluster utilizing
1200 CPU cores with 3 Tb of RAM for 15 days, employing the large-eddy-
simulations (LES) with the wall-adapting local eddy-viscosity subgrid model
WALE of ref. 63. to model subgrid turbulence. The initial and boundary
conditions for the simulations were established based on the international
benchmarking and validation exercise for PDC model64 .
The top boundary of the domain is described as a pressure boundary
that creates a stratied atmosphere. All other boundaries, except the inlet,
follow the no-slip boundaries for the uid phase and partial slip for the solid
phase, following the work of ref. 65. The experimental geometry was
uploaded in MFIX as an STL le upon which we created a roughness of
approximately 0.01 m in height to emulate the basal boundary conditions.
The mesh was cut to follow the geometry using the cutcell method62.
The inlet boundary is treated as a mass inow, where a ux of tem-
perature, uid and solid concentration and velocity were prescribed. The
continuous grain-size distributions of the experimental mixture was dis-
cretized in ve bins of particles with a diameter of 2×103m, 5×104m,
125×106m, 32×106m and 8×106m. This representation preserves the
mean size diameter D[3,2] (surface-mean diameter) and D[4,3] (volume-
mean diameter) consistent with the experimental counterpart. We also
discretized the solid density distribution into the same respective ve bins to
ensure accurate gas-particle coupling in the simulation. In order to match
experiments closely, the inlet used time- and height-variant velocity, con-
centration, grain-size distribution and temperature proles described
by ref. 66.
At the inlet, we assumed thermal and kinetic equilibria, resulting in
uniform temperature and no-slip velocity across all phases in each inlet
computational cell. However, the numerical simulation captured thermal
and kinetic decoupling, dictated by particle Stokes number as shown in
Fig. 8. Using the nite volume method with a second-order solver and
Superbee limiter, the MFIX solves mass, momentum and energy equations.
This approach allows us to derive the temperature, velocity, and con-
centration elds for the uid and each solid phase in each computation cell.
The results were saved as binary VTK les, which were loaded in Paraview®
v5.11 using the Talapas HPC at the University of Oregon and subsequently
exported as images. The input and boundary conditions used in the
simulations are summarized in Supplementary table 2.
Data availability
The data generated in this study are available at https://doi.org/10.5281/
zenodo.10725051.
Code availability
The codeused to produce the data analysesof the experimental data is freely
available at https:/ww.python.org. The code used to conduct the large-eddy
resolving multiphase simulation is freely available at https://mx.netl.doe.
gov/products/mx/.
Received: 20 October 2023; Accepted: 4 March 2024;
References
1. Neri, A., Esposti Ongaro, T., Voight, B. & Widiwijayanti, C.
Pyroclastic Density Current Hazards and Risk, in: Papale, P.,
Shroder, J. (Eds.), Volcanic Hazards, Risks and Disasters. Elsevier
Science (2014).
2. Dufek, J., Esposti Ongaro, T. & Roche, O. Pyroclastic Density
Currents: Processes and Models, in Sigurdsson et al. (Eds.)
Encyclopedia of Volcanoes, 2nd Ed., pp 617-629 (2015).
3. Sulpizio, R., Dellino, P., Doronzo, D. M. & Sarocchi, D. Pyroclastic
density currents: state of the art and perspectives. J. Volcanol.
Geotherm. Res. 283,3665 (2014).
4. Lube, G., Breard, E. C. P., Esposti-Ongaro, T., Dufek, J. & Brand, B.
Multiphase ow behaviour and hazard prediction of pyroclastic
density currents. Nat. Rev. Earth Environ. 1, 348365 (2020).
5. Dufek, J. The uid mechanics of pyroclastic density currents. Ann.
Rev. Fluid Mech. 48, 459485 (2016).
6. Chester, D., Degg, M., Duncan, A. & Guest, J. The increasing exposure
of cities to the effects of volcanic eruptions: a global survey. Glob.
Environ. Change B: Environ. Hazards 2,89103 (2000).
7. Small, C. & Naumann, T. The global distribution of human population
and recent volcanism. Environ. Hazards 3,93109 (2001).
