Next Article in Journal
Recent Changes in Water Discharge in Snow and Glacier Melt-Dominated Rivers in the Tienshan Mountains, Central Asia
Next Article in Special Issue
Multi-Sensor Analysis of a Weak and Long-Lasting Volcanic Plume Emission
Previous Article in Journal
Urbanization-Driven Changes in Land-Climate Dynamics: A Case Study of Haihe River Basin, China
Previous Article in Special Issue
First Measurements of Gas Flux with a Low-Cost Smartphone Sensor-Based UV Camera on the Volcanoes of Northern Chile
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultraviolet Camera Measurements of Passive and Explosive (Strombolian) Sulphur Dioxide Emissions at Yasur Volcano, Vanuatu

1
Department of Geography, University of Sheffield, Winter Street, Sheffield S10 2TN, UK
2
Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EQ, UK
3
BP Institute, University of Cambridge, Madingley Rd, Cambridge CB3 E0EZ, UK
4
INGV, Sezione di Palermo, Via Ugo la Malfa 153, 90146 Palermo, Italy
5
DiSTeM, Università di Palermo, Via Archirafi, 22, 90123 Palermo, Italy
6
School of Geosciences, the University of Sydney, Sydney 2006, Australia
7
Faculty of Health, Engineering and Sciences, University of Southern Queensland, Toowoomba 4350, Australia
8
Geohazards Division, Vanuatu Meteorology and Geo-hazards Department, Lini Highway, Port Vila, Vanuatu
*
Author to whom correspondence should be addressed.
Remote Sens. 2020, 12(17), 2703; https://doi.org/10.3390/rs12172703
Submission received: 29 June 2020 / Revised: 14 August 2020 / Accepted: 17 August 2020 / Published: 20 August 2020
(This article belongs to the Special Issue Ground Based Imaging of Active Volcanic Phenomena)

Abstract

:
Here, we present the first ultraviolet (UV) camera measurements of sulphur dioxide (SO2) flux from Yasur volcano, Vanuatu, for the period 6–9 July 2018. These data yield the first direct gas-measurement-derived calculations of explosion gas masses at Yasur. Yasur typically exhibits persistent passive gas release interspersed with frequent Strombolian explosions. We used compact forms of the “PiCam” Raspberry Pi UV camera system powered through solar panels to collect images. Our daily median SO2 fluxes ranged from 4 to 5.1 kg s−1, with a measurement uncertainty of −12.2% to +14.7%, including errors from the gas cell calibration drift, uncertainties in plume direction and distance, and errors from the plume velocity. This work highlights the use of particle image velocimetry (PIV) for plume velocity determination, which was preferred over the typically used cross-correlation and optical flow methods because of the ability to function over a variety of plume conditions. We calculated SO2 masses for Strombolian explosions ranging 8–81 kg (mean of 32 kg), which to our knowledge is the first budget of explosive gas masses from this target. Through the use of a simple statistical measure using the moving minimum, we estimated that passive degassing is the dominant mode of gas emission at Yasur, supplying an average of ~69% of the total gas released. Our work further highlights the utility of UV camera measurements in volcanology, and particularly the benefit of the multiple camera approach in error characterisation. This work also adds to our inventory of gas-based data, which can be used to characterise the spectrum of Strombolian activity across the globe.

1. Introduction

Strombolian volcanism is one of the more common forms of basaltic explosive activity globally, associated with the rapid ejection of hot pyroclasts from a vent in a single impulsive burst [1,2], with event frequencies ranging from seconds to minutes [3]. Volcanoes with frequent Strombolian activity include the archetypal Stromboli, Italy [4,5]; Pacaya, Guatemala [6,7]; Erebus, Antarctica [8,9,10,11]; and Yasur, Vanuatu [12,13,14], the subject of this study. Other volcanoes also known to produce Strombolian activity include Etna, Italy [15,16,17]; Villarrica, Chile [18]; Arenal, Costa Rica [19,20]; Batu Tara, Indonesia [21,22]; and Shishaldin, USA [23].
Classically, this style of behaviour has been related to the ascent from depth of elongated and overpressured bubbles, termed gas slugs (Taylor bubbles), which rapidly expand in length as they approach the surface [2,24,25,26]. However, recent research has suggested that the causal mechanisms may be far more diverse [27,28,29,30] and that the presence of crystal-rich layers in the magmatic column is important in the mechanism of Strombolian explosions. To test these hypotheses, it is useful to investigate the spectrum of Strombolian activity at volcanoes where this behaviour is typical, including Yasur. In addition, recent studies have highlighted the importance of the eruption frequency in determining the behaviour of ascending gas slugs [21], as well as interslug interactions [15,31]. This has led to the classification of behaviour styles, ranging from rapidly bursting slugs, which may interact with one another during ascent, through to single bursting slugs [3,31].
There are several methods used to obtain information about individual Strombolian explosions based on the capture of seismic [5,32], infrasonic [6,33,34,35], thermal [4,5], and gas-derived [15,36,37,38] data. Here, we focus on gas emission rate measurements using an ultraviolet (UV) camera, an instrument frequently used to quantify gas release from persistently outgassing volcanoes [39,40]. The UV camera is able to resolve rapid fluctuations in the release of sulphur dioxide (SO2) gas. When the camera is used in tandem with an in situ multicomponent gas analyser (MultiGAS) to measure gas ratios within a volcanic plume [41,42], it is possible to estimate the total gas emission rate [43]. An important parameter with respect to causal mechanisms for Strombolian explosions is the ratio of gas released during explosions to that released passively [27,28,44,45,46,47]. This “active-to-passive” degassing ratio also provides information about conduit fluid dynamics [15,21,37,48]. For example, Tamburello et al. [38] discovered that the most efficient mode of degassing at Stromboli was actually the passive degassing, which accounted for ~77% of gases released, demonstrating the dominance of passive gas release [49] and the smaller gas bubbles within a volcanic conduit.
UV-camera-derived SO2 masses from Strombolian explosions [15,34,38,50] can be combined with gas ratio data (e.g., from MultiGAS) to generate total gas masses and volumes for individual explosive events [15,51]. These data provide parameters for analytical and computational models of gas flow in conduits, which yield further information about the activity and mechanisms, for example on the slug length, explosive vigour, and categorisation of the burst behaviour using fluid dynamics [3,24,26,52].

Yasur Volcano

Yasur (Vanuatu) is a basaltic stratovolcano located on the southeast of Tanna Island, which is thought to have been persistently active for at least ~800 years [53]. The main volcanic edifice is a cone with a crater area measuring 350–450 m in diameter, divided by a septum into northern and southern craters, each containing multiple active vents exhibiting Strombolian-style behaviour. A number of studies on Yasur have focused on the characteristics of its Strombolian activity, particularly its dynamism. Multivent basaltic volcanoes are known to exhibit vent-specific behaviours that can change through time (e.g., at Stromboli) [30,54]. Simons et al. [55] discussed systematic changes in behaviour at individual vents within the southern crater at Yasur, showing switching from bomb-rich (incandescent pyroclasts) through to ash-rich explosions. They also discussed conduit branching and the possibility of a single bubble (i.e., gas slug) driving paired explosions from separate vents at Yasur, with the potential for eruption styles to diverge at different vents due to cooling of the magma in the upper conduit branches. Spina et al. [56] observed two decoupled styles of degassing from infrasound data: puffing, which was near-constant; and Strombolian explosions. Meier et al. [57] highlighted the ash-rich and ash-poor (or bomb-rich) styles and their similarity at Yasur to those of Stromboli [4,58,59,60,61]. Kremers et al. [12] were able to calculate the lengths of gas slugs generating the Strombolian explosions on Yasur as ranging from 59 to 244 m, with mean and median values of 112 m and 103 m, respectively.
SO2 fluxes at Yasur ranged from 2.5 to 17.2 kg s−1 from April 2004 to November 2005, with a mean of 7.9 kg s−1 based on differential optical absorption spectroscopy (DOAS) traverses [13]. Between August 2007 and December 2008, SO2 fluxes at Yasur were 1.3 to 11.1 kg s−1, with a mean and median of 7.2 kg s−1 and 7.1 kg s−1, respectively [62]. In October 2007, a mean SO2 flux of 8.0 ± 3.8 kg s−1 across four days of traverses was reported [63]. A satellite-derived SO2 flux of 6.8 to 23.3 kg s−1 was estimated between 2000–2015, with a mean and median of 16.3 kg s−1 and 19.2 kg s−1, respectively [49]. Comparisons of gas flux between different periods of observations and between methods must be treated with caution; in discrete campaigns such as presented in this study, they may not represent broader changes through time.
Here, we demonstrate the use of a portable, solar-chargeable version of the low-cost Raspberry Pi ultraviolet camera [64,65] at Yasur, combined with a new approach to estimate plume velocity using UV camera imagery to obtain SO2 fluxes. We present the first UV camera measurements at Yasur, providing gas-based estimates of explosive Strombolian gas masses, which are key to unravelling information on the spectrum of behaviours for this style of activity globally. Furthermore, we illustrate the use of statistical methods to differentiate between passive and explosive gas release, and finally apply mathematical models to estimate the driving slug dimensions of the Strombolian explosions at Yasur volcano.

