Gas Mass Derived by Infrasound and UV Cameras - Implications For Mass Flow Rate

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Volcanology and Geothermal Research 325 (2016) 169–178

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

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

Gas mass derived by infrasound and UV cameras: Implications for mass


flow rate
D. Delle Donne a,b,⁎, M. Ripepe b, G. Lacanna b, G. Tamburello a, M. Bitetto a, A. Aiuppa a,c
a
DiSTeM, Università di Palermo, Palermo, Italy
b
Dipartimento di Scienze della Terra, Università di Firenze, Firenze, Italy
c
Istituto Nazionale di Geofisica e Vulcanologia, Palermo, Italy

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

Article history: Mass Flow Rate is one of the most crucial eruption source parameter used to define magnitude of eruption and to
Received 23 February 2016 quantify the ash dispersal in the atmosphere. However, this parameter is in general difficult to be derived and no
Received in revised form 20 June 2016 valid technique has been developed yet to measure it in real time with sufficient accuracy. Linear acoustics has
Accepted 20 June 2016
been applied to infrasonic pressure waves generated by explosive eruptions to indirectly estimate the gas
Available online 02 July 2016
mass erupted and then the mass flow rate. Here, we test on Stromboli volcano (Italy) the performance of such
Keywords:
methodology by comparing the acoustic derived results with independent gas mass estimates obtained with
Volcano acoustics UV cameras, and constraining the acoustic source by thermal imagery. We show that different acoustic methods
Sulphur dioxide camera give comparable total gas masses in the 2 to 1425 kg range, which are fully consistent with the gas masses
Thermal imagery derived by UV cameras and previous direct SO2 measurements.
Mass flow rate We show that total erupted gas mass, estimated by infrasound is not simply a function of the initial pressure, but
rather the full infrasonic waveform should be considered. Thermal imagery provides evidence that infrasound is
generated during the entire gas thrust phase. We provide examples to show how total gas masses derived by in-
frasonic signals can be affected by large uncertainties if duration of the signal is neglected. Only when duration of
infrasound is included, the best correlation (0.8) with UV cameras and the 1:1 direct linear proportionality is ob-
tained. Our results open new perspective for remotely derived gas mass and mass flow rates from acoustic
signals.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction the explosive source process could be transported far away from the
source with strong implications for monitoring volcanic activity from
The real-time assessment of the Mass Flow Rate (MFR) during an remote and safe locations.
eruption may have strong implications for the forecasting of the ash- The infrasonic signal associated with volcanic explosions are typical-
cloud spreading in the atmosphere, and therefore is essential to reduce ly characterized by an initial low frequency (b5 Hz) high amplitude
the impact on population and on the air-traffic infrastructure. However, sharp onset, followed by a longer lasting (N10 s) higher frequency
MFR is difficult to derive during an eruption in real time with sufficient (N 10 Hz) coda (Vergniolle and Brandeis, 1996; Ripepe and Marchetti,
accuracy (Mastin et al., 2009; Degruyter and Bonadonna, 2012). In this 2002). While the initial low frequency infrasonic wave is believed to
paper, we explore the reliability of using the infrasonic wave-field to be generated within the conduit by oscillation and/or by the bulging
provide in real time the MFR of small volcanic eruptions. of the magma-free surface (Vergniolle and Brandeis, 1996; Ripepe
According to the linear theory of sound, the infrasonic pressure field et al., 2001; Johnson, 2007; Gerst et al., 2008; Delle Donne & Ripepe,
can be directly related to the mass rate of gas and fine particles ejected 2012; Goto et al., 2014), the high frequency phase is explained as the
at the volcanic vent (see Johnson and Ripepe, 2011 for a review). From a turbulence of the gas leaving the conduit outside the vent and associat-
volcano monitoring point of view, this will have a strong potential be- ed with volcanic jet dynamics (Matoza et al., 2013; Taddeucci et al.,
cause acoustic waves can travel for long distances without any signifi- 2014; Goto et al., 2014). Several evidences indicate that both processes
cant signal distortion due to propagation and thus the information on can simultaneously be active during the same explosive process (Goto
et al., 2014; Genco et al., 2014). This makes the origin of infrasound
⁎ Corresponding author at: DiSTeM, Università di Palermo, Palermo, Italy. more complex than what previously thought and probably linked to
E-mail address: dario.delledonne@unipa.it (D. Delle Donne). multiple sources acting in different places and at different time. For

http://dx.doi.org/10.1016/j.jvolgeores.2016.06.015
0377-0273/© 2016 Elsevier B.V. All rights reserved.
170 D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178

