Charge injection into the atmosphere by explosive volcanic eruptions through triboelectrification and fragmentation charging

https://doi.org/10.1016/j.epsl.2021.117162Get rights and content

Highlights

  • Size-dependent bipolar charging detected in spark-bearing volcanic granular flows.

  • Charging at the vent is likely controlled importantly by triboelectric processes.

  • Ice is not required to produce electrical discharges at the volcanic vent.

  • More experimental research on fragmentation charging is required.

Abstract

Volcanic eruptions are associated with a wide range of electrostatic effects. Increasing evidence suggests that high-altitude discharges (lightning) in maturing plumes are driven by electrification processes that require the formation of ice (analogous to processes underpinning meteorological thunderstorms). However, electrical discharges are also common at or near the volcanic vent. A number of “ice-free” electrification mechanisms have been proposed to account for this activity: fractocharging, triboelectric charging, radioactive charging, and charging through induction. Yet, the degree to which each mechanism contributes to a jet's total electrification and how electrification in the gas-thrust region influences electrostatic processes aloft remains poorly constrained. Here, we use a shock-tube to simulate overpressured volcanic jets capable of producing spark discharges in the absence of ice. These discharges may be representative of the continual radio frequency (CRF) emissions observed at a number of eruptions. Using a suite of electrostatic sensors, we demonstrate the presence of size-dependent bipolar charging (SDBC) in a discharge-bearing flow for the first time. SDBC has been readily associated with triboelectric charging in other contexts and provides direct evidence that contact and frictional electrification play significant roles in electrostatic processes in the vent and near-vent regions of an eruption. Additionally, we find that particles leaving the region where discharges occur remain moderately electrified. This degree of electrification may be sufficient to drive near-vent lightning higher in the column. Thus, near-vent discharges may be underpinned by the same electrification mechanisms driving CRF, albeit involving greater degrees of charge separation.

Introduction

Investigations over the last two decades reveal that electrical activity in volcanic columns may be broadly characterized into plume lightning and vent/near-vent discharges (Thomas et al., 2007; Behnke et al., 2013; Cimarelli et al., 2016; Aizawa et al., 2016; Van Eaton et al., 2020). The first modality comprises large-scale discharges at elevation in maturing plumes and, in many regards, is analogous to meteorological lightning (Prata et al., 2020; Van Eaton et al., 2020). Because of the large energies involved, plume lightning can often be detected with wide-range lightning networks (Van Eaton et al. (2020); Prata et al. (2020)). The second category, vent and near-vent discharges, are electrical events that neutralize lower amounts of charge per event and, as their names suggest, occur closer to the volcanic vent (Thomas et al., 2007; Behnke et al., 2013, Behnke et al., 2018). Although often lumped together, Behnke et al. (2018) showed that vent and near-vent discharges originate from fundamentally distinct breakdown processes. Vent discharges are innumerable streamer discharges that occur within or directly above the vent and are no more than a few tens of meters in length (Thomas et al., 2007; Behnke et al., 2013, Behnke et al., 2018). With current measurement techniques, these minute discharges cannot be detected individually (either at optical or RF wavelengths) or at very great distances. Collectively, however, vent discharges produce a continuous electromagnetic “hum” (commonly referred to as continual radio frequency or CRF) that can be observed with instruments like the lightning mapping array (Thomas et al., 2007; Behnke et al., 2013; Behnke and Bruning, 2015; Behnke et al., 2018). CRF is often detected together with seismic and acoustic signals implying a relationship with explosions and over-pressure conditions at the vent (Note: we will use vent-discharges and CRF sources interchangeably) (Smith et al., 2020). Occurring somewhat higher in the column and later in the eruption, near-vent lightning involves leader discharges that can have lengths between a few hundred meters to several kilometers and can be detected individually (Aizawa et al., 2016; Cimarelli et al., 2016; Behnke et al., 2018). Although larger than vent discharges, near-vent lightning still moves smaller amounts of charge per event than meteorological lightning and, thus, may be invisible to global detection networks (Vossen et al., 2021). Locally, however, it may produce changes to the ambient electric field (Behnke et al., 2018). Aizawa et al. (2016) notes that meteorological/plume lightning shares many characteristics with near-vent lightning, hinting that separating both into two categories may be unnecessary. Nonetheless, an explicit distinction between the two (which we make in the present work) may be warranted given the likely differences in electrification mechanisms underlying near-vent and plume/meteorological lightning.