2. Observed Activity during 5th to 11th July 2018

During the measurement period, an ashy plume was present throughout the week related to ash-rich Strombolian explosions arising from both craters (Figure 1). From the summit, multiple vents displaying incandescence were visible within the southern crater, each exhibiting different styles of explosive behaviour (Figure 1c). Gas release from the summit vents was constantly visible, occasionally including “puffing” (described elsewhere by [3,21,38,48]). The northern crater contained at least two vents, but access to its rim was precluded by safety concerns due to ballistic ejecta from the crater’s Strombolian explosions, which also appeared to be more ash-rich than those from the southern crater. From the southern crater (Figure 1d) we directly observed explosions from at least three vents. Each vent exhibited different behaviours, two with jet-like characteristics (i.e., with a strong vertical component to the trajectory of the ejecta), hinting at the potential influence of the conduit wall during the explosion process, i.e., the explosion (slug burst) happens deeper within the conduit, providing a vertical direction to the released material [66,67]. Another vent exhibited parabolic transport of incandescent pyroclasts (without an initial jet), as though an ascending bubble burst within an over-topped magma column [24] or within a flared conduit geometry [68], i.e., allowing the lateral expansion of bubble prior to burst. Interestingly, these Strombolian explosions also differed in the noise generated, with the hemispherical-shaped (non-jet-like) explosions associated with a deeper booming sound. During 8th to 9th July, explosions were frequently associated with visible shockwaves propagating through the condensed plume. Throughout the measurement period, the morphology of the crater was dynamic, with spatter and ash accumulating around vents leading to changes in the size, shape, and position of vents.

3. UV Camera Methods

Low-cost Raspberry Pi ultraviolet (UV) camera systems (PiCams) were used to measure volcanic SO2 outgassing [64,65]. In this case the units were modified to include “PiJuice” hardware and software (https://github.com/PiSupply/PiJuice) to provide power to the Raspberry Pi boards at the heart of the camera system (see also [30]). The PiJuice units provide continuous supplies of power via lithium-polymer mobile phone batteries, which can be recharged using solar panels. In the field, both 1600 mAh and 2300 mAh batteries were used. With continuous solar charging (via 40 W solar panels for each Pi board), this configuration readily enabled field data acquisition for at least 6–7 h per day in this location. This camera setup omitted the GPS module used in the prior generation of the PiCam system, which automatically provided time synchronisation for the Raspberry Pi computers on start-up. Instead, GPS time synchronisation was performed manually via the command line, expedited by the PiJuices’ on-board real-time clocks. The PiCam systems were equipped with two Edmund Optics Inc. filters (full width at half maximum—10 nm), one for each lens, centred at around 310 and 330 nm, respectively, corresponding to spectral regions where SO2 does and does not absorb incident UV radiation. As detailed further elsewhere, UV imaging systems in volcanic gas monitoring are predicated upon contrasting image intensities in these two wavebands in order to isolate absorption in the image cause by sulphur dioxide absorption; for further details please see [39,69,70,71,72,73].
Two separate PiCams (Camera 1 and Camera 2) were operated simultaneously (enabling assessment of error and comparison of two plume angles), viewing the plume from a position southwest of the summit crater from the treehouse site (~1900 m from the plume) at the jungle oasis, on 6th and 7th July, and from the ash plain site (~2300 m from the plume) to the north-northwest on 8th and 9th July (see Figure 2 for locations). The UV cameras were also operated on 11th July, however inclement weather and grounding of the plume prevented reliable data processing for that day. During the measurement days, the plume direction varied from west to northwest, with dry and predominantly cloud-free weather (bar a brief period of rain on 9th July). Of the five days on which measurements were attempted, we acquired high-quality data on four of the days, amassing 16 h of imagery across these days.
The camera images were captured with acquisition rates of 0.5–0.25 Hz, with additional collection of clear sky images prior to the plume sequences’ capture, which are required in the processing routine to account for vignetting effects. Dark images were also acquired for each sequence to enable subtraction of dark noise. We conducted frequent calibrations using gas cells with known SO2 column densities (0, 412, and 1613 ppm m, with a manufacturer quoted error of 10%) between measurement sequences at least every 1–1.5 h, with more frequent calibration when light conditions changed more rapidly. The data were then processed following the commonly applied protocols, which have already been extensively described in the literature [39,70,74,75] (i.e., aligning images, selecting a clear sky background region and choosing a plume cross-section along which to determine integrated column amounts (ICA), before multiplying by the plume speed to calculate the flux). For the resulting flux data time series, we determined the data distribution statistically with the Kolmogorov–Smirnov normality test to inform on appropriate measures of central tendency. The data were all non-normally distributed, therefore the median was used in the further calculations, which are detailed below. However, we continued to detail both mean and median values.
A goal of this study was to attempt to differentiate degassing fluxes from each of the vents. However, it was not possible to do this rigorously and at all times, given that changes in wind shear and crater-derived eddying led to time-varying separation or overlap of individual plumes [76,77], creating difficulties in resolving emissions from the individual vents. Indeed, the plume predominantly appeared well-mixed on emergence from the summit crater (Figure 2c). However, at times, the view from the ash plain site did allow us to identify gas pulses from two distinct sources, likely corresponding to the two craters associated with explosions, where distinct gas pulses could be spatially resolved (Figure 2d).

3.1. Particle Image Velocimetry (PIV) for Plume Velocity Determination

One of the most significant and yet frequently overlooked errors in UV camera image analysis is that associated with plume velocity determination, for which three main methods are commonly used: cross-correlation [78,79], optical-flow [71,73,80,81,82,83], and manual tracking [84]. The optimal method will largely be determined by the plume conditions, as no single method is ideally suited to all situations. Manual tracking is suitable for stable plumes travelling at slow velocities or for measurements at greater distances from the plume, where cross-correlation and optical-flow are less desirable, as the plume is more dilute and fewer pixels containing SO2 are available for the analysis. Cross-correlation is preferred for broadly homogenous plumes that are well-mixed and undergo little turbulence (e.g., whereby eddying can cause SO2 within parts of the plume to travel backwards relative to the bulk plume vector of motion e.g., the wind direction). Optical flow methods are well suited to high velocity plumes, where the velocity field over the plume profile is non-constant (e.g., due to pulsed gas outputs from craters, associated with Strombolian explosions or puffing) [80,82,83,85].
In this study, we encountered difficulties in using these traditional methods. Indeed, our goal was to find one method to use on all separate acquisitions to maintain consistency. In particular, efficient mechanisms for tracking pulses of gas in a large dataset were required. Indeed, cross-correlation (which tracks the delay time between two integration lines of known distances) sometimes failed, likely because of turbulent motion in the plume altering arrival times in the peaks of the gas release. Furthermore, this approach does not cope well with the transient increases in gas velocity associated with impulsive gas release during Strombolian explosions. Hence, this method is probably the least favourable in this context. A lack of plume structure (increased homogeneity and plumes that filled more of the field of view) appeared to lead to the failure of the Farneback optical flow algorithm for some of the acquisitions [65,73,80]. We, therefore, instead adopted the use of particle image velocimetry (PIV) for plume velocity determination, as this successfully allowed plume velocity analysis across all separate acquisitions, maintaining consistency. PIV was briefly discussed by Kern et al. [74]. Previous use of PIV in a volcanic context has included tracking of lava lake velocity at Masaya [77], and it is similar to the pyroclast tracking velocimetry used in [59,86]. Here, we used PIVlab, a user-friendly MATLAB toolbox and app [87,88]. PIV works by comparing image pairs in sequences and looking for differences between them through two methods: direct cross-correlation and through the correlation of Fourier transforms. Both of these methods are conducted on integration areas (here we used three) with decreasing sizes on each pass. The end result is a velocity grid for the whole plume image, similar to those grids produced during the application of optical flow [73,80] (see Figure 3a).
We found that by using PIV we were able to detect velocity differences in even the more homogenous plumes (i.e., with a quasi-uniform SO2 distribution across most of the plume, except during Strombolian explosions). PIV was used to extract velocity components corresponding to each image pixel perpendicular to the integration line used in the ICA determination. Here, rather than using a single plume speed perpendicular to the integration line, the plume velocity vectors per pixel were multiplied by the pixel’s SO2 column amount, then these ICAs per-pixel were summed over the plume profile (see Figure 3). The PIV analyses show temporal and spatial variability in plume velocity, capturing a heterogeneity that is a real feature of the plume motion, yet which is not captured by cross-correlation or manual tracking. We report errors for the PIV analysis as the length of the integration line at given distance to the plume corresponding to each pixel divided by the lowest image capture frequency. For the ash plain site this equates to an error of 2 ± 0.3 ms−1 or ~±15%, and for the treehouse site this equates to an error of 5 ± 0.6 ms−1 or ~±9%. These error estimates are based on typical plume speeds for each site.