larger eruptions, infrasonic jet noise radiation within the plume is sug-
gested to be predominant as the result of large scale turbulence mecha-
nisms (Matoza et al., 2013) while for small-medium size eruptions, such
as strombolian explosions, the jet noise mechanisms is likely to be less
significant.
Here we consider volcano infrasound as explained by the linear
acoustic theory and we look for the possible link with total erupted
gas mass by comparing infrasonic waveforms with thermal and SO2
measurements associated with 71 explosions recorded at Stromboli vol-
cano, Italy. The recent advent of UV cameras is paving the way to direct
observations and quantification of gas fluxes from individual gas
explosions (see Burton et al., 2015, for a review), and may thus provide
key constraints to interpret volcano infrasound. Although no clear
correlation has been found in previous attempts (McGonigle et al.,
2009; Pering et al., 2015) to compare gas masses derived from
infrasound with those derived from UV cameras, a similar approach re-
sulted into an overall good match at Pacaya volcano (Guatemala)
(Dalton et al., 2010). For each explosion, we also estimated duration
and thermal energy of the gas thrust phase from infrared thermometry
(Delle Donne & Ripepe, 2012). These parameters were then compared
to infrasonic waves, and different methods based on the acoustic theory
to derive gas mass (see Johnson, 2003; Woulff and McGetchin, 1976;
Delle Donne & Ripepe, 2012; Lane et al., 2013, among others) were
applied. We show that the gas masses estimated by acoustics are linear-
ly correlated with UV-based and/or thermal camera-based masses only Fig. 1. Map of Stromboli volcano indicating the position of active SW and NE craters, and
when the full wave-field is used to derive the velocity of the atmospher- the location of the experiment setup at Pizzo Sopra La Fossa (PZZ), and Roccette (ROC).
ic parcels displaced by the injection of gas and fragments mixture in the
atmosphere. 2007), we then converted SO2 in total gas mass involved in each
Strombolian explosion.
2. Instrument setup
3. Volcanic activity and infrasonic signals
The persistent, low-energy strombolian explosive activity makes
Stromboli volcano (Italy) the ideal site for calibrated test experiments, During data acquisition, explosive activity was relatively powerful,
where mass flow rates are also reported in previous work using being characterized by scoria-rich and ash-rich explosions from the
spectroscopic and imaging approaches (McGonigle et al., 2009; NE and SW crater, respectively, ejecting ballistics at ~200 m above the
Tamburello et al., 2012). Infrasound was recorded using an iTem prs- vents. Explosions from the NE crater generally consisted of highly ener-
0100a differential pressure transducer with a sensitivity of 25 mV/Pa getic blasts with large amounts of scoria, whereas explosions at the SW
and flat frequency response between 0.01 and 100 Hz at a full scale crater consisted of sustained emissions of ash, with relatively small
range of 250 Pa. amounts of scoria (Table 1).
Thermal images were recorded with a FLIR-A20 thermal camera The infrasonic waveforms associated with both craters show an ini-
fitted with a 34° × 25° optical lens (9.2 mm), and 0.1 °C thermal resolu- tial sharp positive onset with a first pulse lasting 0.1–0.5 s reaching am-
tion. The thermal sensor is an un-cooled micro-bolometer with focal plitude between 5 and 140 Pa, followed by a higher frequency coda with
plane array of 160 × 120 pixels, which is electronically oversampled at duration between 1 and 12 s (Fig. 2a). Acoustic waveforms were auto-
320 × 240 pixels, and sensitive in the 7.5–13 μm spectral range. Thermal matically extracted using the long term average LTA (30 s) and the
images and infrasound were synchronized using the same GPS clock short term average STA (1 s) amplitude ratio of 2 (Fig. 2b). For each ex-
with an accuracy of ~ 5 ms. The thermal camera was located at PZZ plosion we thus calculated the acoustic peak pressure and the duration
(Fig. 1) in direct line of sight to the NE and SW craters and at a slant of the signal (Table 1).
distance of 345 m and 327 m, respectively, and thus with field of
views of 218 × 153 m and 206 × 145 m. 4. Mass flux and infrasonic pressure
The SO2 flux was measured at ROC (Fig. 1) using a stand-alone UV
camera system, based on UV-absorption techniques (Tamburello et al., Many authors have suggested that infrasound can be used to calcu-
2012), and installed in June 2014 within the framework of ERC project late gas overpressure and/or to estimate the MFR associated with the
“BRIDGE”. The UV camera system is composed of two JAI CM-140GE- eruptive dynamics (Caplan-Auerbach et al., 2010; Ripepe et al., 2013)
UV cameras sensitive to UV-radiation, and one Ocean-Optics using an acoustic source model based on the linear theory of sound
USB2000+ UV spectrometer. These cameras employ a UV sensitive sen- (Lighthill, 1978; Woulff and McGetchin, 1976). For a point source, the
sor for covering UV wavelengths and are equipped with two different acoustic pressure p(t) in a half-space at a distance r is given by
band-pass optical filters with Full Width at Half Maximum (FWHM) of (Lighthill, 1978):
10 nm, and a central wavelength of 310 and 330 nm, respectively. Si-
multaneous image acquisition from these two cameras allows a direct 1
pðt Þ ¼ po ðt Þ−patm ¼ q_ ðt Þ ð1Þ
measure of UV absorption associated with SO2 concentration in the 2πr
plume. UV absorption measurements using UV cameras are calibrated
using an Ocean-Optics USB2000 + UV spectrometer (Lübcke et al., where q(t) is the outflow mass rate from a point source. When the
2012), that allows to get one absolute measurement of the UV spectrum magma surface deforms and expands, it produces an equivalent dis-
of the SO2 slant column densities averaged over a portion of the UV placement of the atmosphere, inducing a volumetric compression and
cameras field of view every ~ 5 s. Considering that SO2 makes up for generating an excess of pressure that mimics the rate of volumetric
Stromboli about 4.2 wt% of the total magmatic gas mass (Burton et al. change of the atmosphere displaced (e.g. Gerst et al., 2008).
D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178 171

Table 1
Infrasound amplitude and duration associated with the analyzed strombolian explosions, gas mass estimates using mass outflow (Method #1, par. 4.1), bulging of the magma free surface
(Method #2, par. 4.2), and gas flow velocity (Method #3, par. 4.3) methods, compared with thermal energy, thrust duration, and gas mass estimated using UV camera.

a) Experiment from PZZ site

Time GMT Infrasonic parameters Gas mass estimates from infrasound [kg] Thermal parameters

Peak pressure [pa] Duration [s] Method #1 Method #2 Method #3 Radiated energy [MJ] Thrust duration [s] Source vent

05-May-2008 16:38:22 136.82 5.48 867 1425 553 16.25 5.42 NE


05-May-2008 17:00:46 40.6 5.13 197 264 335 7.33 4.58 NE
05-May-2008 17:13:03 76.47 6.63 258 609 450 11.57 6.20 NE
05-May-2008 17:26:08 66.67 3.76 304 534 440 2.68 3.50 NE
05-May-2008 17:33:07 71.05 5.65 380 570 403 5.78 5.04 NE
07-May-2008 14:16:23 22.3 11.65 263 133 476 64.90 11.70 SW
07-May-2008 14:36:48 20.69 12.05 203 119 410 44.48 9.22 SW
07-May-2008 14:58:47 21.34 4.85 241 125 229 14.99 5.97 SW
07-May-2008 15:03:17 22.04 7.78 305 140 334 30.31 8.20 SW
07-May-2008 15:28:30 16.34 7.22 217 103 246 23.91 8.40 SW
07-May-2008 16:14:28 24.91 6.44 276 171 242 17.66 6.80 SW
07-May-2008 16:30:17 18.82 5.04 305 112 210 15.38 7.62 SW
07-May-2008 16:31:28 32.85 3.88 421 224 186 11.64 5.72 SW
07-May-2008 16:35:06 30.03 3.59 519 189 212 10.01 6.23 SW
07-May-2008 16:50:23 23.15 3.46 235 137 151 9.73 6.14 SW
07-May-2008 16:53:43 44.14 8.02 578 311 449 53.17 8.69 SW
08-May-2008 15:55:09 20.56 4.46 305 128 266 20.87 6.58 SW
08-May-2008 16:14:30 25.17 1.89 294 168 155 9.52 5.17 SW
08-May-2008 16:18:58 47.1 6.17 126 339 377 22.35 7.73 SW
08-May-2008 16:29:15 36.7 7.3 132 269 384 16.77 7.21 SW
08-May-2008 16:40:14 30.39 9.7 112 211 427 45.66 9.00 SW
08-May-2008 17:18:50 57.74 7.19 341 412 317 26.52 9.30 SW
08-May-2008 17:47:30 71.05 6.7 417 564 490 49.95 9.52 SW
08-May-2008 18:21:51 38.49 9.79 382 261 351 24.65 7.47 SW
08-May-2008 18:39:29 21.24 3.38 141 146 171 8.89 5.02 SW
09-May-2008 17:40:00 17.59 12.25 90 110 476 8.29 8.30 NE
09-May-2008 17:44:41 15.87 6.13 43 96 274 3.77 6.17 NE
09-May-2008 19:01:29 10.82 0.95 77 57 16 0.19 1.52 NE
09-May-2008 19:11:12 4.88 7.88 5 24 156 1.57 7.85 NE
09-May-2008 19:17:26 17.44 5.75 23 98 205 5.24 6.76 NE
09-May-2008 19:48:03 10.23 9.67 50 53 252 4.55 6.44 NE

b) Experiment from ROC site

Time GMT Infrasonic parameters Gas mass estimates from infrasound [kg] Gas mass estimates from UV cameras [kg]

