Next Article in Journal
Bacterial Microbiota of Asthmatic Children and Preschool Wheezers’ Airways—What Do We Know?
Previous Article in Journal
Identification and Construction of Strong Promoters in Yarrowia lipolytica Suitable for Glycerol-Based Bioprocesses
Previous Article in Special Issue
Metagenomic Insights into Microbial Community Structure, Function, and Salt Adaptation in Saline Soils of Arid Land, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Glacial Water: A Dynamic Microbial Medium

1
Centre for Microbial Ecology and Genomics, Department of Biochemistry, Genetics and Microbiology, University of Pretoria, Pretoria 0002, South Africa
2
Rhizosphere Processes Group, Swiss Federal Research Institute WSL, 8903 Birmensdorf, Switzerland
3
Departments of Geology and Geography, Portland State University, Portland, OR 97212, USA
4
Department of Environmental Science, iClimate, Aarhus University, DK-4000 Roskilde, Denmark
*
Authors to whom correspondence should be addressed.
Microorganisms 2023, 11(5), 1153; https://doi.org/10.3390/microorganisms11051153
Submission received: 22 March 2023 / Revised: 15 April 2023 / Accepted: 19 April 2023 / Published: 28 April 2023
(This article belongs to the Special Issue Extremophilic Microorganisms and Their Communities)

Abstract

:
Microbial communities and nutrient dynamics in glaciers and ice sheets continuously change as the hydrological conditions within and on the ice change. Glaciers and ice sheets can be considered bioreactors as microbiomes transform nutrients that enter these icy systems and alter the meltwater chemistry. Global warming is increasing meltwater discharge, affecting nutrient and cell export, and altering proglacial systems. In this review, we integrate the current understanding of glacial hydrology, microbial activity, and nutrient and carbon dynamics to highlight their interdependence and variability on daily and seasonal time scales, as well as their impact on proglacial environments.

1. Introduction

Although the movement of glaciers and ice sheets profoundly affects the landscape, the most immediate protagonist of glacial influences is glacial meltwater [1]. Meltwater accelerates the movement of glaciers, evacuates glacially eroded sediment downstream, contributes to alpine hydrology and riparian zones, regional runoff, and ultimately sea level rise [2,3]. From a microbial perspective, meltwater transfers nutrients and microbes across and within the glacial mass [4,5] and eventually to proglacial ecosystems [6,7,8]. Flow systems within glaciers are complex and poorly known, varying spatially and temporally and differing between glaciers and ice sheets [9,10]. This spatial and temporal variability leads to continuously changing biochemical activities, nutrient cycling, and microbial community dispersion [4,11,12]. Microbial communities adapted to these icy environments transform meltwater chemistry, including both organic and inorganic compounds [11,13,14], and enrich glacial exiting waters with bioavailable molecules [13,15,16], making glaciers and ice sheets low-temperature bioreactors. Glacial microorganisms transform the geochemistry of glacial meltwater through a variety of metabolisms. For example, the microbial communities on the glacial surface are dominated by photoautotrophs and nitrogen fixers, whereas microbial communities in the subglacial waters can be dominated by iron reducers, sulfide oxidizers, sulfate reducers and methanogens [17,18]. Which metabolisms are most active at a specific time and in a specific glacial compartment highly depends on the state of the ice body [19,20]. Such data clearly indicate that glaciers and ice sheets are not inert masses, placing them as integral components of global biogeochemical cycling.
Ice sheets and glaciers occupy 10% of the Earth’s surface and represent about 68% of Earth’s freshwater [21]. Their retreat, and the consequent increase in released meltwater, will have global impacts [22,23,24,25]. Meltwater discharge from glaciers and ice sheets is enormous; the Greenland Ice Sheet alone is expected to release 357 ± 58 Gt water y−1 [26]. Meltwater discharges to a wide range of different proglacial systems, from glacial forefields to glacial lakes or oceans, impacting these environments by contributing microorganisms, nutrients, and sediment [6,27,28]. Cameron et al. (2017) estimated the export of 6.9 × 1022 cells y−1 from Greenland alone [6], and about 1.1 × 1022 cells y−1 are expected from its surface under a medium emission scenario in the next 80 years [29]. This export of nutrients and microorganisms to proglacial systems is, therefore, thought to represent a major contribution to global biogeochemical cycling in proglacial areas [27,30,31,32,33].
Although several recent reviews of the microbiomics of glacial systems are available [4,34,35], this review provides an up-to-date holistic view of the interplay among glacial hydrology, microbial activity, and geochemistry. In fact, through an extensive literature review, we compile the current state of knowledge on glacial hydrology, microbial activity, and geochemical cycling; we explore their interdependence and how differences in glacial meltwater flow systems influence nutrient distributions and microbial processes. In this review, we also explore how glacial processes and patterns influence a variety of proglacial environments, outlining the possible global impacts of increases in meltwater discharge due to climate warming.

2. Glacial Hydrology

Glacial hydrological systems consist of interconnected micro- and macro-pathways, their size and location being influenced by the glacial thermal regime (Box 1) and water availability [36], which, in turn, are related to climate (e.g., latitude, altitude, wind, precipitation, temperature, and radiation) [37,38].
Box 1. Ice temperature and permeability.
 Warm atmospheric temperatures and high precipitation rates lead to temperate ice (i.e., ice temperature above the ice melting temperature), where snow deposition can isolate surface ice from constant subzero temperatures. On the contrary, surface cold ice (i.e., ice temperature below the ice melting temperature) is the result of permanent cold temperatures [39]. Whereas surface glacial ice is directly influenced by the atmospheric temperature, the temperature of the subglacial ice is influenced only indirectly: i.e., the temperature of subglacial ice is conditioned by geothermal heat fluxes and by the amount of heat that is generated by glacial sliding, glacial water flow, and ice deformation, which are deeply conditioned by the supraglacial water dynamics [40]. Ice temperature profiles are used to classify glaciers. Temperate glaciers are those entirely composed of ice with a temperature above the ice melting temperature (i.e., temperate ice), cold glaciers are composed of ice with a temperature below the ice melting temperature (i.e., cold ice), and polythermal glaciers have a more complex thermal structure characterized by different ice temperatures [39]. Polythermal glaciers often consist of temperate ice throughout the glacier body, except the first meters of the glacial surface (typically in the ablation zone), where the ice is directly in contact with the cold atmosphere and maintains a cold temperature across its surface ice. Examples of temperate glaciers can be found between alpine glaciers (e.g., Arolla glacier), polythermal glaciers between Arctic glaciers (e.g., Midtre Lovénbreen), and cold glaciers between polar glaciers (e.g., Larsbreen) [39,41].
 Ice temperature influences glacier and ice sheet functioning, where small changes in ice temperature can deeply change the glacier dynamics [42,43]. For instance, ice temperature determines how permeable the ice is [1]. Primary permeability is the movement of water at a small scale through ice veins and only occurs in temperate ice. The volumetric water content in temperate ice has been estimated to reach 9% of the ice volume [44]. On the contrary, cold ice is not permeable and therefore does not present a liquid matrix between ice crystals [1].
 Contrary to temperate glaciers, water only flows through crevasses, moulins, englacial channels, and conduits (secondary permeability) in cold glaciers [45,46]. Whereas an active englacial system (i.e., with flowing water) has been observed in a wide variety of glacial thermal regimes [36,47,48], temperate glaciers have the most developed water discharging system compared to the other glacier types due to the higher permeability and malleability of temperate ice [1]. Differences in permeability, therefore, lead to the development of different water discharge networks: e.g., because of the less developed englacial channel network in cold glaciers, water is discharged mainly through the supraglacial environment in most systems during the summer [49], where supraglacial features are more developed compared to those in the other ice systems [50].

2.1. Seasonal Variation

The presence of water in glaciers is highly seasonal, as the generation of glacial water depends on the energy available to melt snow and ice. Water input also comes from rain events. For most glaciers in the temperate and subpolar regions, mass accumulates during the accumulation season (e.g., winter), which is characterized by colder temperatures and snow accumulation [1]. By comparison, mass loss, including melt, occurs during the ablation season (i.e., summer), which is characterized by warmer air temperatures and greater solar radiation. During this season, the lower elevation ablation zone of a glacier is gradually exposed as the winter snow cover melts, revealing an ice surface. In the accumulation zone, located at a higher altitude on the glacier, only some of the accumulated snow melts, therefore accumulating and compacting it into firn (transition stage between snow and ice) and ultimately into ice. The amount of snow gained over winter minus the amount of snow and ice lost defines the glacial ice balance [1].

2.1.1. Ablation Season

In spring, the snowpack temperature increases through heat conduction from warm air and via refreezing of meltwater that percolates into the cold snow (Figure 1). Once the snowpack becomes isothermal at 0 °C, snow melt percolates into the entire seasonal snowpack. In the accumulation zone, the meltwater enters the firn, forming a firn aquifer that drains to nearby crevasses [51,52,53] (Figure 2). In the ablation zone, the water accumulates on the ice surface and also drains into nearby crevasses [54]. Once the snow melts off the ice in the ablation zone, a shallow weathering crust of porous ice forms, caused by penetration of solar radiation and enhanced by the flow of meltwater and potentially warm air [55]. The resulting weathering crust represents a photic zone a few centimeters to a few meters deep [55,56,57,58]. Both the weathering crust and the firn aquifers are active parts of the glacial hydrological system and contribute to accumulating, distributing, and regulating water discharge to the englacial system [36,51,52,53,55,57,59,60]. Cryoconite holes can also form on the ice surface as a result of the deposition of sediment and biological materials on the glacial surface. These patches of material melt into the ice faster than the surrounding ice, driven by increased heat absorption by the dark material [61,62,63]. Under the right circumstances, these holes can form glacial ponds and lakes (cryo-lakes) [39,50]. Thus the surface streams, weathering crust, cryoconite holes, and cryo-lakes form the complex surface hydrology of glaciers [64]. This supraglacial hydrology influences the drainage pattern and flux of meltwater into the glacial interior (i.e., englacial system) and bed and to proglacial systems [64].
The main pathways from the glacier surface and firn aquifers to the glacial interior are moulins and crevasses. Moulins constitute a direct connection to an englacial system of conduits, whereas crevasses drain water via a network of fractures thought to connect to englacial conduits [69,70,71]. The englacial system of conduits reaches the bottom of the glacier, where it flows along the ice-substrate interface and, if the substrate is sediment, through the sediment as well.
Most subglacial water originates from the glacier surface, although a small flux may originate englacially and subglacially due to frictional dissipation of heat generated by flowing water, ice deformation, or from geothermal heat at the bottom of the glacier [72,73,74]. The Antarctic ice sheet is a major exception, where the main source of subglacial water is basal melt [75]. Generally, subglacial water flows in two main types of systems: slow flow within a distributed network of linked cavities or confined groundwater flow within a layer of subglacial rock debris (i.e., subglacial till), or in a quick flow system composed of a network of subglacial channels [76,77]. Distributed systems are thought to be water-filled all year round, whereas channels are water-filled only at high meltwater discharge [11,36,78]. Under ice sheets, subglacial lakes may occur as stable subglacial water bodies that can be isolated or hydrologically connected to a subglacial flow system [79,80]. Regardless of its path and residence time within the glacier, water will eventually escape to proglacial systems via streams or groundwater flow.