An ever growing number of observations suggests that vent and near-vent discharges –what we will collectively call proximal discharges– are common during explosive eruptions (Thomas et al., 2007; Behnke et al., 2013; Aizawa et al., 2016; Cimarelli et al., 2016; Behnke et al., 2018; Smith et al., 2020; Vossen et al., 2021). These observations imply that erupted material charged had efficiently within the conduit and in the jet-thrust region. Furthermore, there is evidence that proximal discharges contain valuable information about the source of the eruption. For instance, CRF is only detected with forcing at the vent and occurs within the gas-thrust region (Behnke et al., 2018; Smith et al., 2020). Smith et al. (2020) demonstrated this fact by showing that CRF can be correlated with the acoustic and seismic signals associated with active fragmentation. Experimentally, Méndez Harper et al. (2018a) showed that the location and timing CRF emissions reflect the geometry and temporal evolution of barrel shock structures in supersonic jets. These spatiotemporal constraints suggest CRF is a valuable tool to detect incipient eruptions (Behnke et al., 2018). Regarding near-vent lightning, Cimarelli et al. (2016) indicate that the number of discharges is proportional to the over-pressure at the vent. These authors conclude that the intensity of near-vent electrical activity scales with the energy of eruptions. Furthermore, Aizawa et al. (2016) argue that the volumetric charge density in proximal jets may be much larger than that in thunderstorms. Because of these elevated charged loadings, the proximal regions of the volcanic system may also be interrogated using active methods such as GNSS occultation (Méndez Harper et al., 2019). Using an array of electrostatic instruments, Behnke et al. (2018) report complex feedback mechanisms between CRF sources and larger near-vent discharges, suggesting that both forms of discharge may depend on a shared charge budget.

Nonetheless, the physical, chemical, and dynamical processes that charge pyroclasts within the conduit and the gas-thrust region remain poorly constrained. Ice and graupel are generally absent in any large quantities (Cimarelli et al., 2016; Vossen et al., 2021). Thus, in contrast to volcanic lightning at altitude (Van Eaton et al., 2020; Prata et al., 2020), electrification mechanisms comparable to those in thunderclouds cannot account for electrical activity near the vent. Instead, proximal discharges likely reflect “dry” charging processes operating with varying degrees of efficiency within the conduit and an expanding jet.

Starting at depth, material possibly charges during the brittle failure of the magmatic column and subsequent disruptive clast-clast collisions (James et al., 2000). Fractocharging may involve a number of pathways, including piezoelectricity, pyroelectricity, atomic dislocations, positive-hole activation, and the release and capture of positive and negative ions as new surfaces are created (Dickinson et al., 1981; Xie and Li, 2018). James et al. (2000) fractured pumice through repeated impacts and abrasion and found that fragments carried elevated surface charge densities. Quite recently, Smith et al. (2018) found that eruptions producing more equant grains were associated with CRF, perhaps suggesting that milling (secondary fragmentation) plays a role in vent-discharges. It is worth noting that, although the fracture mechanism is often invoked to account for electrification in the near-vent region, not a single experimental follow up work has been conducted on the matter using natural materials (pumice) in the last 20 years. As such, fractocharging is perhaps the least-well understood “major” charging mechanism in the volcanic context.

Non-disruptive collisions may too lead to electrification through the well-known (but imperfectly understood) triboelectric effect (Hatakeyama and Uchikawa, 1951; Kikuchi and Endoh, 1982; Aplin et al., 2014; Méndez Harper and Dufek, 2016; Méndez Harper et al., 2017, Méndez Harper et al., 2020, Méndez Harper et al., 2021). Importantly, not only does triboelectricity have the ability to produce efficient charging in a granular material, but may separate charges of opposite polarity based on particle size (Hatakeyama and Uchikawa, 1951; Kikuchi and Endoh, 1982; Zhao et al., 2003; Forward et al., 2009; Waitukaitis et al., 2014). Indeed, triboelectric charging often results in smaller, negatively-charged grains and larger grains with generally positive charges. This phenomenological feature, size-dependent bipolar charging (SDBC), may be critical in the production of discharges in proximal volcanic jets (and other dusty planetary environments) as particles of different sizes and opposite charge become separated through hydrodynamics (Cimarelli et al., 2014) or sedimentation (Harrison et al., 2016).