Peak pressure [pa] Duration [s] Method #1 Method #2 Method #3 Gas mass [kg] Source vent

29-Jun.-2014 06:49:28 2.94 5.33 122 95 275 593 NE


29-Jun.-2014 06:51:14 3.14 7.07 156 99 218 643 NE
29-Jun.-2014 06:55:42 2.98 7.05 105 96 251 682 NE
29-Jun.-2014 06:58:10 12.97 2.65 232 177 108 378 NE
29-Jun.-2014 06:59:26 12.88 3.59 129 50 160 555 NE
29-Jun.-2014 07:08:58 10.00 2.73 96 142 85 441 NE
29-Jun.-2014 07:20:56 5.67 7.61 174 112 327 599 NE
29-Jun.-2014 07:34:29 3.52 3.41 151 113 150 541 NE
29-Jun.-2014 07:43:52 13.59 2.92 112 167 130 390 NE
29-Jun.-2014 07:52:55 3.30 3.97 100 97 125 524 NE
29-Jun.-2014 08:06:45 5.00 3.66 116 109 123 335 NE
29-Jun.-2014 08:14:41 3.81 3.82 155 100 186 338 NE
29-Jun.-2014 08:21:15 5.53 2.41 72 109 81 486 NE
29-Jun.-2014 08:22:56 3.72 10.09 193 100 369 605 NE
29-Jun.-2014 08:25:20 20.16 1.82 218 136 121 382 NE
29-Jun.-2014 08:29:28 3.87 3.24 116 96 95 358 NE
29-Jun.-2014 08:39:18 16.12 2.45 166 212 108 691 NE
29-Jun.-2014 08:45:03 3.92 5.41 422 100 326 678 NE
29-Jun.-2014 08:51:28 3.84 3.31 50 91 61 373 NE
29-Jun.-2014 08:52:01 10.91 2.61 156 139 153 427 NE
29-Jun.-2014 08:54:19 3.36 5.23 414 102 176 485 NE
29-Jun.-2014 08:58:37 3.85 9.97 185 105 334 849 NE
29-Jun.-2014 09:03:39 2.85 3.47 81 120 106 447 NE
29-Jun.-2014 09:06:13 7.55 6.54 184 124 90 632 NE
29-Jun.-2014 09:11:59 2.66 3.40 138 95 119 497 NE
29-Jun.-2014 09:13:48 4.09 1.61 211 95 104 339 NE
29-Jun.-2014 09:16:26 14.60 2.83 150 192 120 432 NE
29-Jun.-2014 09:17:05 3.21 9.43 344 93 345 910 NE
29-Jun.-2014 09:26:01 3.70 5.04 162 99 181 693 NE
29-Jun.-2014 09:27:25 4.14 2.63 215 100 69 339 NE
03-Aug.-2014 08:22:26 7.39 4.00 244 120 278 776 NE
03-Aug.-2014 08:30:29 3.24 2.55 215 94 104 478 NE
03-Aug.-2014 08:59:28 36.62 4.63 611 628 467 918 NE
03-Aug.-2014 09:11:44 8.42 0.98 64 128 39 368 NE

(continued on next page)


172 D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178

Table 1 (continued)

b) Experiment from ROC site

Time GMT Infrasonic parameters Gas mass estimates from infrasound [kg] Gas mass estimates from UV cameras [kg]

Peak pressure [pa] Duration [s] Method #1 Method #2 Method #3 Gas mass [kg] Source vent

03-Aug.-2014 09:15:13 11.10 5.18 504 162 259 698 NE


03-Aug.-2014 10:08:14 31.15 3.54 835 396 384 747 NE
03-Aug.-2014 10:24:23 16.82 3.94 406 304 293 601 NE
03-Aug.-2014 10:38:30 6.18 8.69 95 104 81 416 NE
03-Aug.-2014 11:04:32 52.80 2.49 855 999 327 594 NE
03-Aug.-2014 11:06:22 52.44 3.59 836 739 376 753 NE

4.1. Outflow mass rate properties of the medium, such as magma density of the gas/particles
cloud is required. The integration constants C1 and C2 can be derived
Assuming that the mass flux q(t) of the atmosphere displaced is by trial and error methods (Oshima and Maekawa, 2001), but are in
equivalent to the mass flux of the gas involved in the explosive process, general neglected for impulsive explosions with no prolongated emis-
the mass flux q(t) is calculated directly from Eq. (1) by integrating the sion of gas and/or material.
variation of the acoustic pressure recorded at a distance r from the This method has been successfully applied to calculate the gas mass
source (e.g. Firstov and Kravchenko, 1996; Johnson et al., 2004; content of strombolian activity at the Mount Erebus lava lake (Johnson,
Oshima and Maekawa, 2001): 2003; Johnson et al., 2004), Pacaya volcano (Dalton et al., 2010), and of
the violent vulcanian explosions at Sakura-jima volcano (Johnson &
Zt 
r Miller, 2014). We applied this method to the 71 explosions recorded
qðt Þ ¼ 2πr p t þ dt þ C 1 ð2Þ during the experiments. Each acoustic signal was thus integrated
c
0 twice following Eqs. (2) and (3) and the total gas mass is calculated as
the largest value of the gas mass distribution (Fig. 2c). We obtain that
being c the speed of sound in air and C1 the integration constant. Then, the total gas mass ranges between 5 and 866 kg (Table 1), which is
the mass flux is integrated a second time to calculate the gas mass m(t) very close to the range between 50 and 960 kg of the gas masses mea-
associated with the explosion (Fig. 2c): sured by UV cameras (Mori & Burton, 2009; Tamburello et al., 2012).

Zt
mðt Þ ¼ qðt Þdt þ C 2 ð3Þ 4.2. Bulging of the magma free surface
0
Assuming that volcanic explosions are generated by the bursting of
This method has the advantage to be extremely simple, since no a an over-pressurized gas slug, infrasonic pressure can be also related to
priori knowledge on the source (vent) dimensions and on the physical the bulging of the magma free-surface just before the bubble bursting.