8. Auker, M. R., Sparks, R. S. J., Siebert, L., Crosweller, H. S. & Ewert, J. A
statistical analysis of the global historical volcanic fatalities record. J.
Appl. Volcanol. 2, 2 (2013).
9. Valentine, G. A. Stratied ow in pyroclastic surges. Bull. Volcanol. 49,
616630 (1987).
10. Lube, G. et al. Kinematic characteristics of pyroclastic density
currents at Merapi and controls of their avulsion from natural
and engineered channels. Geol. Soc. Am. Bull. 123,
11271140 (2011).
11. Lube, G. et al. Dynamics of surges generated by hydrothermal blasts
during the 6 August 2012 Te Maari eruption, Mt. Tongariro, New.
Zealand. J. Volcanol. Geotherm. Res. 286, 348366 (2014).
12. Cronin, S. J. et al. Insights into the OctoberNovember 2010 Gunung
Merapi eruption (Central Java, Indonesia) from the stratigraphy,
volume and characteristics of its pyroclastic deposits. J. Volcanol.
Geotherm. Res. 261, 244259 (2013).
13. Charbonnier, S. J. et al. Unravelling the dynamics and hazards of the
June 3rd, 2018, pyroclastic density currents at Fuego volcano
(Guatemala). J. Volcanol. Geotherm. Res. 436, 107791 (2023).
14. Huppert, H. E. & Simpson, J. E. The slumping of gravity currents. J.
Fluid Mech. 99, 785799 (1980).
15. Valentine, G. A. Damage to structures by pyroclastic ows and surges,
inferred from nuclear weapons effects. J. Volcanol. Geotherm. Res.
87, 117140 (1998).
16. Andrews, B. J. & Manga, M. Effects of topography on pyroclastic
density current runout and formation of co-ignimbrites. Geology 39,
10991102 (2011).
17. Andrews, B. J. & Manga, M. Experimental study of turbulence,
sedimentation, and coignimbrite mass partitioning in dilute
pyroclastic density currents. J. Volcanol. Geotherm. Res. 225,
3044 (2012).
18. Brosch, E. et al. Characteristics and controls of the runout behaviour
of non-Boussinesq particle-laden gravity currents A large-scale
experimental investigation of dilute pyroclastic density currents. J.
Volcanol. Geotherm. Res. 432, 107697 (2022).
19. Breard, E. C. P., Lube, G., Cronin, S. J. & Valentine, G. A. Transport and
deposition processes of the hydrothermal blast of the 6 August 2012
Te Maari eruption, Mt. Tongariro. Bull. Volcanol. 7, 100 (2015).
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 13
Content courtesy of Springer Nature, terms of use apply. Rights reserved
20. Dellino, P. et al. Experimental evidence links volcanic particle
characteristics to pyroclastic ow hazard. Earth Planet. Sci. Lett. 295,
314320 (2010).
21. Cole, P. D., et al (Eds.), The Encyclopedia of Volcanoes, 943-956,
Academic Press (2015).
22. Dellino, P. et al. The impact of pyroclastic density currents duration on
humans: the case of the AD 79 eruption of Vesuvius. Sci. Rep. 11,
4959 (2021).
23. Baxter, P. J. et al. Human survival in volcanic eruptions: Thermal
injuries in pyroclastic surges, their causes, prognosis and emergency
management. Burns 43, 10511069 (2017).
24. Pensa, A., Capra, L. & Giordano, G. Ash clouds temperature
estimation. Implication on dilute and concentrated PDCs coupling
and topography connement. Sci. Rep. 9, 5657 (2019).
25. Elghobashi, S. On predicting particle-laden turbulent ows. Appl. Sci.
Res. 52, 309329 (1994).
26. Burgisser, A. Physical volcanology of the 2,050 BP caldera-forming
eruption of Okmok volcano, Alaska. Bull. Volcanol. 67,
497525 (2005).
27. Burgisser, A. & Bergantz, G. W. Reconciling pyroclastic ow and
surge: the multiphase physics of pyroclastic density currents. Earth
Planet. Sci. Lett. 202, 405418 (2002).