3.2. Estimation of a Total UV Camera Measurement Error

Here, we highlight the range of possible error sources and perform additional analyses on our data pertaining specifically to the calibration curve drift, plume orientation, and plume distance. The final determined values for error (in Table 1) are our best possible estimates on the basis of the available information and protocols applied in-the-field, which wherever possible were designed to minimise error. The final total of the combined errors was the root-mean-square (RMS) error of all the separate sources, as detailed below.
The effects of light dilution were quantified at a range of volcanic gas plume targets—SO2 mass column amounts may have been underestimated by ~10–60% over a range of distances (2.1 km to 6.5 km) and conditions, from hazy through to very clear [89]. Light dilution has a larger effect during hazier conditions, which were not present during our successful measurement days. Ilanko et al. [84] calculated that at ~10.3 km distance from the plume (during clear conditions at Sabancaya volcano, Peru), SO2 fluxes could be underestimated by 2.5 times, while at 4.25 km SO2 fluxes could be underestimated by 1.5 times (which would correspond to ~1.18 times (18%) at our maximum distance of 2300 m at Yasur). It is important to note that light dilution estimates are very specific to the measurement location and conditions. Given our range of distances to the plume and clear measurement conditions we, therefore, suggest that the error relating to light dilution was < +20%. We also note that the plume was not optically thick, except following ash-rich Strombolian explosions. Unfortunately, exact errors due to scattering of UV by ash were currently not quantifiable, but ash within the plume would likely lead to an underestimation of SO2 column amounts [38,90]. We attempted to minimise this error by integrating away from the summit area, where the plume was visibly less ash-rich and more transparent. We also further note that the peaks in gas flux from Strombolian explosions were well defined within the resulting dataset (Figure 3).
Gas cell calibrations change throughout the day in response to the position of the sun and changing illumination as a result of background clouds, with changes in gas cell calibrations potentially leading to over-estimation in SO2 column densities of up to 60% [91]. Figure 4 shows the changes in the calibration slope coefficient (between regressions of the apparent absorbance coefficient and column density) throughout the day from the time of the first calibration (rather than using UTC), showing variations ranging from 1.22 × 10−4 to 1.46 × 10−4 for this parameter. When taking into account this characterised range for the slope of 2.4 × 10−5 and the broad assumption (for indicative purposes) that there is a linear change between the first point and the highest point (corresponding to maximum solar zenith angle), we arrive over 122 min between these points at a 1.97 × 10−7 increase in the slope coefficient per minute. This would equate to a potential change in error of ~ 0.16% per minute, which expanded over an hour could become 9.6%, or for our maximum intercalibration interval of ~95 min, an error of 15.2%. It is possible, therefore, that any underlying trends in apparent gas emission rates below these thresholds are not differentiable from this error, i.e., an increase or decrease in flux at a rate of <~0.16% per minute. We suggest, therefore, that errors from cell calibration (notwithstanding the ~±10% manufacturer-quoted cell content error) amount to a maximum of ±15% for our measurement period.
We used fixed distances of 1900 and 2300 m from the camera to the plume for our retrieval calculations for data from the treehouse and ash plain sites, respectively. By repeating the same retrievals at different distances, for the ash plain data we determined that a 100 m error in the plume distance leads to a <5% difference in computed gas masses across the plume cross-section (with underestimation of this distance corresponding to underestimation of the gas mass) and a 200 m error in distance to <9% error. Comparisons of the same test dataset with different velocities in PIV analyses corresponding to the different distances showed variations ranging 1–7% with the 100 m error and 5–11% in the 200 m case. The combined effects of distance uncertainties on mass and velocity gave error ranges of 7–10% for fluxes at 100 m distance and 16–18% for 200 m. We, therefore, take the maximum value here of ~18% and apply this conservatively to our entire dataset.
Given the changes in plume direction, the orientation with which the integration line bisects the plume is also relevant in consideration of the measurement uncertainty [92]. To investigate this, we used overlaps between data from two synchronously acquiring cameras at the same site (Figure 5), which had slightly different plume views, and hence slightly different integration line orientations relative to the plume geometry. This simulated the time-dependent effect of the plume moving in response to changing wind conditions with respect to a fixed integrated column amount line. In this case, the two datasets were cross-correlated and shifted by the lag corresponding to the maximum correlation to account for different transport times from the source to the two cameras’ different integration lines. The calculated difference in flux retrieval from the two units based on comparing the acquired median values per unit was ~±5%. To calculate the RMS error, we took the RMS of all error sources, which led to a positive error (an underestimation) or a negative error (an overestimation) separately (see Table 1 for a summary). Our total estimated RMS errors were, therefore, −11.2%/+13.9% for the treehouse site and –12.2%/+14.7% for the ash plain site (i.e., it is more likely that our data were underestimated).
In addition, we also report on the computed flux data in Figure 6 (which documents the retrieved data from the entire campaign) during periods when the plume grounded, e.g., the integration line could not cover the entire plume cross-section, as well as episodes of heavy rainfall. During these periods, median SO2 fluxes were underestimated significantly by ~4.3–4.4 and 5.6–7.3 times, respectively, based on comparisons with median values of retrieved fluxes on either side of these episodes. Whilst the data captured under these circumstances were not used in the foregoing analysis, nor considered representative of the volcanic outgassing, they are reported here to illustrate the significant error to which these effects give rise.

4. Results and Discussion

4.1. SO2 Fluxes and Estimates of the Masses of Gas Emitted during Strombolian Explosions

Time series gas fluxes are shown in Figure 6, with a summary of daily statistics in Table 2. The median flux across the four days of measurements was 4.5 kg s−1 and the mean was 4.9 kg s−1, reflecting the peaks in SO2 flux associated with frequent Strombolian explosions. These gas fluxes correspond to a daily median and mean of 389 and 423 t d−1 across the measurement period. Daily statistics are given in Table 2. The median SO2 fluxes ranged between 4.0 and 5.1 kg s−1 across the measurement days, while the daily means ranged from 4.1 to 5.5 kg s−1. The time series data are suggestive of gradual changes in background SO2 emissions over several hours, but it is not clear whether these were real or a product of artefact error. A shift in activity is, however, plausible based on the observation of large Strombolian explosions with visible ballistics and shockwaves, particularly on 8th and 9th July, when lower fluxes were measured.
The masses of SO2 released during each Strombolian explosion were calculated by integrating beneath the explosive pulse and summing the total SO2 released, following Tamburello et al. [38] (see Figure 6c). However, a challenge here was that the onsets of Strombolian explosions were not visible within the imagery (i.e., vents were at depth within the crater). We used two methods to determine when explosions occurred within the UV camera imagery. Firstly, gas pulses in the camera images must be observed to originate and visibly accelerate above the rim of the summit crater (see Figure 1a) to confirm that an explosion occurred. Secondly, where gas burst traces in the flux time series were noted, showing the characteristic coda (a period of elevated flux following a Strombolian explosions, which gradually declines) detailed in Pering et al. [37] (see Figure 6c). The number of explosions was probably underestimated using these methods; however, the resulting estimation of the SO2 released during each explosion was useful for comparison with literature values (Table 3). Overall, the SO2 masses released were estimated for 135 explosions across five days. The mean masses of SO2 released increased from 6th to 9th July 2018, which is consistent with visual observations of more powerful explosions on 8th and 9th July 2018.
The range of SO2 masses released during Strombolian explosions at Yasur of 8–81 kg (mean 32 kg) is similar to the range estimated by Tamburello et al. [38] at Stromboli, who found a range of 2–55 kg (mean of 20 kg). However, these values are higher than those observed at Etna during mild Strombolian activity [15], which ranged from 0.1 to 14 kg. Gas ratios (SO2, H2S, H2O, and CO2) derived from a combined Fourier transform infrared spectroscopy (FTIR) and MultiGAS study from 6th to 16th July 2018 showed distinct gas compositions during passive and explosive activity [29]. Gases emitted during Strombolian activity had molar ratios of: CO2/SO2 = 2.85 ± 0.17; H2O/SO2 = 315 ± 71.8; SO2/HCl = 1.7 ± 0.22. Using these data, we can estimate total gas slug masses, as shown in Table 3. The mean total gas mass emitted during Strombolian explosions at Yasur was 2960 kg, with a range of 910–5940 kg. These estimates are similar to the range of 170–1674 kg for Strombolian explosions at Pacaya [6], whereas at Stromboli the explosion masses ranged from 44 to 238 kg according to Barnie et al. [93] and from 2 to 1425 kg as determined by Delle Donne et al. [34].

4.2. Simple Statistical Separation of Passive and Explosive Degassing

Others have studied the ratios of explosive-to-passive release during Strombolian explosions on Stromboli [38] and Etna [15]. Here, we attempt to expand on this by using a simple statistical measure involving the moving minimum (which traces the lower values in a dataset over a defined window, similarly to the moving mean) to estimate the passive release of gas through time, which when subtracted from total flux provides an approximate estimate of passive vs. explosive release. This was necessary, as our manual selection of events missed or excluded several Strombolian explosions. Our approach is similar to the automated method of Delle Donne et al. [82], which involved finding local peaks in time series data. For an example period (Figure 7), we highlight the moving minimum, which is set to a window size of 20 s, which is generally the characteristic timeframe of large peaks and troughs associated with Strombolian explosions [37,82]. Note that using this moving minimum method, an oscillation (non-uniform) background is apparent. Delle Donne et al. [82] also showed a fluctuation in the passive background between Strombolian explosions. This background is used as a best estimate solely to extract the explosive contribution. In this instance at Yasur, a moving minimum over a 20 s window proved best, given the higher frequency of explosive events; however, with a greater timeframe between events, the moving median may be a better measure. We also prefer this statistical estimation technique over using our estimated SO2 masses, given that the latter required manual selection of Strombolian explosions. This simple moving minimum approach could be readily and simply automated for routine monitoring of activity from Strombolian-explosion-producing volcanic systems.
Daily estimates of the passive and active degassing contributions are shown in Table 3, with a mean of 69% passive release compared to 31% explosive release. These estimates are similar to those estimated at Stromboli, showing 77% passive release compared to 23% explosive release (termed “active”, which also includes puffing); and at Etna, showing 67% passive release compared to 33% explosive release [15]. These datasets serve to illustrate the dominance of passive degassing in the gas emission budget at volcanoes that exhibit Strombolian activity. On the 8th July 2018, we calculated a higher passive degassing contribution of 78%. This day was characterised by higher SO2 masses emitted during individual explosions but lower overall SO2 fluxes. These features may be consistent with a degassing magma column beneath a thicker, more viscous, impermeable crystal-rich plug, requiring a higher gas mass to drive more powerful explosions [29,55,94], which is consistent with visual observations.