Fig. 2. (a) Infrasonic signal band-pass (0.5–25 Hz) filtered and (b) acoustic power associated with the 5th May 2008, 16:38:22 explosion from NE crater (Table 1). Black lines show the full
scale records while gray lines show the same signal at a different amplitude scale to highlight the low amplitude 5.48 s lasting coda. Yellow area indicates the segment of infrasonic signal
automatically extracted using LTA/STA method. Infrasonic duration is defined by the time between markers 1 and 2, while peak pressure is defined as the maximum positive infrasonic
amplitude within the two markers. Converted gas mass using mass outflow method (par. 4.1) and using the gas flow velocity method (par. 4.3) are shown in (c) and (d) respectively. Gas
mass yielded using the bulging of the magma free surface method (par. 4.2) is achieved by using the peak-to-peak pressure of the signal. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)
D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178 173

^ will depend on the fluid-


In this case, infrasonic peak-to-peak pressure p 5. Comparing infrasound with thermal images
dynamic conditions of the magma controlling the rise and the expan-
sion of the gas slug in the conduit (Lane et al., 2013): Infrasound then seems able to provide gas mass consistent with in-
dependent estimates of the gas mass, regardless of the method used.
  However, the gas mass derived by direct integration of infrasound
^
p ^ 2
p
γ≈1:59 þ 21:2 þ 48:9 ð4Þ (par. 4.1) does not require assuming the diameter of the conduit nor
patm patm
the gas density and thus appears more efficient than the two other
methods (par. 4.2 and 4.3). To better understand these results we com-
where slug stability index γ is an adimensional dynamical property of
pared infrasound source to the explosive process by using the thermal
the rising slug (Del Bello et al., 2012) that depends on the slug volume
images collected by the camera deployed (Fig. 1).
(Vgas), the atmospheric pressure patm, and the proportion of the cross
Analysis of infrared thermal imagery has been widely used to study
section A' occupied by the falling liquid film (Llewellin et al., 2012):
explosive dynamics of strombolian activity and to constrain the source
process (Ripepe et al., 1993; Patrick, 2007; Delle Donne & Ripepe,
ρmagma gV gas A0
γ¼   ð5Þ 2012). The temperature derived from thermal imagery allows one to
πpatm R2v 1−A0
evaluate several parameters, including the mass of fragments involved
in each eruption (Harris et al., 2012, 2013), the style of activity, the
Thus, combining Eqs. (4) and (5), it is possible to derive the slug height reached by the ejecta, and the magma output rate (see Ripepe
volume Vgas from infrasonic peak-to-peak pressure ðp ^Þ by assuming et al., 1993; Patrick, 2007; Ripepe et al., 2008, among others). Here we
A' = 0.4814 (Lane et al., 2013), the vent radius Rv = 2.5 m (Delle use thermal radiance mainly to determine i) the thermal energy and
Donne & Ripepe, 2012), and the magma density ρmagma = ii) the duration of the gas thrust associated with each explosion.
2950 kgm−3 (Pioli et al., 2014).
Using Eq. (4), from the amplitude of the infrasound we can thus cal- 5.1. Thermal energy and gas/particles cloud volume
culate a slug stability index γ ranging between 1.9 and 45 which from
Eq. (5) gives slug volumes Vgas between 46 and 2740 m3. Considering The heat flux dissipated during strombolian explosions depends on
that most of the erupted gas is H2O and CO2 (Allard et al., 1994; temperature of the total ejected solid mass (see Harris et al., 2013 for
Aiuppa et al., 2010), we calculate a mean gas density of ~ 0.52 kg/m3, a review). The total radiated heat of strombolian explosions can be mea-
assuming an atmospheric pressure of 0.93 x 105 Pa and a temperature sured by the integrated temperature over a fixed FOV and by applying
between 100 and 352 °C (Harris and Ripepe, 2007a, b). We then the Stefan-Boltzmann law:
calculate gas masses ranging between 24 and 1425 kg (Table 1),
which are within the same range of previous method (par. 4.1), and in Zt h i
agreement with the independent gas mass measurements (Mori & T 4 ðt Þ−T 4BKG dt
ET ðt Þ ¼ εσAFOV ð9Þ
Burton, 2009; Tamburello et al., 2012).
0

4.3. Infrasound and gas flow velocity


where ET is the apparent radiated thermal energy (in Joules), ε is the ash
If we consider infrasound as produced by the gas flow in a medium emissivity (~0.96), s is the Stefan-Boltzmann constant, AFOV is the area
at rest, for a monopole source (Vergniolle & Brandeis, 1996; Delle of the field of view, T(t) is the integrated temperature within the FOV
Donne & Ripepe, 2012; Iemma et al., 2014; Sánchez et al., 2014), the at the time t, and TBKG is the background temperature in the FOV evalu-
acoustic power Π(t) can be related to the velocity u(t) of the flow ated just before the eruption onset. The thermal energy calculated for
(Woulff and McGetchin, 1976): the explosions during our experiment spans between 0.19 and 65 MJ
(Table 1).
2πρatm R2v u4 ðt Þ Because volcanic gases are characterized by very weak thermal
Π ðt Þ ¼ ð6Þ emissivity, the main contribution to the heat flux is given by the erupted
c
ash and volcanic bombs. Assuming a nearly constant magma tempera-
where ρatm is the density of the atmosphere (~ 1.08 kg/m3 at ture, the radiated thermal energy ET can be thus related to relative var-
~ 900 m a.s.l. and ambient temperature of 25 °C). The acoustic power iation of total mass of fragments erupted (Delle Donne & Ripepe, 2012;
is also related to acoustic pressure: Harris et al., 2013).

2πr 2 2 5.2. Gas thrust duration


Π ðt Þ ¼ p ðt Þ ð7Þ
ρatm c
A typical explosive event can be characterized by two main phases:
the early gas thrust phase, during which volcanic materials are pro-
where p is the acoustic pressure recorded at a distance r. Thus, combin-
pelled in the atmosphere, and the late buoyant phase, when the gas
ing Eq. (6) with Eq. (7) we can first calculate the velocity u(t) and then
still expands at constant velocity in the atmosphere by entraining air
we can calculate the gas mass flux assuming a vent radius Rv = 2.5 m
in the plume. In terms of linear theory of sound, acoustic pressure is
(Delle Donne & Ripepe, 2012) and a gas density ρgas of 0.52 kgm− 3
generated by changes in the mass outflow and thus more related to
(Fig. 2d):
the gas thrust phase rather than to the constant expansion rate of the
 1=2 buoyancy phase (Eq. (1)).
r The gas thrust phase is typically associated with a sharp increase of
qðt Þ ¼ πρgas R1:5
v jpðt Þj ð8Þ
ρatm the temperature within the camera field of view (FOV). The buoyant
phase is instead associated with the temperature waning, since no
Gas mass is the calculated by the integration over time of Eq. (8) and new volcanic material flows into the FOV but rather the plume is slowly
ranges between 15 and 553 kg (Table 1), again in the same order of expanding because of the temperature gradient (Marchetti et al., 2009).
magnitude of the previous methods and perfectly inline with the previ- Duration of the gas thrust can be thus defined as the time difference be-
ous gas measurements (Allard et al., 2008; Mori & Burton, 2009; tween the onset of the explosion and the maximum peak of the temper-
Tamburello et al., 2012). ature gradient (Marchetti et al., 2009). Gas thrust duration of the
174 D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178