A handful of studies have been explicitly designed to investigate triboelectric SDBC using volcanic materials. Forward et al. (2009) employed a fluidized bed to electrify basalt particles under partial vacuum. This study, however, used heavily altered materials to approximate Martian regoliths rather than recently erupted ash. Nonetheless, serendipitous reports of size-dependent bipolar charging in chemically-unmodified volcanic ash exist in the literature. Many of these observations were not placed within the modern framework of triboelectrification simply because the models had not yet been formulated. Hatakeyama and Uchikawa (1951) studied the frictional electrification of Aso and Asama ash samples. Those investigators reported standard SDBC –that is positive large grains, negative small grains– in Aso ash. However, the Asama ash samples displayed inverse SDBC (negative large grains, positive small grains). In these experiments, particles were allowed to contact foreign objects (an aluminum plate, for example), possibly biasing the polarity of the charge in manners that would not be encountered in natural systems. Thirty years later, Kikuchi and Endoh (1982) conducted similar experiments and found standard SDBC in ash particles from the 1977 Usu eruption. At Sakurajima, Miura et al. (2002) measured changes in the atmospheric potential gradient associated with small explosive events and estimated the surface charge density and polarity of ash falling out of plumes using an electrostatic separator (a method similar to the one we describe below). Those authors report particles with surface charge densities approaching the ionization limit (106105 Cm−2) and standard SDBC.

In addition to tribo- and fractoelectric processes, other mechanisms have been proposed to account for proximal discharges. Pahtz et al. (2010) suggests that materials like volcanic ash and mineral dust could charge through the polarizing effects of an ambient electric field. Aplin et al. (2014) provide evidence that the decay of radioactive elements in the magma may lead to “self-charging” of ash. Very recently, Nicoll et al. (2019) deployed sensors into a plume at Stromboli, finding that the gas phase itself is charged.

Building monitoring tools that effectively leverage proximal electrical effects requires a better understanding of the mechanisms that charge pyroclasts. Accomplishing such a feat, however, is complicated by the fact that much uncertainty remains regarding proposed electrification mechanisms themselves. For instance, while triboelectrification has been described since the time of the ancient Greeks, we have yet to unequivocally identify the charge carriers being exchanged during frictional interactions (Lacks and Sankaran, 2011; Lacks and Shinbrot, 2019). These charge carriers could be electrons, ions, or both. Similarly, some authors have presented evidence that triboelectrification arises from surface damage at minute spatial scales, implying that contact and frictional electrification are ultimately forms of fragmentation charging (Pan and Zhang, 2019; Lacks and Shinbrot, 2019).

Beyond questions surrounding the charging mechanisms that putatively drive vent and near-vent discharges, little is known about how proximal electrification influences the long term electrostatic evolution of the eruptive column. One possibility is that pyroclasts advected high into the atmosphere retain charge generated in the conduit and the gas-thrust region. This “pre-charging” may have important consequences for subsequent electrical effects, as some work indicates that charging in a granular material depends on pre-existing electric fields (e.g. Pahtz et al. (2010)). A second possibility is that the abundance of proximal discharges effectively neutralizes charge gained at or near the vent. Recombination in the gas-thrust region would imply that “downstream” lightning storms in mature plumes generally necessitate additional cycles of electrification (perhaps driven by ice) and reflect little about eruption dynamics at the source. Evidence for this second hypothesis exists in field data collected at Augustine (Thomas et al., 2007) and Redoubt (Behnke et al., 2013), which show that electrical activity waned after the initial explosive phases. These periods of electrical inactivity could signify that volcanic columns emerge from the gas-thrust region with weak degrees of charging. The resumption of electrical activity in mature plumes could indicate activation of water-based electrification mechanisms (Prata et al., 2020; Van Eaton et al., 2020).

Here, we use a shock-tube to simulate explosive, overpressured volcanic jets and address a subset of the questions posed above. Our setup allows us to investigate the charge mechanisms that drive CRF sources and make inferences regarding subsequent near-vent lightning. For the first time, we identify nominal size-dependent bipolar charging in a simulated volcanic jet bearing streamer discharges. SDBC in our shock-tube experiments provides direct evidence that tribocharging is a dominant electrification mechanism in the gas-thrust region. Additionally, we find that particles emerging from the supersonic flow carry charge densities comparable to those measured on grains falling out of proximal volcanic columns (e.g. Gilbert et al. (1991); Miura et al. (2002)). Further analysis shows that this amount of charge may be sufficient to drive near-vent lightning. As such, our results indicate that near-vent lightning is likely underpinned by the same electrification mechanisms as CRF sources, but reflects larger scale charge separation in columns.