4.3. Models of Gas Slug Behaviour

Using our determined values for total slug mass, we can estimate slug lengths using the static pressure model of Del Bello et al. [24]. We use fixed values of 2600 kg m−3 and 1000 Pa s for density and viscosity, respectively, with an atmospheric pressure of 101,325 Pa. The only parameter we vary in the model is that of the conduit diameter, which we step up from 3 to 7 m. We use only the mean explosive gas ratios and masses (and not the range obtained when including error) for simplicity. It should be noted that the molar H2O/SO2 ratio is high and variable [29]. As water is the gas that contributes most to the mass of the slug, it is likely that our determined total gas masses are an overestimation. Our results are summarised in Table 4. We determine slug lengths ranging 188–609 m (median and mean of 347 and 366 m, respectively) for a conduit diameter of 3 m. However, this reduces to 76–260 m (median and mean of 146 and 154 m, respectively) for a conduit diameter of 7 m. Kremers et al. [12] calculated lower values of 59–244 m using seismoacoustic data. It would, therefore, seem that a larger conduit diameter may be more plausible at Yasur, which may bifurcate or split at very shallow depths [55].

5. Summary and Conclusions

In this work, we highlighted the utility of using low-cost solar-powered Raspberry Pi UV cameras for prolonged field campaigns. We continuously imaged the volcanic plume to yield both the velocity using a PIV method [87,88] and SO2 fluxes over periods of several hours per day, at temporal resolutions of up to 0.5 Hz, with brief pauses for calibration. SO2 fluxes were determined, with daily means of 4.1–5.5 kg s−1 (medians from 4.0–5.1 kg s−1), which are within the ranges of those measured previously at Yasur using ground-based methods, with values of 2.5 to 17.2 kg s−1 [13,62]. SO2 masses emitted during individual Strombolian explosions ranged from 8 to 81 kg, similarly to events at Stromboli, which were associated with emission values ranging 2–55 kg SO2 [38]. By using a simple statistical measure, we estimate that passive degassing at 69% is the dominant mode of degassing at Yasur, compared to 31% explosive serve. Our observations suggest that periods of lower gas output are associated with conduit sealing and more violent explosions, however a longer dataset would be needed to test this hypothesis substantively. By combining SO2 explosion masses with gas ratios [29], we determined total mean explosion gas masses of 910–5940 kg, which using the model of [24] correspond to slug lengths of 76–260 m if a larger conduit diameter of 7 m is used. Smaller conduit diameters lead to longer slug lengths of ~188–600 m at 3 m diameter, larger than those estimated previously at ~59–244 m [12]. The data presented here represent an important addition to our gas-data-based characterisation of the spectrum of Strombolian activity across the globe.

Author Contributions

Conceptualization, T.D.P.; Data curation, A.A.; Formal analysis, T.I. and R.D.; Investigation, T.I., J.W. and E.G.; Methodology, T.I., T.D.P., T.C.W., R.D., A.J.S.M. and M.E.; Resources, A.A.; Software, T.C.W.; Writing—original draft, T.D.P.; Writing—review & editing, T.D.P., A.A., A.J.S.M. and M.E. All authors have read and agreed to the published version of the manuscript.

Funding

J.W. and M.E. were supported by the Natural Environment Research Council (grant number NE/L002507/1); by the postgraduate travel funds received from Fitzwilliam College; by the Elspeth Matthews grant given by the Royal Geological Society; by the Mary Euphrasia Mosley, Sir Bartle Frere, and Worts travel fund report given by the University of Cambridge; and by the Exzellenzstipendium received by WKO. A.A. acknowledges funding support from the Alfred P. Sloan Foundation via the Deep Carbon Observatory (UniPa-CiW subcontract 10881-1262) and from MIUR (under grant n. PRIN2017-2017LMNLAW). T.D.P. acknowledges the support of the Royal Society (RG170226). T.I. is a Commonwealth Rutherford Fellow, funded by the UK government. A.McG. acknowledges support from the Rolex Institute.

Acknowledgments

We would like to thank the Vanuatu Meteorology and Geohazards Department for permission to conduct this fieldwork, Kelson and Joyce Hosea for their hospitality at the Jungle Oasis, and Roger for his assistance in the field