2.1.2. Seasonal Evolution

During summers with plenty of meltwater, surface streams flow over the ablation zone, and the weathering crust, including cryoconite holes, is full of water [1] (Figure 1). Surface water that accumulates in weathering crust and cryoconite holes drains into crevasses and moulins, and water in surface streams commonly pours into moulins. Compared to supraglacial systems, water accumulated in firn aquifers is less connected to the rest of the glacial system, and it occasionally drains into nearby crevasses [54]. Englacial and subglacial channels (quick flow) are thought to be filled when water discharge reaches its daily maximum and are less full early the following morning when the air temperature is the coolest, and the sun is just beginning to warm the snow and ice [81,82]. The subglacial distributed systems are thought to be continually water filled [11]. In addition to daily cycles of glacial runoff, longer variations occur with the occurrence of warm and cool weather systems [81,82]. Episodic floods may also occur due to the sudden release of ponded water within the glacier or to the sudden drainage of glacier-dammed lakes [83,84,85].
At summer’s end and the beginning of winter’s accumulation season, much less water is present in the glacial system due to the low air temperatures and reduced solar radiation, which largely eliminates surface melt, and while precipitation falls mostly as snow [1]. Some water may remain unfrozen over winter in subsurface weathering crust and firn aquifer due to the insulating effects of a deep snow cover [63]. Over winter, the englacial and subglacial channels collapse due to ice pressure and the absence of a counteracting pressure from flowing water [36,86,87]. However, due to a decrease in water flow, water pockets may form within the englacial or subglacial system [88,89]. Glacial streamflow also often continues through the winter, albeit at very low discharge. This may be due to the drainage of water resident in the glacier as well as meltwater produced by geothermal heat and by deformation within the ice [54,81].

2.2. Hydrological Residence Times

The residence time of water in various parts of the glacial hydrological system varies (Table 1) and influences in situ geochemical processes [90]. The extent to which microbial communities can develop and contribute to geochemical modification depends both on microbial doubling times [31,58] and on how long they can reside in a specific glacial compartment (i.e., water residence time). Furthermore, different residence times create a variety of conditions for microbial metabolism [5,12]. For example, the lower the residence time, the better ventilated a system is with the consequent creation of oxic waters. Oxic waters will promote aerobic or facultative anaerobic microbial metabolism, such as nitrification and iron oxidation, whereas anoxic waters will promote the growth of anaerobic organisms, such as methanogens and sulfate-reducing bacteria and archaea [19].
The residence times of water in a glacier depend on the glacier’s size, where the bigger the glacier, the longer the water takes to flow through the system because of longer routing pathways [67]. Meltwater residence in ice sheets is, therefore, generally higher than in glaciers [9,81]. In the photic zone on and just below the glacier surface, water residence times in the weathering crust are at least several days during the ablation season [58,60]; weathering crust can store some water in winter if covered with a thick snowpack [55]. Ice-lidded cryoconite holes found on ice sheets and polar glaciers store water from days to months, and isolated cryoconite holes that melt within the ice and refreeze annually without connecting to the surface or subsurface hydrologic system may store water for more than a decade [61,91,103]. Open cryoconite holes on temperate and polythermal glaciers are connected to the supraglacial hydrological system and may have residence times of a few minutes to hours [39,63]. Water residence times in supraglacial lakes and ponds vary similarly to those of cryoconite holes [50,61,64]. Supraglacial streams are usually fast water-flowing systems where water fluxes vary based on the glacial system state [64].
Below the photic zone, within the firn aquifer of the ablation zone, water residence times can range from hours to days [54,104]; perennial firn aquifers have also been observed [51,52,92,93,94,105]. Within the englacial realm, two very different water-flowing systems exist. A quick flow system of conduits may have residence times of hours and perhaps up to a day [88,96,97]. A very slow system exists along the boundaries of the ice crystals, where three or more crystals meet, creating a small flow passage (i.e., ice vein) [98]. In ice veins, water flows at very slow rates and is easily blocked by air bubbles. However, these veins can host viable microorganisms [106]. Residence time in these veins is unknown. At the bottom of the glacier, channels residence time can be hours, whereas, in slow-flow distributed subglacial systems, it can be days to months [99,100,101], reaching estimated water permanence times of millions of years in subglacial lakes under ice sheets [79,80,102].
As with meltwater discharge, water residence times change seasonally, with the shortest residence time being in the summer [90]. In winter, after the system closes down and various components are no longer linked due to the lack of water flow, the freezing of passageways or ductile closure of passageways causes residence times to increase [90].

3. Hydrology Influences on Glacial Nutrients and Microbial Communities

3.1. Deposition of Nutrients and Microorganisms on Glacial Surfaces

Dry and wet deposition of atmospheric aerosols transports diverse chemistry and biology to the glacier surface [107,108,109,110,111]. Their chemistry and concentration depend on patterns of atmospheric circulation, distance to source regions, and type of source emissions [112,113,114]. The distribution of chemical compounds across a glacier can be heterogeneous due to variable aeolian deposition [115]. For example, mineral dust is more highly concentrated at the margins of a glacier due to its proximity to bare rock and soils [116].
Bioaerosols (particulates containing microorganisms) may inoculate glacier surfaces, and subsequent microbial community development depends on the surface environment and microbial adaptations [117,118,119]. Such microbial communities typically include both endemic and cosmopolitan microorganisms [120,121,122,123,124,125]. Considering the dynamic and ever-shifting nature of glacial ice sheet systems, these communities will experience a broad range of icy micro-environments and will shift accordingly as they transit through the system [34,35].

3.2. Microbial and Geochemical Dynamics during the Ablation Season

3.2.1. Supraglacial Realm

In the supraglacial environment, organic and inorganic nutrients and carbon are often available in dissolved forms released by the biochemical weathering of deposited particles [13,126]. With the onset of the ablation season, nutrients and carbon deposited during the accumulation season progressively percolate through the snowpack into the weathering crust, cryoconite holes, and glacial ponds (ablation zone) [12,55], and microorganisms are able to resume metabolic activity shortly after thawing [20] (Figure 1).
In the photic layer of the weathering crust, UV- and visible-light-driven chemical transformations lead to the dissolution of iron oxide and silicates in mineral particles and Fe3+ reduction in ice-hosted sediment particles [13,55,127]. Nutrient concentrations, particularly dissolved organic nitrogen (DON) and P are generally higher in weathering crust compared to supraglacial streams and cryoconite holes [128]. Cryoconite holes are widely recognized as hotspots of microbial activity [63,91,129,130] (Figure 2). The waters are generally oxic environments, and ice-lidded cryoconite holes in Antarctica can be supersaturated in O2 [103,131]. However, within thick cryoconite granules (aggregates of microorganisms and organic and inorganic nutrients present in both open and ice-lidded cryoconite holes), the environment can become anoxic [132]. The oxidation state of cryoconite holes is important as it dictates the oxidation state of key nutrients. For example, the most common Fe ions, Fe2+ and Fe3+, are soluble only within certain pH and dO2 conditions, affecting their capacity to associate with other ions (e.g., chloride and hydroxide ions) or to adsorb onto ice crystal surfaces [13].
With the onset of the ablation season, during which the primary nutrient input is from snowmelt, the exposed component of the supraglacial environment (e.g., weathering crust) is dominated by prokaryotic photoautotrophs (e.g., Cyanobacteria in cryoconite holes) and ice algae (e.g., Ancylonema nordenskiöldii, and Mesotaenium berggrenii on the surface of the ice/weathering crust) [111,133]. These organisms can directly affect melt rates and surface morphology (Box 2) and play an important role in fixing atmospheric CO2 and N2 when nitrogen-fixing cyanobacteria, such as Nostoc and Anabaena, are present [134,135]. Microbial activity changes as the ablation season progresses (Figure 1). For example, during initial phases of snowmelt early in the ablation, excess concentrations of inorganic nitrogen (NO3 and NH4+) flush from the snowpack and are utilized by microbial communities, while later in the season (after the inorganic flush and when concentrations are low) microbial activity switches to dinitrogen fixation [136,137,138]. Phototrophic communities help to produce dissolved organic nutrients (i.e., dissolved organic carbon, nitrogen, and phosphorous), enriching the glacial surface with organic and inorganic nutrients that would be otherwise limiting factors for heterotrophic microbial activity [124,128,139]. Phosphorous is typically sourced from supraglacial particles via geochemical and physical processes and from microbial necromass via biological activity [139,140]. Microbial exudates and necromass on the glacial surface are essential for the functioning of the heterotrophic component of the microbial community. In this context, it has also been observed that different exudates are differentially utilized by microorganisms [141,142] within a complex glacial microbiome where heterotrophic organisms present a wide range of metabolisms [17,49,142]. Bradley et al. (2023) observed that more than 50% of bacterial cells (dominated by Actinomycetota, Pseudomonadota, and Planctomycetota) are translationally active on glacial surfaces [20]. Most of the deposited and transformed nutrients in the supraglacial realm are then exported to the rest of the glacial system [31,35]. With the development of warmer conditions due to the progression of the ablation season, weathering crust and cryoconite holes contribute nutrients and microbial cells to the rest of the glacier [29,58,143,144].
Box 2. Albedo and bioalbedo: how nutrients and microorganisms influence glacial hydrology.
 Although the subject of this review is how glacial hydrology influences nutrient and microbial distribution, the contrary is also true. The distribution of nutrients and microbial cells influences supraglacial water flux. Albedo (i.e., the proportion of light that is reflected by a surface) is lower for darker surfaces compared to lighter surfaces. Therefore, higher input of mineral dust and black carbon on the glacial and ice sheet surfaces decreases the albedo, causing higher heat absorption and increased melting, which influences the morphology of the ice surface [145]. Bioalbedo is a new term created to specifically refer to the decrease in albedo provoked by snow and ice algae, which are dark-pigmented, on the glacial surface [146,147]. A change in albedo and bioalbedo, therefore, directly influences glacial hydrology by promoting glacial and ice sheet surface melting.