explosions recorded during our experiment ranges between 1.5 and crater explosions are considered, and on the contrary, higher correla-
12 s. tions of R = 0.75 and R = 0.79 for the impulsive scoria-rich NE crater
explosions (Fig. 3d, e). Only the last method based on the integration
5.3. Infrasonic and thermal correlation over time of the source outflow velocity and thus also considering the
duration of the signal, shows high correlation coefficients of R = 0.79
We compared thermal energy and duration of the gas thrust phase and R = 0.89 for explosions of both the SW and NE craters (Fig. 3f).
with peak amplitude and duration of the infrasonic pressure signal to In conclusion, when explosive dynamics is controlled by sharp re-
verify the link between acoustic signal and explosive parameters lease of highly pressurized gas, duration of the gas thrust phase can be
(Fig. 3). No clear correlation pattern is found between thermal energy neglected, while, when explosions are sustained by the prolonged emis-
and acoustic peak pressure. Correlation is low (R = 0.3) for the SW cra- sion of gas with relatively low initial pressure, like the ash-rich at the
ter explosions, but higher (R = 0.78) for the NE crater explosions SW crater, duration of the infrasound becomes important and should
(Fig. 3a). In contrast, the correlation between thermal energy and dura- be considered. Since thermal energy is mainly controlled by quantity
tion of the acoustic signal is low (R = 0.25) for the NE crater, and high of solid mass (Harris et al., 2013), while infrasound depends mainly
(R = 0.8) for the SW crater. This suggests a distinct control of explosive on the gas, the presence of different linear regression for the two craters
styles on acoustic source. When the explosions are impulsive and can be interpreted as reflecting different gas/mass ratios, with explo-
scoria-rich, such as those generated by the NE crater, the initial peak sions at the NE crater having gas/mass ratios larger than explosions at
pressure pulse accounts for most of the acoustic energy and the dura- the SW crater (Fig. 3f).
tion of the infrasonic signal can be neglected. In this case the peak am-
plitude of the acoustic signal provides a good indication of the total
mass ejected. On the contrary, when explosions are ash-rich and 6. Gas mass from UV cameras
sustained like those one generated by the SW crater, the infrasonic en-
ergy of the low initial pressure pulse can be neglected and the duration The gas mass derived by infrasound is here compared to the direct
of the signal becomes important. This is confirmed by the very good cor- measurements of the SO2 mass derived by UV cameras, which provide
relation of R = 0.84 and R = 0.89 between the duration of the thermal SO2 column densities in two dimensions (Tamburello et al., 2012 for a
gas thrust and the duration of the acoustic signal (Fig. 3c) for both ex- review). Estimating the total SO2 mass produced by an explosion re-
plosions associated with the NE and SW craters, respectively, suggesting quires integration of the cumulative amount of SO2 over the image
a close link between the origin of the acoustic signal and the gas thrust FOV. This calculation does not require knowledge on the plume speed,
phase (Marchetti et al., 2009; Delle Donne and Ripepe, 2012). and thus overcomes the inherent difficulty in measuring the ascent
Considering the thermal energy as a proxy for the ejected mass rate of the explosive gas thrust. However, large amounts of ash scatter
(Delle Donne & Ripepe, 2012; Harris et al., 2013), we compared thermal UV light or even prevent light to pass through the plume, thus making
energies with the acoustic derived gas masses and we find that the first the method less efficient in measuring SO2 contents in ash-rich explo-
two methods (par. 4.1 and 4.2) show low correlation coefficients (of sions. We thus explore the correlation with the infrasound only for
R = 0.13 and R = 0.3 respectively) when the long lasting ash-rich SW the gas-rich (and ash-poor) explosions of the NE crater (Figs. 1 and 5).

Fig. 3. Comparison between thermal energy and duration with acoustic peak pressure and duration (a, b, c), and gas mass estimates (d, e, f) calculated for the explosions of the SW (circles)
and NE (squares) craters. Thermal energy associated with NE craters explosions are well correlated with peak pressure but not with acoustic duration, while thermal energy associated
with SW crater explosions show a good correlation only with acoustic duration and not with peak pressure. A good correlation at both craters is achieved only when comparing
thermal and acoustic durations. Gas mass estimates using three different infrasonic methods show different correlation with thermal energy. The best correlation of is achieved if gas
mass is calculated using the gas flow velocity method (par. 4.3).
D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178 175