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wilkes, T.; McGonigle, A.; Pering, T.; Taggart, A.; White, B.; Bryant, R.; Willmott, J.; Wilkes, T.C.; McGonigle, A.J.S.; Pering, T.D.; et al. Ultraviolet Imaging with Low Cost Smartphone Sensors: Development and Application of a Raspberry Pi-Based UV Camera. Sensors 2016, 16, 1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wilkes, T.; Pering, T.; McGonigle, A.; Tamburello, G.; Willmott, J.; Wilkes, T.C.; Pering, T.D.; McGonigle, A.J.S.; Tamburello, G.; Willmott, J.R. A Low-Cost Smartphone Sensor-Based UV Camera for Volcanic SO2 Emission Measurements. Remote Sens. 2017, 9, 27. [Google Scholar] [CrossRef] [Green Version]
  3. Blackburn, E.A.; Wilson, L.; Sparks, R.S.J. Mechanisms and dynamics of strombolian activity. J. Geol. Soc. Lond. 1976, 132, 429–440. [Google Scholar] [CrossRef]
  4. Taddeucci, J.; Edmonds, M.; Houghton, B.; James, M.R.; Vergniolle, S. Hawaiian and Strombolian Eruptions. In The Encyclopedia of Volcanoes; Elsevier: Amsterdam, The Netherlands, 2015; pp. 485–503. [Google Scholar]
  5. Pering, T.D.; McGonigle, A.J.S. Combining Spherical-Cap and Taylor Bubble Fluid Dynamics with Plume Measurements to Characterize Basaltic Degassing. Geosciences 2018, 8, 42. [Google Scholar] [CrossRef] [Green Version]
  6. Patrick, M.R.; Harris, A.J.L.; Ripepe, M.; Dehn, J.; Rothery, D.A.; Calvari, S. Strombolian explosive styles and source conditions: Insights from thermal (FLIR) video. Bull. Volcanol. 2007, 69, 769–784. [Google Scholar] [CrossRef]
  7. Ripepe, M.; Harris, A.J.L.; Carniel, R. Thermal, seismic and infrasonic evidences of variable degassing rates at Stromboli volcano. J. Volcanol. Geotherm. Res. 2002, 118, 285–297. [Google Scholar] [CrossRef]
  8. Dalton, M.P.; Waite, G.P.; Watson, I.M.; Nadeau, P.A. Multiparameter quantification of gas release during weak Strombolian eruptions at Pacaya Volcano, Guatemala. Geophys. Res. Lett. 2010, 37, L09303. [Google Scholar] [CrossRef] [Green Version]
  9. Battaglia, A.; Bitetto, M.; Aiuppa, A.; Rizzo, A.L.; Chigna, G.; Watson, I.M.; D’Aleo, R.; Juárez Cacao, F.J.; de Moor, M.J. The Magmatic Gas Signature of Pacaya Volcano, With Implications for the Volcanic CO2 Flux From Guatemala. Geochem. Geophys. Geosyst. 2018, 19, 667–692. [Google Scholar] [CrossRef]
  10. Johnson, J.B.; Aster, R.C. Relative partitioning of acoustic and seismic energy during Strombolian eruptions. J. Volcanol. Geotherm. Res. 2005, 148, 334–354. [Google Scholar] [CrossRef]
  11. Oppenheimer, C.; Lomakina, A.S.; Kyle, P.R.; Kingsbury, N.G.; Boichu, M. Pulsatory magma supply to a phonolite lava lake. Earth Planet. Sci. Lett. 2009, 284, 392–398. [Google Scholar] [CrossRef]
  12. Sweeney, D.; Kyle, P.R.; Oppenheimer, C. Sulfur dioxide emissions and degassing behavior of Erebus volcano, Antarctica. J. Volcanol. Geotherm. Res. 2008, 177, 725–733. [Google Scholar] [CrossRef]
  13. Ilanko, T.; Oppenheimer, C.; Burgisser, A.; Kyle, P. Transient degassing events at the lava lake of Erebus volcano, Antarctica: Chemistry and mechanisms. GeoRes. J. 2015, 7, 43–58. [Google Scholar] [CrossRef] [Green Version]
  14. Kremers, S.; Wassermann, J.; Meier, K.; Pelties, C.; van Driel, M.; Vasseur, J.; Hort, M. Inverting the source mechanism of Strombolian explosions at Mt. Yasur, Vanuatu, using a multi-parameter dataset. J. Volcanol. Geotherm. Res. 2013, 262, 104–122. [Google Scholar] [CrossRef]
  15. Bani, P.; Lardy, M. Sulphur dioxide emission rates from Yasur volcano, Vanuatu archipelago. Geophys. Res. Lett. 2007, 34, L20309. [Google Scholar] [CrossRef]
  16. Oppenheimer, C.; Bani, P.; Calkins, J.A.; Burton, M.R.; Sawyer, G.M. Rapid FTIR sensing of volcanic gases released by Strombolian explosions at Yasur volcano, Vanuatu. Appl. Phys. B 2006, 85, 453–460. [Google Scholar] [CrossRef]
  17. Pering, T.D.; Tamburello, G.; McGonigle, A.J.S.; Aiuppa, A.; James, M.R.; Lane, S.J.; Sciotto, M.; Cannata, A.; Patanè, D. Dynamics of mild strombolian activity on Mt. Etna. J. Volcanol. Geotherm. Res. 2015, 300, 103–111. [Google Scholar] [CrossRef] [Green Version]
  18. Aiuppa, A.; Coco, E.L.; Liuzzo, M.; Giudice, G. Terminal Strombolian activity at Etna’s central craters during summer 2012: The most CO2-rich volcanic gas ever recorded at Mount Etna. Geochem. J. 2016, 50, 123–138. [Google Scholar] [CrossRef]
  19. Branca, S.; del Carlo, P. Types of eruptions of Etna volcano AD 1670–2003: Implications for short-term eruptive behaviour. Bull. Volcanol. 2005, 67, 732–742. [Google Scholar] [CrossRef]
  20. Shinohara, H.; Witter, J.B. Volcanic gases emitted during mild Strombolian activity of Villarrica volcano, Chile. Geophys. Res. Lett. 2005, 32, L20308. [Google Scholar] [CrossRef]
  21. Garcés, M.A.; Hagerty, M.T.; Schwartz, S.Y. Magma acoustics and time-varying melt properties at Arenal Volcano, Costa Rica. Geophys. Res. Lett. 1998, 25, 2293–2296. [Google Scholar] [CrossRef]
  22. Szramek, L.; Gardner, J.E.; Larsen, J. Degassing and microlite crystallization of basaltic andesite magma erupting at Arenal Volcano, Costa Rica. J. Volcanol. Geotherm. Res. 2006, 157, 182–201. [Google Scholar] [CrossRef]
  23. Gaudin, D.; Taddeucci, J.; Scarlato, P.; del Bello, E.; Ricci, T.; Orr, T.; Houghton, B.; Harris, A.; Rao, S.; Bucci, A. Integrating puffing and explosions in a general scheme for Strombolian-style activity. J. Geophys. Res. Solid Earth 2017, 122, 1860–1875. [Google Scholar] [CrossRef]
  24. Laiolo, M.; Massimetti, F.; Cigolini, C.; Ripepe, M.; Coppola, D. Long-term eruptive trends from space-based thermal and SO2 emissions: A comparative analysis of Stromboli, Batu Tara and Tinakula volcanoes. Bull. Volcanol. 2018, 80, 1–19. [Google Scholar] [CrossRef]
  25. Vergniolle, S.; Boichu, M.; Caplan-Auerbach, J. Acoustic measurements of the 1999 basaltic eruption of Shishaldin volcano, Alaska 1. Origin of Strombolian activity. J. Volcanol. Geotherm. Res. 2004, 137, 109–134. [Google Scholar] [CrossRef]
  26. Del Bello, E.; Llewellin, E.W.; Taddeucci, J.; Scarlato, P.; Lane, S.J. An analytical model for gas overpressure in slug-driven explosions: Insights into Strombolian volcanic eruptions. J. Geophys. Res. Solid Earth 2012, 117, B02206. [Google Scholar] [CrossRef]
  27. Seyfried, R.; Freundt, A. Experiments on conduit flow and eruption behavior of basaltic volcanic eruptions. J. Geophys. Res. 2000, 105, 23727. [Google Scholar] [CrossRef]
  28. James, M.R.; Lane, S.J.; Corder, S.B. Modelling the rapid near-surface expansion of gas slugs in low-viscosity magmas. Geol. Soc. Lond. Spec. Publ. 2008, 307, 147–167. [Google Scholar] [CrossRef]
  29. Barth, A.; Edmonds, M.; Woods, A. Valve-like dynamics of gas flow through a packed crystal mush and cyclic strombolian explosions. Sci. Rep. 2019, 9, 821. [Google Scholar] [CrossRef] [Green Version]
  30. Suckale, J.; Keller, T.; Cashman, K.V.; Persson, P.O. Flow-to-fracture transition in a volcanic mush plug may govern normal eruptions at Stromboli. Geophys. Res. Lett. 2016, 43, 12071–12081. [Google Scholar] [CrossRef] [Green Version]
  31. Woitischek, J.; Woods, A.W.; Edmonds, M.; Oppenheimer, C.; Aiuppa, A.; Pering, T.D.; Ilanko, T.; D’Aleo, R.; Garaebiti, E. Strombolian eruptions and dynamics of magma degassing at Yasur Volcano (Vanuatu). J. Volcanol. Geotherm. Res. 2020, 398, 106869. [Google Scholar] [CrossRef]
  32. Pering, T.D.; Liu, E.J.; Wood, K.; Wilkes, T.C.; Aiuppa, A.; Tamburello, G.; Bitetto, M.; Richardson, T.; McGonigle, A.J.S. Combined ground and aerial measurements resolve vent-specific gas fluxes from a multi-vent volcano. Nat. Commun. 2020, 11, 3039. [Google Scholar] [CrossRef] [PubMed]
  33. Pering, T.D.; McGonigle, A.J.S.; James, M.R.; Capponi, A.; Lane, S.J.; Tamburello, G.; Aiuppa, A. The dynamics of slug trains in volcanic conduits: Evidence for expansion driven slug coalescence. J. Volcanol. Geotherm. Res. 2017, 348, 26–35. [Google Scholar] [CrossRef]
  34. Chouet, B.; Dawson, P.; Ohminato, T.; Martini, M.; Saccorotti, G.; Giudicepietro, F.; de Luca, G.; Milana, G.; Scarpa, R. Source mechanisms of explosions at Stromboli Volcano, Italy, determined from moment-tensor inversions of very-long-period data. J. Geophys. Res. Solid Earth 2003, 108, 2019. [Google Scholar] [CrossRef]
  35. Marchetti, E.; Ripepe, M.; Harris, A.J.L.; Delle Donne, D. Tracing the differences between Vulcanian and Strombolian explosions using infrasonic and thermal radiation energy. Earth Planet. Sci. Lett. 2009, 279, 273–281. [Google Scholar] [CrossRef]
  36. Delle Donne, D.; Ripepe, M.; Lacanna, G.; Tamburello, G.; Bitetto, M.; Aiuppa, A. Gas mass derived by infrasound and UV cameras: Implications for mass flow rate. J. Volcanol. Geotherm. Res. 2016, 325, 169–178. [Google Scholar] [CrossRef]
  37. Johnson, J.B.; Ripepe, M. Volcano infrasound: A review. J. Volcanol. Geotherm. Res. 2011, 206, 61–69. [Google Scholar] [CrossRef]
  38. McGonigle, A.J.S.; Aiuppa, A.; Ripepe, M.; Kantzas, E.P.; Tamburello, G. Spectroscopic capture of 1 Hz volcanic SO2 fluxes and integration with volcano geophysical data. Geophys. Res. Lett. 2009, 36, 1–5. [Google Scholar] [CrossRef]
  39. Pering, T.D.; McGonigle, A.J.S.; James, M.R.; Tamburello, G.; Aiuppa, A.; Delle Donne, D.; Ripepe, M. Conduit dynamics and post explosion degassing on Stromboli: A combined UV camera and numerical modeling treatment. Geophys. Res. Lett. 2016, 43, 5009–5016. [Google Scholar] [CrossRef] [Green Version]
  40. Tamburello, G.; Aiuppa, A.; Kantzas, E.P.; Mcgonigle, A.J.S.; Ripepe, M. Passive vs. active degassing modes at an open-vent volcano (Stromboli, Italy). Earth Planet. Sci. Lett. 2012, 359–360, 106–116. [Google Scholar] [CrossRef] [Green Version]
  41. McGonigle, A.J.S.; Pering, T.D.; Wilkes, T.C.; Tamburello, G.; D’Aleo, R.; Bitetto, M.; Aiuppa, A.; Willmott, J.R. Ultraviolet Imaging of Volcanic Plumes: A New Paradigm in Volcanology. Geosciences 2017, 7, 68. [Google Scholar] [CrossRef] [Green Version]
  42. Pering, T.D.; Ilanko, T.; Liu, E.J. Periodicity in Volcanic Gas Plumes: A Review and Analysis. Geosciences 2019, 9, 394. [Google Scholar] [CrossRef] [Green Version]
  43. Shinohara, H.; Ohminato, T.; Takeo, M.; Tsuji, H. Monitoring of volcanic gas composition at Asama volcano, Japan, during 2004–2014. J. Volcanol. 2015, 303, 199–208. [Google Scholar] [CrossRef]
  44. Aiuppa, A.; Federico, C.; Giudice, G.; Gurrieri, S. Chemical mapping of a fumarolic field: La Fossa Crater, Vulcano Island (Aeolian Islands, Italy). Geophys. Res. Lett. 2005, 32, L13309. [Google Scholar] [CrossRef] [Green Version]
  45. Pering, T.D.; Tamburello, G.; McGonigle, A.J.S.; Aiuppa, A.; Cannata, A.; Giudice, G.; Patanè, D. High time resolution fluctuations in volcanic carbon dioxide degassing from Mount Etna. J. Volcanol. Geotherm. Res. 2014, 270, 115–121. [Google Scholar] [CrossRef]
  46. Jaupart, C.; Vergniolle, S. Laboratory models of Hawaiian and Strombolian eruptions. Nature 1988, 331, 58–60. [Google Scholar] [CrossRef]
  47. Jaupart, C.; Vergniolle, S. The generation and collapse of a foam layer at the roof of a basaltic magma chamber. J. Fluid Mech. 1989, 203, 347–380. [Google Scholar] [CrossRef]
  48. Parfitt, E.A. A discussion of the mechanisms of explosive basaltic eruptions. J. Volcanol. Geotherm. Res. 2004, 134, 77–107. [Google Scholar] [CrossRef]
  49. Vergniolle, S.; Jaupart, C. Separated two-phase flow and basaltic eruptions. J. Geophys. Res. Solid Earth 1986, 91, 12842–12860. [Google Scholar] [CrossRef]