3.2.2. Englacial Realm

Snowmelt also enters into the firn aquifer. No studies reporting biological processes in firn aquifers are available. However, there are indications of microbial activity: Holland et al. (2022) observed similar NO3 concentrations between the ice-snow interface meltwater and the snowpack [128]. However, DON and NH4+ concentrations were variable, possibly indicating microbial activity at the ice-snow interface. The chemical composition of the meltwater that enters the englacial system varies across the ablation season. Whereas meltwater at the beginning of the ablation season is likely to reflect the chemical composition of the snow, microbial processes and particle weathering ensure that meltwater that enters the englacial system later in the season is enriched with carbon, macronutrients (e.g., N, P, and Si) and ions such the dissolved inorganic forms of sulfur (e.g., SO42+) [148,149]. The englacial realm, with its network of water pathways, transfers cells and nutrients within the glacial system. It is unclear whether the englacial realm also has a role in nutrient and carbon transformation and is characterized by a microbial community specific to englacial pathways conditions [150]. Microbial nutrient cycling observed in fast-flowing supraglacial streams [49] suggests that microbial processes may be significant in fast-flowing englacial conduits. In addition, ice cores collected from englacial systems indicate that microorganisms are not quiescent but maintain an active metabolism [151]. These active metabolisms could be ascribed to chemoautotrophic organisms [151] or could rely on simple carbon substrates (e.g., acetic and formic acids) using NO3 and SO42+, which are abundant in meltwater deriving from supraglacial systems, as terminal acceptors [19].
In addition to englacial conduits, ice veins, with their high nutrient concentrations, offer a favorable habitat for microbial activity [68,106,152,153], and high cell concentrations have been measured in this environment [154]. Liquid flow in veins has long residence times, suggesting low oxygen concentrations and anaerobic metabolism [68,155], which can be mediated by methanogens such as Methanosphaerula and Methanococcus [156]. Despite the high nutrient concentrations and favorable conditions for active communities, microbial structure and function in the englacial realm are poorly understood [150,151].

3.2.3. Subglacial Realm

In subglacial systems, meltwater contacts bedrock (and subglacial till) and is consequently enriched in compounds released by rock comminution and dissolution (e.g., H2 and FeS2), creating environments where redox conditions may vary widely [65,66]. Chemical compositions of subglacial water highly vary due to (i) supraglacial and englacial hydrology, which controls the pattern, discharge, and biogeochemistry of meltwaters reaching the bottom of the glacier; (ii) subglacial hydrology, which controls the discharge and residence time of waters along the bed and influences patterns of erosion and regelation; and (iii) the geology of the subglacial substrate [11,90,121,143,157].
Depending on the mineral composition of the glacier and ice sheet beds, subsurface meltwater can be influenced by the weathering of pyrite (i.e., pyrite oxidation) and/or carbonates (i.e., carbonate dissolution) [27,158]. The dissolution of pyrite releases protons, and the dissolution of carbonate rocks releases dissolved inorganic carbon (e.g., CO2), which then creates carbonic acid in aqueous environments [18]. Pyrite oxidation is the prevalent form of mineral weathering in subglacial environments and has been observed to drive subglacial microbial metabolism [18,157]. The acidic environment resulting from pyrite oxidation also drives carbonate and silicate weathering [159]. The presence of subglacial tills also influences meltwater chemical compositions [36,160,161], and glacial beds can also be connected via aquifers to subterranean water sources [162]. All these factors further influence meltwater chemical composition and shape microbial community input to the subglacial environment [11,162,163].
As in the supraglacial and englacial realms, subglacial microbial communities are largely composed of heterotrophic microorganisms. However, contrary to the supraglacial realm, the primary producers of subglacial communities are chemolithotrophs. These organisms rely on nutrients transported from the glacial surface but also those released by rock weathering, and they accelerate mineral weathering of the glacial bed and chemical transformations within the subglacial system [13]. Biotic pyrite (FeS2) dissolution is rapid both in oxic and anoxic conditions [13,164], where O2 and Fe3+ can be used as sulfide oxidants [66]. Sulphide oxidation in oxic conditions uses pyrite, oxygen, and water to produce H+, Fe(OH)3 (iron (oxyhydr)oxides), and SO42. Fe(OH)3 dissociates to Fe3+ in the acidic subglacial environment (created by a high concentration of H+ due to rock dissolution) [13]. Anoxic pyrite dissociation can then occur: pyrite, Fe3+, and water react to form Fe2+, SO42−, and H+. Anoxic pyrite dissociation is faster than oxic dissociation because of the higher H+ production, which accelerates rock dissolution and weathering [13,165]. These weathering reactions are mediated by iron/sulfur-oxidizing bacteria, such as Thiobacillus and Sideroxydans species, and iron-reducing bacteria, such as Desulfosporosinus, Geobacter and Rhodoferax species [18,157,165,166]. Other microbial-mediated processes in the subsurface environment include denitrification, Mn4+ reduction, SO42− reduction, methanogenesis, and nitrification [19,108,167,168,169]. In this environment, complex microbial interactions occur where, under anaerobic conditions, SO42− reducing bacteria compete with methanogens for carbon substrates [170].
While meltwater that reaches subglacial systems is mostly oxygenated, as it is sourced from well-ventilated environments (e.g., supraglacial streams, moulins, and fast-flowing englacial conduits), oxygen levels in the subglacial environment can significantly vary based on the morphological characteristics of the system and, consequently, on water residence times [11,159]; whereas channelized subglacial systems are characterized by oxic waters thanks to their fast-flowing waters [11], anoxia and higher rock dissolution rates are observed with longer water permanence times and typically indicate higher rates of microbial activity and respiration [11,159]. In distributed drainage systems, there is a progressive development of anoxic conditions due to slow water flow, which creates favorable conditions for the uptake and use of organic matter by heterotrophic organisms via oxidative cellular respiration [67]. Despite these shifts in water oxygen levels, in general, microbial metabolism is thought to be driven by the mineralization of organic carbon and nitrogen under oxic conditions [19], whereas microbial communities performing sulfate reduction and methanogenesis prevail with the development of anoxic conditions [171].

3.3. Microbial and Geochemical Dynamics during the Accumulation Season

Most of the studies on glacial microbial communities take place during the summer ablation season at temperate and polythermal glaciers, where geochemical and microbial processes are most active due to the presence of meltwater and nutrients in the system [172]. Consequently, little information is available on microbial processes during winter.
Even in winter, water in ice veins, englacial pockets, and subglacial regions [11,19,68] may retain sufficiently high solute concentrations to sustain basal microbial metabolism (e.g., DNA repair mechanisms) [173] or even microbial growth. During the accumulation season, the prevalent microbial metabolism is likely to be chemoautotrophy in all glacial environments; the newly secreted nutrients and remnant nutrients from ablation seasons could also sustain metabolism in heterotrophic organisms. Active microbial communities have been identified in systems that are only minimally influenced by ablation/accumulation seasonal differences, such as subglacial lakes in ice sheets [174,175,176]. Furthermore, in glaciers during the accumulation season (when nutrient input from the surface is largely absent), nutrients can be sourced from glacial bedrock where H2 is released abiotically from rock comminution and can be oxidized in both aerobic and anaerobic conditions, fueling microbial chemolithotrophy [65,177,178,179]. Thus, even during the accumulation season, glaciers can harbor biogeochemical transformations thanks to microbial-mediated processes. This is also supported by observations of active heterotrophic communities in snowpacks incubated in cold and dark conditions [180].

4. Glaciers and Ice Sheets as Bioreactors

Considering the array of microbial processes active on, in, and under glaciers and ice sheets, these systems can be thought of as bioreactors. Nutrient concentration and bioavailability increase from supraglacial input waters to exiting subglacial waters. Exiting subglacial waters are enriched with dissolved iron (i.e., Fe2+ and Fe3+) and iron nanoparticulates (e.g., Fe (oxyhydr)oxides and Fe2+-bound compounds), which are bioavailable for microbial uptake [13,181]. Similarly, other metals are more bioavailable in glacial flour (i.e., fine rock particles formed by rock comminution) than in other dust [182]. Organic carbon (DOC) also follows similar trends [183]. The DOC proteinaceous component (DOC that is microbially produced and secreted) is higher compared to humic and fulvic acids (DOC transported by aerosol deposition and terrestrial transport) in glacial meltwater compared to other water systems such as rivers and lakes [16], indicating a strong involvement of glacial microbial communities in carbon cycling, and in the export of labile carbon. Kellerman et al. (2020) observed a shift in protein-like fluorescence with the progression of the ablation period indicating microbial communities as a source of DOC [184]. Barker et al. (2006) also showed that supraglacial and subglacial water had similar DOC concentrations but different fluorescence signals in three different glaciers, clearly indicating shifts in DOC quality due to microbial transformation [15]; this further points to a microbial role in DOM release [7,185]. An increase in microbial cell concentrations in glacial water flowing from the surface to the subsurface of glaciers has also been reported [6] (Table 2). Cell concentrations in subglacial water are usually in the order of 105 cells mL−1 compared to concentrations in supraglacial water which are an order of magnitude lower (with the exception of ice veins where observed concentrations were 106–108 cells mL−1) (Table 2). These clear patterns in the enrichment of cell, nutrient, and bioavailable compounds show how glaciers serve as bioreactors for a range of biogeochemical transformations, which in turn can deeply influence proglacial systems.