28. Brandt, L. & Coletti, F. Particle-laden turbulence: progress and
perspectives. Ann. Rev. Fluid Mech. 54, 159189 (2021).
29. Breard, E. C. P. et al. Coupling of turbulent and non-turbulent ow
regimes within pyroclastic density currents. Nat. Geosci. 9,
767771 (2016).
30. Breard, E. C. P. & Lube, G. Inside pyroclastic density currents
Uncovering the enigmatic ow structure and transport behaviour in
large-scale experiments. Earth Planet. Sci. Lett. 458,2236 (2017).
31. Weit, A., Roche, O., Dubois, T. & Manga, M. Experimental
measurement of the solid particle concentration in geophysical
turbulent gas-particle mixtures. J. Geophys. Res. 123,
27472761 (2018).
32. Breard, E. C. P., Dufek, J. & Lube, G. Enhanced mobility in
concentrated pyroclastic density currents: An examination of a self-
uidization mechanism. Geophys. Res. Lett.45:https://doi.org/10.
1002/2017GL075759 (2018).
33. Valentine, G. & Sweeney, M. R. Compressible ow phenomena at
inception of lateral density currents fed by collapsing gas-particle
mixtures. J. Volcanol. Geotherm. Res. 123, 12861302 (2018).
34. Lube, G. et al. Pyroclastic ows generate their own air-lubrication. Nat.
Geosci. 12, 381386 (2019).
35. Brosch, E. & Lube, G. Spatiotemporal sediment transport and
deposition in experimental dilute pyroclastic density currents. J.
Volcanol. Geotherm. Res. 401,116 (2020).
36. Brosch, E. et al. Destructiveness of pyroclastic surges controlled by
turbulent uctuations. Nat. Comm. 12, 7306 (2021).
37. Clarke, A. B. & Voight, B. Pyroclastic current dynamic pressure from
aerodynamics of tree or pole blow-down. J. Volcanol. Geotherm. Res.
100, 395412 (2000).
38. Baxter, P. J. et al. The impacts of pyroclastic surges on buildings at the
eruption of the Soufriere Hills volcano, Montserrat. Bull. Volcanol. 67,
292313 (2005).
39. Jenkins, S. et al. The Merapi 2010 eruption: An interdisciplinary
impact assessment methodology for studying pyroclastic
density current dynamics. J. Volcanol. Geotherm. Res. 261,
316329 (2013).
40. Gardner, J. E., Nazworth, C., Helper, M. A. & Andrews, B. J. Inferring
the nature of pyroclastic density currents from tree damage: The 18
May 1980 blast surge of Mount St. Helens, USA. Geology 46,
795798 (2018).
41. Dellino, P., Mele, D., Sulpizio, R., La Volpe, L. & Braia, G. A method for
the calculation of the impact parameters of dilute pyroclastic density
currents based on deposit particle characteristics. J. Geophys. Res.-
Solid Earth 113, B07206 (2008).
42. Dioguardi, F. & Mele, D. PYFLOW_2.0: a computer program for
calculating ow properties and impact parameters of past dilute
pyroclastic density currents based on eld data. Bull. Volcanol. 80,
28 (2018).
43. Mele, D. et al. Hazard of pyroclastic density currents at the Campi
Flegrei Caldera (Southern Italy) as deduced from the combined use of
facies architecture, physical modeling and statistics of the impact
parameters. J. Volcanol. Geotherm. Res. 299,3553 (2015).
44. Lube, G., Breard, E. C. P., Cronin, S. J. & Jones, J. Synthesizing large-
scale pyroclastic ows: Experimental design, scaling, and rst
results from PELE. J. Geophys. Res.-Solid Earth 120,
14871502 (2015).
45. Dellino, P. et al. Large-scale experiments on the mechanics of
pyroclastic ows: Design, engineering, and rst results. J. Geophys.
Res.-Solid Earth 112, B04202 (2007).
46. Sulpizio, R., Castioni, D., Rodriguez-Sedano, L. A., Sarocchi, D. &
Lucchi, F. The inuence of slope-angle ratio on the dynamics of
granular ows: insights from laboratory experiments. Bull. Volcanol.