The SO2 mass from each explosion was calculated always within the agreement with observations of a stationary flux component during
same integration window, chosen to contain the targeted SO2 cloud and the explosive events, typically showing a post-explosive coda of en-
part of the background. The peak in SO2 mass is typically reached in tens hanced passive degassing, associated with Strombolian explosions
of seconds after the blast, and reflects the expansion velocity of the gas (Tamburello et al., 2012; Pering et al., 2016).
in the atmosphere. By subtracting the SO2 contribution observed before
the explosion from the corresponding observed maximum value, we
obtained the total SO2 mass released during the explosion. 7. Conclusive remarks
According to previous measurements made with UV cameras
(Tamburello et al., 2012), SO2 mass ranges between 14 and 37 kg. Infrasound associated with explosive eruptions is generally pro-
This, assuming an SO2 concentration of 4.2 wt% (Burton et al., 2007), duced by the rapid expansion of the gas and therefore are linked to
converts into total magmatic gas mass between 335 and 918 kg the dynamics of its mass outflow. This links volcano acoustics to the
(Fig. 4, Table 1). magma fragmentation process and, in general, to any other phenome-
This range is fully consistent with the gas mass derived from non well coupled with the atmosphere (any source inducing a volume
acoustics (between 39 and 999 kg), but shows quite distinct correlation change in the atmosphere is a potential producer of infrasonic waves)
coefficients of R = 0.5, R = 0.33 and R = 0.8 when the different making infrasound a highly potential tool for the remote monitoring
methods described at par. 4.1, 4.2, and 4.3 are considered. Correlation of active volcanoes.
analysis is thus indicating that, in spite of the similar gas mass ranges, The linear theory of sound explains acoustic pressure as generated
the method considering the integration over time of the gas flow by the fluctuations of the mass outflow and it has been used to calculate
velocity (4.3) is the most reliable, and the calculated linear regression either the velocity of the gas and/or directly the mass of gas during ex-
fit has slope of 1.2: plosive eruptions (Woulff and McGetchin, 1976; Vergniolle and
Caplan-Auerbach, 2006; Caplan-Auerbach et al., 2010; Johnson et al.,
MUV ½kg ¼ 1:2  M ACU ½kg þ 324 ð10Þ 2004; Dalton et al., 2010; Lane et al., 2013; Ripepe et al., 2013; Lamb
et al., 2015). In spite of the uncertainty on the nature of the source, prop-
when a vent radius Rv of 2.5 m is considered, but has an offset of 324 kg, agation effects and geometry of the vent, in most cases the acoustic
which is indicating that the gas mass calculated by the acoustics is method has provided mass and MFR in harmony with other techniques
somehow underestimating the total gas mass. This misfit can be due (Dalton et al., 2010; Ripepe et al., 2013). However, the limits and the ef-
to uncertainty in one of the parameters used to convert acoustic ficiency of the acoustic method to retrieve eruption source parameters
pressure into gas flux using Eq. (8). have not been thoroughly tested.
Considering that gas mass is calculated from the infrasound as the We show that duration of the infrasound is highly correlated
integral of the gas flux which is a function of the gas velocity, a possible (R = 0.84–0.89) with the duration of the gas thrust phase detected
source of error may arise from our uncertainties in the gas density and by IR camera. This suggests that the infrasonic source is active during
in the vent radius Rv. However, changes of the vent radius Rv between the gas thrust phase of the explosive process and thus duration of the
0.5 and 5 m and of the gas density between 0.3 and 1 kg m−3 do not af- infrasonic signal is reflecting the duration of the gas thrust, and indi-
fect the offset but rather the slope of the linear fit (which varies between cating that the unsteady gas flux in the conduit is responsible for the
13 and 0.4, respectively). Thus, uncertainties on the vent radius and gas origin of the infrasound. Acoustic signals of comparable peak ampli-
density cannot be responsible for the underestimation of gas mass tude but different duration can result in significantly different mass
derived by the acoustic method, but only for the proportionality ratio if their duration is not considered. This is clear in the example of
between the two methods. In fact, considering a vent radius Rv = Fig. 6 showing infrasonic and thermal signals associated with two
2.8 m and a gas density of 0.6 kgm− 3 the 1:1 direct proportionality distinct explosions from the SW crater occurred at 14:16:23 and
ratio between UV (MUV) and acoustic (MACU) based gas masses is 16:30:17 GMT of 7th May 2008 (Table 1). Thermal images evidence
achieved, but the offset of 324 kg still remains unchanged. how duration of the infrasonic signals of 11.65 s and 5.04 s, respec-
A possible mechanism that could justify this offset is the gas released tively, are comparable to the durations of the gas thrust (11.7 s and
during a phase of stationary flux, or the increase of the passive 7.62 s), and how infrasonic amplitudes die out after the explosive
degassing triggered by the decompression of the magmatic column dur- cloud has reached the maximum height, marking the end of the gas
ing the explosive process (Genco and Ripepe, 2010). This is also in thrust phase (Fig. 6).

Fig. 4. Gas mass estimates using three acoustic methods: a) mass outflow (par. 4.1); b) bulging of the magma surface (par. 4.2), and c) gas flow velocity (par. 4.3), and compared with UV
camera gas mass measurements of explosions of the NE crater. Acoustic gas mass achieved using the gas flow velocity method (par. 4.3) shows the best correlation with UV camera
measurements (R = 0.8), with a slope of the linear fit equal to 1.2 for vent radius of 2.5 m. Correlation is reduced when gas mass from UV-camera is compared with the gas mass
derived using the other two acoustic methods.
176 D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178

Fig. 5. A typical strombolian explosion detected by (a) UV camera, (b) thermal images and (c) the associated infrasound (blue line) with the acoustic-derived based gas mass using gas flow
velocity method (black line). Explosive onset is marked by an infrasonic pulse that reaches the amplitude of ~10 Pa. UV images show the gas is feeding the plume during the gas thrust
phase, while during the waning phase of the thermal the plume is detached from the vent and no gas is released. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

In spite of the similarity of the acoustic onset, thermal imagery re- are considered for the two explosions (Table 1). Only the third method
veals that these two explosions with similar waveform and comparable (par. 4.3), based on the integration over time of gas flow velocity gives
initial peak pressure (22.3 Pa and 18.82 Pa respectively) can have differ- significantly different gas mass estimates of 476 kg and 210 kg, between
ent thermal energies of 64.9 MJ and 15.38 MJ, respectively, and the the two explosions, reflecting the different gas thrust duration and ther-
plume reaching different elevation above the vent (Fig. 6). The similar- mal energy (Table 1). We conclude that acoustic duration cannot be
ity of the acoustic signal brings to comparable gas mass estimates when neglected when gas mass and/or MFR is derived by infrasound, and
the first method (304 and 264 kg) and the second one (132 and 112 kg) that the method based on the integration over time of the gas flow

Fig. 6. Thermal and infrasonic signals of two explosions from the same crater (SW) occurred at 14:16:23 and 16:30:17 GMT on 7th May 2008 (Table 1). a) Infrasound associated with these
two explosions has almost the same peak amplitude but different duration of 11.65 s and 5.04 s, respectively. (b) The main amplitude first infrasonic onset shows comparable peak
pressure of 22.3 Pa and 18.82 Pa for both explosions and similar waveform, suggesting similar gas content. (c) Thermal images acquired at 2 s interval, show instead that the two
explosions are significantly different both for the duration of 11.7 s and 7.62 s, and also for the amount of volcanic ejecta released, which gives thermal energies of 64.9 MJ and
15.38 MJ, respectively, in line with the gas mass of 476 kg and 210 kg each.
D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178 177