5. Proglacial Systems

Glacial meltwater is an important source of cells and nutrients in outflow systems, carrying bioavailable Fe, DOC, N, P, Si, and rare metals, together with sediments and glacial flour [183,192,193,194,195,196,197]. How the proglacial system impacts downstream systems [198,199,200,201] depends on different factors, where the morphology of the proglacial system (i.e., land or maritime terminating) and its nutrient state (e.g., oligotrophic vs. eutrophic) play important roles [202].
The retreat of land-terminating glaciers exposes soil in the proximity of the glacier (i.e., forefield), where soils show a gradient in texture and chemical characteristics from the ice edge: newly exposed soils are usually characterized by low nutrient levels and little or no vegetation, and are impacted by microorganisms and nutrient from glacial meltwater [135,203,204,205]. Subglacial water from land-terminating glaciers also flows into rivers and streams, which will then export nutrients to other water bodies (e.g., lakes, rivers, and seas) [206]. Glacial streams and lakes typically show different dissolved inorganic nitrogen and phosphate concentrations than similar water bodies without glacial water input, and silica concentrations are lower in glacial streams compared to non-glacial streams [200]. Streams and lakes are also enriched with nitrogen deriving from glacial water in North America [207,208], where an increase in nitrogen concentration influences microbial structure and diversity (e.g., planktonic diatoms) [209]. Warner et al. (2017) also observed higher algal biomass in glacial lakes (c.f. non-glacial lakes) [199]. Lake connectivity to glacial water influences microbial communities due to the import of nutrients and the shift between turbidity and clear water conditions [210].
During the transport in the river, glacial meltwater undergoes chemical modifications. For instance, Fe and DOC are sequestered by precipitation and adsorption during the glacial meltwater transport in rivers and streams in Svalbard [211]. Nitrogen was also observed to decline in downstream glacial lakes in North America [199]. However, DOC increased during its flow in proglacial streams in Iceland, probably due to carbon mineralization and microbial transformation [212]. These contrasting results show how the processes in the proglacial streams are regulated by many local factors. Even if exported nutrient concentrations (e.g., Fe) decrease during their flow through the rivers and estuaries, glacial-fed streams are still able to export nutrients to the ocean [213]. Glacial meltwater from land-terminating glaciers enters the sea/ocean through estuaries. Here, the glacial water creates a top layer of sediment-rich water which can inhibit primary productivity in the proximity of the land because of the decrease of light filtration in the system [214,215] (Figure 3A). Whereas this is often the case in the Arctic, land-terminating glaciers in the Antarctic can have positive effects because water is so oligotrophic that even the import of low nutrient concentrations can increase primary production [202,213,216,217].
Subglacial water from marine-terminating glaciers provides carbon, macronutrients (e.g., nitrogen), and micronutrients (e.g., iron) to the surrounding marine system [214,218,219]. Organic carbon and other macronutrients are generally less concentrated in subglacial water than in seawater but more bioavailable [183,192,220]. In addition, glacial meltwater causes an upwelling of nutrient-rich deep seawater due to the buoyancy of the glacial cold water (Figure 3B), therefore enriching the surface seawater column with deep sea macronutrients [220,221,222,223]. Micronutrients (e.g., Fe) are generally more concentrated in subglacial waters compared to seawater [193], although the destiny and availability of these nutrients are still under debate [193,195,221].
The impact that subglacial discharge has on the coastal system depends on two main factors: the seawater depth at which subglacial water is discharged and the nutrient state of the marine system [224,225]. Meltwater from marine-terminating glaciers enters the marine environment at different depths in the seawater column depending on the kind of glacier (e.g., tidewater or ice shelf) and glacier thickness [225]. If the discharge of cold subglacial waters creates an upwelling of deep nutrient-rich waters that reach the photic zone of the seawater column, an increase in primary productivity (e.g., higher chlorophyll concentration) is observed thanks to the import of nitrogen, ammonium, phosphate, and silicate in the Arctic marine coastal environment where the seawater is generally nitrogen depleted [220,221,222,224]. However, the upwelling of nutrient-rich waters might not reach the photic zone in the case of deep or shallow marine-terminating glaciers, causing limited primary productivity in Arctic waters [225]. Contrary to the Arctic seawater, which is generally nitrogen depleted, the Southern Ocean is mainly limited by low concentrations of bioavailable Fe (e.g., Fe2+ and colloidal Fe) [224,225], and it, therefore, relies less on the inputs of macronutrients by deep seawater upwelling, but rather more on the import of micronutrients from glacial waters themselves [195,226,227].
Generally, nutrient export increases with meltwater export to the proglacial systems [148]. This is the result of the release of the nutrients trapped in glaciers and ice sheets, which act as storage for different elements such as carbon, and also of the higher presence of water in the system, which promotes geochemical and microbial processes in supraglacial and subglacial realms [183]. Whereas it is clear that an increase in nutrient release affects geochemical cycles, trophic chains, and microbial diversity [23,228,229], an increase in nutrient export from the glacial system does not always correspond to an increase in the proglacial system primary production [221,230].
Changes in water residence times in glaciers could cause drastic shifts in proglacial environments. In particular, a decrease in glacial residence times with a consequent increase of glacial water released to outlet systems is observed due to global warming [9,231,232]. For example, a change in the glacial water runoff and nutrient export can correspond to a change in water acidity in the coastal environment caused by a variation in carbonate concentrations, sediment input and burial, water stratigraphy and turbidity, and nutrient import into the system, leading to a change in environmental dynamics and primary productivity in proglacial systems [148,233,234,235,236,237,238] (Figure 3C).

6. Conclusions

Supraglacial and subglacial microbial communities have been widely studied and characterized in the last few decades, with many studies showing how biotic and abiotic processes in these systems are linked to global biogeochemical cycles [4,34,35]. It is also common knowledge that climate warming affects the interconnectivity among glacial hydrology, microbial community, and geochemistry and that increases in meltwater discharge from glaciers and ice sheets impact the proglacial systems [20,198,199,200,201]. Therefore, whereas it is clear that climate warming affects and will affect dynamics within glaciers and ice sheets and in their out-stream environments, little information is present on how glacial microbial communities will be impacted. In order to fully understand these impacts, we believe a more holistic understanding of how microbial communities, nutrient cycling, and glacial meltwater hydrology function and interact is necessary. For example, nutrient cycling and microbial activity in the yet unexplored englacial channels and firn aquifers should be explored, together with biotic and abiotic processes during the accumulation season. Further, better estimates of water residence times in different compartments of glaciers and ice sheets are needed to more confidently estimate microbial activity and conditions for nutrient cycling. Whereas more comprehensive information on glacial and ice sheet hydrology will inform on microbial activity and nutrient composition, the opposite is also true, where a deeper understanding of microbial and nutrient distribution and transformations could further inform on water paths within glaciers.

Author Contributions

Conceptualization, G.V., P.H.L. and A.G.F.; writing—review and editing, G.V., A.G.F., P.H.L., B.F., A.M.A. and D.A.C.; visualization, G.V.; funding acquisition, D.A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Swiss Polar Institute (grant No. TEG-2022-002) and the National Research Foundation (grant No. 129227).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

