scispace - formally typeset
Search or ask a question
Posted ContentDOI

Halogen activation in the plume of Masaya volcano: field observations and box model investigations

TL;DR: In this article, gas diffusion denuder sampling followed by GC-MS analysis was performed on the plume of the Santiago Crater of Masaya volcano in Nicaragua to determine the reactive species of bromine, chlorine and iodine.
Abstract: . Volcanic emissions are a source of halogens to the atmosphere. Rapid reactions convert the initially emitted hydrogen halides (HCl, HBr, HI) into reactive species e.g. BrO, Br2, BrCl, ClO, OClO and IO. The activation reaction mechanisms in the plume consume ozone (O3), which is entrained by in-mixed ambient air. In this study, we present observations of the oxidation of bromine, chlorine and iodine during the first 11 minutes after emission, investigating the plume of Santiago Crater of Masaya volcano in Nicaragua. Two field campaigns were conducted, in July 2016 and September 2016. The sum of the reactive species of the respective halogens were determined by gas diffusion denuder sampling followed by GC-MS analysis, while the total amounts of halogens and sulfur amounts were obtained by alkaline trap sampling with subsequent IC and ICP-MS measurements. Both ground and airborne sampling with an unmanned aerial vehicle (including a denuder sampler in combination with an electrochemical SO2 sensor) was performed at different distances from the crater rim. The in-situ measurements were accompanied by remote sensing observations (DOAS). For bromine, the reactive fraction increased from 0.20 ± 0.13 at the crater rim to 0.76 ± 0.26 at 2.8 km downwind, while chlorine showed an increase of the reactive fraction from (2.7 ± 0.7) × 10−4 to (11 ± 3) × 10−4 in the first 750 m. Additionally, a reactive iodine fraction of 0.3 at the crater rim and 0.9 at 2.8 km was measured. No significant increase in BrO / SO2 molar ratios was observed with the estimated age of the observed plume ranging from 1.4 min to 11.1 min. This study presents a comprehensive gas diffusion denuder data set on reactive halogen species and compares BrO / SO2 ratios with the sum of all reactive Br species. With the observed field data, a chemistry box model (CAABA/MECCA) enabled the reproduction of the observed progression of the reactive bromine to total bromine ratio. An observed contribution of BrO to the reactive bromine fraction of about 10 % was reproduced in the first minutes of the model run. The model results emphasize the importance of ozone entrainment into the plume for the reproduction of the measured reactive bromine formation and the dependence on the availability of HXOY and NOX.

Summary (5 min read)

1 Introduction

  • The relatively simple detection of BrO by differential optical absorption spectroscopy (DOAS) promoted research on the origin and fate of BrO in volcanic plumes.
  • Both the transformation of halogen species in the plume and their fate in the atmosphere are of interest.

2 Volcanic plume halogen chemistry

  • Besides numerous field surveys at various volcanoes, several atmospheric modeling studies have been conducted that have improved their understanding of the complex chemical reactions in volcanic plumes marking the interface between volcanic trace gases (and aerosols) and ambient air.
  • The models used in this study (MISTRA; PlumeChem; and Chemistry As A Boxmodel Application Module Efficiently Calculating the Chemistry of the Atmosphere, CAABA/MECCA) are initialized with the gas composition of a so-called “effective source region”.
  • The suitability of the HSC model to represent hightemperature chemistry in the plume has been debated, and it has recently been shown that the chemistry should be better represented by a kinetics model (Martin et al., 2012; Roberts et al., 2019).
  • Under ozone consumption, A subsequent uptake into aerosol enables the conversion of HBr into Br2, which partitions into the gas phase and is photolyzed to give two Br radicals and start the cycle again (Reaction R5a).
  • The major reaction pathways that involve the formation and degradation of BrO in volcanic plumes are shown in Table 1.

3.1 Site description and flight/sample strategy

  • The lava lake has been associated with large volcanic gas emissions, making it one of the largest contributors of SO2 emissions in the Central American volcanic arc (Martin et al., 2010; de Moor et al., 2017; Aiuppa et al., 2018).
  • Due to the high emission rates and the low-altitude ground-hugging plume, Masaya volcano has a severe environmental impact on the downwind areas, affecting human and animal health and vegetation (Delmelle et al., 2002; van Manen, 2014).
  • These halogen values are considered to be on the high end observed in magmas and plumes, yet they are rather typical for arc volcanism (Aiuppa, 2009; Gutmann et al., 2018).
  • UAV-based sampling was conducted in the plume hovering over the Nindirí crater and above the caldera bottom and caldera rim (red points in Fig. 1d).

3.2 Alkaline traps

  • Total halogen amounts were obtained by ground-based sampling using alkaline traps (Raschig tubes, RTs, and Drechsel bottles, DBs) (Liotta et al., 2012; Wittmer et al., 2014) at the locations marked in Fig.
  • The alkaline solution quantitatively captures acidic gas species, due to an acid–base reaction, and enables the determination of total halogens (F, Cl, Br, and I) and sulfur (S) concentrations.
  • Total volume data logging enabled mixing ratio calculation of the RT samples.
  • These samples were used for gas ratio comparisons over a longer time period.

3.3 Gas diffusion denuder sampling

  • Reactive halogen species (RHS) were sampled by gas diffusion denuder samplers using 1,3,5-trimethoxybenzene as a reactive coating (Rüdiger et al., 2017) on borosilicate brown- Atmos.
  • It is not yet experimentally evidenced that the denuder technique is completely naïve to halogen radicals, although following the chemical reaction mechanism, the reaction of the denuder coating with halogen radicals is unlikely.
  • For the UAV-based sampling, a remotely controlled sampler (called Black Box) was used and is described in detail in Rüdiger et al. (2018).
  • Furthermore, the SO2 sensor signal was transmitted to the remote control, which helped to identify regions of high SO2 concentrations in real time.

3.4 Unoccupied aerial vehicle sampler

  • The UAV used for this study is a small four-rotor multicopter with foldable arms (Black Snapper, Globe Flight, Germany) called RAVEN (Rüdiger et al., 2018).
  • Global Positioning System (GPS) data of the flights were recorded on board using the micro-SD data logger (Core 2, Flytrex, Aviation, Tel Aviv, Israel) with a 2 Hz time resolution.
  • The four batteries of the UAV were charged in the field with a car battery, enabling up to eight flights per day.

3.5 DOAS

  • DOAS measurements of SO2 and BrO were performed by a scanning-DOAS station from NOVAC (Galle et al., 2010), which is located approximately 1.5 km west-southwest of Santiago crater at an altitude of 387 m a.s.l. (above sea level, Aiuppa et al., 2018).
  • This UV spectrometer records the intensity spectra of the diffuse solar radiation over a wavelength range from 280 to 450 nm for different viewing angles by scanning the sky from horizon to horizon at steps of 3.6◦.
  • Due to a data gap caused by an instrument outage, DOAS data for July 2016 were not available; therefore, the times series only covers the later part of the field study.
  • This distance was used to estimate plume ages for BrO /SO2 ratios, by dividing 1.4 km by the wind speed obtained for the respective BrO /SO2 ratios.
  • This wind speed was based on ground-based measurements with a handheld anemometer that were taken during the field campaigns at the rims of Santiago and Nindirí craters.

4 Modeling

  • In order to compare the results of the field measurements of RHS with theoretical predictions, the CAABA/MECCA (Sander et al., 2011) box model was used.
  • In its base configuration, CAABA/MECCA simulates the chemistry of an atmospheric air parcel.
  • The atmospheric box model was initialized with the gas composition of the effective source region that was calculated by the thermodynamic equilibrium model HSC (HSC 6.1, Outotec, Finland) and then quenched with ambient air to start the atmospheric model, similar to earlier works (e.g., Gerlach, 2004; Bobrowski et al., 2007; Roberts et al., 2009, 2014).

4.1 Thermodynamic equilibrium model (HSC)

  • Data from field measurements in 2016 determined the initial conditions for the model runs.
  • The sums of all gas mixing ratios were set to 100 % to estimate the magmatic gas composition.
  • The high-temperature magmatic gas composition was mixed with different percentages of ambient atmospheric background air resulting in different atmospheric– magmatic gas ratios (VA : VM), according to the calculations of Martin et al. (2006).

4.2 Atmospheric box model (CAABA/MECCA)

  • The start point of the atmospheric chemistry box model was set to be within Santiago crater, so the plume reaching the crater rim has already experienced chemical reactions.
  • Thus, the authors were able to compare their field measurement results with the model output.
  • In the box model, the dilution was achieved by adding an amount of ambient air, mixing it, and then removing the same amount of mixed plume at a rate that achieves dilution to 1/e (0.37) over the dilution times listed in Table 3.
  • The aerosol chemical composition was set to be a 1 : 1 sulfuric acid / sulfate aerosol composition with ion concentrations according to the Köhler equation (Laaksonen et al., 1998) and given radii, temperature, and relative humidity.
  • The progressions of the respective reactive bromine species (BrX and r-Br) were fitted over the estimated plume age, and the fit coefficients were compared with the coefficients from the field data to find the best agreement by minimizing the deviation of the respective fit parameters.

5 Results and discussion

  • For samples taken on the ground, for example at the Santiago or Nindirí crater rims, the data include denuder (reactive halogens such as Br2, BrCl, or, in general, BrX and ClX) and RT samples (total halogens such as total bromine, Br, chlorine, Cl), whereas for aerial samples (e.g., caldera valley), RT data are not available.
  • A comparison of an RT sample simultaneously taken with Multi-GAS measurements resulted in 4.18± 0.22 ppm of SO2 in the RT sample and an average mixing ratio of 3.95± 0.20 ppm SO2 for the Multi-GAS data.
  • Therefore, the authors regard the alkaline trap sulfur and the electrochemically sensed SO2 as equivalent and use measured SO2 mixing ratios as a plume dilution marker.
  • The uncertainties of the obtained ratios are derived by the propagation of the errors of the analytical procedure and the sampling parameters.
  • This includes the errors of gas chromatography–mass spectrometry (GCMS), IC, and ICP-MS measurements as well as uncertainties in the sampling flow rate and the solution volume.

5.1 Total halogens

  • The alkaline trap samples were analyzed by IC and ICP-MS: sulfur, fluorine, and chlorine amounts were obtained by IC, and bromine and iodine amounts were derived from ICP-MS analysis.
  • During this time, precipitation events might have affected the incorporated plume Hal /S ratios due to the different scavenging efficiencies for each halogen compound and the water solubilities of the respective gases.
  • A significant change in the I /S ratio is observable between July and September (higher values in September), whereas the other Hal /S ratios do not change largely over this period (see Fig. S6).
  • Based on the CO2 /SO2 ratio, Aiuppa et al. (2018) reported increased CO2 emissions associated with the increased level and extension of the lava lake.

5.2 Reactive halogens

  • Reactive halogens were measured by gas diffusion denuder sampling.
  • The reactive halogen data are categorized by their sampling location, and the median of the species ratios for each location was calculated along with their propagated uncertainties (Table 6).
  • The uncertainties in the distance were estimated for each location based on the spatial distribution of the respective samples.
  • Therefore, SO2 might be underestimated, and the BrX /S ratio might be overestimated in the diluted plume.

5.3 Nighttime sample anomaly

  • One simultaneous denuder and RT sample was taken during a nighttime visit to the Santiago crater rim in 2016 (sampling duration approx.
  • This sample shows an anomalous value for reactive chlorine (see Fig. 4).
  • The values for reactive bromine and iodine are similar to those measured at the same location during the day.
  • Therefore, contamination by a calibration standard during the analysis can be ruled out.
  • A fraction of the HalX species is already formed by high-temperature reactions on the surface of the lava lake (e.g., Br2 and Cl2) (Martin et al., 2006) and can be measured at the Santiago crater rim without involving photochemistry.

5.4 Comparison with box model results

  • A two-stage chemistry modeling approach (see Sect. 1) was applied to analyze the field observations.
  • The two major objectives were (1) investigating the field data for plausibility and (2) applying the CAABA/MECCA box model in the field of volcanic plume chemistry.
  • Each model run simulated the progression of the set of model species during the first 25 min.
  • The routine compares the fit parameters of the progression of the measurement data with the respective model speciation output for bromine (see Sect. 4.2 and Fig. S2).
  • Only the BrX /Br progressions are discussed in more detail for the sake of clarity.

5.4.1 Bromine chemistry

  • For the BrX /Br progression, the best matching scenarios are presented in Fig. 6 (dashed lines).
  • Therefore, two scenarios with magmatic and atmospheric HxOy /NOx composition are investigated as extremes, representing the HSC output and the atmospheric background composition, respectively.
  • For the BrX /S ratios (Fig. 6b), the model underpredicts the field observations.
  • The measured BrO /SO2 progression (Fig. 6c) could be reproduced by various model runs with different VA : VM ratios, HxOy /NOx mixing ratios, and initial start concentrations.
  • As the modeled outgassing of Br2 is faster than the uptake of HBr and HOBr, bromide is depleted from the particles.

5.4.2 Input parameter sensitivity analysis

  • The so-called base run of the CAABA/MECCA box model, which encompasses the set of parameters from Table 3 that produced the most proximal model recreation of the field observations of the BrX /Br was chosen to study the sensitivity of the model with respect to changes in the initial start conditions.
  • The model run shown in Fig. 7c was chosen as the model base run.
  • Less quenching, indicated by higher SO2 mixing ratios, leads to less r-Br formation (relative to total Br) and increases the fraction of Br radicals.
  • The shape of the BrX species’ progression in the scenarios with less quenching (30 to 500 ppmv SO2) and slower dilution time (30 and 60 min) is related to a substantial consumption of ozone in the modeled plume (see Fig. S5).
  • After reaching a certain threshold, enough aerosol particles are present such that the Br activation is not limited by this parameter.

5.4.3 Chlorine and iodine chemistry

  • The HSC model produces reactive chlorine and iodine species.
  • ClONO2 and OClO are formed over the model runtime, and the measured reactive chlorine species are of the order of the model predic- Atmos.
  • Diatomic iodine species are formed during the first minute of the box model simulation alongside HOI by reactions analogous to Reactions (R1) and (R2).
  • For a more distant position, the measurements and the model diverge (Fig. 9c), although a smaller loss rate of OIO and further aqueous chemistry of IxOy(aq) could potentially explain the measured reactive iodine species downwind.
  • Furthermore, ultrafine and newly formed particles (< 10 nm) consisting of IxOy could also diffuse to the denuder walls and react with the coating, thereby inducing a false reactive iodine signal.

6 Conclusion and outlook

  • Additionally, the application of different techniques allowed for the most detailed observation of changes in the halogen speciation during the first 11 min following the gas https://doi.org/10.5194/acp-21-3371-2021.
  • In addition to their field studies, the authors applied the CAABA/MECCA box model to test if their current understanding of bromine chemistry in volcanic plumes fits their experimental results, assuming reasonable input data for the species not measured.
  • For iodine, the implementations of iodine chemistry, such as the knowledge on iodine oxide particle formation, into the model are necessary to enable a qualified comparison with the observed iodine data.
  • JR, AG, NB, JMdM, MI, MM, AF, and JS performed field sample and data collection.
  • This paper was edited by Rolf Müller and reviewed by Christoph Kern, Tjarda Roberts, and one anonymous referee.

Did you find this useful? Give us your feedback

Figures (17)

Content maybe subject to copyright    Report

Atmos. Chem. Phys., 21, 3371–3393, 2021
https://doi.org/10.5194/acp-21-3371-2021
© Author(s) 2021. This work is distributed under
the Creative Commons Attribution 4.0 License.
Halogen activation in the plume of Masaya volcano: field
observations and box model investigations
Julian Rüdiger
1,2
, Alexandra Gutmann
1
, Nicole Bobrowski
3,4
, Marcello Liotta
5
, J. Maarten de Moor
6
, Rolf Sander
4
,
Florian Dinger
3,4
, Jan-Lukas Tirpitz
3
, Martha Ibarra
7
, Armando Saballos
7
, María Martínez
6
, Elvis Mendoza
7
,
Arnoldo Ferrufino
7
, John Stix
8
, Juan Valdés
9
, Jonathan M. Castro
10
, and Thorsten Hoffmann
1
1
Institute of Inorganic and Analytical Chemistry, Johannes Gutenberg University, Mainz, Germany
2
Environmental Chemistry and Air Research, Technical University Berlin, Berlin, Germany
3
Institute of Environmental Physics, University of Heidelberg, Heidelberg, Germany
4
Max Planck Institute for Chemistry, Mainz, Germany
5
Istituto Nazionale di Geofisica e Vulcanologia, Sezione di Palermo, Italy
6
Observatorio Vulcanológico y Sismológico de Costa Rica Universidad Nacional, Heredia, Costa Rica
7
Instituto Nicaragüense de Estudios Territoriales, Managua, Nicaragua
8
Department of Earth and Planetary Sciences, McGill University, Montreal, Canada
9
Laboratorio de Química de la Atmósfera, Universidad Nacional, Heredia, Costa Rica
10
Institute of Geosciences, Johannes Gutenberg University Mainz, Mainz, Germany
Correspondence: Thorsten Hoffmann (t.hoffmann@uni-mainz.de)
Received: 25 March 2020 Discussion started: 22 June 2020
Revised: 9 January 2021 Accepted: 26 January 2021 Published: 4 March 2021
Abstract. Volcanic emissions are a source of halogens in the
atmosphere. Rapid reactions convert the initially emitted hy-
drogen halides (HCl, HBr, and HI) into reactive species such
as BrO, Br
2
, BrCl, ClO, OClO, and IO. The activation reac-
tion mechanisms in the plume consume ozone (O
3
), which
is entrained by ambient air that is mixed into the plume.
In this study, we present observations of the oxidation of
bromine, chlorine, and iodine during the first 11 min fol-
lowing emission, examining the plume from Santiago crater
of the Masaya volcano in Nicaragua. Two field campaigns
were conducted: one in July 2016 and one in September
2016. The sum of the reactive species of each halogen was
determined by gas diffusion denuder sampling followed by
gas chromatography–mass spectrometry (GC-MS) analysis,
whereas the total halogens and sulfur concentrations were
obtained by alkaline trap sampling with subsequent ion chro-
matography (IC) and inductively coupled plasma mass spec-
trometry (ICP-MS) measurements. Both ground and airborne
sampling with an unoccupied aerial vehicle (carrying a de-
nuder sampler in combination with an electrochemical SO
2
sensor) were conducted at varying distances from the crater
rim. The in situ measurements were accompanied by remote
sensing observations (differential optical absorption spec-
troscopy; DOAS). The reactive fraction of bromine increased
from 0.20 ± 0.13 at the crater rim to 0.76 ± 0.26 at 2.8 km
downwind, whereas chlorine showed an increase in the re-
active fraction from (2.7 ± 0.7) × 10
4
to (11 ± 3) × 10
4
in
the first 750 m. Additionally, a reactive iodine fraction of 0.3
at the crater rim and 0.9 at 2.8 km downwind was measured.
No significant change in BrO / SO
2
molar ratios was ob-
served with the estimated age of the observed plume ranging
from 1.4 to 11.1 min. This study presents a large complemen-
tary data set of different halogen compounds at Masaya vol-
cano that allowed for the quantification of reactive bromine
in the plume of Masaya volcano at different plume ages. With
the observed field data, a chemistry box model (Chemistry
As A Boxmodel Application Module Efficiently Calculat-
ing the Chemistry of the Atmosphere; CAABA/MECCA) al-
lowed us to reproduce the observed trend in the ratio of the
reactive bromine to total bromine ratio. An observed contri-
bution of BrO to the reactive bromine fraction of about 10 %
was reproduced in the first few minutes of the model run.
Published by Copernicus Publications on behalf of the European Geosciences Union.

3372 J. Rüdiger et al.: Halogen activation in the plume of Masaya volcano
1 Introduction
Volcanoes are known to be important emitters of atmo-
spheric trace gases and aerosols, both through explosive
eruptions and persistent quiescent degassing (von Glasow et
al., 2009). The most abundant gases in volcanic emissions
are water, carbon dioxide, sulfur compounds, and hydrogen
halides (Symonds et al., 1994). Typically, halogen emissions
are largely dominated by chlorine (HCl) and fluorine (HF),
whereas bromine (HBr) and iodine (HI) are 3 and 5 orders
of magnitude less abundant than chlorine and fluorine, re-
spectively (e.g., Aiuppa et al., 2005; Pyle and Mather, 2009).
Despite their low abundance, the heavy halogens (bromine
and iodine) can have a significant impact on the chemistry
of the atmosphere (e.g., von Glasow, 2010; Saiz-Lopez and
von Glasow, 2012; Platt and Bobrowski, 2015). The chemi-
cal composition of volcanic plumes is the subject of a large
number of studies, which are often aimed at gaining insights
into subsurface processes, such as the degassing of magma in
connection with changes in volcanic activity (e.g., Aiuppa et
al., 2007). In addition, the effects of volcanic gases on the at-
mosphere and biosphere at local, regional, and global scales
are also of interest, including acid deposition (wet and dry),
nutrient input (e.g., Delmelle, 2003), aerosol formation, and
the effects on the solar radiation balance (e.g., Mather et al.,
2013; Malavelle et al., 2017).
Volcanic halogen emissions have been studied for years
(e.g., Noguchi and Kamiya, 1963; Giggenbach, 1975), and
the determination of chlorine and sulfur is a common proce-
dure in such gas geochemical investigations. Bromine only
attracted more attention in later years, when the reactive
bromine species BrO was observed in volcanic plumes (e.g.,
Bobrowski et al., 2003; Oppenheimer et al., 2006). This also
proved that not only sulfur species (H
2
S, SO
2
) undergo ox-
idation by ambient reactants (such as OH and O
3
), and it
laid the foundation for various studies on oxidized halogen
species (such as BrO, ClO, OClO, and IO) (e.g., Lee et al.,
2005; Theys et al., 2014; General et al., 2015; Gliß et al.,
2015; Schönhardt et al., 2017). Despite the low abundance of
bromine in volcanic gas emissions, the relatively simple de-
tection of BrO by differential optical absorption spectroscopy
(DOAS) promoted research on the origin and fate of BrO in
volcanic plumes. Based on thermodynamic modeling, Ger-
lach (2004) hypothesized that BrO is not primarily emitted
by volcanoes and is instead formed only after the initial emis-
sions are mixed with entrained ambient air. As SO
2
can also
be easily measured by DOAS, and the oxidation of SO
2
plays
a minor role over a period of minutes to hours (McGonigle
et al., 2004), the ratio of BrO to SO
2
is used as a dilution-
compensated observation parameter.
An increase in the BrO / SO
2
ratio with increasing dis-
tance from the emitting vent has been observed at various
volcanoes (e.g., Bobrowski et al., 2007; Vogel, 2012; Gliß et
al., 2015) as have variations in BrO / SO
2
in a lateral plume
dimension with higher ratios at the edges of the plume (e.g.,
Table 1. Overview of halogen reactions in volcanic plumes (X = Cl,
Br).
Br
(g)
+ O
3(g)
BrO
(g)
+ O
2(g)
(R1)
BrO
(g)
+ HO
2(g)
HOBr
(g)
+ O
2(g)
(R2)
HBr
(g)
Br
(aq)
+ H
+
(aq)
(R3)
HOBr
(aq)
+ Br
(aq)
+ H
+
(aq)
Br
2(g)
+ H
2
O
(aq)
(R4a)
HOBr
(aq)
+ HCl
(aq)
BrCl
(g)
+ H
2
O
(aq)
(R4b)
Br
2(g)
hv
2Br
(g)
(R5a)
BrCl
(g)
hv
Br
(g)
+ Cl
(g)
(R5b)
Cl
(g)
+ O
3(g)
ClO
(g)
+ O
2(g)
(R6)
BrO
(g)
+ BrO
(g)
2Br
(g)
+ O
2(g)
(R7a)
BrO
(g)
+ BrO
(g)
Br
2(g)
+ O
2(g)
(R7b)
BrO
(g)
+ ClO
(g)
OClO
(g)
+ Br
(g)
(R8)
BrO
(g)
+ NO
2(g)
BrONO
2(g)
(R9)
BrONO
2(g)
+ H
2
O
(aq)
HOBr
(aq)
+ HNO
3(aq, g)
(R10)
Bobrowski et al., 2007; Louban et al., 2009; General et al.,
2015; Kern and Lyons, 2018). This has been explained by
a limited transfer of atmospheric O
3
to the center of the
plume, which is thought to promote the formation of BrO in
a chain reaction mechanism involving heterogeneous chem-
istry. Shortly after the discovery of the reactive bromine
species BrO, reactive chlorine species, ClO and OClO, were
also observed using the same DOAS techniques (e.g., Lee
et al., 2005; Bobrowski et al., 2007; Donovan et al., 2014;
Theys et al., 2014; General et al., 2015; Gliß et al., 2015;
Kern and Lyons, 2018). It was found that the abundance of
ClO and OClO is of the same order of magnitude as BrO, in
contrast to total chlorine, which is typically 3 orders of mag-
nitude more abundant than bromine. The formation of reac-
tive chlorine species is considered to be a secondary product
of the activation cycle of bromine (see Table 1). Recently,
reactive iodine species have also been detected by satellite
observations in the plume of Kasatochi (Schönhardt et al.,
2017), but they could not be confirmed by ground-based
measurements to date.
Both the transformation of halogen species in the plume
and their fate in the atmosphere are of interest. In particu-
lar, clarification regarding the amounts emitted into the at-
mosphere and the distribution of the halogens emitted by
quiescent (i.e., passive, non-eruptive) and eruptive degassing
is of interest. The global volcanic SO
2
flux has been esti-
mated as 23 Tg yr
1
for the period from 2004 to 2016 (Carn
et al., 2017), resulting in estimated halogen fluxes of the
same order for chlorine and 3 orders of magnitude lower for
bromine, taking global mean sulfur / halogen ratios into ac-
count (Aiuppa et al., 2009).
Bromine from various sources (e.g., polar regions, salt
lakes, and volcanoes) is involved in tropospheric and strato-
spheric ozone depletion (e.g., Wennberg, 1999; Rose et al.,
Atmos. Chem. Phys., 21, 3371–3393, 2021 https://doi.org/10.5194/acp-21-3371-2021

J. Rüdiger et al.: Halogen activation in the plume of Masaya volcano 3373
2006; Simpson et al., 2007). Tropospheric ozone depletion
has also been observed in volcanic plumes (e.g., Hobbs et
al., 1982; Kelly et al., 2013; Surl et al., 2015; Roberts, 2018),
which supports the proposed reaction mechanisms for BrO
formation via autocatalytic chain reactions. Recent observa-
tions of halogen oxides by satellites (e.g., Theys et al., 2009;
Carn et al., 2016) and aircraft missions (Heue et al., 2011)
confirm that some large volcanic eruptions may inject a pro-
portion of their volcanic halogens into the free troposphere
or even to the stratosphere, thereby confirming their poten-
tial impact on stratospheric ozone. In addition to the effects
of volcanic degassing on atmospheric chemistry, measure-
ments of volcanic emission have become an important and
well-established tool in the assessment of volcanic hazard,
and gas monitoring is used at many volcanoes around the
world (e.g., Carroll and Holloway, 1994; Aiuppa et al., 2007;
de Moor et al., 2016).
It has been also observed that the BrO / SO
2
gas ra-
tio changes with the activity of volcanoes. Bobrowski and
Giuffrida (2012) observed lower BrO / SO
2
ratios in Etna’s
plume during eruptive phases. Moveover, long-term DOAS
observations by Lübcke et al. (2014), who used stationary
spectrometers within the Network for Observation of Vol-
canic and Atmospheric Change (NOVAC; Galle et al., 2010),
showed a decrease in the BrO / SO
2
ratio before and during
explosive activity at Nevado del Ruiz volcano. More recently,
a study by Dinger et al. (2018) at the Cotopaxi volcano
(Ecuador) showed low BrO / SO
2
ratios at the beginning of
eruptive activity compared with higher ratios present during
declining volcanic activity. Finally, Warnach et al. (2019)
found a low BrO / SO
2
ratio during high explosive periods
and an increased BrO / SO
2
ratio during less explosive peri-
ods at Tungurahua volcano.
However, it is not yet clear whether the BrO / SO
2
ratio
can be used as a robust diagnostic tool for forecasting vol-
canic activity. As BrO is a reactive secondary gas species,
its concentration in a volcanic plume potentially depends
on atmospheric variables such as humidity, oxidant abun-
dance, solar radiation, and aerosol surface. The BrO / SO
2
ratio might not always or may only partially be controlled
by the total bromine emission at a particular volcano under
study (Roberts et al., 2018). Therefore, further knowledge of
the chemistry that drives halogen activation is required.
Following the Introduction, an overview of the volcanic
plume halogen chemistry and related model studies is given
in Sect. 2. The comprehensive data set obtained during two
field campaigns at Masaya volcano using several measure-
ment techniques, including DOAS (e.g, Bobrowski et al.,
2003), Multi-GAS (Shinohara, 2005; Aiuppa et al., 2006), al-
kaline traps (e.g., Wittmer et al., 2014), and gas diffusion de-
nuder sampling (e.g., Rüdiger et al., 2017) is then presented
in Sect. 3. In Sect. 4, we outline the use of an unoccupied
aerial vehicle (UAV; e.g., Rüdiger et al., 2018; Stix et al.,
2018) that enabled the sampling of a downwind plume for
the investigation of halogen-induced plume-aging processes;
this process is then reproduced by the atmospheric modeling
of plume halogen chemistry. In Sect. 5, the outcome of the
modeling is compared with the field measurement results be-
fore Sect. 6 draws the conclusions and provides an outlook
for future studies.
2 Volcanic plume halogen chemistry
Besides numerous field surveys at various volcanoes, several
atmospheric modeling studies have been conducted that have
improved our understanding of the complex chemical reac-
tions in volcanic plumes marking the interface between vol-
canic trace gases (and aerosols) and ambient air. Two differ-
ent models have been developed by researchers to simulate
the in-plume chemistry, the microphysical marine boundary
layer model (MISTRA; Bobrowski et al., 2007; von Glasow,
2010) and PlumeChem (Roberts et al., 2009; Roberts et al.,
2014). While MISTRA is a 1-D box model including mul-
tiphase chemistry, PlumeChem additionally includes plume
dispersion and 3-D simulation by employing a multiple grid
box mode, and the multiphase chemistry is parameterized us-
ing uptake coefficients rather than being modeled explicitly
in the aqueous phase. The models used in this study (MIS-
TRA; PlumeChem; and Chemistry As A Boxmodel Appli-
cation Module Efficiently Calculating the Chemistry of the
Atmosphere, CAABA/MECCA) are initialized with the gas
composition of a so-called “effective source region”. This
gas composition is typically derived from a thermodynamic
equilibrium model (HSC, Outotec, Finland; e.g., Gerlach,
2004; Martin et al., 2006). Different mixtures of magmatic
gas and ambient air yield the hot gas mixture of the effec-
tive source region, which is quenched to ambient tempera-
ture and then mixed with ambient air including O
3
, OH, and
NO
x
. The suitability of the HSC model to represent high-
temperature chemistry in the plume has been debated, and it
has recently been shown that the chemistry should be better
represented by a kinetics model (Martin et al., 2012; Roberts
et al., 2019). The limitations of HSC with respect to repre-
senting the high-temperature chemistry include the assump-
tion of thermodynamic equilibria in the hot plume region,
which is quite improbable. The choices for temperature, a
mixing ratio of volcanic and magmatic gas, and a “quench-
ing factor” are rather arbitrary and do not necessarily reflect
reality. The kinetics model studies show that the timescale for
substantial NO
x
to be formed thermally appears longer than
a reasonable lifetime of a hot plume (Martin et al., 2012).
Moreover, Roberts et al. (2019) showed differences in the
predicted formation of H
x
O
y
in a kinetics model compared
with those assumed by thermodynamics, in terms of magni-
tudes and speciation. Therefore, conclusions drawn from at-
mospheric modeling that depend on the HSC initializations
are inherently limited by the uncertainties and limitation of
HSC. Nevertheless, we follow the former studies using HSC,
despite its limitations, because there is currently no kinetics
https://doi.org/10.5194/acp-21-3371-2021 Atmos. Chem. Phys., 21, 3371–3393, 2021

3374 J. Rüdiger et al.: Halogen activation in the plume of Masaya volcano
model with a more comprehensive chemistry available as an
alternative.
The initially emitted HBr is converted into reactive species
via an autocatalytic mechanism, involving multiphase reac-
tions, which constitute a so-called “bromine explosion” (von
Glasow et al., 2009). Under ozone consumption, Br radicals
formed by high-temperature dissociation in the effective
source region react to BrO (Table 1, Reaction R1), which
in turn reacts with HO
2
or NO
2
to form HOBr (Reaction R2)
or BrNO
3
, respectively. A subsequent uptake into aerosol en-
ables the conversion of HBr into Br
2
, which partitions into
the gas phase and is photolyzed to give two Br radicals and
start the cycle again (Reaction R5a). The self-reaction of
two BrO to give two Br (with Br
2
as a secondary product)
and O
2
is suggested to be the major ozone-depleting channel
at high bromine concentrations, as in a young plume (von
Glasow, 2009). Once HBr becomes depleted, the uptake of
HOBr / BrNO
3
may promote the formation of BrCl, which
also consumes O
3
and forms reactive chlorine species such
as ClO and OClO (Reactions R6, R8). The major reaction
pathways that involve the formation and degradation of BrO
in volcanic plumes are shown in Table 1.
An extensive review of the advances of bromine specia-
tion in volcanic plumes including a comparison of different
model approaches has recently been presented by Gutmann
et al. (2018). In this study, we present in situ measurements
along with remote sensing data on the activation of Br, Cl,
and I in the volcanic plume of Masaya and further investigate
the involved halogen species by atmospheric model simula-
tions using the CAABA/MECCA box model. Although flu-
orine has been measured as well, it is not discussed in detail
in this study due to the high water solubility and the non-
reactivity of fluoride towards oxidation.
3 Measurements
3.1 Site description and flight/sample strategy
Masaya volcano in Nicaragua is a shield volcano with a
caldera size of 6 km × 11 km. The caldera hosts a set of vents,
of which the Santiago pit crater, which formed in 1858–1859,
is currently active (McBirney, 1956). Since mid-November
2015, the Santiago crater has contained a persistent superfi-
cial lava lake ( 40 m × 40 m). The lava lake has been asso-
ciated with large volcanic gas emissions, making it one of the
largest contributors of SO
2
emissions in the Central Ameri-
can volcanic arc (Martin et al., 2010; de Moor et al., 2017;
Aiuppa et al., 2018). Due to the high emission rates and the
low-altitude ground-hugging plume, Masaya volcano has a
severe environmental impact on the downwind areas, affect-
ing human and animal health and vegetation (Delmelle et al.,
2002; van Manen, 2014). With its easy accessibility by car
and low altitude, the emissions of Masaya volcano have been
studied extensively throughout the last decades. Of particular
note is the establishment of molar halogen-to-sulfur ratios,
determined to be of the order of 0.3–0.7 for chlorine and
3 × 10
4
for bromine (e.g., Witt et al., 2008; Martin et al.,
2010; de Moor et al., 2013). These halogen values are consid-
ered to be on the high end observed in magmas and plumes,
yet they are rather typical for arc volcanism (Aiuppa, 2009;
Gutmann et al., 2018). Measurements of reactive bromine
species (BrO) have been reported in the past (Bobrowski and
Platt, 2007; Kern et al., 2009). Continuous composition mon-
itoring (by Multi-GAS) has been realized since 2014, and gas
data for the onset of the superficial lava lake were presented
by Aiuppa et al. (2018).
In our field campaigns in 2016, UAV-based and ground-
based sampling approaches were undertaken to study the
plume of Masaya volcano with a focus on halogen emissions
and atmospheric reactions of the emitted halogens. Samples
were taken on the ground level at the edge of Santiago crater
(Fig. 1) at two locations (lookout south and pole site), at the
top of the rim of Nindirí crater (Nindirí rim) and at Cerro
Ventarrón. UAV-based sampling was conducted in the plume
hovering over the Nindirí crater and above the caldera bottom
and caldera rim (red points in Fig. 1d). Using ground-based
and UAV-based methods, the plume was sampled over a dis-
tance of about 2.8 km, covering an estimated age of 10 min,
depending on the wind velocity.
3.2 Alkaline traps
Total halogen amounts were obtained by ground-based sam-
pling using alkaline traps (Raschig tubes, RTs, and Drechsel
bottles, DBs) (Liotta et al., 2012; Wittmer et al., 2014) at
the locations marked in Fig. 1. The alkaline solution quanti-
tatively captures acidic gas species, due to an acid–base re-
action, and enables the determination of total halogens (F,
Cl, Br, and I) and sulfur (S) concentrations. The sampled
solutions were measured by ion chromatography (IC) and
inductively coupled plasma mass spectrometry (ICP-MS) at
the Geochemistry Laboratories of the Istituto Nazionale di
Geofisica e Vulcanologia, Palermo (Italy). A 1 M NaOH so-
lution was used in the RTs and a 4 M NaOH solution was
used in the DBs. Both solutions were made from 99 % pu-
rity NaOH (Merck, Germany) in 18.2 M cm
1
water. The
plume samples were pumped through the RTs using a GilAir
Plus pump (Sensidyne, USA) for about 1 h at 2.8–4 L min
1
.
Total volume data logging enabled mixing ratio calculation
of the RT samples. A custom-built pump (without data log-
ging) was used to pump approx. 1 L min
1
of gas through the
DBs for between 18 and 30 h each. These samples were used
for gas ratio comparisons over a longer time period.
3.3 Gas diffusion denuder sampling
Reactive halogen species (RHS) were sampled by gas dif-
fusion denuder samplers using 1,3,5-trimethoxybenzene as a
reactive coating (Rüdiger et al., 2017) on borosilicate brown-
Atmos. Chem. Phys., 21, 3371–3393, 2021 https://doi.org/10.5194/acp-21-3371-2021

J. Rüdiger et al.: Halogen activation in the plume of Masaya volcano 3375
Figure 1. (a) Location of the Masaya volcano in Central America Google Maps). (b) The Masaya pit crater system in the Masaya
caldera. The flight area (green patch and red points) and sampling locations marked by colored circles in a sketched map (c) and 3-D plot
(d).
glass tubes with a diameter of 0.9 cm. An electrophilic sub-
stitution reaction occurs within this coating, effectively trap-
ping halogen species with an oxidation state (OS) of +1 or 0
(HalX, e.g., Br
2
(OS 0) or BrCl, BrNO3, or BrO (OS + 1)),
which are considered as reactive species in contrast to the -1
OS species Br
(aq)
or HBr
(g)
. However, it is not yet experi-
mentally evidenced that the denuder technique is completely
naïve to halogen radicals, although following the chemical
reaction mechanism, the reaction of the denuder coating with
halogen radicals is unlikely. Ground-based denuder measure-
ments employed a serial setup of two denuders (2 × 50 cm)
at a flow rate of 250 mL min
1
using a GilAir Plus pump
and were conducted simultaneously to the RT sampling for
60 min to give the ratios of reactive species to total halo-
gens (e.g., HalX / Br) or total sulfur (e.g., HalX / S). For the
UAV-based sampling, a remotely controlled sampler (called
Black Box) was used and is described in detail in Rüdi-
ger et al. (2018). The typical sampling flow rate was about
180 mL min
1
for 5 to 15 min. The Black Box enabled log-
ging of the sampling duration and SO
2
mixing ratios via
the built-in SO
2
electrochemical sensor (CiTiceL 3MST/F,
City Technology, Portsmouth, United Kingdom). Further-
more, the SO
2
sensor signal was transmitted to the remote
control, which helped to identify regions of high SO
2
con-
centrations in real time. The SO
2
signal of the sensor was
time integrated over the sampling period of the denuders to
derive the HalX / S ratios at the location where the UAV hov-
ered during sampling.
3.4 Unoccupied aerial vehicle sampler
The UAV used for this study is a small four-rotor multicopter
with foldable arms (Black Snapper, Globe Flight, Germany)
called RAVEN (Rüdiger et al., 2018). We achieved flight
times of up to 15 min with a payload of approximately 1 kg,
depending on the sampling setup. Global Positioning System
(GPS) data of the flights were recorded on board using the
micro-SD data logger (Core 2, Flytrex, Aviation, Tel Aviv,
Israel) with a 2 Hz time resolution. The four batteries of the
UAV were charged in the field with a car battery, enabling up
to eight flights per day.
3.5 DOAS
DOAS measurements of SO
2
and BrO were performed by a
scanning-DOAS station from NOVAC (Galle et al., 2010),
which is located approximately 1.5 km west-southwest of
Santiago crater at an altitude of 387 m a.s.l. (above sea level,
Aiuppa et al., 2018). This UV spectrometer records the inten-
sity spectra of the diffuse solar radiation over a wavelength
range from 280 to 450 nm for different viewing angles by
scanning the sky from horizon to horizon at steps of 3.6
. For
most of the time, the volcanic plume transects the scan plane
nearly orthogonally. The slant column densities (SCDs) are
retrieved from these spectra via the DOAS method (Platt and
Stutz, 2008). Due to the rather high BrO detection limit,
spectral and arithmetical averaging is required for a reli-
able retrieval of the BrO SCDs and, ultimately, the calcu-
lated BrO / SO
2
molar ratios. As a drawback, the temporal
resolution of the BrO and BrO / SO
2
data is reduced to a
data point roughly every 30 min. For a detailed methodolog-
ical description, see Lübcke et al. (2014) and Dinger (2019).
Due to a data gap caused by an instrument outage, DOAS
data for July 2016 were not available; therefore, the times
series only covers the later part of the field study. The ob-
tained BrO / SO
2
ratios were investigated for a period be-
tween 6 August and 30 September 2016. The plume age was
estimated by employing wind speed data obtained at the air-
https://doi.org/10.5194/acp-21-3371-2021 Atmos. Chem. Phys., 21, 3371–3393, 2021

Citations
More filters
Journal ArticleDOI
TL;DR: In this article, a multi-rotor drone has been adapted for studies of volcanic gas plumes, which includes improved capacity for high-altitude and long-range, real-time SO2 concentration monitoring, longrange manual control, remotely activated bag sampling and plume speed measurement capabilities.
Abstract: . A multi-rotor drone has been adapted for studies of volcanic gas plumes. This adaptation includes improved capacity for high-altitude and long-range, real-time SO2 concentration monitoring, long-range manual control, remotely activated bag sampling and plume speed measurement capability. The drone is capable of acting as a stable platform for various instrument configurations, including multi-component gas analysis system (MultiGAS) instruments for in situ measurements of SO2 , H2S , and CO2 concentrations in the gas plume and portable differential optical absorption spectrometer (MobileDOAS) instruments for spectroscopic measurement of total SO2 emission rate, remotely controlled gas sampling in bags and sampling with gas denuders for posterior analysis on the ground of isotopic composition and halogens. The platform we present was field-tested during three campaigns in Papua New Guinea: in 2016 at Tavurvur, Bagana and Ulawun volcanoes, in 2018 at Tavurvur and Langila volcanoes and in 2019 at Tavurvur and Manam volcanoes, as well as in Mt. Etna in Italy in 2017. This paper describes the drone platform and the multiple payloads, the various measurement strategies and an algorithm to correct for different response times of MultiGAS sensors. Specifically, we emphasize the need for an adaptive flight path, together with live data transmission of a plume tracer (such as SO2 concentration) to the ground station, to ensure optimal plume interception when operating beyond the visual line of sight. We present results from a comprehensive plume characterization obtained during a field deployment at Manam volcano in May 2019. The Papua New Guinea region, and particularly Manam volcano, has not been extensively studied for volcanic gases due to its remote location, inaccessible summit region and high level of volcanic activity. We demonstrate that the combination of a multi-rotor drone with modular payloads is a versatile solution to obtain the flux and composition of volcanic plumes, even for the case of a highly active volcano with a high-altitude plume such as Manam. Drone-based measurements offer a valuable solution to volcano research and monitoring applications and provide an alternative and complementary method to ground-based and direct sampling of volcanic gases.

13 citations

Journal ArticleDOI
TL;DR: In this article, the authors presented a continuous time series of SO2 emission fluxes and BrO/ SO2 molar ratios in the gas plume of Masaya from March 2014 to March 2020.
Abstract: . Masaya volcano (Nicaragua, 12.0° N, 86.2° W, 635 m a.s.l.) is one of the few volcanoes hosting a lava lake, today. We present continuous time series of SO2 emission fluxes and BrO / SO2 molar ratios in the gas plume of Masaya from March 2014 to March 2020. This study has two foci: (1) discussing the state of the art of long-term SO2 emission flux monitoring on the example of Masaya and (2) the provision and discussion of a continuous dataset on volcanic gas data unique in its temporal coverage, which poses a major extension of the empirical data base for studies on the volcanologic as well as atmospheric bromine chemistry. Our SO2 emission flux retrieval is based on a comprehensive investigation of various aspects of the spectroscopic retrievals, the wind conditions, and the plume height. Our retrieved SO2 emission fluxes are on average a factor of 1.4 larger than former estimates based on the same data. We furthermore observed a correlation between the SO2 emission fluxes and the wind speed when several of our retrieval extensions are not applied. We make plausible that such a correlation is not expected and present a partial correction of this artefact via applying dynamic estimates for the plume height as a function of the wind speed (resulting in a vanishing correlation for wind speeds larger than 10 m/s). Our empirical data set covers the three time periods (1) before the lava lake elevation, (2) period of high lava lake activity (December 2015–May 2018), (3) after the period of high lava lake activity. For these three time periods, we report average SO2 emission fluxes of 1000 ± 200 t d−1, 1000 ± 300 t d−1, and 700 ± 200 t d−1 and average BrO / SO2 molar ratios of (2.9 ± 1.5) × 10−5, (4.8 ± 1 : 9) ×10−5, and (5.5 ± 2–6) × 10−5. These variations indicate that the two gas proxies provide complementary information: the BrO / SO2 molar ratios were susceptible in particular for the transition between the two former periods while the SO2 emission fluxes were in particular susceptible for the transition between the two latter time periods. We observed an extremely significant annual cyclicity for the BrO / SO2 molar ratios (amplitudes between 1–4–2–6 × 10−5) with a weak semi-annual modulation. We suggest that this cyclicity might be a manifestation of meteorological cycles. We found an anti-correlation between the BrO / SO2 molar ratios and the atmospheric water concentration (correlation coefficient of −47 %) but in contrast to that neither a correlation with the ozone mixing ratio (+21 %) nor systematic dependencies between the BrO / SO2 molar ratios and the atmospheric plume age for an age range of 2–20 min after the release from the volcanic edifice. The two latter observations indicate an early stop of the autocatalytic partial transformation of bromide Br− solved in aerosol particles to atmospheric BrO. Further patterns in the BrO / SO2 time series were (1) a step increase by 0.7 × 10−5 in late 2015, (2) a linear trend of 1.2 × 10−5 per year from December 2015 to March 2018, and (3) a linear trend of −0.8 × 10−5 per year from June 2018 to March 2020. The step increase in 2015 coincided with the 55 elevation of the lava lake and was thus most likely caused by a change in the magmatic system. The linear trend between late 2015 and early 2018 may indicate the evolution of the magmatic gas phase during the ascent of juvenile gas-rich magma whereas the linear trend from June 2018 on may indicate a decreasing bromine abundance in the magma.

11 citations

Journal ArticleDOI
TL;DR: In this paper, the authors used nested grids to model the plume close to the volcano at 1 km and found that ozone loss rate was 1.3 × 10 −5 molecules of O 3 per second per molecule of SO2.
Abstract: Volcanoes emit halogens into the atmosphere that undergo chemical cycling in plumes and cause destruction of ozone. The impacts of volcanic halogens are inherently difficult to measure at volcanoes, and the complexity of the chemistry, coupled with the mixing and dispersion of the plume, makes the system challenging to model numerically. We present aircraft observations of the Mount Etna plume in the summer of 2012, when the volcano was passively degassing. Measurements of SO 2-an indicator of plume intensity-and ozone were made in the plume a few 10s of km from the source, revealing a strong negative correlation between ozone and SO 2 levels. From these observations we estimate a mean in-plume ozone loss rate of 1.3 × 10 −5 molecules of O 3 per second per molecule of SO2. This value is similar to observation-derived estimates reported very close to the Mount Etna vents, indicating continual ozone loss in the plume up to at least 10's km downwind. The chemically reactive plume is simulated using a new numerical 3D model "WRF-Chem Volcano" (WCV), a version of WRF-Chem we have modified to incorporate volcanic emissions (including HBr and HCl) and multi-phase halogen chemistry. We used nested grids to model the plume close to the volcano at 1 km. The focus is on the early evolution of passively degassing plumes aged less than one hour and up to 10's km downwind. The model reproduces the so-called 'bromine explosion': the daytime bromine activation process by which HBr in the plume is converted to other more reactive forms that continuously cycle in the plume. These forms include the radical BrO, a species whose ratio with SO 2 is commonly measured in volcanic plumes as an indicator of halogen ozone-destroying chemistry. We track the modelled partitioning of bromine between its forms. The model yields in-plume BrO/SO 2 ratios (around 10 −4 mol/mol) similar to those observed previously in Etna plumes. The modelled BrO/SO 2 is lower in plumes which are more dilute (e.g. at greater windspeed). It is also slightly lower in plumes in the middle of the day compared than in the morning and evening, due to BrO's reaction with diurnally varying HO 2. Sensitivity simulations confirm the importance of near-vent products from high temperature chemistry, notably bromine radicals, in initiating the ambient temperature plume halogen cycling. Note also that heterogeneous reactions that activate bromine also activate a small fraction of the emitted chlorine; the resulting production of chlorine radical Cl causes a strong reduction in the methane lifetime and increasing formation of HCHO in the plume.

7 citations


Cites background or methods or result from "Halogen activation in the plume of ..."

  • ...A more recent study by Rüdiger et al. (2021) used the CABBA/MECCA box model....

    [...]

  • ...Modern techniques now allow for a degree of speciation in the bromine observed through these methods (Rüdiger et al., 2017, 2021)....

    [...]

  • ...Note that only a small fraction of chlorine emissions undertakes reactive chemistry in plumes, and this small fraction is mostly activated as a result of bromine chemistry (Rüdiger et al., 2021, and references therein)....

    [...]

  • ...We exclude BrNO2 as previous studies have found it to be a negligible component (Roberts et al., 2014; Rüdiger et al., 2021)....

    [...]

Journal ArticleDOI
TL;DR: In this paper, the authors presented compositions of water stable isotopes in precipitation and atmospheric nitrate (δ18O-H2O, δ2H-H 2O,, δ15N-NO3-, and δ 18O-NO 3-) collected daily between August 2018 and November 2019 in a tropical urban atmosphere of central Costa Rica.
Abstract: Increasing energy consumption and food production worldwide results in anthropogenic emissions of reactive nitrogen into the atmosphere. To date, however, little information is available on tropical urban environments where inorganic nitrogen is vastly transported and deposited through precipitation on terrestrial and aquatic ecosystems. To fill this gap, we present compositions of water stable isotopes in precipitation and atmospheric nitrate (δ18O-H2O, δ2H-H2O, δ15N-NO3-, and δ18O-NO3-) collected daily between August 2018 and November 2019 in a tropical urban atmosphere of central Costa Rica. Rainfall generation processes (convective and stratiform rainfall fractions) were identified using stable isotopes in precipitation coupled with air mass back trajectory analysis. A Bayesian isotope mixing model using δ15N-NO3- compositions and corrected for potential 15N fractionation effects revealed the contribution of lightning (25.9 ± 7.1%), biomass burning (21.8 ± 6.6%), gasoline (19.1 ± 6.4%), diesel (18.4 ± 6.0%), and soil biogenic emissions (15.0 ± 2.6%) to nitrate wet deposition. δ18O-NO3- values reflect the oxidation of NOx sources via the ·OH + RO2 pathways. These findings provide necessary baseline information about the combination of water and nitrogen stable isotopes with atmospheric chemistry and hydrometeorological techniques to better understand wet deposition processes and to characterize the origin and magnitude of inorganic nitrogen loadings in tropical regions.

5 citations

Journal ArticleDOI
TL;DR: In this article, major, minor and trace element concentrations of single rainfall events were investigated at Masaya volcano (Nicaragua) in order to determine the relative contributions of volcanogenic elements.

5 citations

References
More filters
Journal ArticleDOI
06 Jun 2002-Nature
TL;DR: It is suggested that marine iodocarbon emissions have a potentially significant effect on global radiative forcing.
Abstract: The formation of marine aerosols and cloud condensation nuclei—from which marine clouds originate—depends ultimately on the availability of new, nanometre-scale particles in the marine boundary layer. Because marine aerosols and clouds scatter incoming radiation and contribute a cooling effect to the Earth's radiation budget, new particle production is important in climate regulation. It has been suggested that sulphuric acid—derived from the oxidation of dimethyl sulphide—is responsible for the production of marine aerosols and cloud condensation nuclei. It was accordingly proposed that algae producing dimethyl sulphide play a role in climate regulation, but this has been difficult to prove and, consequently, the processes controlling marine particle formation remains largely undetermined. Here, using smog chamber experiments under coastal atmospheric conditions, we demonstrate that new particles can form from condensable iodine-containing vapours, which are the photolysis products of biogenic iodocarbons emitted from marine algae. Moreover, we illustrate, using aerosol formation models, that concentrations of condensable iodine-containing vapours over the open ocean are sufficient to influence marine particle formation. We suggest therefore that marine iodocarbon emissions have a potentially significant effect on global radiative forcing.

696 citations

Journal ArticleDOI
TL;DR: In the polar regions, unique photochemistry converts inert halide salt ions (e.g. Br−) into reactive halogen species that deplete ozone in the boundary layer to near zero levels as discussed by the authors.
Abstract: . During springtime in the polar regions, unique photochemistry converts inert halide salt ions (e.g. Br−) into reactive halogen species (e.g. Br atoms and BrO) that deplete ozone in the boundary layer to near zero levels. Since their discovery in the late 1980s, research on ozone depletion events (ODEs) has made great advances; however many key processes remain poorly understood. In this article we review the history, chemistry, dependence on environmental conditions, and impacts of ODEs. This research has shown the central role of bromine photochemistry, but how salts are transported from the ocean and are oxidized to become reactive halogen species in the air is still not fully understood. Halogens other than bromine (chlorine and iodine) are also activated through incompletely understood mechanisms that are probably coupled to bromine chemistry. The main consequence of halogen activation is chemical destruction of ozone, which removes the primary precursor of atmospheric oxidation, and generation of reactive halogen atoms/oxides that become the primary oxidizing species. The different reactivity of halogens as compared to OH and ozone has broad impacts on atmospheric chemistry, including near complete removal and deposition of mercury, alteration of oxidation fates for organic gases, and export of bromine into the free troposphere. Recent changes in the climate of the Arctic and state of the Arctic sea ice cover are likely to have strong effects on halogen activation and ODEs; however, more research is needed to make meaningful predictions of these changes.

581 citations

Journal Article

541 citations


"Halogen activation in the plume of ..." refers background in this paper

  • ...The most abundant gases in volcanic emissions are water, carbon dioxide, sulfur compounds, and hydrogen halides (Symonds et al., 1994)....

    [...]

Journal ArticleDOI
TL;DR: Atmospheric Chemistry of Iodine Alfonso Saiz-Lopez,* John M. C. Plane,* Alex R. Baker, Lucy J. Carpenter, Roland von Glasow, Juan C. G omez Martín, Gordon McFiggans, and Russell W. Smith.
Abstract: Atmospheric Chemistry of Iodine Alfonso Saiz-Lopez,* John M. C. Plane,* Alex R. Baker, Lucy J. Carpenter, Roland von Glasow, Juan C. G omez Martín, Gordon McFiggans, and Russell W. Saunders Laboratory for Atmospheric and Climate Science (CIAC), CSIC, Toledo, Spain School of Chemistry, University of Leeds, Leeds, LS2 9JT, United Kingdom School of Environmental Sciences, University of East Anglia, Norwich NR4 7TJ, United Kingdom Department of Chemistry, University of York, Heslington, York YO10 5DD, United Kingdom School of Earth, Atmospheric & Environmental Sciences, University of Manchester, Manchester, M13 9PL, United Kingdom

429 citations


"Halogen activation in the plume of ..." refers background or methods in this paper

  • ...They attributed this observation to the incomplete activation of Br in the plume center and dominance of Br2 formation (R4a) over BrCl formation (R4b) with an undepleted reservoir of particulate Br. However, Kern et al. (2009) did not detect ClO or OClO in the plume of Masaya close to the vent, but presented a detection limit for ClO/SO2 and OClO/SO2 of 5×10 and 7×10, respectively. Since the ClX/S ratios potentially include ClO and OClO as reactive chlorine species, we applied a calculation by Kern et al. (2009) to compare our results with 5 their detection limit (for long path DOAS). Under the assumption that ClX is made up by ClO, a potential OClO/SO2 ratio of 6.5×10 was calculated by employing the rate constant and photolysis frequency for the formation and depletion of OClO at an average SO2 mixing ratio of 6 ppmv at the crater rim measurement site. With this calculated ratio being below the estimated detection limit for OClO by Kern et al. (2009) our observations are in agreement with their DOAS measurements conducted in 2007....

    [...]

  • ...They attributed this observation to the incomplete activation of Br in the plume center and dominance of Br2 formation (R4a) over BrCl formation (R4b) with an undepleted reservoir of particulate Br. However, Kern et al. (2009) did not detect ClO or OClO in the plume of Masaya close to the vent, but presented a detection limit for ClO/SO2 and OClO/SO2 of 5×10 and 7×10, respectively. Since the ClX/S ratios potentially include ClO and OClO as reactive chlorine species, we applied a calculation by Kern et al. (2009) to compare our results with 5 their detection limit (for long path DOAS)....

    [...]

  • ...They attributed this observation to the incomplete activation of Br in the plume center and dominance of Br2 formation (R4a) over BrCl formation (R4b) with an undepleted reservoir of particulate Br. However, Kern et al. (2009) did not detect ClO or OClO in the plume of Masaya close to the vent, but presented a detection limit for ClO/SO2 and OClO/SO2 of 5×10 and 7×10, respectively....

    [...]

Related Papers (5)
Frequently Asked Questions (2)
Q1. What contributions have the authors mentioned in the paper "Halogen activation in the plume of masaya volcano: field observations and box model investigations" ?

In this study, the authors present observations of the oxidation of bromine, chlorine, and iodine during the first 11 min following emission, examining the plume from Santiago crater of the Masaya volcano in Nicaragua. The sum of the reactive species of each halogen was determined by gas diffusion denuder sampling followed by gas chromatography–mass spectrometry ( GC-MS ) analysis, whereas the total halogens and sulfur concentrations were obtained by alkaline trap sampling with subsequent ion chromatography ( IC ) and inductively coupled plasma mass spectrometry ( ICP-MS ) measurements. This study presents a large complementary data set of different halogen compounds at Masaya volcano that allowed for the quantification of reactive bromine in the plume of Masaya volcano at different plume ages. 

However, more detailed speciation within the reactive fraction is a desirable topic for future research. Detailed measurements in the field and further studies in controlled environments, like atmospheric simulation chambers, will help to further assess bromine activation in volcanic plumes. For this purpose, other selective denuder coatings will be developed and applied to further distinguish between species such as Br2, BrCl, or BrNO3 and Br radicals. As BrO detection is possible with DOAS spectrometers and has already been conducted at numerous volcanoes, the influencing factors on the extent of its formation need to be studied further, in particular with respect to its potential as a volcanic forecasting parameter and its use to estimate total bromine emissions.