velocity gives the most reliable results. In fact, we demonstrated, how in Goto, A., Ripepe, M., Lacanna, G., 2014. Wideband acoustic records of explosive volcanic
eruptions at Stromboli: new insights on the explosive process and the acoustic
this case the total gas mass shows a linear correlation 1:1 ratio with the source. Geophys. Res. Lett. 41 (2014), 3851–3857.
SO2 mass measured by UV camera for a vent radius of 2.8 m and gas Harris, A.J.L., Ripepe, M., 2007a. Synergy of multiple geophysical approaches to unravel
density of 0.6 kgm−3. This strongly supports the idea that, when the du- explosive eruption conduit and source dynamics. A case study for Stromboli. Chem.
Erde-Geochem. 2008.
ration is considered, acoustic waves can be used to obtain a remote and Harris, A.J.L., Ripepe, M., 2007b. Temperature and dynamics of degassing at Stromboli.
reliable estimate of the gas mass from the velocity, and the MFR during J. Geophys. Res. 112, B03205. http://dx.doi.org/10.1029/2006JB004393.
strombolian eruptions can be derived. Infrasound could then determine Harris, A.J.L., Ripepe, M., Hughes, E.E., 2012. Detailed analysis of particle launch velocities,
size distributions and gas densities during normal explosions at Stromboli.
a step toward improved and fast estimates of eruption source parame-
J. Volcanol. Geotherm. Res. 231-232, 109–131.
ters that can be integrated into the current real time numerical models Harris, A.J.L., Delle Donne, D., Dehn, J., 2013. Ripepe M, and Worden AK, 2013. Volcanic
used to estimate ash dispersal in the atmosphere (Degruyter and plume and bomb field masses from thermal infrared camera imagery. Earth Planet.
Sci. Lett. 365, 77–85.
Bonadonna, 2012; Woodhouse et al., 2013) with immediate impact on
Iemma, U., Di Paolo, F., Ripepe, M., 2014. Characterization of the gas–magmatic outflow at
improving our risk assessment. a volcanic vent through integral-equation based inverse acoustics. Appl. Acoust. 87,
Our experiment also provides a robust evidence of an acoustic 123–130.
source active during the full injection of gas and fine particles into the Johnson, J.B., 2003. Generation and propagation of infrasonic airwaves from volcanic ex-
plosions. J. Volcanol. Geotherm. Res. 121, 1–14.
atmosphere. This is consistent with high-speed camera image analysis Johnson, J., Aster, R., Kyle, P., 2004. Volcanic eruptions observed with infrasound. Geophys.
(Genco et al., 2014) which show how infrasonic wavefield is radiated Res. Lett. 31, L14604. http://dx.doi.org/10.1029/2004GL020020.
from the vent during the entire explosive process, indicating a clear Johnson, J.B., 2007. On the relation between infrasound, seismicity, and small pyroclastic
explosions at Karymsky volcano. J. Geophys. Res. 112, B08203.
link between mass discharged and infrasonic waves. Even if we cannot Johnson, J.B., Ripepe, M., 2011. Volcano infrasound: a review. J. Volcanol. Geotherm. Res.
exclude that turbulence within the plume could also generate acoustic 206 (3–4), 61–69.
radiation similar to the large-scale turbulence jet noise spectra (e.g. Johnson, J.B., Miller, A.J.C., 2014. Application of the monopole source to quantify ex-
plosive flux during vulcanian explosions at Sakurajima volcano (Japan). Seismol.
Matoza et al., 2013; Taddeucci et al. 2014), our results are supporting Res. Lett. 85, 6.
the idea that infrasound associated with strombolian explosions can Lamb, O., De Angelis, S., Lavalleè, Y., 2015. Using infrasound to constrain ash plume rise.
be mainly explained by linear acoustics conditions of a monopole source J. Appl. Volcanol. 4 (2015), 20.
Lane, S.J., James, M.R., Corder, S.B., 2013. Volcano infrasonic signals and magma degassing:
and can be directly linked to the velocity of the mass discharge rate. first-order experimental insights and application to Stromboli. Earth Planet. Sci. Lett.
377–378, 169–179.
Lighthill, M.J., 1978. Waves in Fluids. Cambridge Univ. Press, New York (504 pages.
Acknowledgements
Reissue).
Llewellin, E.W., Del Bello, E., Taddeucci, J., Scarlato, P., Lane, S.J., 2012. The thickness of the
The research leading to these results has received funding from the falling film of liquid around a Taylor bubble. Proc. R. Soc. Lond. Ser. A, Math. Phys. Sci.
European Community's Seventh Framework Program under Grant 468 (2140), 1041–1064.
Lübcke, P., Bobrowski, N., Illing, S., Kern, C., Alvarez Nieves, J.M., Vogel, L., Zielcke, J.,
Agreements No. 305377 (Project BRIDGE) and No. 308377 (Project Delgado Granados, H., Platt, U., 2012. On the absolute calibration of SO2 cameras.
FUTUREVOLC). Atmos. Meas. Tech. Discuss. 5 (5), 6183–6240.
Marchetti, E., Ripepe, M., Harris, A.J.L., Delle Donne, D., 2009. Tracing the differences be-
tween vulcanian and Strombolian explosions using infrasonic and thermal radiation
References energy. Earth Planet. Sci. Lett. 279, 273–281.
Mastin, L., Guffanti, M., Servanckx, R., Webley, P., Barostti, S., Dean, K., Denlinger, R.,
Aiuppa, A., Bertagnini, A., Métrich, N., Moretti, R., Di Muro, A., Liuzzo, M., Tamburello, G., Durant, A., Ewert, J., Gardner, C., Holliday, A., Neri, A., Rose, W., Schneider, D.,
2010. A model of degassing for Stromboli volcano. Earth Planet. Sci. Lett. 295 Siebert, L., Stunder, B., Swanson, G., Tupper, A., Volentik, A., Waythomas, A.,
(2010), 195–204. 2009. A multidisciplinary effort to assign realistic source parameters to models
Allard, P.A., Carbonelle, J., Métrich, N., Loyer, H., Zettwoog, P., 1994. Sulphur output and of volcanic ash-cloud transport and dispersion during eruptions. J. Volcanol.
magma degassing budget of Stromboli volcano. Nature 368, 326–330. Geotherm. Res. 186, 10–21 (special issue on Volcanic Ash Clouds; Mastin, L.,
Allard, P.A., Aiuppa, A., Burton, M., Caltabiano, T., Federico, C., Salerno, G., La Spina, A., Webley, P.W. (Eds.)).
2008. Crater gas emissions and the magma feeding system of Stromboli volcano. Matoza, R.S., Fee, D., Neilsen, T.B., Gee, K.L., Ogden, D.E., 2013. Aeroacoustics of volcanic
The Stromboli Volcano: An Integrated Study of the 2002–2003. Eruption Geophysical jets: acoustic power estimation and jet velocity dependence. J. Geophys. Res. 118
Monograph Series 182. (12).
Burton, M., Allard, P., Murè, F., La Spina, A., 2007. Magmatic gas composition reveals the McGonigle, A.J.S., Aiuppa, A., Ripepe, M., Kantzas, E.P., Tamburello, G., 2009. Spectroscopic
source depth of slug-driven Strombolian explosive activity. Science 317, 227–230. capture of 1 Hz volcanic SO2 fluxes and integration with volcano geophysical data.
Burton, M.R., Salerno, G., D'Auria, L., Caltabiano, T., Murè, F., Maugeri, R., 2015. SO2 flux Geophys. Res. Lett. 31, L21309.
monitoring at Stromboli with the new permanent INGV SO2 camera system: a com- Mori, T., Burton, M., 2009. Quantification of the gas mass emitted during single
parison with the FLAME network and seismological data. J. Volcanol. Geotherm. Res. explosions on Stromboli with the SO2 imaging camera. J. Volcanol. Geotherm.
300, 95–102. Res. 188, 395–400.
Caplan-Auerbach, J., Bellesiles, A., Fernandes, J.K., 2010. Estimates of eruption velocity and Oshima, H., Maekawa, T., 2001. Excitation process of infrasonic waves associated with
plume height from infrasonic recordings of the 2006 eruption of Augustine volcano, Merapi-type pyroclastic flow as revealed by a new recording system. Geophys. Res.
Alaska. J. Volcanol. Geotherm. Res. 189, 12–18. Lett. 28, 1099–1102.
Dalton, M.P., Waite, G.P., Watson, I.M., Nadeau, P.A., 2010. Multiparameter quantification Patrick, M.R., 2007. Dynamics of Strombolian ash plumes from thermal video: motion,
of gas release during weak Strombolian eruptions at Pacaya volcano, Guatemala. morphology, and air entrainment. J. Geophys. Res. 112, B06202. http://dx.doi.org/
Geophys. Res. Lett. 37 (9). 10.1029/2006JB004387.
Degruyter, W., Bonadonna, C., 2012. Improving on mass flow rate estimates of volcanic Pering, T.D., Tamburello, G., McGonigle, A.J.S., Aiuppa, A., James, M.R., Lane, S.J., Sciotto, M.,
eruptions. Geophys. Res. Lett. 39, L16308. http://dx.doi.org/10.1029/2012GL052566. Cannata, A., Patanè, D., 2015. Dynamics of mild strombolian activity on Mt. Etna.
Del Bello, E., Llewellin, E.W., Taddeucci, J., Scarlato, P., Lane, S.J., 2012. An analytical model J. Volcanol. Geotherm. Res. 300, 103–111.
for gas overpressure in slug-driven explosions: insights into Strombolian volcanic Pering, T.D., McGonigle, A.J.S., James, M.R., Tamburello, G., Aiuppa, A., Delle Donne,
eruptions. J. Geophys. Res. 117, B02206. http://dx.doi.org/10.1029/2011JB008747. D., Ripepe, M., 2016. Conduit dynamics and post explosion degassing on Strom-
Delle Donne, D., Ripepe, M., 2012. High frame rate thermal imagery of Strombolian explo- boli: a combined UV camera and numerical treatment. Geophys. Res. Lett. 43,
sions: implications for explosive and infrasonic source dynamics. J. Geophys. Res. 117, 5009–5016.
B9. Pioli, L., Pistolesi, M., Rosi, M., 2014. Transient explosions at open-vent volcanoes: the case
Firstov, P.P., Kravchenko, N.M., 1996. Estimation of the amount of explosive gas released of Stromboli (Italy). Geology 42 (10), 863–866. http://dx.doi.org/10.1130/G35844.1.
in volcanic eruptions using air waves. Volcanol. Seismol. 17, 547–560. Ripepe, M., Rossi, M., Saccorotti, G., 1993. Image processing of explosive activity at Strom-
Genco, R., Ripepe, M., 2010. Inflation‐deflation cycles revealed by tilt and seismic records boli. J. Volcanol. Geotherm. Res. 54, 335–351.
at Stromboli volcano. Geophys. Res. Lett. 37 (12). Ripepe, M., Ciliberto, S., Della Schiava, M., 2001. Time constraints for modeling source dy-
Genco, R., Ripepe, M., Marchetti, E., Bonadonna, C., Biass, S., 2014. Acoustic wavefield and namics of volcanic explosions at Stromboli. J. Geophys. Res. 106 (B5), 8713–8727.
Mach wave radiation of flashing arcs in strombolian explosion measured by image lu- Ripepe, M., Marchetti, E., 2002. Array tracking of infrasonic sources at Stromboli volcano.
minance. Geophys. Res. Lett. 41 (20). Geophys. Res. Lett. 29 (22), 2076. http://dx.doi.org/10.1029/2002GL015452.
Gerst, A., Hort, M., Kyle, P.R., Vodge, M., 2008. 4D velocity of Strombolian eruptions and Ripepe, M., Delle Donne, D., Harris, A.J.L., Marchetti, E., Ulivieri, G., 2008. Dynamics of
man-made explosions derived from multiple Doppler radar instruments. Strombolian Activity. In: Calvari, S., Inguaggiato, S., Puglisi, G., Ripepe, M., Rosi, M.
J. Volcanol. Geotherm. Res. 177 (3) (10 November 2008, Pages 648-660, ISSN 0377- (Eds.), The Stromboli Volcano: An Integrated Study of the 2002–2003 Eruption.
0273, http://dx.doi.org/10.1016/j.jvolgeores.2008.05.022). AGU Geophysical Monograph Series 182, pp. 39–48.
178 D. Delle Donne et al. / Journal of Volcanology and Geothermal Research 325 (2016) 169–178