78, 77 (2016).
47. Wilson, C. J. N. The Taupo Eruption, New Zealand. II. The Taupo
Ignimbrite. Philos. Trans. R. Soc. Lond. Ser. A Math. Phys. Sci. 314,
229310 (1985).
48. Bosse, T., Kleiser, L. & Meiburg, E. Small particles in homogeneous
turbulence: Settling velocity enhancement by two-way coupling.
Phys. Fluids 18, 027102 (2006).
49. Strouhal, V. Ueber eine besondere Art der Tonerregung. Ann. Phys.
241, 216251 (1878).
50. Turner, J. S. Buoyant plumes and thermals. Annu. Rev. Fluid Mech. 1,
2944 (1969).
51. Crow, S. C. & Champagne, F. H. Orderly structure in jet turbulence. J.
Fluid Mech. 48, 547591 (1971).
52. Martin, J. E. & Meiburg, E. The accumulation and dispersion of heavy-
particles in forced 2-dimensional mixing layers. 1. The fundamental
and subharmonic cases. Phys. Fluids 6, 11161132 (1994).
53. Raju, N. & Meiburg, E. Dynamics of small, spherical particles in vortical
and stagnation point elds. Phys. Fluids 9, 229314 (1997).
54. Dioguardi, F. & Mele, D. A new shape dependent drag correlation
formula for non-spherical rough particles. Experiments and results.
Powder Technol. 277, 222230 (2015).
55. Spence, R. et al. Residential building and occupant vulnerability to
pyroclastic density currents in explosive eruptions. Nat. Hazards Earth
Syst. Sci. 7, 219230 (2007).
56. Sovilla, B., McElwaine, J. N. & Köhler, A. The intermittency regions of
powder snow avalanches. J. Geophys. Res.: Earth Surf. 123,
25252545 (2018).
57. Zhang, F., Frost, D. L., Thibault, P. A. & Murray, S. B. Explosive
dispersal of solid particles. Shock Waves 10, 431443 (2001).
58. Balachandar, S. & Eaton, J. K. Turbulent dispersed multiphase ow.
Annu. Rev. Fluid Mech. 42, 111133 (2010).
59. Capecelatro, J. & Wagner, J. L. Gas-Particle Dynamics in High-Speed
Flows. Annu. Rev. Fluid Mech.3https://doi.org/10.1146/annurev-
uid-121021-015818. (2024)
60. Thielicke, W. & Sonntag, R. Particle Image Velocimetry for MATLAB:
Accuracy and Algorithms in PIVlab. J. Open Res. Softw. 9.https://doi.
org/10.5334/jors.334 (2021)
61. Harris, C. R. et al. Array programming with NumPy. Nature 585,
357362 (2020).
62. Musser, J., Vaidheeswaran, A. & Clarke, M. A. MFIX Documentation
Volume 3: Verication and Validation Manual; 3rd ed. NETL-PUB-
22050; NETL Technical Report Series; U.S. Department of
Energy, National Energy Technology Laboratory: Morgantown,
WV. (2021).
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 14
Content courtesy of Springer Nature, terms of use apply. Rights reserved
63. Ducros, F., Nicoud, F. & Poinsot, T. Wall-adapting local eddy-
viscosity models for simulations in complex geometries. Numer.
Methods Fluid Dyn. VI 6, 293299 (1998).
64. Esposti Ongaro, T., Cerminara, M., Charbonnier, S. J., Lube, G. &
Valentine, G. A. A framework for validation and benchmarking of
pyroclastic current models. Bull. Volcanol. 82, 51 (2020).
65. Johnson, P. C. & Jackson, R. Frictional-collisional constitutive
relations for the granular materials, with application to plane shearing.
J. Fluid. Mech 176, 223233 (1987).
66. Cerminara, M., Brosch, E. & Lube, G. A theoretical framework and the
experimental dataset for benchmarking numerical models of dilute
pyroclastic density currents. arXiv preprint https://arxiv.org/abs/
2106.14057. (2021).