  50. Gaudin, D.; Taddeucci, J.; Scarlato, P.; Harris, A.; Bombrun, M.; del Bello, E.; Ricci, T. Characteristics of puffing activity revealed by ground-based, thermal infrared imaging: The example of Stromboli Volcano (Italy). Bull. Volcanol. 2017, 79, 24. [Google Scholar] [CrossRef]
  51. Carn, S.A.; Fioletov, V.E.; Mclinden, C.A.; Li, C.; Krotkov, N.A. A decade of global volcanic SO2 emissions measured from space. Sci. Rep. 2017, 7, 1–12. [Google Scholar] [CrossRef] [Green Version]
  52. Mori, T.; Burton, M. Quantification of the gas mass emitted during single explosions on Stromboli with the SO2 imaging camera. J. Volcanol. Geotherm. Res. 2009, 188, 395–400. [Google Scholar] [CrossRef]
  53. Burton, M.; Allard, P.; Mure, F.; La Spina, A. Magmatic Gas Composition Reveals the Source Depth of Slug-Driven Strombolian Explosive Activity. Science 2007, 317, 227–230. [Google Scholar] [CrossRef] [Green Version]
  54. James, M.R.; Lane, S.J.; Wilson, L.; Corder, S.B. Degassing at low magma-viscosity volcanoes: Quantifying the transition between passive bubble-burst and Strombolian eruption. J. Volcanol. Geotherm. Res. 2009, 180, 81–88. [Google Scholar] [CrossRef]
  55. Firth, C.W.; Handley, H.K.; Cronin, S.J.; Turner, S.P. The eruptive history and chemical stratigraphy of a post-caldera, steady-state volcano: Yasur, Vanuatu. Bull. Volcanol. 2014, 76, 1–23. [Google Scholar] [CrossRef]
  56. Salvatore, V.; Silleni, A.; Corneli, D.; Taddeucci, J.; Palladino, D.M.; Sottili, G.; Bernini, D.; Andronico, D.; Cristaldi, A. Parameterizing multi-vent activity at Stromboli Volcano (Aeolian Islands, Italy). Bull. Volcanol. 2018, 80, 64. [Google Scholar] [CrossRef]
  57. Simons, B.C.; Jolly, A.D.; Eccles, J.D.; Cronin, S.J. Spatiotemporal Relationships between Two Closely-spaced Strombolian-style Vents, Yasur, Vanuatu. Geophys. Res. Lett. 2020, 47, e85687. [Google Scholar] [CrossRef]
  58. Spina, L.; Taddeucci, J.; Cannata, A.; Gresta, S.; Lodato, L.; Privitera, E.; Scarlato, P.; Gaeta, M.; Gaudin, D.; Palladino, D.M. Explosive volcanic activity at Mt. Yasur: A characterization of the acoustic events (9–12th July 2011). J. Volcanol. Geotherm. Res. 2016, 322, 175–183. [Google Scholar] [CrossRef]
  59. Meier, K.; Hort, M.; Wassermann, J.; Garaebiti, E. Strombolian surface activity regimes at Yasur volcano, Vanuatu, as observed by Doppler radar, infrared camera and infrasound. J. Volcanol. Geotherm. Res. 2016, 322, 184–195. [Google Scholar] [CrossRef]
  60. Taddeucci, J.; Scarlato, P.; Capponi, A.; del Bello, E.; Cimarelli, C.; Palladino, D.M.; Kueppers, U. High-speed imaging of Strombolian explosions: The ejection velocity of pyroclasts. Geophys. Res. Lett. 2012, 39, 1–6. [Google Scholar] [CrossRef] [Green Version]
  61. Gaudin, D.; Taddeucci, J.; Scarlato, P.; Moroni, M.; Freda, C.; Gaeta, M.; Palladino, D.M. Pyroclast Tracking Velocimetry illuminates bomb ejection and explosion dynamics at Stromboli (Italy) and Yasur (Vanuatu) volcanoes. J. Geophys. Res. Solid Earth 2014, 119, 5384–5397. [Google Scholar] [CrossRef]
  62. Ripepe, M.; Harris, A.J.L.; Marchetti, E. Coupled thermal oscillations in explosive activity at different craters of Stromboli volcano. Geophys. Res. Lett. 2005, 32, 1–4. [Google Scholar] [CrossRef]
  63. Ripepe, M.; Marchetti, E. Array tracking of infrasonic sources at Stromboli volcano. Geophys. Res. Lett. 2002, 29, 33.1–33.4. [Google Scholar] [CrossRef]
  64. Bani, P.; Oppenheimer, C.; Allard, P.; Shinohara, H.; Tsanev, V.; Carn, S.; Lardy, M.; Garaebiti, E. First estimate of volcanic SO 2 budget for Vanuatu island arc. J. Volcanol. Geotherm. Res. 2012, 211–212, 36–46. [Google Scholar] [CrossRef] [Green Version]
  65. Métrich, N.; Allard, P.; Aiuppa, A.; Bani, P.; Bertagnini, A.; Shinohara, H.; Parello, F.; Di Muro, A.; Garaebiti, E.; Belhadj, O.; et al. Magma and Volatile Supply to Post-collapse Volcanism and Block Resurgence in Siwi Caldera (Tanna Island, Vanuatu Arc). J. Pet. 2011, 52, 1077–1105. [Google Scholar] [CrossRef] [Green Version]
  66. Delle Donne, D.; Ripepe, M. High-frame rate thermal imagery of strombolian explosions: Implications for explosive and infrasonic source dynamics. J. Geophys. Res. Solid Earth 2012, 117, 1–12. [Google Scholar] [CrossRef]
  67. Salvatore, V.; Cigala, V.; Taddeucci, J.; Arciniega-Ceballos, A.; Peña Fernández, J.J.; Alatorre-Ibargüengoitia, M.A.; Gaudin, D.; Palladino, D.M.; Kueppers, U.; Scarlato, P. Gas-Pyroclast Motions in Volcanic Conduits During Strombolian Eruptions, in Light of Shock Tube Experiments. J. Geophys. Res. Solid Earth 2020, 125, e19182. [Google Scholar] [CrossRef]
  68. Dibble, R.R.; Kyle, P.R.; Rowe, C.A. Video and seismic observations of Strombolian eruptions at Erebus volcano, Antarctica. J. Volcanol. Geotherm. Res. 2008, 177, 619–634. [Google Scholar] [CrossRef]
  69. Mori, T.; Burton, M. The SO 2 camera: A simple, fast and cheap method for ground-based imaging of SO 2 in volcanic plumes. Geophys. Res. Lett. 2006, 33, L24804. [Google Scholar] [CrossRef]
  70. Kantzas, E.P.; McGonigle, A.J.S.; Tamburello, G.; Aiuppa, A.; Bryant, R.G. Protocols for UV camera volcanic SO2 measurements. J. Volcanol. Geotherm. Res. 2010, 194, 55–60. [Google Scholar] [CrossRef] [Green Version]
  71. Kern, C.; Lübcke, P.; Bobrowski, N.; Campion, R.; Mori, T.; Smekens, J.-F.; Stebel, K.; Tamburello, G.; Burton, M.; Platt, U.; et al. Intercomparison of SO2 camera systems for imaging volcanic gas plumes. J. Volcanol. Geotherm. Res. 2015, 300, 22–36. [Google Scholar] [CrossRef]
  72. Platt, U.; Lübcke, P.; Kuhn, J.; Bobrowski, N.; Prata, F.; Burton, M.; Kern, C. Quantitative imaging of volcanic plumes—Results, needs, and future trends. J. Volcanol. Geotherm. Res. 2015, 300, 7–21. [Google Scholar] [CrossRef]
  73. Gliß, J.; Stebel, K.; Kylling, A.; Dinger, A.; Sihler, H.; Sudbø, A. Pyplis–A Python Software Toolbox for the Analysis of SO2 Camera Images for Emission Rate Retrievals from Point Sources. Geosciences 2017, 7, 134. [Google Scholar] [CrossRef] [Green Version]
  74. Kern, C.; Sutton, J.; Elias, T.; Lee, L.; Kamibayashi, K.; Antolik, L.; Werner, C. An automated SO2 camera system for continuous, real-time monitoring of gas emissions from Kīlauea Volcano’s summit Overlook Crater. J. Volcanol. Geotherm. Res. 2014, 300, 81–94. [Google Scholar] [CrossRef]
  75. D’Aleo, R.; Bitetto, M.; Delle Donne, D.; Tamburello, G.; Battaglia, A.; Coltelli, M.; Patanè, D.; Prestifilippo, M.; Sciotto, M.; Aiuppa, A. Spatially resolved SO2 flux emissions from Mt Etna. Geophys. Res. Lett. 2016, 43, 7511–7519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Tamburello, G.; Aiuppa, A.; McGonigle, A.J.S.; Allard, P.; Cannata, A.; Giudice, G.; Kantzas, E.P.; Pering, T.D. Periodic volcanic degassing behavior: The Mount Etna example. Geophys. Res. Lett. 2013, 40, 4818–4822. [Google Scholar] [CrossRef]
  77. Pering, T.D.; Ilanko, T.; Wilkes, T.C.; England, R.A.; Silcock, S.R.; Stanger, L.R.; Willmott, J.R.; Bryant, R.G.; McGonigle, A.J.S. A Rapidly Convecting Lava Lake at Masaya Volcano, Nicaragua. Front. Earth Sci. 2019, 6, 241. [Google Scholar] [CrossRef]
  78. Williams-Jones, G.; Horton, K.A.; Elias, T.; Garbeil, H.; Mouginis-Mark, P.J.; Sutton, A.J.; Harris, A.J.L. Accurately measuring volcanic plume velocity with multiple UV spectrometers. Bull. Volcanol. 2006, 68, 328–332. [Google Scholar] [CrossRef]
  79. McGonigle, A.J.S.; Hilton, D.R.; Fischer, T.P.; Oppenheimer, C. Plume velocity determination for volcanic SO2 flux measurements. Geophys. Res. Lett. 2005, 32, 1–4. [Google Scholar] [CrossRef]
  80. Peters, N.; Hoffmann, A.; Barnie, T.; Herzog, M.; Oppenheimer, C. Use of motion estimation algorithms for improved flux measurements using SO2 cameras. J. Volcanol. Geotherm. Res. 2015, 300, 58–69. [Google Scholar] [CrossRef]
  81. Peters, N.; Oppenheimer, C. Plumetrack: Flux calculation software for UV cameras. Comput. Geosci. 2018, 118, 86–90. [Google Scholar] [CrossRef]
  82. Delle Donne, D.; Tamburello, G.; Aiuppa, A.; Bitetto, M.; Lacanna, G.; D’Aleo, R.; Ripepe, M. Exploring the explosive-effusive transition using permanent ultraviolet cameras. J. Geophys. Res. Solid Earth 2017, 122, 4377–4394. [Google Scholar] [CrossRef]
  83. Delle Donne, D.; Aiuppa, A.; Bitetto, M.; D’aleo, R.; Coltelli, M.; Coppola, D.; Pecora, E.; Ripepe, M.; Tamburello, G. Changes in SO2 Flux Regime at Mt. Etna Captured by Automatically Processed Ultraviolet Camera Data. Remote Sens. 2019, 11, 1201. [Google Scholar] [CrossRef] [Green Version]
  84. Ilanko, T.; Pering, T.; Wilkes, T.; Apaza Choquehuayta, F.; Kern, C.; Díaz Moreno, A.; de Angelis, S.; Layana, S.; Rojas, F.; Aguilera, F.; et al. Degassing at Sabancaya volcano measured by UV cameras and the NOVAC network. Volcanica 2019, 2, 239–252. [Google Scholar] [CrossRef]
  85. Liu, E.J.; Wood, K.; Mason, E.; Edmonds, M.; Aiuppa, A.; Giudice, G.; Bitetto, M.; Francofonte, V.; Burrow, S.; Richardson, T.; et al. Dynamics of Outgassing and Plume Transport Revealed by Proximal Unmanned Aerial System (UAS) Measurements at Volcán Villarrica, Chile. Geochem. Geophys. Geosyst. 2019, 20, 730–750. [Google Scholar] [CrossRef] [Green Version]
  86. Gaudin, D.; Moroni, M.; Taddeucci, J.; Scarlato, P.; Shindler, L. Pyroclast Tracking Velocimetry: A particle tracking velocimetry-based tool for the study of Strombolian explosive eruptions. J. Geophys. Res. Solid Earth 2014, 119, 5369–5383. [Google Scholar] [CrossRef]
  87. Thielicke, W.; Stamhuis, E.J. PIVlab—Towards User-friendly, Affordable and Accurate Digital Particle Image Velocimetry in MATLAB. J. Open Res. Softw. 2014, 2. [Google Scholar] [CrossRef] [Green Version]
  88. Thielicke, W. The flapping Flight of Birds: Analysis and Application. Ph.D. Thesis, University of Groningen, Groningen, The Netherlands, 2014. [Google Scholar]
  89. Campion, R.; Delgado-Granados, H.; Mori, T. Image-based correction of the light dilution effect for SO2 camera measurements. J. Volcanol. Geotherm. Res. 2015, 300, 48–57. [Google Scholar] [CrossRef]
  90. Kern, C.; Werner, C.; Elias, T.; Sutton, A.J.; Lübcke, P. Applying UV cameras for SO2 detection to distant or optically thick volcanic plumes. J. Volcanol. Geotherm. Res. 2013, 262, 80–89. [Google Scholar] [CrossRef]
  91. Lübcke, P.; Bobrowski, N.; Illing, S.; Kern, C.; Alvarez Nieves, J.M.; Vogel, L.; Zielcke, J.; Delgado Granados, H.; Platt, U. On the absolute calibration of SO2 cameras. Atmos. Meas. Tech. 2013, 6, 677–696. [Google Scholar] [CrossRef] [Green Version]
  92. Klein, A.; Lübcke, P.; Bobrowski, N.; Kuhn, J.; Platt, U. Plume propagation direction determination with SO2 cameras. Atmos. Meas. Tech. 2017, 10, 979–987. [Google Scholar] [CrossRef]
  93. Barnie, T.; Bombrun, M.; Burton, M.R.; Harris, A.; Sawyer, G. Quantification of gas and solid emissions during Strombolian explosions using simultaneous sulphur dioxide and infrared camera observations. J. Volcanol. Geotherm. Res. 2015, 300, 167–174. [Google Scholar] [CrossRef]
  94. Polacci, M.; Baker, D.R.D.; La Rue, A.; Mancini, L.; Allard, P. Degassing behaviour of vesiculated basaltic magmas: An example from Ambrym volcano, Vanuatu Arc. J. Volcanol. Geotherm. Res. 2012, 233–234, 55–64. [Google Scholar] [CrossRef]