We thank the University of Pretoria for providing postdoctoral bursaries for G.V. and P.L. We also thank the Swiss Polar Institute (grant No. TEG-2022-002) and the National Research Foundation for funding support (grant No. 129227).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hooke, R. Principles of Glacier Mechanics, 3rd ed.; Cambridge University Press: Cambridge, UK, 2020. [Google Scholar]
  2. Jones, D.B.; Harrison, S.; Anderson, K.; Whalley, W.B. Rock Glaciers and Mountain Hydrology: A Review. Earth-Sci. Rev. 2019, 193, 66–90. [Google Scholar] [CrossRef]
  3. Zemp, M.; Huss, M.; Thibert, E.; Eckert, N.; McNabb, R.; Huber, J.; Barandun, M.; Machguth, H.; Nussbaumer, S.U.; Gärtner-Roer, I.; et al. Global Glacier Mass Changes and Their Contributions to Sea-Level Rise from 1961 to 2016. Nature 2019, 568, 382–386. [Google Scholar] [CrossRef] [PubMed]
  4. Hotaling, S.; Hood, E.; Hamilton, T.L. Microbial Ecology of Mountain Glacier Ecosystems: Biodiversity, Ecological Connections and Implications of a Warming Climate. Environ. Microbiol. 2017, 19, 2935–2948. [Google Scholar] [CrossRef] [PubMed]
  5. Cameron, K.A.; Müller, O.; Stibal, M.; Edwards, A.; Jacobsen, C.S. Glacial Microbiota Are Hydrologically Connected and Temporally Variable. Environ. Microbiol. 2020, 22, 3172–3187. [Google Scholar] [CrossRef] [PubMed]
  6. Cameron, K.A.; Stibal, M.; Hawkings, J.R.; Mikkelsen, A.B.; Telling, J.; Kohler, T.J.; Gözdereliler, E.; Zarsky, J.D.; Wadham, J.L.; Jacobsen, C.S. Meltwater Export of Prokaryotic Cells from the Greenland Ice Sheet: Microbial Export from the Greenland Ice Sheet. Environ. Microbiol. 2017, 19, 524–534. [Google Scholar] [CrossRef]
  7. Fegel, T.; Boot, C.M.; Broeckling, C.D.; Baron, J.S.; Hall, E.K. Assessing the Chemistry and Bioavailability of Dissolved Organic Matter From Glaciers and Rock Glaciers. J. Geophys. Res. Biogeosci. 2019, 124, 1988–2004. [Google Scholar] [CrossRef]
  8. Kohler, T.J.; Vinšová, P.; Falteisek, L.; Žárský, J.D.; Yde, J.C.; Hatton, J.E.; Hawkings, J.R.; Lamarche-Gagnon, G.; Hood, E.; Cameron, K.A.; et al. Patterns in Microbial Assemblages Exported From the Meltwater of Arctic and Sub-Arctic Glaciers. Front. Microbiol. 2020, 11, 669. [Google Scholar] [CrossRef]
  9. Chu, V.W. Greenland Ice Sheet Hydrology: A Review. Prog. Phys. Geogr. Earth Environ. 2014, 38, 19–54. [Google Scholar] [CrossRef]
  10. Simkins, L.M.; Greenwood, S.L.; Munevar Garcia, S.; Eareckson, E.A.; Anderson, J.B.; Prothro, L.O. Topographic Controls on Channelized Meltwater in the Subglacial Environment. Geophys. Res. Lett. 2021, 48, e2021GL094678. [Google Scholar] [CrossRef]
  11. Tranter, M.; Skidmore, M.; Wadham, J. Hydrological Controls on Microbial Communities in Subglacial Environments. Hydrol. Process. 2005, 19, 995–998. [Google Scholar] [CrossRef]
  12. Dubnick, A.; Kazemi, S.; Sharp, M.; Wadham, J.; Hawkings, J.; Beaton, A.; Lanoil, B. Hydrological Controls on Glacially Exported Microbial Assemblages: Subglacial Hydrology and Microbiology. J. Geophys. Res. Biogeosci. 2017, 122, 1049–1061. [Google Scholar] [CrossRef]
  13. Raiswell, R.; Hawkings, J.; Elsenousy, A.; Death, R.; Tranter, M.; Wadham, J. Iron in Glacial Systems: Speciation, Reactivity, Freezing Behavior, and Alteration During Transport. Front. Earth Sci. 2018, 6, 222. [Google Scholar] [CrossRef]
  14. Margesin, R.; Collins, T. Microbial Ecology of the Cryosphere (Glacial and Permafrost Habitats): Current Knowledge. Appl. Microbiol. Biotechnol. 2019, 103, 2537–2549. [Google Scholar] [CrossRef]
  15. Barker, J.D.; Sharp, M.J.; Fitzsimons, S.J.; Turner, R.J. Abundance and Dynamics of Dissolved Organic Carbon in Glacier Systems. Arct. Antarct. Alp. Res. 2006, 38, 163–172. [Google Scholar] [CrossRef]
  16. Dubnick, A.; Barker, J.; Sharp, M.; Wadham, J.; Lis, G.; Telling, J.; Fitzsimons, S.; Jackson, M. Characterization of Dissolved Organic Matter (DOM) from Glacial Environments Using Total Fluorescence Spectroscopy and Parallel Factor Analysis. Ann. Glaciol. 2010, 51, 111–122. [Google Scholar] [CrossRef]
  17. Anesio, A.M.; Laybourn-Parry, J. Ecology of Arctic Glaciers. In Arctic Ecology; Thomas, D.N., Ed.; Wiley: Hoboken, NJ, USA, 2021; pp. 133–158. ISBN 978-1-118-84654-4. [Google Scholar]
  18. Boyd, E.S.; Hamilton, T.L.; Havig, J.R.; Skidmore, M.L.; Shock, E.L. Chemolithotrophic Primary Production in a Subglacial Ecosystem. Appl. Environ. Microbiol. 2014, 80, 6146–6153. [Google Scholar] [CrossRef]
  19. Hodson, A.; Anesio, A.M.; Tranter, M.; Fountain, A.; Osborn, M.; Priscu, J.; Laybourn-Parry, J.; Sattler, B. Glacial Ecosystems. Ecol. Monogr. 2008, 78, 41–67. [Google Scholar] [CrossRef]
  20. Bradley, J.A.; Trivedi, C.B.; Winkel, M.; Mourot, R.; Lutz, S.; Larose, C.; Keuschnig, C.; Doting, E.; Halbach, L.; Zervas, A.; et al. Active and Dormant Microorganisms on Glacier Surfaces. Geobiology 2023, 21, 244–261. [Google Scholar] [CrossRef]
  21. Jain, S. Fundamentals of Physical Geology; Springer Geology; Springer India: New Delhi, India, 2014; ISBN 978-81-322-1538-7. [Google Scholar]
  22. Buytaert, W.; Moulds, S.; Acosta, L.; De Bièvre, B.; Olmos, C.; Villacis, M.; Tovar, C.; Verbist, K.M.J. Glacial Melt Content of Water Use in the Tropical Andes. Environ. Res. Lett. 2017, 12, 114014. [Google Scholar] [CrossRef]
  23. Milner, A.M.; Khamis, K.; Battin, T.J.; Brittain, J.E.; Barrand, N.E.; Füreder, L.; Cauvy-Fraunié, S.; Gíslason, G.M.; Jacobsen, D.; Hannah, D.M.; et al. Glacier Shrinkage Driving Global Changes in Downstream Systems. Proc. Natl. Acad. Sci. USA 2017, 114, 9770–9778. [Google Scholar] [CrossRef]
  24. Biemans, H.; Siderius, C.; Lutz, A.F.; Nepal, S.; Ahmad, B.; Hassan, T.; von Bloh, W.; Wijngaard, R.R.; Wester, P.; Shrestha, A.B.; et al. Importance of Snow and Glacier Meltwater for Agriculture on the Indo-Gangetic Plain. Nat. Sustain. 2019, 2, 594–601. [Google Scholar] [CrossRef]
  25. Catania, G.A.; Stearns, L.A.; Moon, T.A.; Enderlin, E.M.; Jackson, R.H. Future Evolution of Greenland’s Marine-Terminating Outlet Glaciers. J. Geophys. Res. Earth Surf. 2020, 125, e2018JF004873. [Google Scholar] [CrossRef]
  26. Slater, T.; Shepherd, A.; McMillan, M.; Leeson, A.; Gilbert, L.; Muir, A.; Munneke, P.K.; Noël, B.; Fettweis, X.; van den Broeke, M.; et al. Increased Variability in Greenland Ice Sheet Runoff from Satellite Observations. Nat. Commun. 2021, 12, 6069. [Google Scholar] [CrossRef] [PubMed]
  27. Torres, M.A.; Moosdorf, N.; Hartmann, J.; Adkins, J.F.; West, A.J. Glacial Weathering, Sulfide Oxidation, and Global Carbon Cycle Feedbacks. Proc. Natl. Acad. Sci. USA 2017, 114, 8716–8721. [Google Scholar] [CrossRef]
  28. Delaney, I.; Adhikari, S. Increased Subglacial Sediment Discharge in a Warming Climate: Consideration of Ice Dynamics, Glacial Erosion, and Fluvial Sediment Transport. Geophys. Res. Lett. 2020, 47, e2019GL085672. [Google Scholar] [CrossRef]
  29. Stevens, I.T.; Irvine-Fynn, T.D.L.; Edwards, A.; Mitchell, A.C.; Cook, J.M.; Porter, P.R.; Holt, T.O.; Huss, M.; Fettweis, X.; Moorman, B.J.; et al. Spatially Consistent Microbial Biomass and Future Cellular Carbon Release from Melting Northern Hemisphere Glacier Surfaces. Commun. Earth Environ. 2022, 3, 275. [Google Scholar] [CrossRef]
  30. Anderson, S.P. Biogeochemistry of Glacial Landscape Systems. Annu. Rev. Earth Planet. Sci. 2007, 35, 375–399. [Google Scholar] [CrossRef]
  31. Anesio, A.M.; Sattler, B.; Foreman, C.; Telling, J.; Hodson, A.; Tranter, M.; Psenner, R. Carbon Fluxes through Bacterial Communities on Glacier Surfaces. Ann. Glaciol. 2010, 51, 32–40. [Google Scholar] [CrossRef]
  32. Hawkings, J.; Wadham, J.; Tranter, M.; Telling, J.; Bagshaw, E.; Beaton, A.; Simmons, S.-L.; Chandler, D.; Tedstone, A.; Nienow, P. The Greenland Ice Sheet as a Hot Spot of Phosphorus Weathering and Export in the Arctic: THE GREENLAND ICE SHEET P CYCLE. Glob. Biogeochem. Cycles 2016, 30, 191–210. [Google Scholar] [CrossRef]
  33. Wadham, J.L.; Hawkings, J.R.; Tarasov, L.; Gregoire, L.J.; Spencer, R.G.M.; Gutjahr, M.; Ridgwell, A.; Kohfeld, K.E. Ice Sheets Matter for the Global Carbon Cycle. Nat. Commun. 2019, 10, 3567. [Google Scholar] [CrossRef]
  34. Boetius, A.; Anesio, A.M.; Deming, J.W.; Mikucki, J.A.; Rapp, J.Z. Microbial Ecology of the Cryosphere: Sea Ice and Glacial Habitats. Nat. Rev. Microbiol. 2015, 13, 677–690. [Google Scholar] [CrossRef]
  35. Anesio, A.M.; Lutz, S.; Chrismas, N.A.M.; Benning, L.G. The Microbiome of Glaciers and Ice Sheets. npj Biofilms Microbiomes 2017, 3, 10. [Google Scholar] [CrossRef]
  36. Fountain, A.G.; Walder, J.S. Water Flow through Temperate Glaciers. Rev. Geophys. 1998, 36, 299–328. [Google Scholar] [CrossRef]