Acknowledgements
The authorswould like to thank KevinKreutz and Anja Moebisfor assistance
with running the large-scale experiments. This study was partiallysupported
by the Royal Society of New Zealand Marsden Fund(contract no. MAU1902
to GL), the Resilience to Natures Challenges National Science Challenge
Fund New Zealand (GNSRNC047 to GL), the New Zealand Ministry of
Business, Innovation and Employments Endeavour Fund (contract no.
GNS-MBIE00216 to GL), theKAVLI Institute of Theoretical Physics Program
Multiphase Flows in Geophysics and the Environment(contract no. NSF
PHY-1748958 to EM, GLand JD), the NationalScience Foundation(contract
no. NSF EAR-1926025 to JD) and the National Environment Research
Council UKRI fund (contract no. NERC-IRF NE/V014242/1 to ECPB).
Author contributions
D.U., J.A.,E.B., L.C. and G.L.designed and conducted the experiments.G.L.
and D.U.co-wrote the manuscript, which wasthen revised by allthe authors.
D.U. led theanalysis and interpretationof the experimentaldata with the help
of G.L., E.C.P.B., E.M. and J.R.J., while E.C.P.B. and J.D. conducted and
post-processed the numerical simulations. E.B. and J.A. assisted with the
P.I.V., ow density and grain-size analyses. S.F.J. and G.L. analysed
pockmarks of the Merapi 2010 eruption. G.L. developed the PELE facility
together with J.R.J., E.B. and E.C.P.B.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information The online version contains
supplementary material available at
https://doi.org/10.1038/s43247-024-01305-x.
Correspondence and requests for materials should be addressed to
Daniel H. Uhle.
Peer review information Communications Earth & Environment thanks
Pierfrancesco Dellino, Karim Kelfoun and the other, anonymous, reviewer(s)
for their contribution to the peer review of this work. Primary Handling
Editors: Domenico Doronzo and Joe Aslin. A peer review le is available
Reprints and permissions information is available at
http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, 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 articles Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in the
articles 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://creativecommons.org/licenses/by/4.0/.
© The Author(s) 2024
https://doi.org/10.1038/s43247-024-01305-x Article
Communications Earth & Environment | (2024) 5:245 15
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
One of the most dangerous aspects of explosive volcanism is the occurrence of dilute pyroclastic density currents that move at high velocities of tens to about a hundred of metres per second outwards from volcanic vents. Predicting the runout behaviour of these turbulent flows of hot particles and air is complicated by strong changes in the flow density resulting from entrainment of ambient air, sedimentation of particles, as well as heating and expansion of the gas phase. Current hazard models that are based on the behaviour of aqueous gravity currents cannot capture all aspects of the flow dynamics, and thus pyroclastic density current dynamics remain comparatively poorly understood. Here we interrogate the runout behaviour of dilute pyroclastic density currents in large-scale experiments using hot volcanic material and gas. We demonstrate that the flows transition through four dynamic regimes with distinct density and force characteristics. The first, inertial regime is characterized by strong deceleration under high density differences between the flow and ambient air where suspended particles carry a main proportion of the flows' momentum. When internal gravity waves start to propagate from the flow body into the advancing flow front, the currents transition into a second, inertia-buoyancy regime while flow density continues to decline. In this regime, subsequent arrivals of fast-moving internal gravity waves into the front replenish momentum and lead to sudden short-lived front accelerations. In the third regime, when the density ratio between flow and ambient air decreases closer to a value of unity, buoyancy forces become negligible, but pressure drag forces are large and constitute the main flow retarding force. In this inertia-pressure drag regime, internal gravity waves cease to reach the front. Finally, and with the density ratio decreasing below 1, the current transitions into a buoyantly rising thermal in regime 4. Unlike for aqueous gravity currents, the Froude number is not constant and viscous forces are negligible in these gas-particle gravity currents. We show that, in this situation, existing Boussinesq and non-Boussinesq gravity current models strongly underpredict the front velocity for most of the flow runout for at least half of the flow propagation. These results are not only important for hazard mitigation of pyroclastic density currents but are also relevant for other turbulent gas-particle gravity currents, such as powder snow avalanches and dust storms.