Figure 1. Activity at Yasur volcano, Vanuatu, during the July 2018 field campaign. (a) Image of the gas plume rising from the summit crater. Large gas pulses are associated with explosions. (b) A nighttime view, with the south crater in the foreground and incandescence from the north crater in the background. Several vents are visible in the south crater, with one producing a Strombolian explosion. (c) Ash-rich gas plumes formed by Strombolian explosions occurred from the north crater and ash-poor gas plumes from explosions from the south crater. (d) A daytime view into the north crater, showing the crater floor topography divided by a septum into northern and southern craters.
Figure 1. Activity at Yasur volcano, Vanuatu, during the July 2018 field campaign. (a) Image of the gas plume rising from the summit crater. Large gas pulses are associated with explosions. (b) A nighttime view, with the south crater in the foreground and incandescence from the north crater in the background. Several vents are visible in the south crater, with one producing a Strombolian explosion. (c) Ash-rich gas plumes formed by Strombolian explosions occurred from the north crater and ash-poor gas plumes from explosions from the south crater. (d) A daytime view into the north crater, showing the crater floor topography divided by a septum into northern and southern craters.
Remotesensing 12 02703 g001
Figure 2. (a) An elevation-based perspective of the low summit of Yasur volcano, along with measurement positions and the prevailing plume transport direction, with the inset (b) showing a close-up of the summit crater. (c) A typical view of the plume, with red colours representing higher concentrations of SO2, showing clear mixing between plumes from different vents. (d) An example of where it was possible to differentiate between emissions from both craters. Imagery is from Google Earth®.
Figure 2. (a) An elevation-based perspective of the low summit of Yasur volcano, along with measurement positions and the prevailing plume transport direction, with the inset (b) showing a close-up of the summit crater. (c) A typical view of the plume, with red colours representing higher concentrations of SO2, showing clear mixing between plumes from different vents. (d) An example of where it was possible to differentiate between emissions from both craters. Imagery is from Google Earth®.
Remotesensing 12 02703 g002
Figure 3. (a) Example plume vectors generated during PIV analysis for movement from one frame to the next, superimposed over an SO2 absorption image. (b) Example median plume velocity across the integration line and SO2 fluxes for a time interval of 2500 s on 7th July 2018, showing clear accelerations in plume velocities during Strombolian explosions.
Figure 3. (a) Example plume vectors generated during PIV analysis for movement from one frame to the next, superimposed over an SO2 absorption image. (b) Example median plume velocity across the integration line and SO2 fluxes for a time interval of 2500 s on 7th July 2018, showing clear accelerations in plume velocities during Strombolian explosions.
Remotesensing 12 02703 g003
Figure 4. Example calibration slope coefficients (derived from regressions of gas cell concentrations against apparent absorbance) from two days of data. Timings are from the time of the first calibration on each of these measurement days. Note that the coefficient peaks towards solar noon.
Figure 4. Example calibration slope coefficients (derived from regressions of gas cell concentrations against apparent absorbance) from two days of data. Timings are from the time of the first calibration on each of these measurement days. Note that the coefficient peaks towards solar noon.
Remotesensing 12 02703 g004
Figure 5. (a) Example period of overlapping data from two separately acquiring synchronous cameras, which viewed the plume from slightly different orientations. One dataset has been shifted by the lag value, which generated the maximum correlation coefficient, following the cross-correlation between the two series in an attempt to best temporally match the data. Note that there are differences in the magnitudes of peaks and troughs in the different data series, even when shifted relative to one another in this way, due to smoothing or turbulence during plume movement through the atmosphere and the slightly different views of the units through the plume. (b) A linear regression model (R2 = 0.4) is shown, demonstrating the best fit between time series data from the two cameras, as well as confidence intervals. The statistical parameters are similar, but there are differences in peaks and troughs between the two datasets.
Figure 5. (a) Example period of overlapping data from two separately acquiring synchronous cameras, which viewed the plume from slightly different orientations. One dataset has been shifted by the lag value, which generated the maximum correlation coefficient, following the cross-correlation between the two series in an attempt to best temporally match the data. Note that there are differences in the magnitudes of peaks and troughs in the different data series, even when shifted relative to one another in this way, due to smoothing or turbulence during plume movement through the atmosphere and the slightly different views of the units through the plume. (b) A linear regression model (R2 = 0.4) is shown, demonstrating the best fit between time series data from the two cameras, as well as confidence intervals. The statistical parameters are similar, but there are differences in peaks and troughs between the two datasets.
Remotesensing 12 02703 g005
Figure 6. (ad) Retrieved gas fluxes, where clear peaks correspond to Strombolian explosions, for all the image data captured during the observation period; also highlighted are periods where the plume was grounded (a) and heavy rainfall was encountered (d). (c) An example of the explosive mass determination, where integration occurs below the explosion peak (e.g., shaded blue area), above the background levels.
Figure 6. (ad) Retrieved gas fluxes, where clear peaks correspond to Strombolian explosions, for all the image data captured during the observation period; also highlighted are periods where the plume was grounded (a) and heavy rainfall was encountered (d). (c) An example of the explosive mass determination, where integration occurs below the explosion peak (e.g., shaded blue area), above the background levels.
Remotesensing 12 02703 g006
Figure 7. (a) Separation of passive and explosion gas release for a period on the 7th July with (inset) an enhanced illustration of the simple statistical moving-minimum-based model, showing oscillation in the passive background degassing overtopped by explosive contributions. (b) The ratio of passive-to-explosive degassing and (c) a cumulative plot showing the division between passive and explosive gas release. Here, passive and explosive releases have been cumulatively summed to show the change through time at the sampling frequency. The passive-to-explosive ratio is the ratio of the final sum of the gas release.
Figure 7. (a) Separation of passive and explosion gas release for a period on the 7th July with (inset) an enhanced illustration of the simple statistical moving-minimum-based model, showing oscillation in the passive background degassing overtopped by explosive contributions. (b) The ratio of passive-to-explosive degassing and (c) a cumulative plot showing the division between passive and explosive gas release. Here, passive and explosive releases have been cumulatively summed to show the change through time at the sampling frequency. The passive-to-explosive ratio is the ratio of the final sum of the gas release.
Remotesensing 12 02703 g007
Table 1. A summary of errors from UV camera measurements of SO2 fluxes at Yasur volcano in July 2018, including short comments and the total root-mean-square (RMS) error. The bold emphasise different bits of data below.
Table 1. A summary of errors from UV camera measurements of SO2 fluxes at Yasur volcano in July 2018, including short comments and the total root-mean-square (RMS) error. The bold emphasise different bits of data below.
Treehouse Ash PlainComments
Distance1900 m2300 m
DescriptionErrorError
Light Dilution+20%+20%Underestimation only, low given plume proximity.
Gas Cell Concentration±10%±10%Manufacturer quoted
Calibration drift±15%±15%Changing calibration conditions (see text)
Plume Velocity±9±15%Based on pixel size (see text)
Plume Direction±5%±5%Based on coincident UV camera data
Plume Distance±18%±18%Based on plume deviation of 200 m
Ash content--Underestimation, not quantifiable
RMS Error−11.2%/−12.2%/Note the higher error related to underestimation (positive error)
+13.9%+14.7%
Table 2. A summary of measurement durations and SO2 flux statistics for daily UV camera measurements at Yasur Volcano in July 2018. UTC is Coordinated Universal Time.
Table 2. A summary of measurement durations and SO2 flux statistics for daily UV camera measurements at Yasur Volcano in July 2018. UTC is Coordinated Universal Time.
Date (UTC)05–06/07/1806–07/07/1808/07/1808–09/07/18
Date (Local)06/07/1807/07/1808/08/1809/08/18
Time series duration (hh:mm)4:15 04:42 03:5403:33
Total time (hh:mm)05:0105:3104:1404:17
Mean (kg/s)5.25.54.54.1
Median (kg/s)4.75.14.24.0
Standard Deviation2.22.51.71.5
Coefficient of Variation42%45%38%37%
Table 3. A breakdown of daily SO2 explosion mass data. Also displayed are percentages indicating the partitioning of gas masses between passive and explosive degassing. Below this the total gas slug masses are shown, where lower and upper ratios refer to the ranges indicated in determined active molar ratios by Woitischek et al. (Table 3; CO2/SO2 = 2.85 ± 0.17; H2O/SO2 = 315 ± 71.8; SO2/HCl = 1.6 ± 0.22).
Table 3. A breakdown of daily SO2 explosion mass data. Also displayed are percentages indicating the partitioning of gas masses between passive and explosive degassing. Below this the total gas slug masses are shown, where lower and upper ratios refer to the ranges indicated in determined active molar ratios by Woitischek et al. (Table 3; CO2/SO2 = 2.85 ± 0.17; H2O/SO2 = 315 ± 71.8; SO2/HCl = 1.6 ± 0.22).
Date (UTC)05/07/201806/07/201807/07/201808/07/201809/07/2019Total
Explosions Counted84339369135
Mean (Error)
SO2 Min (kg)10.28.9812109.8 (±1.4)
SO2 Mean (kg)26.922.127394532.0 (±4.7)
SO2 Max (kg)64.144.962816964.2 (±9.4)
Passive 1 %666470786869
Explosive 1 %343630223231
Total Slug Masses from Lower Ratios
Min (kg)712625575822721691
Mean (kg)188415501892271331642241
Max (kg)302031484370567848134206
Total Slug Masses from Mean Ratios
Min (kg)9408247591084952912
Mean (kg)248620462497357941752957
Max (kg)592941535767749363505938
Total Slug Masses from Upper Ratios
Min (kg)11671024942134711821132
Mean (kg)308825413102444651863673
Max (kg)736551597163930778887376
1 These values are from the statistical calculation of this paper (Section 4.2).
Table 4. A summary of gas slug volumes and lengths using the model of Del Bello et al. (2012) and based on gas flux and composition data acquired during Strombolian activity at Yasur volcano in July 2018.
Table 4. A summary of gas slug volumes and lengths using the model of Del Bello et al. (2012) and based on gas flux and composition data acquired during Strombolian activity at Yasur volcano in July 2018.
StatisticSlug Volume (m3)D = 3 mD = 4 mD = 5 mD = 6 mD = 7 m
Minimum (m)42861881391109076
Median (m)14055347259205170146
Mean (m)15556366272217180154
Maximum (m)42337609455364303260
D is the conduit diameter.