Section snippets

Methodology

We use the shock-tube setup described previously by Cimarelli et al. (2014) to produce artificial volcanic discharges (Fig. 1a). The shock-tube (24.5 cm in length, 2.8 cm in diameter) was loaded with approximately 75 ml of volcanic ash and then pressurized to 10 MPa with argon gas. Exceeding this pressure ruptures a set of two diaphragms, causing the granular material to be ejected from the tube by explosive decompression. The dynamics of the decompression event and the resulting supersonic jet

Electrical discharges

All experimental jets produced electrical discharges. A typical electrical spark in an experiment using the 90–300 μm sample is rendered in Fig. 3. We observed discharges exclusively within the region of rarefying jet expansion as described by Méndez Harper et al. (2018a) (we note that discharges could have also occurred within the shock-tube itself, but we were unable to image these). As reported by Cimarelli et al. (2014) and Gaudin and Cimarelli (2019), we find that the number of discharges

Charging mechanisms in the near vent region

Although size-dependent bipolar charging has been described before in mobilized volcanic ash (Hatakeyama and Uchikawa, 1951; Kikuchi and Endoh, 1982; Miura et al., 2002), our experiments are the first (as far as we are aware) to detect this charge partitioning in spark discharge-bearing granular flows. The presence of SDBC in our shock-tube experiments, designed to replicate the dynamical conditions in the conduit and gas-thrust region, provides direct evidence that tribocharging is a primary

Conclusions

We have simulated volcanic jets and proximal electrical phenomena using a shock-tube. To the best of our knowledge, this work demonstrates the presence of size-dependent bipolar charging in a spark-bearing granular flow for the first time. The segregation of charge based on particle size has been associated with triboelectric charging in numerous granular flows over the last 30 years. The detection of SDBC in our experiments, together with investigations at Sakurajima, suggests that frictional

CRediT authorship contribution statement

Joshua Méndez: Conceptualization, Conducted experiments, Hardware, Data Analysis, Writing. Corrado Cimarelli: Conceptualization, Conducted experiments. Valeria Cigala: Conducted experiments. Ulrich Kueppers: Conducted experiments. Josef Dufek: Conceptualization, Writing – Reviewing and Editing.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

Joshua Méndez Harper: Conceptualization, Conducted experiments, Hardware, Data Analysis, Writing. Corrado Cimarelli: Conceptualization, Conducted experiments, Reviewing and Editing. Valeria Cigala: Conducted experiments. Ulrich Kueppers: Conducted experiments. Josef Dufek: Conceptualization, Writing- Reviewing, and Editing.

C.C. acknowledges the support of Deutsche Forschungsgemeinschaft project CI 254/2-1. J.M.H. and J.D. acknowledge the support of NASA SSW 80NSSC19K1211 and NASA IDS 1911057Z4.

References (50)

  • T. Takahashi

    Precipitation particle charge distribution and evolution of East Asian rainbands

    Atmos. Res.

    (2012)
  • C.E. Vossen et al.

    Long-term observation of electrical discharges during persistent vulcanian activity

    Earth Planet. Sci. Lett.

    (2021)
  • G. Wurm et al.

    A challenge for martian lightning: limits of collisional charging at low pressure

    Icarus

    (2019)
  • Y. Xie et al.

    Triboluminescence: recalling interest and new aspects

    Chem

    (2018)
  • K.L. Aplin et al.

    Electrical charging of ash in Icelandic volcanic plumes

  • H.T. Baytekin et al.

    The mosaic of surface charge in contact electrification

    Science

    (2011)
  • S.A. Behnke et al.

    Changes to the turbulent kinematics of a volcanic plume inferred from lightning data

    Geophys. Res. Lett.

    (2015)
  • S.A. Behnke et al.

    Investigating the origin of continual radio frequency impulses during explosive volcanic eruptions

    J. Geophys. Res., Atmos.

    (2018)
  • C. Cimarelli et al.

    Multiparametric observation of volcanic lightning: Sakurajima volcano, Japan

    Geophys. Res. Lett.

    (2016)
  • C. Cimarelli et al.

    Experimental generation of volcanic lightning

    Geology

    (2014)
  • E. Del Bello et al.

    Effect of particle volume fraction on the settling velocity of volcanic ash particles: insights from joint experimental and numerical simulations

    Sci. Rep.

    (2017)
  • J. Dickinson et al.

    The emission of electrons and positive ions from fracture of materials

    J. Mater. Sci.

    (1981)
  • J. Dufek et al.

    Granular disruption during explosive volcanic eruptions

    Nat. Geosci.

    (2012)
  • K.M. Forward et al.

    Particle-size dependent bipolar charging of Martian regolith simulant

    Geophys. Res. Lett.

    (2009)
  • J.S. Gilbert et al.

    Charge measurements on particle fallout from a volcanic plume

    Nature

    (1991)
  • Cited by (19)

    View all citing articles on Scopus
    View full text