Article
Full-text available
Pyroclastic surges are lethal hazards from volcanoes that exhibit enormous destructiveness through dynamic pressures of 10 ⁰ –10 ² kPa inside flows capable of obliterating reinforced buildings. However, to date, there are no measurements inside these currents to quantify the dynamics of this important hazard process. Here we show, through large-scale experiments and the first field measurement of pressure inside pyroclastic surges, that dynamic pressure energy is mostly carried by large-scale coherent turbulent structures and gravity waves. These perpetuate as low-frequency high-pressure pulses downcurrent, form maxima in the flow energy spectra and drive a turbulent energy cascade. The pressure maxima exceed mean values, which are traditionally estimated for hazard assessments, manifold. The frequency of the most energetic coherent turbulent structures is bounded by a critical Strouhal number of ~0.3, allowing quantitative predictions. This explains the destructiveness of real-world flows through the development of c. 1–20 successive high-pressure pulses per minute. This discovery, which is also applicable to powder snow avalanches, necessitates a re-evaluation of hazard models that aim to forecast and mitigate volcanic hazard impacts globally.
Article
Full-text available
PIVlab is a free toolbox and app for MATLAB®. It is used to perform Particle Image Velocimetry (PIV) with image data: A light sheet illuminates particles that are suspended in a fluid. A digital camera records a series of images of the illuminated particles. The input images are divided into sub-images (interrogation areas), and for each of these, a cross-correlation is performed. The resulting correlation matrix is used to estimate the most probable displacement within each interrogation area. PIV is extensively used for flow analyses where a thin laser sheet illuminates suspended particles in the fluid, but also for other moving textures, like cell migration or ultrasonic images. This paper presents several improvements that were implemented in PIVlab, enhancing the robustness of displacement estimates. The benefit of these improvements is evaluated using experimental images and synthetic images of particle and non-particle textures. Linear correlation and repeated correlation increase the robustness and decrease bias and root-mean-square (RMS) error of the displacement estimates. Particle images have a significantly lower bias and RMS error than non-particle images.
Article
Full-text available
Pyroclastic density currents are ground hugging gas-particle flows that originate from the collapse of an eruption column or lava dome. They move away from the volcano at high speed, causing devastation. The impact is generally associated with flow dynamic pressure and temperature. Little emphasis has yet been given to flow duration, although it is emerging that the survival of people engulfed in a current strongly depends on the exposure time. The AD 79 event of Somma-Vesuvius is used here to demonstrate the impact of pyroclastic density currents on humans during an historical eruption. At Herculaneum, at the foot of the volcano, the temperature and strength of the flow were so high that survival was impossible. At Pompeii, in the distal area, we use a new model indicating that the current had low strength and low temperature, which is confirmed by the absence of signs of trauma on corpses. Under such conditions, survival should have been possible if the current lasted a few minutes or less. Instead, our calculations demonstrate a flow duration of 17 min, long enough to make lethal the breathing of ash suspended in the current. We conclude that in distal areas where the mechanical and thermal effects of a pyroclastic density currents are diminished, flow duration is the key for survival.
Article
Full-text available
Array programming provides a powerful, compact and expressive syntax for accessing, manipulating and operating on data in vectors, matrices and higher-dimensional arrays. NumPy is the primary array programming library for the Python language. It has an essential role in research analysis pipelines in fields as diverse as physics, chemistry, astronomy, geoscience, biology, psychology, materials science, engineering, finance and economics. For example, in astronomy, NumPy was an important part of the software stack used in the discovery of gravitational waves1 and in the first imaging of a black hole2. Here we review how a few fundamental array concepts lead to a simple and powerful programming paradigm for organizing, exploring and analysing scientific data. NumPy is the foundation upon which the scientific Python ecosystem is constructed. It is so pervasive that several projects, targeting audiences with specialized needs, have developed their own NumPy-like interfaces and array objects. Owing to its central position in the ecosystem, NumPy increasingly acts as an interoperability layer between such array computation libraries and, together with its application programming interface (API), provides a flexible framework to support the next decade of scientific and industrial analysis.