  37. St Germain, S.L.; Moorman, B.J. Long-Term Observations of Supraglacial Streams on an Arctic Glacier. J. Glaciol. 2019, 65, 900–911. [Google Scholar] [CrossRef]
  38. Cook, S.J.; Swift, D.A.; Kirkbride, M.P.; Knight, P.G.; Waller, R.I. The Empirical Basis for Modelling Glacial Erosion Rates. Nat. Commun. 2020, 11, 759. [Google Scholar] [CrossRef]
  39. Irvine-Fynn, T.D.L.; Hodson, A.J.; Moorman, B.J.; Vatne, G.; Hubbard, A.L. Polythermal glacier hydrology: A review. Rev. Geophys. 2011, 49. [Google Scholar] [CrossRef]
  40. Clason, C.; Mair, D.W.F.; Burgess, D.O.; Nienow, P.W. Modelling the Delivery of Supraglacial Meltwater to the Ice/Bed Interface: Application to Southwest Devon Ice Cap, Nunavut, Canada. J. Glaciol. 2012, 58, 361–374. [Google Scholar] [CrossRef]
  41. Hambrey, M.J.; Glasser, N.F. Discriminating Glacier Thermal and Dynamic Regimes in the Sedimentary Record. Sediment. Geol. 2012, 251–252, 1–33. [Google Scholar] [CrossRef]
  42. Pittard, M.L.; Galton-Fenzi, B.K.; Roberts, J.L.; Watson, C.S. Organization of Ice Flow by Localized Regions of Elevated Geothermal Heat Flux. Geophys. Res. Lett. 2016, 43, 3342–3350. [Google Scholar] [CrossRef]
  43. Marmoni, G.M.; Martino, S.; Salvatore, M.C.; Gaeta, M.; Perinelli, C.; Scarascia Mugnozza, G.; Baroni, C. Numerical Modelling of Geothermal Heat Flux and Ice Velocity Influencing the Thermal Conditions of the Priestley Glacier Trough (Northern Victoria Land, Antarctica). Geomorphology 2021, 394, 107959. [Google Scholar] [CrossRef]
  44. Pettersson, R.; Jansson, P.; Blatter, H. Spatial Variability in Water Content at the Cold-Temperate Transition Surface of the Polythermal Storglaciären, Sweden: Spatial Variability in Water Content. J. Geophys. Res. 2004, 109, F02009. [Google Scholar] [CrossRef]
  45. Vatne, G.; Irvine-Fynn, T.D.L. Morphological Dynamics of an Englacial Channel. Hydrol. Earth Syst. Sci. 2016, 20, 2947–2964. [Google Scholar] [CrossRef]
  46. Badgeley, J.A.; Pettit, E.C.; Carr, C.G.; Tulaczyk, S.; Mikucki, J.A.; Lyons, W.B.; MIDGE Science Team. An Englacial Hydrologic System of Brine within a Cold Glacier: Blood Falls, McMurdo Dry Valleys, Antarctica. J. Glaciol. 2017, 63, 387–400. [Google Scholar] [CrossRef]
  47. Temminghoff, M.; Benn, D.I.; Gulley, J.D.; Sevestre, H. Characterization of the Englacial and Subglacial Drainage System in a High Arctic Cold Glacier by Speleological Mapping and Ground-Penetrating Radar. Geogr. Ann. Ser. A Phys. Geogr. 2019, 101, 98–117. [Google Scholar] [CrossRef]
  48. Hansen, L.U.; Piotrowski, J.A.; Benn, D.I.; Sevestre, H. A Cross-Validated Three-Dimensional Model of an Englacial and Subglacial Drainage System in a High-Arctic Glacier. J. Glaciol. 2020, 66, 278–290. [Google Scholar] [CrossRef]
  49. Bergstrom, A.; Gooseff, M.N.; Singley, J.G.; Cohen, M.J.; Welch, K.A. Nutrient Uptake in the Supraglacial Stream Network of an Antarctic Glacier. J. Geophys. Res. Biogeosci. 2020, 125, e2020JG005679. [Google Scholar] [CrossRef]
  50. Bagshaw, E.A.; Tranter, M.; Wadham, J.L.; Fountain, A.G.; Basagic, H. Dynamic Behaviour of Supraglacial Lakes on Cold Polar Glaciers: Canada Glacier, McMurdo Dry Valleys, Antarctica. J. Glaciol. 2010, 56, 366–368. [Google Scholar] [CrossRef]
  51. Miller, O.L.; Solomon, D.K.; Miège, C.; Koenig, L.S.; Forster, R.R.; Montgomery, L.N.; Schmerr, N.; Ligtenberg, S.R.M.; Legchenko, A.; Brucker, L. Hydraulic Conductivity of a Firn Aquifer in Southeast Greenland. Front. Earth Sci. 2017, 5, 38. [Google Scholar] [CrossRef]
  52. Montgomery, L.N.; Schmerr, N.; Burdick, S.; Forster, R.R.; Koenig, L.; Legchenko, A.; Ligtenberg, S.; Miège, C.; Miller, O.L.; Solomon, D.K. Investigation of Firn Aquifer Structure in Southeastern Greenland Using Active Source Seismology. Front. Earth Sci. 2017, 5, 10. [Google Scholar] [CrossRef]
  53. Kendrick, A.K.; Schroeder, D.M.; Chu, W.; Young, T.J.; Christoffersen, P.; Todd, J.; Doyle, S.H.; Box, J.E.; Hubbard, A.; Hubbard, B.; et al. Surface Meltwater Impounded by Seasonal Englacial Storage in West Greenland. Geophys. Res. Lett. 2018, 45, 10474–10481. [Google Scholar] [CrossRef]
  54. Fountain, A.G. Effect of snow and firn hydrology on the physical and chemical characteristics of glacial runoff. Hydrol. Process. 1996, 10, 509–521. [Google Scholar] [CrossRef]
  55. Cook, J.M.; Hodson, A.J.; Irvine-Fynn, T.D.L. Supraglacial Weathering Crust Dynamics Inferred from Cryoconite Hole Hydrology: Weathering Crust Hydrology Observed in Cryoconite Holes. Hydrol. Process. 2016, 30, 433–446. [Google Scholar] [CrossRef]
  56. Hoffman, M.J.; Fountain, A.G.; Liston, G.E. Near-Surface Internal Melting: A Substantial Mass Loss on Antarctic Dry Valley Glaciers. J. Glaciol. 2014, 60, 361–374. [Google Scholar] [CrossRef]
  57. Christner, B.C.; Lavender, H.F.; Davis, C.L.; Oliver, E.E.; Neuhaus, S.U.; Myers, K.F.; Hagedorn, B.; Tulaczyk, S.M.; Doran, P.T.; Stone, W.C. Microbial Processes in the Weathering Crust Aquifer of a Temperate Glacier. Cryosphere 2018, 12, 3653–3669. [Google Scholar] [CrossRef]
  58. Irvine-Fynn, T.D.L.; Edwards, A.; Stevens, I.T.; Mitchell, A.C.; Bunting, P.; Box, J.E.; Cameron, K.A.; Cook, J.M.; Naegeli, K.; Rassner, S.M.E.; et al. Storage and Export of Microbial Biomass across the Western Greenland Ice Sheet. Nat. Commun. 2021, 12, 3960. [Google Scholar] [CrossRef]
  59. Christianson, K.; Kohler, J.; Alley, R.B.; Nuth, C.; Pelt, W.J.J. Dynamic Perennial Firn Aquifer on an Arctic Glacier. Geophys. Res. Lett. 2015, 42, 1418–1426. [Google Scholar] [CrossRef]
  60. Stevens, I.T.; Irvine-Fynn, T.D.L.; Porter, P.R.; Cook, J.M.; Edwards, A.; Smart, M.; Moorman, B.J.; Hodson, A.J.; Mitchell, A.C. Near-Surface Hydraulic Conductivity of Northern Hemisphere Glaciers. Hydrol. Process. 2018, 32, 850–865. [Google Scholar] [CrossRef]
  61. Fountain, A.G.; Tranter, M.; Nylen, T.H.; Lewis, K.J.; Mueller, D.R. Evolution of Cryoconite Holes and Their Contribution to Meltwater Runoff from Glaciers in the McMurdo Dry Valleys, Antarctica. J. Glaciol. 2004, 50, 35–45. [Google Scholar] [CrossRef]
  62. MacDonell, S.; Fitzsimons, S. The Formation and Hydrological Significance of Cryoconite Holes. Progress. Phys. Geogr. Earth Environ. 2008, 32, 595–610. [Google Scholar] [CrossRef]
  63. Cook, J.; Edwards, A.; Takeuchi, N.; Irvine-Fynn, T. Cryoconite: The Dark Biological Secret of the Cryosphere. Progress. Phys. Geogr. Earth Environ. 2016, 40, 66–111. [Google Scholar] [CrossRef]
  64. Pitcher, L.H.; Smith, L.C. Supraglacial Streams and Rivers. Annu. Rev. Earth Planet. Sci. 2019, 47, 421–452. [Google Scholar] [CrossRef]
  65. Dunham, E.C.; Dore, J.E.; Skidmore, M.L.; Roden, E.E.; Boyd, E.S. Lithogenic Hydrogen Supports Microbial Primary Production in Subglacial and Proglacial Environments. Proc. Natl. Acad. Sci. USA 2021, 118, e2007051117. [Google Scholar] [CrossRef] [PubMed]
  66. Bottrell, S.H.; Tranter, M. Sulphide Oxidation under Partially Anoxic Conditions at the Bed of the Haut Glacier d’Arolla, Switzerland. Hydrol. Process. 2002, 16, 2363–2368. [Google Scholar] [CrossRef]
  67. Wadham, J.L.; Tranter, M.; Skidmore, M.; Hodson, A.J.; Priscu, J.; Lyons, W.B.; Sharp, M.; Wynn, P.; Jackson, M. Biogeochemical Weathering under Ice: Size Matters: Glacial Biogeochemical Weathering. Glob. Biogeochem. Cycles 2010, 24, GB3025. [Google Scholar] [CrossRef]
  68. Maccario, L.; Sanguino, L.; Vogel, T.M.; Larose, C. Snow and Ice Ecosystems: Not so Extreme. Res. Microbiol. 2015, 166, 782–795. [Google Scholar] [CrossRef]
  69. Fountain, A.G.; Jacobel, R.W.; Schlichting, R.; Jansson, P. Fractures as the Main Pathways of Water Flow in Temperate Glaciers. Nature 2005, 433, 618–621. [Google Scholar] [CrossRef]
  70. McGrath, D.; Colgan, W.; Steffen, K.; Lauffenburger, P.; Balog, J. Assessing the Summer Water Budget of a Moulin Basin in the Sermeq Avannarleq Ablation Region, Greenland Ice Sheet. J. Glaciol. 2011, 57, 954–964. [Google Scholar] [CrossRef]
  71. Lampkin, D.J.; Amador, N.; Parizek, B.R.; Farness, K.; Jezek, K. Drainage from Water-filled Crevasses along the Margins of Jakobshavn Isbræ: A Potential Catalyst for Catchment Expansion. J. Geophys. Res. Earth Surf. 2013, 118, 795–813. [Google Scholar] [CrossRef]
  72. Llubes, M.; Lanseau, C.; Rémy, F. Relations between Basal Condition, Subglacial Hydrological Networks and Geothermal Flux in Antarctica. Earth Planet. Sci. Lett. 2006, 241, 655–662. [Google Scholar] [CrossRef]
  73. Fisher, A.T.; Mankoff, K.D.; Tulaczyk, S.M.; Tyler, S.W.; Foley, N.; The WISSARD Science Team. High Geothermal Heat Flux Measured below the West Antarctic Ice Sheet. Sci. Adv. 2015, 1, e1500093. [Google Scholar] [CrossRef]