Ripepe, M., Bonadonna, C., Folch, A., Delle Donne, D., Lacanna, G., Marchetti, E., Vergniolle, S., Brandeis, G., 1996. Strombolian explosions. 1. A large bubble breaking at the
Höskuldsson, A., 2013. Ash-plume dynamics and eruption source parameters by surface of a lava column as a source of sound. J. Geophys. Res. 101 (B9) (20,433-
infrasound and thermal imagery: the 2010 Eyjafjallajökull eruption. Earth Planet. 20,447).
Sci. Lett. 366, 112–121. Vergniolle, S., Caplan-Auerbach, J., 2006. Basaltic thermals and subplinian plumes: con-
Sánchez, C., Álvarez, B., Melo, F., Vidal, V., 2014. Experimental modeling of infrasound straints from acoustic measurements at Shishaldin volcano, Alaska. Bull. Volcanol.
emission from slug bursting on volcanoes. Geophys. Res. Lett. 41 (19). 68, 611–630.
Taddeucci, J., Sesterhenn, J., Scarlato, P., Stampka, K., Del Bello, E., Pena Fernandez, J.J., Gaudin, Woulff, G., McGetchin, T.R., 1976. Acoustic noise from volcanoes: theory and experiments.
D., 2014. High-speed imaging, acoustic features, and aeroacoustic computations of jet Geophys. J. R. Astron. Soc. 45, 601–616.
noise from Strombolian (and vulcanian) explosions. Geophys. Res. Lett. 41 (9). Woodhouse, M.J., Hogg, A.J., Phillips, J.C., Sparks, R.S.J., 2013. Interaction between volcanic
Tamburello, G., Aiuppa, A., Kantzas, E.P., McGonigle, A.J.S., Ripepe, M., 2012. Passive vs. ac- plumes and wind during the 2010 Eyjafjallajokull eruption, Iceland. J. Geophys. Res.
tive degassing modes at an open-vent volcano (Stromboli, Italy). Earth Planet. Sci. Solid Earth 118. http://dx.doi.org/10.1029/2012JB009592 (92–109).
Lett. 359–360, 106–116.

You might also like