Article
Full-text available
Pyroclastic density currents (PDCs) are dangerous multiphase flows originating from volcanic eruptions. PDCs cause more than a third of volcanic fatalities globally and, therefore, development of robust PDC hazard models is a priority in volcanology and natural hazard science. However, the complexity of gas–particle interactions inside PDCs, as well as their hostile nature, makes quantitative measurements of internal flow properties, and the validation of hazard models, challenging. Within the last decade, major advances from large-scale experiments, field observations and computational and theoretical models have provided new insights into the enigmatic internal structure of PDCs and identified key processes behind their fluid-like motion. Recent developments have also revealed important links between newly recognized processes of mesoscale turbulence and PDC behaviour. In this Review, we consider how recent advances in PDC research close the gaps towards more robust hazard modelling, outline the need to measure the internal properties of natural flows using geophysical methods and identify critical future research challenges. Greater understanding of PDCs will also provide insights into the dynamics of other natural gravity currents and high-energy turbulent multiphase flows, such as debris avalanches and turbidity currents.
Article
High-speed disperse multiphase flows are present in numerous environmental and engineering applications with complex interactions between turbulence, shock waves, and particles. Compared with its incompressible counterpart, compressible two-phase flows introduce new scales of motion that challenge simulations and experiments. This review focuses on gas–particle interactions spanning subsonic to supersonic flow conditions. An overview of existing Mach-number-dependent drag laws is presented, with origins from eighteenth-century cannon firings and new insights from particle-resolved numerical simulations. The equations of motion and phenomenology for a single particle are first reviewed. Multiparticle systems spanning dusty gases to dense suspensions are then discussed from numerical and experimental perspectives. Expected final online publication date for the Annual Review of Fluid Mechanics, Volume 56 is January 2024. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
This review is motivated by the fast progress in our understanding of the physics of particle-laden turbulence in the last decade, partly due to the tremendous advances of measurement and simulation capabilities. The focus is on spherical particles in homogeneous and canonical wall-bounded flows. The analysis of recent data indicates that conclusions drawn in zero gravity should not be extrapolated outside of this condition, and that the particle response time alone cannot completely define the dynamics of finite-size particles. Several breakthroughs have been reported, mostly separately, on the dynamics and turbulence modifications of small inertial particles in dilute conditions and of large weakly buoyant spheres. Measurements at higher concentrations, simulations fully resolving smaller particles, and theoretical tools accounting for both phases are needed to bridge this gap and allow for the exploration of the fluid dynamics of suspensions, from laminar rheology and granular media to particulate turbulence. Expected final online publication date for the Annual Review of Fluid Mechanics, Volume 54 is January 2022. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Numerical models of pyroclastic currents are widely used for fundamental research and for hazard and risk modeling that supports decision-making and crisis management. Because of their potential high impact, the credibility and adequacy of models and simulations needs to be assessed by means of an established, consensual validation process. To define a general validation framework for pyroclastic current models, we propose to follow a similar terminology and the same methodology that was put forward by Oberkampf and Trucano (Prog Aerosp Sci, 38, 2002) for the validation of computational fluid dynamics (CFD) codes designed to simulate complex engineering systems. In this framework, the term validation is distinguished from verification (i.e., the assessment of numerical solution quality), and it is used to indicate a continuous process, in which the credibility of a model with respect to its intended use(s) is progressively improved by comparisons with a suite of ad hoc experiments. The methodology is based on a hierarchical process of comparing computational solutions with experimental datasets at different levels of complexity, from unit problems (well-known, simple CFD problems), through benchmark cases (complex setups having well constrained initial and boundary conditions) and subsystems (decoupled processes at the full scale), up to the fully coupled natural system. Among validation tests, we also further distinguish between confirmation (comparison of model results with a single, well-constrained dataset) and benchmarking (inter-comparison among different models of complex experimental cases). The latter is of particular interest in volcanology, where different modeling approaches and approximations can be adopted to deal with the large epistemic uncertainty of the natural system.