  74. Lösing, M.; Ebbing, J.; Szwillus, W. Geothermal Heat Flux in Antarctica: Assessing Models and Observations by Bayesian Inversion. Front. Earth Sci. 2020, 8, 105. [Google Scholar] [CrossRef]
  75. Beem, L.H.; Jezek, K.C.; Van Der Veen, C.J. Basal Melt Rates beneath Whillans Ice Stream, West Antarctica. J. Glaciol. 2010, 56, 647–654. [Google Scholar] [CrossRef]
  76. Nanni, U.; Gimbert, F.; Roux, P.; Lecointre, A. Observing the Subglacial Hydrology Network and Its Dynamics with a Dense Seismic Array. Proc. Natl. Acad. Sci. USA 2021, 118, e2023757118. [Google Scholar] [CrossRef]
  77. Scholzen, C.; Schuler, T.V.; Gilbert, A. Sensitivity of Subglacial Drainage to Water Supply Distribution at the Kongsfjord Basin, Svalbard. Cryosphere 2021, 15, 2719–2738. [Google Scholar] [CrossRef]
  78. Ronayne, M.J.; Houghton, T.B.; Stednick, J.D. Field Characterization of Hydraulic Conductivity in a Heterogeneous Alpine Glacial Till. J. Hydrol. 2012, 458–459, 103–109. [Google Scholar] [CrossRef]
  79. Siegert, M.J.; Ross, N.; Le Brocq, A.M. Recent Advances in Understanding Antarctic Subglacial Lakes and Hydrology. Philos. Trans. R. Soc. A 2016, 374, 20140306. [Google Scholar] [CrossRef]
  80. Bowling, J.S.; Livingstone, S.J.; Sole, A.J.; Chu, W. Distribution and Dynamics of Greenland Subglacial Lakes. Nat. Commun. 2019, 10, 2810. [Google Scholar] [CrossRef]
  81. Jansson, P.; Hock, R.; Schneider, T. The Concept of Glacier Storage: A Review. J. Hydrol. 2003, 282, 116–129. [Google Scholar] [CrossRef]
  82. Vaňková, I.; Voytenko, D.; Nicholls, K.W.; Xie, S.; Parizek, B.R.; Holland, D.M. Vertical Structure of Diurnal Englacial Hydrology Cycle at Helheim Glacier, East Greenland. Geophys. Res. Lett. 2018, 45, 8352–8362. [Google Scholar] [CrossRef]
  83. Dunse, T.; Schellenberger, T.; Hagen, J.O.; Kääb, A.; Schuler, T.V.; Reijmer, C.H. Glacier-Surge Mechanisms Promoted by a Hydro-Thermodynamic Feedback to Summer Melt. Cryosphere 2015, 9, 197–215. [Google Scholar] [CrossRef]
  84. Cooper, M.G.; Smith, L.C.; Rennermalm, A.K.; Miège, C.; Pitcher, L.H.; Ryan, J.C.; Yang, K.; Cooley, S.W. Meltwater Storage in Low-Density near-Surface Bare Ice in the Greenland Ice Sheet Ablation Zone. Cryosphere 2018, 12, 955–970. [Google Scholar] [CrossRef]
  85. Thøgersen, K.; Gilbert, A.; Schuler, T.V.; Malthe-Sørenssen, A. Rate-and-State Friction Explains Glacier Surge Propagation. Nat. Commun. 2019, 10, 2823. [Google Scholar] [CrossRef] [PubMed]
  86. Harrison, W.D. Temperature of a Temperate Glacier. J. Glaciol. 1972, 11, 15–29. [Google Scholar] [CrossRef]
  87. Harrison, W.D.; Raymond, C.F. Impurities and Their Distribution in Temperate Glacier Ice. J. Glaciol. 1976, 16, 173–181. [Google Scholar] [CrossRef]
  88. Phillips, T.; Rajaram, H.; Steffen, K. Cryo-Hydrologic Warming: A Potential Mechanism for Rapid Thermal Response of Ice Sheets: Rapid Thermal Response of Ice Sheets. Geophys. Res. Lett. 2010, 37. [Google Scholar] [CrossRef]
  89. Pitcher, L.H.; Smith, L.C.; Gleason, C.J.; Miège, C.; Ryan, J.C.; Hagedorn, B.; As, D.; Chu, W.; Forster, R.R. Direct Observation of Winter Meltwater Drainage From the Greenland Ice Sheet. Geophys. Res. Lett. 2020, 47, e2019GL086521. [Google Scholar] [CrossRef]
  90. Hindshaw, R.S.; Tipper, E.T.; Reynolds, B.C.; Lemarchand, E.; Wiederhold, J.G.; Magnusson, J.; Bernasconi, S.M.; Kretzschmar, R.; Bourdon, B. Hydrological Control of Stream Water Chemistry in a Glacial Catchment (Damma Glacier, Switzerland). Chem. Geol. 2011, 285, 215–230. [Google Scholar] [CrossRef]
  91. Foreman, C.M.; Sattler, B.; Mikucki, J.A.; Porazinska, D.L.; Priscu, J.C. Metabolic Activity and Diversity of Cryoconites in the Taylor Valley, Antarctica: Activity in Antarctic Cryoconites. J. Geophys. Res. 2007, 112, G04S32. [Google Scholar] [CrossRef]
  92. Forster, R.R.; Box, J.E.; van den Broeke, M.R.; Miège, C.; Burgess, E.W.; van Angelen, J.H.; Lenaerts, J.T.M.; Koenig, L.S.; Paden, J.; Lewis, C.; et al. Extensive Liquid Meltwater Storage in Firn within the Greenland Ice Sheet. Nat. Geosci. 2014, 7, 95–98. [Google Scholar] [CrossRef]
  93. Koenig, L.S.; Miège, C.; Forster, R.R.; Brucker, L. Initial in Situ Measurements of Perennial Meltwater Storage in the Greenland Firn Aquifer: Measurements of Greenland Aquifer. Geophys. Res. Lett. 2014, 41, 81–85. [Google Scholar] [CrossRef]
  94. Chu, W.; Schroeder, D.M.; Siegfried, M.R. Retrieval of Englacial Firn Aquifer Thickness From Ice-Penetrating Radar Sounding in Southeastern Greenland. Geophys. Res. Lett. 2018, 45, 11770–11778. [Google Scholar] [CrossRef]
  95. Miller, O.; Solomon, D.K.; Miège, C.; Koenig, L.; Forster, R.; Schmerr, N.; Ligtenberg, S.R.M.; Legchenko, A.; Voss, C.I.; Montgomery, L.; et al. Hydrology of a Perennial Firn Aquifer in Southeast Greenland: An Overview Driven by Field Data. Water Resour. Res. 2020, 56, e2019WR026348. [Google Scholar] [CrossRef]
  96. Seaberg, S.Z.; Seaberg, J.Z.; Hooke, R.L.; Wiberg, D.W. Character of the Englacial and Subglacial Drainage System in the Lower Part of the Ablation Area of Storglaciären, Sweden, as Revealed by Dye-Trace Studies. J. Glaciol. 1988, 34, 217–227. [Google Scholar] [CrossRef]
  97. Nienow, P.; Sharp, M.; Willis, I. Temporal Switching Between Englacial and Subglacial Drainage Pathways: Dye Tracer Evidence from the Haut Glacier D’arolla, Switzerland. Geogr. Ann. Ser. A Phys. Geogr. 1996, 78, 51–60. [Google Scholar] [CrossRef]
  98. Nye, J.F. The Geometry of Water Veins and Nodes in Polycrystalline Ice. J. Glaciol. 1989, 35, 17–22. [Google Scholar] [CrossRef]
  99. Fountain, A.G. Geometry and Flow Conditions of Subglacial Water at South Cascade Glacier, Washington State, U.S.A.; an Analysis of Tracer Injections. J. Glaciol. 1993, 39, 143–156. [Google Scholar] [CrossRef]
  100. Werder, M.A.; Loye, A.; Funk, M. Dye Tracing a Jökulhlaup: I. Subglacial Water Transit Speed and Water-Storage Mechanism. J. Glaciol. 2009, 55, 889–898. [Google Scholar] [CrossRef]
  101. Bartholomew, I.; Nienow, P.; Mair, D.; Hubbard, A.; King, M.A.; Sole, A. Seasonal Evolution of Subglacial Drainage and Acceleration in a Greenland Outlet Glacier. Nat. Geosci. 2010, 3, 408–411. [Google Scholar] [CrossRef]
  102. Wingham, D.J.; Siegert, M.J.; Shepherd, A.; Muir, A.S. Rapid Discharge Connects Antarctic Subglacial Lakes. Nature 2006, 440, 1033–1036. [Google Scholar] [CrossRef]
  103. Tranter, M.; Fountain, A.G.; Fritsen, C.H.; Berry Lyons, W.; Priscu, J.C.; Statham, P.J.; Welch, K.A. Extreme Hydrochemical Conditions in Natural Microcosms Entombed within Antarctic Ice. Hydrol. Process. 2004, 18, 379–387. [Google Scholar] [CrossRef]
  104. Fountain, A.G. The Storage of Water in, and Hydraulic Characteristics of, the Firn of South Cascade Glacier, Washington State, USA. Ann. Glaciol. 1989, 13, 69–75. [Google Scholar] [CrossRef]
  105. van Wessem, J.M.; Steger, C.R.; Wever, N.; van den Broeke, M.R. Modelling Perennial Firn Aquifers in the Antarctic Peninsula (1979–2016); Snow/Snow Physics: Seattle, WA, USA, 2020. [Google Scholar]
  106. Barletta, R.E.; Priscu, J.C.; Mader, H.M.; Jones, W.L.; Roe, C.H. Chemical Analysis of Ice Vein Microenvironments: II. Analysis of Glacial Samples from Greenland and Antarctica. J. Glaciol. 2012, 58, 1109–1118. [Google Scholar] [CrossRef]
  107. Neff, J.C.; Holland, E.A.; Dentener, F.J.; McDowell, W.H.; Russell, K.M. The Origin, Composition and Rates of Organic Nitrogen Deposition: A Missing Piece of the Nitrogen Cycle? In The Nitrogen. Cycle at Regional to Global Scales; Boyer, E.W., Howarth, R.W., Eds.; Springer: Dordrecht, The Netherlands, 2002; pp. 99–136. ISBN 978-90-481-6086-0. [Google Scholar]
  108. Hodson, A.J.; Mumford, P.N.; Kohler, J.; Wynn, P.M. The High Arctic Glacial Ecosystem: New Insights from Nutrient Budgets. Biogeochemistry 2005, 72, 233–256. [Google Scholar] [CrossRef]
  109. Zhang, Q.; Huang, J.; Wang, F.; Mark, L.; Xu, J.; Armstrong, D.; Li, C.; Zhang, Y.; Kang, S. Mercury Distribution and Deposition in Glacier Snow over Western China. Environ. Sci. Technol. 2012, 46, 5404–5413. [Google Scholar] [CrossRef] [PubMed]
  110. Björkman, M.; Kühnel, R.; Partridge, D.; Roberts, T.; Aas, W.; Mazzola, M.; Viola, A.; Hodson, A.; Ström, J.; Isaksson, E. Nitrate Dry Deposition in Svalbard. Tellus B Chem. Phys. Meteorol. 2013, 65, 19071. [Google Scholar] [CrossRef]
  111. McCutcheon, J.; Lutz, S.; Williamson, C.; Cook, J.M.; Tedstone, A.J.; Vanderstraeten, A.; Wilson, S.A.; Stockdale, A.; Bonneville, S.; Anesio, A.M.; et al. Mineral Phosphorus Drives Glacier Algal Blooms on the Greenland Ice Sheet. Nat. Commun. 2021, 12, 570. [Google Scholar] [CrossRef]