Share and Cite

MDPI and ACS Style

Ilanko, T.; Pering, T.D.; Wilkes, T.C.; Woitischek, J.; D’Aleo, R.; Aiuppa, A.; McGonigle, A.J.S.; Edmonds, M.; Garaebiti, E. Ultraviolet Camera Measurements of Passive and Explosive (Strombolian) Sulphur Dioxide Emissions at Yasur Volcano, Vanuatu. Remote Sens. 2020, 12, 2703. https://doi.org/10.3390/rs12172703

AMA Style

Ilanko T, Pering TD, Wilkes TC, Woitischek J, D’Aleo R, Aiuppa A, McGonigle AJS, Edmonds M, Garaebiti E. Ultraviolet Camera Measurements of Passive and Explosive (Strombolian) Sulphur Dioxide Emissions at Yasur Volcano, Vanuatu. Remote Sensing. 2020; 12(17):2703. https://doi.org/10.3390/rs12172703

Chicago/Turabian Style

Ilanko, Tehnuka, Tom D Pering, Thomas Charles Wilkes, Julia Woitischek, Roberto D’Aleo, Alessandro Aiuppa, Andrew J S McGonigle, Marie Edmonds, and Esline Garaebiti. 2020. "Ultraviolet Camera Measurements of Passive and Explosive (Strombolian) Sulphur Dioxide Emissions at Yasur Volcano, Vanuatu" Remote Sensing 12, no. 17: 2703. https://doi.org/10.3390/rs12172703

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop