CARBONOOO

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

Global and Planetary Change 213 (2022) 103837

Contents lists available at ScienceDirect

Global and Planetary Change


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

Black carbon, organic carbon, and mineral dust in South American tropical
glaciers: A review
S. Gilardoni a, B. Di Mauro a, *, P. Bonasoni b
a
Institute of Polar Sciences, National Research Council, Milan, Italy
b
Institute of Atmospheric Sciences and Climate, National Research Council, Bologna, Italy

A R T I C L E I N F O A B S T R A C T

Editor: Dr. Liviu Matenco Tropical glaciers are extremely sensitive to changes of climate variables. Their response to climate change is
complex and depends on multiple mechanisms affecting their mass and energy balance, including deposition of
Keywords: light absorbing particles (LAPs) from the atmosphere on snow and ice. Such particles can reduce glaciers surface
Light absorbing particles albedo, thus enhancing the melting process. LAPs include carbonaceous particles (black carbon - BC and organic
Black carbon
carbon - OC) and mineral dust aerosol (MDA). Although their relevance in global cryosphere, LAPs observations
Organic carbon
in the Andes tropical glacier areas are limited and sparse. This review aims at providing a critical evaluation of
Mineral dust
Tropical glaciers available data on LAPs in South America tropical glaciers, and highlights research gaps that will help to improve
Climate change our understanding of natural processes and anthropogenic emissions impacts on the cryosphere of the region.
In South American tropical glaciers, LAPs measurements in surface snow are mainly focused in the Cordillera
Blanca and limited information are available about their chemical composition (carbonaceous or mineral
components), while dust ice core records have been investigated in several sectors of the Andes, including in the
Cordillera Blanca, Cordillera Oriental, and Cordillera Real. Remote and field observations in South American
tropical glaciers indicate that LAPs might explain a significant fraction of snow albedo variability, however snow
albedo reduction from modelling studies varies significantly depending on LAPs concentration and composition.
Carbonaceous LAPs sources in South America are dominated by BC emissions from open fires, linked to agri­
cultural and land clearing activities, peaking in the southern hemisphere dry season (August–October). Natural
and anthropogenic dust emissions are potentially relevant contributors of LAPs on the Andes glaciers, as well.
Satellite and in-situ measurements were deployed to investigate transport episodes of carbonaceous and mineral
particles from lower altitudes towards the Andean glaciers. Nevertheless, the small number of atmospheric re­
cords of BC, OC, and MDA does not allow a systematic understanding of transport and deposition processes of
such species in the region.

1. Introduction Andean tropical glaciers are particularly sensitive to climate change


and since the 1970s they are retreating at an unprecedented rate,
South American tropical glaciers represent the main source of water although with geographical differences (Dussaillant et al., 2019; Kaser,
for 77 million people (Palut et al., 2007), who inhabit the Andean region 1999; López-Moreno et al., 2014; Pörtner et al., 2019; Seehaus et al.,
and the cities located in the nearby areas below. These glaciers are 2019; Veettil and Kamp, 2019).
essential elements for life and economy in several countries. Tropical The tropical glacier retreat in the Andes is closely linked to the in­
glaciers represent water reservoirs in the dry season for humans and crease of sea surface temperature anomalies and near-surface tempera­
wildlife, and they supply water to the agricultural sector, fostering food ture at high elevation (Francou et al., 2003; Vuille et al., 2018).
production. They also feed rivers used by hydroelectric production Nevertheless, several uncertainties persist in model predictions of
plants, and offer opportunities for recreational activities and tourism. In glacier evolution in the near future. One element of uncertainty is rep­
addition, glaciers in South America are fundamental components of the resented by the impact of light absorbing particles (LAPs) on the optical
local community culture and spiritual life (Veettil and Kamp, 2019). properties of the glacier surface and the glacier energy balance. At

* Corresponding author.
E-mail address: [email protected] (B. Di Mauro).

https://doi.org/10.1016/j.gloplacha.2022.103837
Received 30 June 2021; Received in revised form 8 April 2022; Accepted 2 May 2022
Available online 5 May 2022
0921-8181/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

global scale, the role of LAPs on snow and glaciers has been acknowl­ relevant for their radiative impact on the cryosphere.
edged as relevant (Skiles et al., 2018). LAPs are transported for days or BC is the most strongly absorbing component of fine atmospheric
weeks through the atmosphere from the local or regional emission aerosol particles (Bond and Bergstrom, 2006). It is emitted by incom­
sources to the glacier area. After deposition, such particles can absorb plete combustion and found mainly in anthropized environments.
solar radiation and reduce snow and ice albedo, promoting surface Nevertheless, BC-containing particles can survive in the atmosphere for
melting processes (Pörtner et al., 2019; Rozwalak et al., 2022). Black hours up to about a few days before being removed by wet and dry
carbon (BC), organic carbon (OC), and mineral dust aerosol (MDA) are deposition (Cape et al., 2012; Ogren and Charlson, 1983; Samset et al.,
the LAPs commonly investigated for their cryospheric potential impact 2014). Such residence time allows BC-particle transport over long dis­
(Di Mauro et al., 2021; Pörtner et al., 2019; Skiles et al., 2018). While the tances, from the emission sources to remote sites, including mountain
impacts of LAPs on the cryosphere in Greenland (Dumont et al., 2014), and glacier areas (Bonasoni et al., 2010; Lim et al., 2017).
Arctic (Khan et al., 2017), Antarctica (Cordero et al., 2022a), Southern OC is composed by hundreds of thousands of different chemical
Andes (Cordero et al., 2022b), the Alps (Di Mauro et al., 2015), and the species, but the OC component more interesting from a glacier
Asian High Mountains (Takeuchi et al., 2001; Sarangi et al., 2020) have perspective is brown carbon (BrC), which corresponds to the OC fraction
been investigated, a limited knowledge exists on the tropical Andean able to absorb solar radiation below 700 nm (Beres et al., 2020; Kirch­
glaciers. stetter et al., 2004). BrC is ubiquitous as it is co-emitted with BC during
To assess the role of LAPs in this fragile equilibrium the scientific low temperature combustion (primary BrC), and it is formed in the at­
community needs a deep knowledge of LAPs origin and their occurrence mosphere through gas-phase and aqueous-phase chemical reactions
in the cryosphere and the atmosphere in the glacier regions. In addition, (secondary BrC) (Gilardoni et al., 2016; Laskin et al., 2015; Saleh et al.,
a better understanding of the relative role of the different LAPs and their 2018).
climate-relevant properties is essential to improve model prediction of MDA is found both in urban and remote areas, since it is produced by
glacier evolution in the future. This review aims at summarizing avail­ anthropogenic activities (traffic, mining, and some agricultural prac­
able knowledge of LAPs in the South American tropical glacier area and tices) and wind re-suspension of soil particles, including desert dust. In
highlighting knowledge gaps that need to be filled to predict the syn­ fact, arid and desert areas are able to mobilize large quantities of sand
ergic impacts of climate change and natural/anthropogenic emission which can then reach the glaciers and deposit on their surface (Maho­
changes on the Andes cryosphere. Fig. 1 shows the map of the study wald et al., 2013).
region and satellite glacier images of some highlighted mountain ranges, The optical properties of LAPs strongly influence their role in
later cited in the manuscript. decreasing snow and ice albedo. Particle absorption efficiency varies
with their chemical composition (Bond and Bergstrom, 2006; Laskin
2. Relevant properties of LAPs in the global cryosphere et al., 2015), size distribution (Wang et al., 2013), and post-depositional
processes.
The effect of LAPs on the cryosphere energy balance depends on Mass absorption efficiency of freshly emitted BC varies between 4
transport processes able to mobilize the absorbing aerosol from emission and 10 m2 g− 1 at 550 nm (Bond and Bergstrom, 2006), although a larger
areas to Andean glaciers and on subsequent deposition on the glacier variability is reported for aged particles due to changes in particle size,
surface. Hereafter, we briefly review the properties of BC, OC and MDA morphology, and mixing state (Romshoo et al., 2021). BrC absorbs

Fig. 1. (a) Map of the study area. Tropical Andean glaciers are reported as gray triangles (data from National Snow and Ice Data Center, NSIDC), while colors
indicated the main mountain ranges discussed in the manuscript: Cordillera Real in orange, Cordillera Apolobamba in red, Cordillera Carabaya in blue, Cordillera
Vilcanota in green, Cordillera Huaytapallana in pink, Cordillera Blanca in purple, and Sierra Nevada de Santa Marta in brown. Satellite images of the Cordillera
Blanca (b), Cordillera Vilcanota (c) and Cordillera Real (d). Glacier polygons refer to the Randolph Glacier Inventory (RGI, 2017). (For interpretation of the ref­
erences to colour in this figure legend, the reader is referred to the web version of this article.)

2
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

mainly below 400 nm and less efficiently than BC (Kirchstetter et al., the different LAP components, i.e. dust and carbonaceous particles. SP2
2004; Tuzet et al., 2019), while fine MDA absorption efficiency ranges allows the quantification of rBC mass, but does not give any information
between 0.07 and 0.13 m2 g− 1 at 407 nm (Caponi et al., 2017). about the other light absorbing species, while thermal-optical method
Post-depositional processes modify LAP optical properties as well. quantifies BC and OC, requiring correction for mineral dust artifacts
For example, over bare ice LAPs can aggregate creating characteristics (Wang et al., 2012).
cryoconite, that is a mixture of organic and inorganic particles (Di Table 1 reports the spatial distribution of LAP measurements in snow
Mauro et al., 2017; Takeuchi et al., 2001). Cryoconite can be distributed across the Andean tropical glaciers. All observations here reported were
on the glacier surface by supraglacial streams or can be concentrated in performed with the LAHM technique and concentrations are expressed
the so called ‘cryoconite holes’ (Rozwalak et al., 2022). In addition, as nanograms of effective BC per gram of snow. Measurements were
LAPs deposited on fresh snow have a lower radiative impact because performed in surface snow samples collected over the period 2011–2017
smaller snow grain sizes are able to reflect solar radiation more effi­ (Schmitt et al., 2015; Soto Carrion et al., 2021; Torres et al., 2018;
ciently. On the contrary, when LAPs are deposited or resurfaced on Wallis, 2016). A large variability is reported, and generally effective BC
melting snow the albedo feedback is enhanced (Di Mauro et al., 2019; concentration decreases with elevation (Schmitt et al., 2015) or distance
Tuzet et al., 2020). The coating of LAPs is also important in determining from urban areas. (Wallis, 2016).
their optical properties (He et al., 2018; Pu et al., 2021). In particular, The deployment of specific techniques to quantify MDA in South
modelling results recently showed that internally mixed mineral dust in American tropical glaciers has not been observed in literature. Literature
snow produces an enhanced light absorption and thus a higher reduction studies generally use the total absorption produced by insoluble parti­
in snow albedo (Shi et al., 2021). cles in snow and ice to derive LAPs concentration, with no distinction
In addition to BC, OC and MDA, biogenic particles can also be about their chemical nature. Still the comparison between the results
considered LAPs (Skiles et al., 2018). They are involved in a biological obtained by specific measurement techniques and total absorption
albedo reduction process, also referred to as “bio-albedo feedback” measurements can be used to derive conclusions about the relative
(Cook et al., 2020). The most relevant category of biogenic particles is contribution of dust and carbonaceous particles. For example, a subset of
represented by photosynthetic organisms such as snow algae (Takeuchi snow samples collected on the Cordillera Blanca were analyzed simul­
et al., 2006) and glacier algae (Di Mauro et al., 2020). The quantitative taneously with LAHM and SP2. Since LAHM quantifies the contribution
analysis of biogenic particles in tropical glaciers is still lacking. A recent of dust, BC to light absorption, with no distinction among them, the
review on snow and glacier algae (Hoham and Remias, 2020) also comparison between LAHM and SP2 helped to highlight that refractory
highlights the lack of scientific literature in the tropical Andes. BC dominates light absorption in the southern part of the Cordillera
Blanca (nearby Huaraz), while dust accounts for the largest share of
3. LAPs in South American tropical cryosphere LAPs in the northern area (Schmitt et al., 2015), confirming the higher
contribution of anthropogenic LAPs in the proximity of urban
3.1. In-situ observations agglomerates.

Three different approaches are proposed in literature for the mea­ 3.2. LAPs ice core records
surement of LAPs in snow samples. Two methods require LAP collection
on a filter medium after filtration of the melted snow. In the first filter- The most comprehensive archive of information regarding LAPs in
based method, the absorption spectrum of LAPs is measured with an South American glaciers comes from the ice cores (Thompson et al.,
integrated sphere sandwich spectrometer (ISSW) (Clarke, 1982; Grenfell 2003). In fact, tropical glaciers are important archives of past LAPs
et al., 2011) or by a light absorption heating method (LAHM) (Schmitt deposition in this area, and they have been used to reconstruct past
et al., 2015). The second filter analysis quantifies elemental carbon (a global climate change. The measurements of LAPs in ice cores strongly
proxy of BC) and OC by thermal-optical analysis (Forsström et al., 2009; depend on the nature of the particles. In fact, mineral dust particles
Lavanchy et al., 1999). The third approach measures the mass of re­ concentration and size distribution are determined with optical counters
fractory black carbon (rBC) in aerosolized melted snow, using laser (Thompson et al., 1995), black carbon is determined with single-particle
induced incandescence analysis with a Single Particle Soot Photometer soot photometers (Osmont et al., 2019), and organic carbon is deter­
(SP2) (Lim et al., 2014; Schwarz et al., 2012). ISSW and LAHM measure mined through chemical analysis such as photochemical oxidation and
the light absorption produced by particles, without distinction between electrolytic quantification of the CO2 produced during oxidation

Table 1
Spatial variability of effective BC concentrations measured in surface snow samples collected on the tropical Andean glaciers; symbols indicate range of annual average
(*), range of monthly average (◦ ), and campaign average (§).
Glacier/area Lat Lon Effective BC (ng/g) Year Region Sampling period Reference

Region1 − 9.10 − 77.70 4–9* 2012 Cordillera Blanca Dry season Schmitt et al., 2015
Region 2 − 8.95 − 77.70 7–8* 2012 Cordillera Blanca Dry season Schmitt et al., 2015
Region 3 − 8.60 − 77.45 14–16* 2012 Cordillera Blanca Dry season Schmitt et al., 2015
Region 4 − 8.60 − 77.45 15–25* 2012 Cordillera Blanca Dry season Schmitt et al., 2015
Region 5 − 8.55 − 77.35 12–21* 2012 Cordillera Blanca Dry season Schmitt et al., 2015
Yanapaccha − 9.02 − 77.58 13–1092 ◦
2016 Cordillera Blanca Dry and wet seasons Sánchez Rodríguez and Schmitt, 2018
Shallap − 9.47 − 77.33 91–1047 ◦
2016 Cordillera Blanca Dry and wet seasons Sánchez Rodríguez and Schmitt, 2018
Tocllaraju − 9.35 − 77.40 <60 ◦ 2017 Cordillera Blanca Dry and wet seasons Sánchez Rodríguez and Schmitt, 2018
Vallanaraju − 9.42 − 77.46 2–781 ◦ 2017 Cordillera Blanca Dry and wet seasons Sánchez Rodríguez and Schmitt, 2018
Huaytapallana − 11.89 − 75.00 31 § 2015–16 C. Huaytapallana November, October Torres et al., 2018
Ampay − 13.56 − 72.91 21–65 ◦ 2017 C. de Vilcabamba February–November Soto Carrion et al., 2021
Ishinca − 9.39 − 77.41 2015 Cordillera Blanca June–August Wallis, 2016
Tocllaraju − 9.35 − 77.41 2015 Cordillera Blanca June–August Wallis, 2016
80 §
Urus Este − 9.35 − 77.43 2015 Cordillera Blanca June–August Wallis, 2016
Vallanaraju − 9.42 − 77.46 2015 Cordillera Blanca June–August Wallis, 2016
Alpamayo − 8.88 − 77.65 2015 Cordillera Blanca June–August Wallis, 2016
Pisco − 9.01 − 77.64 41 §
2015 Cordillera Blanca June–August Wallis, 2016
Yanapaccha − 9.02 − 77.58 2015 Cordillera Blanca June–August Wallis, 2016

3
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

(Federer et al., 2008). Important ice cores have been extracted from the presence of 100 and 1000 ppm of dust could lead to 3.8%–9.6% and
Quelccaya Ice Cap (Cordillera Oriental) (Thompson et al., 1995) and the 11.5%–22% snow albedo reduction, respectively. The co-emission of BC
col. of Huascarán (Cordillera Blanca) (Thompson et al., 1986). These and mineral dust from Amazon fires could enhance glacier mass loss by
climate records allowed scientists to study climate variability and 3% - 5% (Magalhães et al., 2019).
environmental changes of the last 1500-year and provided the first ev­
idence of the Little Ice Age in the Southern Hemisphere Tropics. In 4.1. Remote-sensing observations
particular, the analysis of mineral dust revealed valuable information on
the atmospheric transport during the Late Glacial Stage. Mineral dust The presence of impurities in snow and ice induces a decrease in
records from the Nevado Illimani (Cordillera Real) showed a strong surface albedo (Skiles et al., 2018). When snow and ice get darker
geochemical and mineralogical variability and provided information on because of atmospheric impurities, they absorb further solar radiation,
summertime deep convection over the Bolivian Altiplano (Lindau et al., and the melting process is enhanced (Warren and Wiscombe, 1980).
2021). Reis et al. (2021) analyzed dust concentration in the Quelccaya Remote sensing data can be very useful in determining the impact and
Ice Cap (Cordillera Blanca), reconstructing the link between dust variability of impurities in snow and ice (Painter et al., 2013). For
deposition and the Pacific Decadal Oscillation during the last 12 years. example, repeated acquisitions with space borne sensors such as
BC and OC variability in glacier ice has been less studied in the Sentinel 2–3 (ESA) and Landsat 8 (NASA) provide fundamental infor­
tropics. Recent analysis of ice cores from the Illimani glacier in the mation on impurities in snow and ice (Di Mauro et al., 2015; Dumont
Bolivian Andes (Cordillera Real) allowed to reconstruct the climate et al., 2020). These satellite missions feature several spectral bands in
variability of the entire Holocene back to the last deglaciation 13,000 the visible and near infrared wavelengths, and the analysis of reflectance
years ago (Osmont et al., 2019). BC was measured in this ice core and in these bands has been used for monitoring darkening processes of the
revealed a strong seasonality marked by low concentrations during the cryosphere worldwide (Cook et al., 2020; Di Mauro et al., 2015; Dumont
wet season and high concentration in the dry season. The variability in et al., 2014; Kokhanovsky et al., 2019). While BC effect on snow can be
BC concentration was linked to biomass burning in the Amazon Basin difficult to estimate with remote sensing data (Warren, 2013), mineral
and precipitation changes at the Illimani site (Osmont et al., 2019). dust and cryospheric algae induce a distinctive signature on the spectral
albedo of snow and ice (Di Mauro et al., 2020; Painter et al., 2013). This
4. Impacts of LAPs on snow and ice albedo makes possible to determine the abundance and radiative forcing of
different impurities from satellite data (Painter et al., 2012; Sarangi
Deposition of LAPs on snow and glaciers induces a darkening of et al., 2020). Both multi- and hyperspectral sensors orbit the Earth,
snow- and ice-covered surfaces, which accelerates their melting (Di allowing to exploit different source of data for monitoring the spatial
Mauro, 2020). LAPs in fact can accelerate the melting of snow cover over and temporal variability of impurities in snow and ice. In particular,
glacier surfaces, leaving exposed bare ice to solar radiation. Glacier ice high spectral resolution is needed in the visible and near infra-red por­
features a lower albedo with respect to snow (Warren, 2019), so this tions of the spectrum to retrieve snow and ice properties. Current and
feedback mechanism is expected to exert a negative impact on glacier future hyperspectral satellite missions (i.e., DESIS, PRISMA, EnMap,
mass balance. Modelling tools have been developed to estimate snow EMIT) are fundamental tools for LAPs estimation from space. Recent
albedo reduction and melting based on snow properties (grains size and research exploited satellite data to study the temporal variability of
liquid water content), and LAP features such as the nature and con­ glacier albedo in the last decades. Decreasing trends of albedo have been
centration of absorbing particles, and also their size distribution and identified both in the Alps (Fugazza et al., 2019; Naegeli et al., 2019)
coating. and in the Chilean Andes (Shaw et al., 2021). This latter study high­
Several studies investigated the impact of LAPs on surface albedo in lighted a high sensitivity of glacier albedo to drought periods and
the tropical Andean glaciers employing the SNICAR (Snow, Ice, and remarked on the impact of present and future glacier melt on water
Aerosol radiation) model (Flanner and Zender, 2005). Sanchez Rodri­ security in the region.
guez (2016) analyzed the effect of LAPs on the Yanapaccha and Shallap Also, laboratory and field experiments have been developed to
glaciers (Cordillera Blanca) in 2015–2016. The effective BC concentra­ demonstrate the impact of BC on snow albedo. Hadley and Kirchstetter
tion measured in surface snow ranged between 10 and 500 ng/m3, (2012) created a set up for artificially including BC particles in snow
corresponding to an average albedo reduction of 9% at the Yanapaccha crystal and measured the spectral albedo with a laboratory spectrom­
glacier and 15% at the Shallap glacier. Torres et al. (2018) investigated eter. Brandt et al. (2011) developed a controlled snowmaking experi­
surface snow samples collected on the Huaytapallana glacier (Cordillera ment to test the relation between black carbon content and reduction of
Huaytapallana). The measured effective BC concentrations averaged 31 snow albedo. Peltoniemi et al. (2015) organized a field experiment to
ng/g, and led to an albedo reduction ranging between 0.6% and 5%. The deposit soot on snow and studied the bidirectional reflectance factor of
highest albedo reductions were observed during the dry season. The contaminated snow.
aforementioned studies measured LAP concentration in snow using a All the aforementioned experiments faced the difficulties of repro­
LAHM approach. ducing the natural process of LAPs deposition on snow and ice. To gain
Recent studies highlight that snow albedo reduction of the tropical knowledge about the nature and radiative forcing of LAPs in snow and
glaciers in South America are not linked exclusively to BC. Bolaño-Ortiz ice, extensive field sampling campaigns and reflectance measurements
et al. (2020) studied the correlation between snow albedo, atmospheric are needed. Field spectrometers operating continuously have been
LAP concentrations, and meteorological variables from satellite obser­ installed in snow-covered areas, and they provided important informa­
vations in the Sierra Nevada de Santa Marta (Colombia). During the tion on LAPs deposition, resurfacing and radiative forcing (Dumont
investigated period (2000− 2020), dust explained 37% of snow albedo et al., 2017; Kokhanovsky et al., 2021).
variability, while BC and OC did not show any correlation with snow Snow and ice sample collection is challenging in remote moun­
albedo. Nevertheless, the time scale of satellite observations hampered tainous areas, and citizen science initiatives have been launched in
short-term events, such as the transport of biomass burning plumes, and Europe in order to involve hikers and ski mountaineers in scientific
thus BC and OC impact was likely underestimated (Bolaño-Ortiz et al., projects. For example, French researchers asked people to collect snow
2020). Magalhães et al. (2019) analyzed the effect of LAP emitted by the after a massive Saharan dust deposition in 2021,1 and Italian researchers
Amazon fires on the Zongo glacier (Cordillera Real), simulating BC asked the local communities to collect snow and ice samples visibly
transport and deposition using the global fire database emissions as
input. SNICAR simulations indicated that the albedo reduction due to BC
in snow ranged between 1.8% and 7.2%, while the simultaneous 1
https://labo.obs-mip.fr/multitemp/aidez-nous-a-mesurer-la-neige-orange/

4
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

colored in red because of cryospheric algal bloom.2 These initiatives are in the outer tropic impact area. Savanna, grassland, and shrubland fires
useful to engage local communities and to raise awareness on glacier dominate open burning carbonaceous emissions in the inner tropics (16
melting and climate change impacts on the society. kton of BC per year), while deforestation and degradation dominate in
the outer tropics (125 kton of BC per year). Although primary OC
5. LAPs sources emissions are almost one order of magnitude larger than BC, their
impact on cryosphere albedo is likely less significant, due to the smaller
Understanding spatial and temporal variability of LAPs in the cryo­ mass absorption coefficient of OC compared to BC (Tuzet et al., 2019).
sphere requires a detailed knowledge of LAPs sources and their temporal Nevertheless, the contribution of secondary OC is here not reported and
evolution. The following section describes natural and anthropogenic OC optical properties are still characterized by large uncertainties
emission data of BC, OC, and MDA in the tropical Andes. (Laskin et al., 2015).
Chen et al. (2013) observed that in the period 2007–2012, the
5.1. Black carbon and organic carbon number of active fires decreased at a slower rate compared to defores­
tation, suggesting that the contribution of agricultural fires and under­
Although a few countries in the Andes are working on the develop­ story forest fires might become more prominent compared to the past.
ment of national BC emission inventories, regional coherent inventories Chuvieco et al. (2008) reported that in 2004 in all South America a
are lacking (Huneeus et al., 2020). For this reason, here we discuss surface larger than 150.000 km2 was burned, and that a significant
carbonaceous aerosol emissions using global inventories. Open burning fraction (17%) was agricultural area. Similarly, using satellite images
data are derived from the Global Fire Emission Database with small fires collected over a 12-year period, Cardozo et al. (2014) reported that fires
(GFED4.1 s, http://globalfiredata.org) with a spatial resolution of in Rondônia occurred more frequently in areas converted from forest to
0.25◦ x 0.25◦ (van der Werf et al., 2017). Emissions are calculated by livestock or crop fields, confirming the widespread use of open burning
multiplying mass of burned dried matter by the emission factors of the as agricultural practice. In the GFED4.1 s inventory, agricultural fire
different vegetation types, according to Giglio et al. (2013). Primary contribution to BC emissions varies between a few percentage points up
carbonaceous aerosol emissions from other anthropogenic activities are to 25% (43%) in the inner tropic impact area (outer tropic impact area),
from the Emission Database for Global Atmospheric Research EDGAR depending on the month. On the annual time-scale, agriculture BC share
v5.0 (https://edgar.jrc.ec.europa.eu/dataset_ap50#sources) (Crippa is less than 4% in the two areas, although fires classified as “non-agri­
et al., 2019). Here we use annual sector specific BC and OC emissions, cultural” generally start on farm fields (Pearson et al., 2017), suggesting
gridded at a resolution of 0.1◦ x 0.1◦ . a stronger impact of the agricultural sector on total open burning
It is worth noting that global emission inventories rely on many emissions.
datasets, including emission factors, activity data, assumptions about The use of fires in agriculture (slash and burn) is a very common
the implementation of control policies, and proxies used for spatial and practice in the Andes. Although farmers have believed for a long time
temporal disaggregation. The use of different datasets and assumptions that burning enriched the soil, it is now clear that repeated burns
might lead to large discrepancies between inventories, thus source decrease soil productivity and the ability to retain water, favoring
apportionment studies are recommended for their validation in South erosion.
America (Huneeus et al., 2020).
5.1.2. Other anthropogenic sources
5.1.1. Open burning In addition to open burning, carbonaceous aerosol is emitted by
Globally, open burning is the largest source of BC and primary OC combustion for energy production, industrial processes, residential
(Bond et al., 2013). In South America open burning peaks during the dry purposes, and transportation, as well as solid waste incineration. The
season: in August–September in the southern hemisphere, and in transport of such anthropogenic emissions from urban and densely
February in the northern hemisphere (Chuvieco et al., 2008). Open populated regions to the free troposphere and to the cryosphere is not
burning in South America includes forest fires, grassland and woodland well characterized, especially in the southern hemisphere (Wie­
burning, as well as agricultural waste burning (Cochrane, 2002). densohler et al., 2018). During daytime, the mountain breeze can
Gas and particles produced by open burning impact air quality at transport polluted air masses from the boundary layer to the free
regional scale, both in the northern and southern South America (Castro troposphere, moving along the mountain slopes, through the so-called
Videla et al., 2013; Mendez-Espinosa et al., 2019; Rincón-Riveros et al., mountain chimney effect (Chen et al., 2009). Bonasoni et al. (2010)
2020), and can be lifted up to 3–5 km height and transported into the described the transport of BC aerosol from the Kathmandu valley up to
atmosphere on a regional scale (Baars et al., 2012; Bourgeois et al., the Nepal Climate Observatory – Pyramid, a monitoring site located at
2015), potentially reaching the Andean glaciers (Magalhães et al., about 5079 m altitude and at 200 km from the main anthropogenic
2019). Magalhães et al. (2019) observed that from August to September, emission areas. Gramsch et al. (2020) observed the transport of BC from
when the fire count in the Amazon is the highest, the prevailing wind urban and industrial areas nearby Santiago (Chile) up to a mountain site
direction is from east to west, promoting BC and OC transport from the located at about 2800 m a.s.l. A similar transport mechanism might
fire region to the outer tropical glaciers. The main source region is explain the increasing LAP concentration observed in the Cordillera
identified between the equator and 20◦ S. Conversely, inner tropical Blanca glaciers at decreasing distance from the city of Huaraz (Wallis,
glaciers can potentially be impacted mainly by open burning emissions 2016).
in the northern South America, due to the decoupled wind circulation Using the EDGAR v5.0 emission inventory, we calculated the annual
pattern (Castro Videla et al., 2013). BC and OC emissions, for the year 2015, within different distances from
Fig. 2 shows the map of burned area in 2016, together with the the tropical glaciers (Fig. 3). We report emissions for 2015 as it is the
historical monthly means of BC and OC emissions for the period reference year for several Nationally Determined Contributions (NDCs),
2006–2016. Emissions are calculated from the Global Fire Emission as introduced by the Paris Agreement. The main sectors accounting for
Database (GFED4.1 s, http://globalfiredata.org) for the two source re­ BC and OC emissions are road transport, combustion in manufacturing
gions potentially impacting inner tropical and outer tropical glaciers, as industry, and residential combustion. For BC, the relative contribution
indicated in Fig. 2a. Average BC (OC) annual emissions are equal to of the three sectors varies with distance (glacier footprint area). In the
about 20 (180) kton in the inner tropic impact area and 210 (1700) kton area within 500 km from the glaciers, the main BC source is the indus­
trial sector, while within 100 km or smaller distances, the contributions
of transportation and residential combustion become more prominent.
2
https://blogs.egu.eu/divisions/cr/2021/10/08/snow-algae-alps/ Conversely, residential combustion dominates OC emissions at any

5
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

Fig. 2. Map of burned area in 2016 (panel a), and historical monthly means of BC (panel b) and OC (panel c) emissions for the period 2006–2016 in the inner tropic
impact area (rectangle A) in green and the outer tropic impact area (rectangle B) in orange. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

Wiedensohler et al., 2018).


Residential sector includes emissions from wood, agricultural waste,
dung, and coal combustion for cooking and heating purposes (Artaxo
and Raga, 2016). In Latin America, cooking dominates residential
combustion emissions (Klimont et al., 2017), and is often performed
using a biomass-fueled fire surrounded by stones or within a U-shaped
enclosure.

5.2. Mineral dust aerosol

Estimating emissions of mineral dust aerosol and their impact on the


cryosphere is challenging. Recent improvements in mineral dust emis­
sion estimates and deposition fluxes have been made by integrating
ensemble global model simulations with observation constraints, using
inverse modelling techniques (Kok et al., 2021). Although inverse
modelling improved remarkably the description of dust emission and
deposition in the northern hemisphere, the limited number of field ob­
servations and the inaccuracy of model and measurements did not allow
significant progresses in the southern hemisphere (Gaiero et al., 2013;
Kok et al., 2017).
The most important natural dust source region in the Andes is the
Altiplano, or Andean Plateau (Prospero et al., 2002). The Altiplano is
located mainly in Bolivia and reaches Peru to the north and Argentina to
the south, extending over 1000 km. It lies at an altitude of 3750–4000 m
and hosts densely populated regions, including the cities of Puno, La
Fig. 3. Annual emissions of black carbon (BC, panel a) and organic carbon (OC,
Paz, and Oruro. Dust transport events from the Altiplano to the Andes
panel b) in ktons from EDGARv5.0 emission inventory calculated for different
have been reported (Chauvigné et al., 2019; Lindau et al., 2021).
distances from the glaciers. Reported emission sectors are combustion for
manufacturing (IND), road transportation (TRS), residential combustion (RCO), In addition, the west coasts of South America are characterized by
oil refineries and transformation industries (REF), and power industry (ENE). extensive arid regions that can potentially act as dust source for the
tropical glacier areas. In particular, the Atacama Desert extends from
distance from the glaciers. 15◦ S to 30◦ S, with an elevation between sea level and 3500 m. Instead,
Emissions from the road transport sector are mainly due to diesel cars the Sechura desert is located in northern Peru and its altitude rarely
and tracks. The high BC emissions from this sector are due to the lack of exceeds 100 m. Dust devils and convective plumes have been frequently
technical and regulatory tools, necessary to support the use of control observed in the Atacama and Sechura deserts (Jemmett-Smith et al.,
technologies and the implementation of traffic emission certificates and 2015). These structures can form thermals able to uplift dust up to the
vehicle standards (Artaxo and Raga, 2016). The old vehicle fleet and the boundary layer top and the free troposphere (Ansmann et al., 2009; Hess
presence of high emitters contribute to the high atmospheric BC con­ and Spillane, 1990).
centrations observed especially in urban areas (Krecl et al., 2019; Using the gridded dust emissions from Kok et al. (2021), Table 2

6
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

Table 2 (MAAP) since 2011 (Chauvigné et al., 2019). Assuming an absorption


Mean natural dust emission fluxes in Tg per season in the whole Andean region efficiency of 7.5 m2 g− 1 at 550 nm (Bond and Bergstrom, 2006), during
(22◦ S-14◦ N / 65◦ W-81◦ W) and in the Altiplano area (22◦ S-14◦ S / 65◦ W-68◦ W); the period 2012–2015 the annual eBC concentration averaged 85 ng
one-sigma uncertainty range is reported between brackets. m− 3 (Chauvigné et al., 2019). A clear seasonality was observed and eBC
JFM AMJ JAS OND concentration in the dry season (May–July) was on average 40% higher
10.6 10.0 9.4 10.0 compared to the wet season (December –March). The Chacaltaya record
Whole region shows the highest eBC levels during the biomass burning months, be­
(10.6–12.8) (10.0–15.5) (9.4–16.1) (10.0–15.6)

Altiplano
0.6 4.8 4.2 2.6 tween July and November, clearly highlighting the impact of regional
(0.6–1.2) (4.8–10.1) (4.2–10.9) (2.6–7.6) emissions from Amazon fires on the higher troposphere (Chauvigné
et al., 2019).
reports natural dust emission fluxes for the Andean area (22◦ S-14◦ N / Continuous aerosol measurements were performed from March 2007
64◦ W-81◦ W), which includes Sechura desert, Nazca district, northern to May 2008 at the Pico Espejo Atmospheric Research Station in
Atacama Desert, and the Altiplano. In addition, the emission fluxes Venezuela, at 4765 m asl (Hamburger et al., 2013), at about 10 km form
calculated only for the Altiplano are indicated. Total fluxes of PM20dust the city of Mérida (1630 m asl). eBC was derived from light attenuation
(dust particles with aerodynamic diameter smaller than 20 μm) are measurements performed with a Particle Soot Absorption Photometer
equal to 40 Tg per year. The Altiplano is characterized by intense pre­ (PSAP). When the site was in the lower free troposphere the impact of
cipitation in the summer months (December–February) (Lindau et al., the nearby valley was negligible, and eBC averaged 20 ng m− 3 and 50
2021), which lead to significant reduction in emission fluxes in the area. ng m− 3, in the wet and in the dry season, respectively. Conversely, an
Conversely, emission fluxes in the deserts along the west coast of South average eBC concentration of 120 ng m− 3 was observed when the site
America are the highest during summer, when the high insolation pro­ was within the boundary layer and reached by open biomass burning
motes dust re-suspension by convective motion (Jemmett-Smith et al., emissions from the Amazon (Hamburger et al., 2013). Higher eBC levels
2015). were recorded from June to December 2015 in Huancayo (3300 m a.s.
Besides natural sources, dust in the Andes is emitted also by l.), in the Peruvian Andes, at about 12 km from Huancayo City, a
anthropogenic activities, including agriculture and mining (Barraza growing urban agglomerate. eBC, measured with a 7-wavelength
et al., 2021; Brenning and Azócar Sandoval, 2010; Rolando et al., 2017). aethalometer (AE33), was characterized by a campaign average of
In addition to mineral dust, specific mining activities can also contribute 810 ng m− 3. The seasonal trend showed higher values during the dry
to coal dust emissions. Coal dust tracers have been observed in snow season, when the monthly average was about 1.2 μg m− 3 (Suarez et al.,
nearby coal mines in the Arctic (Svalbard Islands, Norway) (Aamaas 2017).
et al., 2011; Khan et al., 2017). Although coal mines are widely spread BC column mass density can be retrieved by satellite image analysis.
around the tropical glacier areas (Karlsen et al., 2006) no regional or Using MODIS daily data, Bolaño-Ortiz et al. (2020) reported a BC col­
national emission inventories are currently available for coal dust. umn mass density varying from 0.5 to 2.5 mg m− 2 over the Sierra
Nevada de Santa Marta (Colombia). Although these data can be useful to
investigate spatial variability of relative column BC, they can hardly be
6. Atmospheric LAP spatial distribution and seasonal variability
compared to surface BC concentrations.
While a few studies investigated OC aerosol in and nearby densely
The origin of most of the LAPs we observe in snow is atmospheric
populated regions of the Andes, observations in remote regions are not
deposition. LAPs concentration in the cryosphere depends on the effi­
reported. Rincón-Riveros et al. (2020) studied for the first time brown
ciency of atmospheric LAP wet and dry deposition, after they have been
carbon in the outskirt of Bogota, in a site representative of regional
transported from the emission region to the high altitude areas. Trans­
background conditions, at an elevation of 3150 m. The authors were
port mechanisms and deposition efficiency are still characterized by
able to link the observed brown carbon to the Amazon fires based on
large uncertainties, especially in remote areas. Data on LAP atmosphere
back-trajectories analysis and active fire satellite observations. Brown
concentrations are essential to understand such transport and deposition
carbon absorption coefficient at 470 nm averaged 2 Mm− 1, with peaks of
processes. The following sections illustrate available data on atmo­
4 Mm− 1 during the biomass burning season (January–March), when it
spheric concentrations of BC, OC, and MDA, and highlight knowledge
accounted for 15% of total aerosol absorption (the sum of BC and brown
gaps in the South American tropical glacier region.
carbon).

6.1. Black carbon and organic carbon 6.2. Mineral dust aerosol

Long-term atmospheric BC and OC measurements in the tropical Together with BC and OC, mineral dust contributes to light absorbing
Andes are limited (Molina et al., 2015) (Table 3). The longest time re­ aerosol burden in the tropical Andes atmosphere. Chauvigné et al.
cord of high-altitude BC concentration is available at the site of Caha­ (2019) inferred dust presence at the Chacaltaya Observatory from
caltaya (Bolivia). The Cahacaltaya Observatory is part of the Global aerosol optical measurements, and reported that mineral dust from the
Atmospheric Watch (GAW) Program network and is located at 5240 m Altiplano impacted the observatory especially during the dry season,
asl, at about 17 km from the urban agglomerate of La Paz (3200 m asl). due to the efficient rain scavenging and reduced soil re-suspension in the
Equivalent BC (eBC) concentration has been measured by a 7-wave­ wet season. Impact of local dust is not excluded, and was mainly
length aethalometer (AE31) and a Multi Angle Absorption Photometer observed when the site was affected by the mix-layer and experienced

Table 3
Atmospheric BC concentrations records available at high altitude observatories in the Andes region.
Site Altitude (m) Country Period Mean concentration
(ng m− 3)

Cahacaltaya Observatory 5240 Bolivia 2012–2015 85


Pico Espejo Atmospheric Research Station 4765 Venezuela March 2007–May 2008 20 (wet season)
50 (dry season)
120 (biomass burning impact)
Huancayo Observatory 3300 Peru June 2015 – December 2015 810

7
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

turbulent atmospheric conditions. assumptions related to emissions, transport, atmospheric processing,


Gaiero et al. (2013) investigated two dust storm episodes in 2009 and and deposition fluxes. In this context, field observations of meteoro­
2010, over the Puna-Altiplano plateau (PAP), integrating field obser­ logical parameters, together with atmosphere and cryosphere compo­
vations with satellite data (from MODIS and CALIOP tools) and back- sition measurements, are essential to understand glacier footprint area
trajectory analysis. The storm episodes were observed during the and to apportion LAPs into different sources. Detailed data on concen­
austral winter and were promoted by intense drought associated with el tration and properties of single LAP type in snow will help to link the
Niño. Vertical profiles from CALIOP showed that dust might travel at an LAP concentration to radiative impact, improving model ability to
altitude up to 8000 m, with a plume thickness of a few kilometers. At the describe spatial and temporal variability of LAPs, snow albedo reduc­
receptor sites, dust storms were characterized by strong wind velocity tion, and the corresponding forcing.
and reduced visibility, and dust aerosol showed a unimodal size distri­ Furthermore, optical satellite data can provide fundamental infor­
bution peaking between 11 and 14 μm, typical of long-ranged trans­ mation about the spatial and temporal variability of LAPs in the tropical
ported dust particles (Gaiero et al., 2013). cryosphere. Advances in satellite products are extremely helpful to
The impact of mineral dust from the Altiplano to the high altitude investigate case studies of long-range transport, for example tracking
Andes is confirmed by long-term columnar aerosol measurements per­ biomass plumes from the Amazon basin to the Andean range. An inte­
formed by the AERONET network. Microphysical properties of aerosol gration of field and satellite observations with modelling will be
particles over La Paz indicate that coarse dust particles, i.e. particles necessary for a comprehensive understanding of the impact of LAPs on
with aerodynamic diameter larger than 1 μm, dominate aerosol size snow and ice in the tropical Andes. Combination of satellite data with
distribution throughout the whole year (Pérez-Ramírez et al., 2017). field observations will increase time coverage and spatial resolution of
In addition to long-ranged dust transport events, local dust transport dynamic data, giving further insights into the frequency and
contamination in the Andes can be triggered by open-pit mining activ­ the characteristics of long-range transport events, and their link with
ities. During the twentieth century the Andes experienced a rapid snow albedo changes. The integration of field data with modelling tools
industrialization, through mining, and nowadays the economy of the will then help to reduce uncertainties in global emission inventories, as
seven Andean countries heavily relies on mining activities (Bebbington well.
et al., 2008). Mineral and coal mining activities are expanding signifi­ LAPs emitted by regional and continental sources, such as open
cantly in South America (Bebbington et al., 2008; Karlsen et al., 2006) burning, can reach the Andes and affect cryosphere energy balance. In
and a recent map of mining fields (Maus et al., 2020) shows that serval addition, local sources can be significant as well, considering that
open mines are located at a few kilometers from tropical glaciers. For several glaciers are located at short distances from densely populated
example, Antamina mine (surface area 39 km2) seats at about 20 km cities, such as Lima, Huaraz, and Quito, and from rural settlements and
from the Huantsan glacier (Cordillera Blanca), Pierina mine (surface industrial activities, including mining. We observed that emission sec­
area 9 km2) and Tomomocho mine (surface area 29 km2) are located at tors of LAPs impacting the cryosphere vary significantly depending on
about 10 km from the Vallunaraju glacier (Cordillera Blanca) and the the glacier footprint region.
Rajuntai glacier (Cordillera Central), respectively, while La Rinconada Furthermore, LAPs atmospheric measurements are vital to under­
mine (25 km2) is sited next to the Auchita glacier (Cordillera Apol­ stand particle transport mechanisms from their emission regions to the
obamba) (https://www.fineprint.global/viewer). Dust and coal resus­ glaciers. Transport mechanism not only determines the LAP burden in
pended during mining operations could potentially reach a glacier in a the cryosphere, but also affects particle mixing state and size distribu­
few hours or even less and, once deposited on its surface, cause positive tion, altering optical properties and deposition rate (Bond and Berg­
radiative forcing and enhanced melting due to the reduction of surface strom, 2006).
albedo (Wittmann et al., 2017). South American tropical glaciers are “water towers” that foster
Unfortunately, in-situ observations are lacking and no quantitative human and animal health, guarantee water availability for agricultural
assessment of the atmospheric and then cryospheric impacts have been and industrial activities, and favor the production of hydroelectric en­
reported in literature. Transport and deposition of mining emissions on ergy. They also embody traditions and spirituality for local populations
tropical glacier deserve analysis in the future, to better quantify the (Veettil and Kamp, 2019). Change in water availability can lead to
anthropogenic impact of local and regional scale activities. conflicts and migration, compromising peace and stability for the entire
region. For all these reasons, and considering the spatial domain of LAP
7. Discussion and concluding remarks sources, which extends behind national borders, a coordinated inter­
national network of atmospheric and cryosphere field observations at
Deposition of atmospheric LAPs contribute to snow albedo reduc­ high mountain sites could be able to provide near-real-time measure­
tion, changing glacier surface microphysical properties and accelerating ments useful to support future climate mitigation and adaptation solu­
snow melting. Currently, several studies are investigating the potential tions, able to reduce anthropogenic pressure on tropical glaciers.
impacts of absorbing particles on snow cover changes, especially in
sensitive climate regions, such as the tropical glaciers (Pörtner et al., Declaration of Competing Interest
2019). However, describing the impact of LAPs on glacier surface energy
balance in the tropical Andes is still challenging due to the limited ob­ The authors declare that they have no known competing financial
servations available in the considered areas; in addition limited are also interests or personal relationships that could have appeared to influence
the knowledge of LAP local and regional sources, transport mechanisms, the work reported in this paper.
and chemical and microphysical properties of absorbing particles,
including their mixing state with other non-absorbing components and Acknowledgements
their interaction with snow and ice (Tuzet et al., 2019). Here we follow
three main LAP classes (black carbon, organic carbon, and mineral dust) This work was performed in the framework of the Cooperation in the
from their emission sources to the tropical glaciers in South America, field of Climate Change Vulnerability, Risk Assessment, Adaptation and
and revise current knowledge gaps. Mitigation between the Italian Ministry for the Environment, Land and
LAP impacts on snow albedo vary significantly, depending on the Sea (IMELS) and the Ministry of the Environment of Peru (MINAM)
chemical nature and microphysical properties of particles in snow. Most (005-2020-MINAMS/IMELS, ALPACA project). We gratefully acknowl­
direct measurements of LAP concentration in snow in South American edge the EDGAR team for emission inventory data and technical sup­
tropical glaciers do not discriminate the single contribution of BC, OC, port. We acknowledge M. Corongiu for the help with the graphical
and mineral dust that are derived from specific models, based on abstract.

8
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

References Dumont, M., et al., 2017. In situ continuous visible and near-infrared spectroscopy of an
alpine snowpack. Cryosphere 11 (3), 1091–1110.
Dumont, M., et al., 2020. Accelerated snow melt in the Russian Caucasus mountains after
Aamaas, B., et al., 2011. Elemental carbon deposition to Svalbard snow from Norwegian
the Saharan dust outbreak in March 2018. J. Geophys. Res. Earth Surf. 125 (9)
settlements and long-range transport. Tellus Ser. B Chem. Phys. Meteorol. 63 (3),
e2020JF005641.
340–351.
Dussaillant, I., et al., 2019. Two decades of glacier mass loss along the Andes. Nat.
Ansmann, A., et al., 2009. Vertical profiling of convective dust plumes in southern
Geosci. 12 (10), 802–808.
Morocco during SAMUM. Tellus Ser. B Chem. Phys. Meteorol. 61 (1), 340–353.
Federer, U., Kaufmann, P.R., Hutterli, M.A., Schüpbach, S., Stocker, T.F., 2008.
Artaxo, P., Raga, G., 2016. Integrated Assessment of Short-Lived Climate Pollutants and
Continuous flow analysis of total organic carbon in polar ice cores. Environ. Sci.
Latin America and the Caribbean.
Technol. 42 (21), 8039–8043.
Baars, H., et al., 2012. Aerosol profiling with lidar in the Amazon Basin during the wet
Flanner, M.G., Zender, C.S., 2005. Snowpack radiative heating: influence on Tibetan
and dry season. J. Geophys. Res.-Atmos. 117 (D21).
Plateau climate. Geophys. Res. Lett. 32 (6).
Barraza, F., et al., 2021. Major atmospheric particulate matter sources for glaciers in
Forsström, S., Ström, J., Pedersen, C.A., Isaksson, E., Gerland, S., 2009. Elemental carbon
Coquimbo Region, Chile. Environ. Sci. Pollut. Res. 28 (27), 36817–36827.
distribution in Svalbard snow. J. Geophys. Res.-Atmos. 114 (D19).
Bebbington, A., et al., 2008. Mining and social movements: struggles over livelihood and
Francou, B., Vuille, M., Wagnon, P., Mendoza, J., Sicart, J.-E., 2003. Tropical climate
rural territorial development in the Andes. World Dev. 36 (12), 2888–2905.
change recorded by a glacier in the Central Andes during the last decades of the
Beres, N., Sengupta, D., Samburova, V., Khlystov, A., Moosmuller, H., 2020. Deposition
twentieth century: Chacaltaya, Bolivia, 16◦ S. J. Geophys. Res.-Atmos. 108 (D5).
of brown carbon onto snow: changes in snow optical and radiative properties.
Fugazza, D., et al., 2019. New evidence of glacier darkening in the Ortles-Cevedale group
Atmos. Chem. Phys. 20 (10), 6095–6114.
from Landsat observations. Glob. Planet. Chang. 178, 35–45.
Bolaño-Ortiz, T.R., Diaz-Gutiérrez, V.L., Camargo-Caicedo, Y., 2020. ENSO and light-
Gaiero, D.M., et al., 2013. Ground/satellite observations and atmospheric modeling of
absorbing impurities and their impact on snow Albedo in the Sierra Nevada de Santa
dust storms originating in the high Puna-Altiplano deserts (South America):
Marta, Colombia. Geosciences 10 (11).
implications for the interpretation of paleo-climatic archives. J. Geophys. Res.-
Bonasoni, P., et al., 2010. Atmospheric Brown Clouds in the Himalayas: first two years of
Atmos. 118 (9), 3817–3831.
continuous observations at the Nepal Climate Observatory-Pyramid (5079 m).
Giglio, L., Randerson, J.T., van der Werf, G.R., 2013. Analysis of daily, monthly, and
Atmos. Chem. Phys. 10 (15), 7515–7531.
annual burned area using the fourth-generation global fire emissions database
Bond, T.C., Bergstrom, R.W., 2006. Light absorption by carbonaceous particles: an
(GFED4). J. Geophys. Res. Biogeosci. 118 (1), 317–328.
investigative review. Aerosol Sci. Technol. 40 (1), 27–67.
Gilardoni, S., et al., 2016. Direct observation of aqueous secondary organic aerosol from
Bond, T., et al., 2013. Bounding the role of black carbon in the climate system: A
biomass-burning emissions. Proc. Natl. Acad. Sci. U. S. A. 113 (36), 10013–10018.
scientific assessment. J. Geophys. Res. Atmos. 118 (11), 5380–5552.
Gramsch, E., et al., 2020. Black carbon transport between Santiago de Chile and glaciers
Bourgeois, Q., Ekman, A.M.L., Krejci, R., 2015. Aerosol transport over the Andes from the
in the Andes mountains. Atmos. Environ. 232 (117546).
Amazon Basin to the remote Pacific Ocean: a multiyear CALIOP assessment.
Grenfell, T.C., Doherty, S.J., Clarke, A.D., Warren, S.G., 2011. Light absorption from
J. Geophys. Res.-Atmos. 120 (16), 8411–8425.
particulate impurities in snow and ice determined by spectrophotometric analysis of
Brandt, R.E., Warren, S.G., Clarke, A.D., 2011. A controlled snowmaking experiment
filters. Appl. Opt. 50 (14), 2037–2048.
testing the relation between black carbon content and reduction of snow albedo.
Hadley, O.L., Kirchstetter, T.W., 2012. Black-carbon reduction of snow albedo. Nat. Clim.
J. Geophys. Res.-Atmos. 116 (D8).
Chang. 2 (6), 437–440.
Brenning, A., Azócar Sandoval, G., 2010. Mining and rock glaciers: environmental
Hamburger, T., et al., 2013. Long-term in situ observations of biomass burning aerosol at
impacts, political and legal situation, and future perspectives, pp. 143–158.
a high altitude station in Venezuela – sources, impacts and interannual variability.
Cape, J., Coyle, M., Dumitrean, P., 2012. The atmospheric lifetime of black carbon.
Atmos. Chem. Phys. 13 (19), 9837–9853.
Atmos. Environ. 59, 256–263.
He, C., et al., 2018. Black carbon-induced snow albedo reduction over the Tibetan
Caponi, L., et al., 2017. Spectral- and size-resolved mass absorption efficiency of mineral
Plateau: uncertainties from snow grain shape and aerosol–snow mixing state based
dust aerosols in the shortwave spectrum: a simulation chamber study. Atmos. Chem.
on an updated SNICAR model. Atmos. Chem. Phys. 18 (15), 11507–11527.
Phys. 17 (11), 7175–7191.
Hess, G.D., Spillane, K.T., 1990. Characteristics of dust devils in Australia. J. Appl.
Cardozo, F.D., Pereira, G., Shimabukuro, Y.E., Moraes, E.C., 2014. Analysis and
Meteorol. Climatol. 29 (6), 498–507.
assessment of the spatial and temporal distribution of burned areas in the Amazon
Hoham, R.A.-O., Remias, D.A.-O.X., 2020. Snow and glacial algae: a review(1). J. Phycol.
Forest. Remote Sens. 6 (9).
56 (2), 264–282.
Castro Videla, F., Barnaba, F., Angelini, F., Cremades, P., Gobbi, G.P., 2013. The relative
Huneeus, N., et al., 2020. Evaluation of anthropogenic air pollutant emission inventories
role of Amazonian and non-Amazonian fires in building up the aerosol optical depth
for South America at national and city scale. Atmos. Environ. 235, 117606.
in South America: A five year study (2005–2009). Atmos. Res. 122, 298–309.
Jemmett-Smith, B.C., Marsham, J.H., Knippertz, P., Gilkeson, C.A., 2015. Quantifying
Chauvigné, A., et al., 2019. Biomass burning and urban emission impacts in the Andes
global dust devil occurrence from meteorological analyses. Geophys. Res. Lett. 42
Cordillera region based on in situ measurements from the Chacaltaya observatory,
(4), 1275–1282.
Bolivia (5240 m a.s.l.). Atmos. Chem. Phys. 19 (23), 14805–14824.
Karlsen, A.W., Tewalt, S.J., Bragg, L.J., Finkelman, R.B., 2006. The World Coal Quality
Chen, Y., et al., 2009. Aircraft study of mountain chimney effect of Beijing, China.
Inventory: South America.
J. Geophys. Res.-Atmos. 114 (D8).
Kaser, G., 1999. A review of the modern fluctuations of tropical glaciers. Glob. Planet.
Chen, Y., et al., 2013. Long-term trends and interannual variability of forest, savanna and
Chang. 22 (1), 93–103.
agricultural fires in South America. Carbon Manag. 4 (6), 617–638.
Khan, A.L., et al., 2017. Impacts of coal dust from an active mine on the spectral
Chuvieco, E., et al., 2008. Global burned-land estimation in Latin America using MODIS
reflectance of Arctic surface snow in Svalbard, Norway. J. Geophys. Res.-Atmos. 122
composite data. Ecol. Appl. 18 (1), 64–79.
(3), 1767–1778.
Clarke, A.D., 1982. Integrating sandwich: a new method of measurement of the light
Kirchstetter, T., Novakov, T., Hobbs, P., 2004. Evidence that the spectral dependence of
absorption coefficient for atmospheric particles. Appl. Opt. 21 (16), 3011–3020.
light absorption by aerosols is affected by organic carbon. J. Geophys. Res. Atmos.
Cochrane, M.A., 2002. Spreading like Wildfire, Tropical Forest Fires in Latin America and
109 (D21).
the Caribbean.
Klimont, Z., et al., 2017. Global anthropogenic emissions of particulate matter including
Cook, J.M., et al., 2020. Glacier algae accelerate melt rates on the South-Western
black carbon. Atmos. Chem. Phys. 17 (14), 8681–8723.
Greenland Ice Sheet. Cryosphere 14 (1), 309–330.
Kok, J.F., et al., 2017. Smaller desert dust cooling effect estimated from analysis of dust
Cordero, R.R., Sepúlveda, E., Feron, S., et al., 2022a. Black carbon footprint of human
size and abundance. Nat. Geosci. 10 (4), 274–278.
presence in Antarctica. Nat. Commun. 13, 984. https://doi.org/10.1038/s41467-
Kok, J.F., et al., 2021. Improved representation of the global dust cycle using
022-28560-w.
observational constraints on dust properties and abundance. Atmos. Chem. Phys. 21
Cordero, R.R., et al., 2022b. Black carbon in the Southern Andean snowpack. Environ.
(10), 8127–8167.
Res. Lett. 17 (4), 044042 https://doi.org/10.1088/1748-9326/AC5DF0.
Kokhanovsky, A., et al., 2019. Retrieval of snow properties from the Sentinel-3 ocean and
Crippa, M., et al., 2019. Fossil CO2 and GHG Emissions of all World Countries.
land colour instrument. Remote Sens. 11 (19).
Di Mauro, B., 2020. A darker cryosphere in a warming world. Nat. Clim. Chang. 10 (11),
Kokhanovsky, A., Di Mauro, B., Garzonio, R., Colombo, R., 2021. Retrieval of dust
979–980.
properties from spectral snow reflectance measurements. Front. Environ. Sci. 9, 42.
Di Mauro, B., et al., 2015. Mineral dust impact on snow radiative properties in the
Krecl, P., Targino, A.C., Ketzel, M., Cipoli, Y.A., Charres, I., 2019. Potential to reduce the
European Alps combining ground, UAV, and satellite observations. J. Geophys. Res.-
concentrations of short-lived climate pollutants in traffic environments: A case study
Atmos. 120 (12), 6080–6097.
in a medium-sized city in Brazil. Transp. Res. Part D: Transp. Environ. 69, 51–65.
Di Mauro, B., et al., 2017. Impact of impurities and cryoconite on the optical properties
Laskin, A., Laskin, J., Nizkorodov, S.A., 2015. Chemistry of atmospheric brown carbon.
of the Morteratsch Glacier (Swiss Alps). Cryosphere 11 (6), 2393–2409.
Chem. Rev. 115 (10), 4335–4382.
Di Mauro, B., et al., 2019. Saharan dust events in the European Alps: role in snowmelt
Lavanchy, V.M.H., Gäggeler, H.W., Schotterer, U., Schwikowski, M., Baltensperger, U.,
and geochemical characterization. Cryosphere 13 (4), 1147–1165.
1999. Historical record of carbonaceous particle concentrations from a European
Di Mauro, B., et al., 2020. Glacier algae foster ice-albedo feedback in the European Alps.
high-alpine glacier (Colle Gnifetti, Switzerland). J. Geophys. Res.-Atmos. 104 (D17),
Sci. Rep. 10 (1), 4739.
21227–21236.
Di Mauro, B., et al., 2021. Light-absorbing particles in snow and ice: a brief journey
Lim, S., et al., 2014. Refractory black carbon mass concentrations in snow and ice:
across latitudes. In: Kokhanovsky, A. (Ed.), Springer Series in Light Scattering,
method evaluation and inter-comparison with elemental carbon measurement.
Volume 7: Light Absorption and Scattering in Turbid Media. Springer International
Atmos. Meas. Tech. 7 (10), 3307–3324.
Publishing, Cham, pp. 1–29.
Lim, S., et al., 2017. Black carbon variability since preindustrial times in the eastern part
Dumont, M., et al., 2014. Contribution of light-absorbing impurities in snow to
of Europe reconstructed from Mt. Elbrus, Caucasus, ice cores. Atmos. Chem. Phys. 17
Greenland’s darkening since 2009. Nat. Geosci. 7 (7), 509–512.
(5), 3489–3505.

9
S. Gilardoni et al. Global and Planetary Change 213 (2022) 103837

Lindau, F.G.L., et al., 2021. Giant dust particles at Nevado Illimani: a proxy of Sarangi, C., et al., 2020. Dust dominates high-altitude snow darkening and melt over
summertime deep convection over the Bolivian Altiplano. Cryosphere 15 (3), high-mountain Asia. Nat. Clim. Chang. 10 (11), 1045–1051.
1383–1397. Schmitt, C.G., et al., 2015. Measurements of light-absorbing particles on the glaciers in
López-Moreno, J.I., et al., 2014. Recent glacier retreat and climate trends in Cordillera the Cordillera Blanca, Peru. The Cryosphere 9 (1), 331–340.
Huaytapallana, Peru. Glob. Planet. Chang. 112, 1–11. Schwarz, J.P., et al., 2012. Assessing Single Particle Soot Photometer and Integrating
Magalhães, N.D., Evangelista, H., Condom, T., Rabatel, A., Ginot, P., 2019. Amazonian Sphere/Integrating Sandwich Spectrophotometer measurement techniques for
Biomass burning Enhances Tropical Andean Glaciers Melting. Sci. Rep. 9 (1), 16914. quantifying black carbon concentration in snow. Atmos. Meas. Tech. 5 (11),
Mahowald, N., Albani, S., Kok, J.F., Engelstaeder, S., Scanza, R., Ward, D.S., Flanner, M. 2581–2592.
G., 2013. The size distribution of desert dust aerosols and its impact on the Earth Seehaus, T., et al., 2019. Changes of the tropical glaciers throughout Peru between 2000
system. Aeolian Res. https://doi.org/10.1016/j.aeolia.2013.09.002. and 2016 – mass balance and area fluctuations. Cryosphere 13 (10), 2537–2556.
Maus, V., et al., 2020. A global-scale data set of mining areas. Scientific Data 7 (1), 289. Shaw, T.E., et al., 2021. Glacier albedo reduction and drought effects in the extratropical
Mendez-Espinosa, J.F., Belalcazar, L.C., Morales Betancourt, R., 2019. Regional air Andes, 1986–2020. J. Glaciol. 67 (261), 158–169.
quality impact of northern South America biomass burning emissions. Atmos. Shi, T., et al., 2021. Enhanced light absorption and reduced snow albedo due to
Environ. 203, 131–140. internally mixed mineral dust in grains of snow. Atmos. Chem. Phys. 21 (8),
Molina, L.T., et al., 2015. Pollution and its impacts on the South American cryosphere. 6035–6051.
Earth’s Future 3 (12), 345–369. Skiles, S.M., Flanner, M., Cook, J.M., Dumont, M., Painter, T.H., 2018. Radiative forcing
Naegeli, K., Huss, M., Hoelzle, M., 2019. Change detection of bare-ice albedo in the Swiss by light-absorbing particles in snow. Nat. Clim. Chang. 8 (11), 964–971.
Alps. Cryosphere 13 (1), 397–412. Soto Carrion, C., et al., 2021. Quantitative estimation of black carbon in the Glacier
Ogren, J.A., Charlson, R.J., 1983. Elemental carbon in the atmosphere: cycle and Ampay-Apurimac. J Sustain. Develop. Energy Water Environ. Syst. 9 (1).
lifetime. Tellus Ser. B Chem. Phys. Meteorol. 35 (4), 241–254. Suarez, L.C.T., Helmig, D., Hueber, J., 2017. Medición y análisis del aerosol de carbono
Osmont, D., Sigl, M., Eichler, A., Jenk, T.M., Schwikowski, M., 2019. A Holocene black negro en el observatorio de Huancayo, Peru. Revista Boliviana de Física 30 (30),
carbon ice-core record of biomass burning in the Amazon Basin from Illimani, 9–17.
Bolivia. Clim. Past 15 (2), 579–592. Takeuchi, N., Kohshima, S., Seko, K., 2001. Structure, formation, and darkening process
Painter, T.H., Bryant, A.C., Skiles, S.M., 2012. Radiative forcing by light absorbing of Albedo-reducing material (cryoconite) on a Himalayan Glacier: a Granular Algal
impurities in snow from MODIS surface reflectance data. Geophys. Res. Lett. 39 (17). Mat growing on the Glacier. Arct. Antarct. Alp. Res. 33 (2), 115–122.
Painter, T.H., Seidel, F.C., Bryant, A.C., McKenzie Skiles, S., Rittger, K., 2013. Imaging Takeuchi, N., Dial, R., Kohshima, S., Segawa, T., Uetake, J., 2006. Spatial distribution
spectroscopy of albedo and radiative forcing by light-absorbing impurities in and abundance of red snow algae on the Harding Icefield, Alaska derived from a
mountain snow. J. Geophys. Res.-Atmos. 118 (17), 9511–9523. satellite image. Geophys. Res. Lett. 33 (21).
Palut, M.L., Palutikof, J.P., Canziani, O., 2007. Contribution of Working Group II to the Thompson, L.G., Mosley-Thompson, E., Dansgaard, W., Grootes, P.M., 1986. The Little
Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Ice Age as recorded in the stratigraphy of the Tropical Quelccaya Ice Cap. Science
Pearson, P.B., Gittelson, S., Kinney, A., McCarty, S., Stevenson, G., Albertengo, J., 2017. 234 (4774), 361.
Fire in the fields: Moving beyond the damage of open agricultural burning on Thompson, L.G., et al., 1995. Late Glacial Stage and Holocene Tropical Ice core records
communities, soil, and the cryosphere. from Huascarán, Peru. Science 269 (5220), 46.
Peltoniemi, J.I., et al., 2015. Soot on Snow experiment: bidirectional reflectance factor Thompson, L.G., et al., 2003. Tropical Glacier and Ice Core evidence of climate change on
measurements of contaminated snow. Cryosphere 9 (6), 2323–2337. annual to millennial time scales. In: Diaz, H.F. (Ed.), Climate Variability and Change
Pérez-Ramírez, D., et al., 2017. Multi year aerosol characterization in the tropical Andes in High Elevation Regions: Past, Present & Future. Springer, Netherlands, Dordrecht,
and in adjacent Amazonia using AERONET measurements. Atmos. Environ. 166, pp. 137–155.
412–432. Torres, C., Suarez, L., Schmitt, C., Estevan, R., Helmig, D., 2018. Measurement of light
Pörtner, H.-O., et al., 2019. IPCC 2019: Special Report on the Ocean and Cryosphere in a absorbing particles in the snow of the Huaytapallana glacier in the Central Andes of
Changing Climate. Peru and their effect on albedo and radiative forcing Optica Pura y. Aplicada 51 (4),
Prospero, J.M., Ginoux, P., Torres, O., Nicholson, S.E., Gill, T.E., 2002. Environmental 1–14.
characterization of global sources of atmospheric soil dust identified with the Tuzet, F., et al., 2019. Influence of light-absorbing particles on snow spectral irradiance
NIMBUS 7 Total Ozone Mapping Spectrometer (TOMS) absorbing aerosol product. profiles. Cryosphere 13 (8), 2169–2187.
Rev. Geophys. 40 (1), 2-1-2-31. Tuzet, F., et al., 2020. Quantification of the radiative impact of light-absorbing particles
Pu, W., et al., 2021. Enhancement of snow albedo reduction and radiative forcing due to during two contrasted snow seasons at Col du Lautaret (2058 m a.s.l., French Alps).
coated black carbon in snow. Cryosphere 15 (5), 2255–2272. Cryosphere 14 (12), 4553–4579.
Reis, R.S.D., et al., 2021. Relationships between Andean Glacier Ice-Core Dust Records van der Werf, G.R., et al., 2017. Global fire emissions estimates during 1997–2016. Earth
and Amazon Basin Riverine Sediments. Cryosphere Discuss. 2021, 1–15. Syst. Sci. Data 9 (2), 697–720.
RGI, 2017. Randolph Glacier Inventory – A Dataset of Global Glacier Outlines: Version Veettil, B.K., Kamp, U., 2019. Global disappearance of tropical mountain glaciers:
6.0: Technical Report, Global Land Ice Measurements from Space, Colorado, USA. observations, causes, and challenges. Geosciences 9 (5).
Digital Media. In: NASA (Ed.). Vuille, M., et al., 2018. Rapid decline of snow and ice in the tropical Andes – impacts,
Rincón-Riveros, J.M., et al., 2020. Long-term brown carbon and smoke tracer uncertainties and challenges ahead. Earth Sci. Rev. 176, 195–213.
observations in Bogotá, Colombia: association with medium-range transport of Wallis, L.K., 2016. Spatial Variability of Snow Chemistry of High Altitude Glaciers in the
biomass burning plumes. Atmos. Chem. Phys. 20 (12), 7459–7472. Peruvian Andes.
Rolando, J.L., et al., 2017. Key ecosystem services and ecological intensification of Wang, M., et al., 2012. The Influence of Dust on Quantitative Measurements of Black
agriculture in the tropical high-Andean Puna as affected by land-use and climate Carbon in Ice and Snow when using a thermal Optical Method. Aerosol Sci. Technol.
changes. Agric. Ecosyst. Environ. 236, 221–233. 46 (1), 60–69.
Romshoo, B., et al., 2021. Optical properties of coated black carbon aggregates: Wang, X., Doherty, S.J., Huang, J., 2013. Black carbon and other light-absorbing
numerical simulations, radiative forcing estimates, and size-resolved impurities in snow across Northern China. J. Geophys. Res.-Atmos. 118 (3),
parameterization scheme. Atmos. Chem. Phys. 21 (17), 12989–13010. 1471–1492.
Rozwalak, P., et al., 2022. Cryoconite – from minerals and organic matter to Warren, S.G., 2013. Can black carbon in snow be detected by remote sensing?
bioengineered sediments on glacier’s surfaces. Sci. Total Environ. 807, 150874. J. Geophys. Res.-Atmos. 118 (2), 779–786.
Saleh, R., Cheng, Z., Atwi, K., 2018. The Brown–Black Continuum of Light-Absorbing Warren, S.G., 2019. Optical properties of ice and snow. Phil. Trans. R. Soc. A 377 (2146).
Combustion Aerosols. Environ. Sci. Technol. Lett. 5 (8), 508–513. Warren, S.G., Wiscombe, W.J., 1980. A model for the spectral albedo of snow 2. Snow
Samset, B.H., et al., 2014. Modelled black carbon radiative forcing and atmospheric containing atmospheric aerosols. J. Atmos. Sci. 37 (12), 2734–2745.
lifetime in AeroCom phase II constrained by aircraft observations. Atmos. Chem. Wiedensohler, A., et al., 2018. Black carbon emission and transport mechanisms to the
Phys. 14 (22), 12465–12477. free troposphere at the La Paz/El Alto (Bolivia) metropolitan area based on the Day
Sanchez Rodriguez, W.E., 2016. Informe de risultando: Estimacion de la contribucion del of Census (2012). Atmos. Environ. 194, 158–169.
carbono negro a la fusion de nieve de los glaciares Yanapaccha y Shallap, para el Wittmann, M., et al., 2017. Impact of dust deposition on the albedo of Vatnajökull ice
periodo entro Octobre 2015 hasta Agosto 2016. cap, Iceland. Cryosphere 11 (2), 741–754.
Sánchez Rodríguez, W., Schmitt, C., 2018. Partículas Absorbentes de Luz durante El Niño
y El Niño Costero en los Glaciares de la Cordillera Blanca, Perú. Rev. Glaciares y
Ecosistemas Montaña 4, 9–22. https://doi.org/10.36580/rgem.i4.9-22.

10

You might also like