  112. Nanus, L.; Campbell, D.H.; Ingersoll, G.P.; Clow, D.W.; Alisa Mast, M. Atmospheric Deposition Maps for the Rocky Mountains. Atmos. Environ. 2003, 37, 4881–4892. [Google Scholar] [CrossRef]
  113. Ginoux, P.; Prospero, J.M.; Gill, T.E.; Hsu, N.C.; Zhao, M. Global-Scale Attribution of Anthropogenic and Natural Dust Sources and Their Emission Rates Based on MODIS Deep Blue Aerosol Products: Anthropogenic and Natural Dust Sources. Rev. Geophys. 2012, 50, RG3005. [Google Scholar] [CrossRef]
  114. Kühnel, R.; Björkman, M.P.; Vega, C.P.; Hodson, A.; Isaksson, E.; Ström, J. Reactive Nitrogen and Sulphate Wet Deposition at Zeppelin Station, Ny-Ålesund, Svalbard. Polar Res. 2013, 32, 19136. [Google Scholar] [CrossRef]
  115. Fortner, S.K.; Lyons, W.B. Dissolved Trace and Minor Elements in Cryoconite Holes and Supraglacial Streams, Canada Glacier, Antarctica. Front. Earth Sci. 2018, 6, 31. [Google Scholar] [CrossRef]
  116. Stibal, M.; Lawson, E.C.; Lis, G.P.; Mak, K.M.; Wadham, J.L.; Anesio, A.M. Organic Matter Content and Quality in Supraglacial Debris across the Ablation Zone of the Greenland Ice Sheet. Ann. Glaciol. 2010, 51, 1–8. [Google Scholar] [CrossRef]
  117. Takeuchi, N.; Fujisawa, Y.; Kadota, T.; Tanaka, S.; Miyairi, M.; Shirakawa, T.; Kusaka, R.; Fedorov, A.N.; Konstantinov, P.; Ohata, T. The Effect of Impurities on the Surface Melt of a Glacier in the Suntar-Khayata Mountain Range, Russian Siberia. Front. Earth Sci. 2015, 3. [Google Scholar] [CrossRef]
  118. Stibal, M.; Bradley, J.A.; Box, J.E. Ecological Modeling of the Supraglacial Ecosystem: A Process-Based Perspective. Front. Earth Sci. 2017, 5, 52. [Google Scholar] [CrossRef]
  119. Tedstone, A.J.; Bamber, J.L.; Cook, J.M.; Williamson, C.J.; Fettweis, X.; Hodson, A.J.; Tranter, M. Dark Ice Dynamics of the South-West Greenland Ice Sheet. Cryosphere 2017, 11, 2491–2506. [Google Scholar] [CrossRef]
  120. Franzetti, A.; Tatangelo, V.; Gandolfi, I.; Bertolini, V.; Bestetti, G.; Diolaiuti, G.; D’Agata, C.; Mihalcea, C.; Smiraglia, C.; Ambrosini, R. Bacterial Community Structure on Two Alpine Debris-Covered Glaciers and Biogeography of Polaromonas Phylotypes. ISME J. 2013, 7, 1483–1492. [Google Scholar] [CrossRef]
  121. Rime, T.; Hartmann, M.; Frey, B. Potential Sources of Microbial Colonizers in an Initial Soil Ecosystem after Retreat of an Alpine Glacier. ISME J. 2016, 10, 1625–1641. [Google Scholar] [CrossRef]
  122. Darcy, J.L.; King, A.J.; Gendron, E.M.S.; Schmidt, S.K. Spatial Autocorrelation of Microbial Communities atop a Debris-Covered Glacier Is Evidence of a Supraglacial Chronosequence. FEMS Microbiol. Ecol. 2017, 93. [Google Scholar] [CrossRef]
  123. Liu, Y.; Vick-Majors, T.J.; Priscu, J.C.; Yao, T.; Kang, S.; Liu, K.; Cong, Z.; Xiong, J.; Li, Y. Biogeography of Cryoconite Bacterial Communities on Glaciers of the Tibetan Plateau. FEMS Microbiol. Ecol. 2017, 93. [Google Scholar] [CrossRef]
  124. Lutz, S.; Anesio, A.M.; Edwards, A.; Benning, L.G. Linking Microbial Diversity and Functionality of Arctic Glacial Surface Habitats: Arctic Glacial Surface Habitats. Environ. Microbiol. 2017, 19, 551–565. [Google Scholar] [CrossRef]
  125. Segawa, T.; Yonezawa, T.; Edwards, A.; Akiyoshi, A.; Tanaka, S.; Uetake, J.; Irvine-Fynn, T.; Fukui, K.; Li, Z.; Takeuchi, N. Biogeography of Cryoconite Forming Cyanobacteria on Polar and Asian Glaciers. J. Biogeogr. 2017, 44, 2849–2861. [Google Scholar] [CrossRef]
  126. Havig, J.R.; Hamilton, T.L. Snow Algae Drive Productivity and Weathering at Volcanic Rock-Hosted Glaciers. Geochim. Et Cosmochim. Acta 2019, 247, 220–242. [Google Scholar] [CrossRef]
  127. Kim, K.; Choi, W.; Hoffmann, M.R.; Yoon, H.-I.; Park, B.-K. Photoreductive Dissolution of Iron Oxides Trapped in Ice and Its Environmental Implications. Environ. Sci. Technol. 2010, 44, 4142–4148. [Google Scholar] [CrossRef] [PubMed]
  128. Holland, A.T.; Williamson, C.J.; Sgouridis, F.; Tedstone, A.J.; McCutcheon, J.; Cook, J.M.; Poniecka, E.; Yallop, M.L.; Tranter, M.; Anesio, A.M.; et al. Dissolved Organic Nutrients Dominate Melting Surface Ice of the Dark Zone (Greenland Ice Sheet). Biogeosciences 2019, 16, 3283–3296. [Google Scholar] [CrossRef]
  129. Wharton, R.A.; McKay, C.P.; Simmons, G.M.; Parker, B.C. Cryoconite Holes on Glaciers. BioScience 1985, 35, 499–503. [Google Scholar] [CrossRef]
  130. Bagshaw, E.A.; Tranter, M.; Fountain, A.G.; Welch, K.; Basagic, H.J.; Lyons, W.B. Do Cryoconite Holes Have the Potential to Be Significant Sources of C, N, and P to Downstream Depauperate Ecosystems of Taylor Valley, Antarctica? Arct. Antarct. Alp. Res. 2013, 45, 440–454. [Google Scholar] [CrossRef]
  131. Bagshaw, E.A.; Tranter, M.; Wadham, J.L.; Fountain, A.G.; Mowlem, M. High-Resolution Monitoring Reveals Dissolved Oxygen Dynamics in an Antarctic Cryoconite Hole. Hydrol. Process. 2011, 25, 2868–2877. [Google Scholar] [CrossRef]
  132. Poniecka, E.A.; Bagshaw, E.A.; Tranter, M.; Sass, H.; Williamson, C.J.; Anesio, A.M.; Black and Bloom Team. Rapid Development of Anoxic Niches in Supraglacial Ecosystems. Arct. Antarct. Alp. Res. 2018, 50, S100015. [Google Scholar] [CrossRef]
  133. Williamson, C.J.; Cameron, K.A.; Cook, J.M.; Zarsky, J.D.; Stibal, M.; Edwards, A. Glacier Algae: A Dark Past and a Darker Future. Front. Microbiol. 2019, 10, 524. [Google Scholar] [CrossRef]
  134. Nash, M.V.; Anesio, A.M.; Barker, G.; Tranter, M.; Varliero, G.; Eloe-Fadrosh, E.A.; Nielsen, T.; Turpin-Jelfs, T.; Benning, L.G.; Sánchez-Baracaldo, P. Metagenomic Insights into Diazotrophic Communities across Arctic Glacier Forefields. FEMS Microbiol. Ecol. 2018, 94. [Google Scholar] [CrossRef]
  135. Varliero, G.; Anesio, A.M.; Barker, G.L.A. A Taxon-Wise Insight Into Rock Weathering and Nitrogen Fixation Functional Profiles of Proglacial Systems. Front. Microbiol. 2021, 12, 627437. [Google Scholar] [CrossRef]
  136. Telling, J.; Anesio, A.M.; Tranter, M.; Irvine-Fynn, T.; Hodson, A.; Butler, C.; Wadham, J. Nitrogen Fixation on Arctic Glaciers, Svalbard. J. Geophys. Res. 2011, 116, G03039. [Google Scholar] [CrossRef]
  137. Telling, J.; Stibal, M.; Anesio, A.M.; Tranter, M.; Nias, I.; Cook, J.; Bellas, C.; Lis, G.; Wadham, J.L.; Sole, A.; et al. Microbial Nitrogen Cycling on the Greenland Ice Sheet. Biogeosciences 2012, 9, 2431–2442. [Google Scholar] [CrossRef]
  138. Clark, S.C.; Barnes, R.T.; Oleksy, I.A.; Baron, J.S.; Hastings, M.G. Persistent Nitrate in Alpine Waters with Changing Atmospheric Deposition and Warming Trends. Environ. Sci. Technol. 2021, 55, 14946–14956. [Google Scholar] [CrossRef]
  139. Stibal, M.; Anesio, A.M.; Blues, C.J.D.; Tranter, M. Phosphatase Activity and Organic Phosphorus Turnover on a High Arctic Glacier. Biogeosciences 2009, 6, 913–922. [Google Scholar] [CrossRef]
  140. Stibal, M.; Tranter, M.; Telling, J.; Benning, L.G. Speciation, Phase Association and Potential Bioavailability of Phosphorus on a Svalbard Glacier. Biogeochemistry 2008, 90, 1–13. [Google Scholar] [CrossRef]
  141. Smith, H.J.; Foster, R.A.; McKnight, D.M.; Lisle, J.T.; Littmann, S.; Kuypers, M.M.M.; Foreman, C.M. Microbial Formation of Labile Organic Carbon in Antarctic Glacial Environments. Nat. Geosci. 2017, 10, 356–359. [Google Scholar] [CrossRef]
  142. Sanyal, A.; Antony, R.; Samui, G.; Thamban, M. Microbial Communities and Their Potential for Degradation of Dissolved Organic Carbon in Cryoconite Hole Environments of Himalaya and Antarctica. Microbiol. Res. 2018, 208, 32–42. [Google Scholar] [CrossRef]
  143. Andrews, M.G.; Jacobson, A.D.; Osburn, M.R.; Flynn, T.M. Dissolved Carbon Dynamics in Meltwaters From the Russell Glacier, Greenland Ice Sheet. J. Geophys. Res. Biogeosci. 2018, 123, 2922–2940. [Google Scholar] [CrossRef]
  144. Samui, G.; Antony, R.; Thamban, M. Chemical Characteristics of Hydrologically Distinct Cryoconite Holes in Coastal Antarctica. Ann. Glaciol. 2018, 59, 69–76. [Google Scholar] [CrossRef]
  145. Zhang, Y.; Gao, T.; Kang, S.; Sprenger, M.; Tao, S.; Du, W.; Yang, J.; Wang, F.; Meng, W. Effects of Black Carbon and Mineral Dust on Glacial Melting on the Muz Taw Glacier, Central Asia. Sci. Total Environ. 2020, 740, 140056. [Google Scholar] [CrossRef]
  146. Benning, L.G.; Anesio, A.M.; Lutz, S.; Tranter, M. Biological Impact on Greenland’s Albedo. Nat. Geosci. 2014, 7, 691. [Google Scholar] [CrossRef]
  147. Hotaling, S.; Lutz, S.; Dial, R.J.; Anesio, A.M.; Benning, L.G.; Fountain, A.G.; Kelley, J.L.; McCutcheon, J.; Skiles, S.M.; Takeuchi, N.; et al. Biological Albedo Reduction on Ice Sheets, Glaciers, and Snowfields. Earth-Sci. Rev. 2021, 220, 103728. [Google Scholar] [CrossRef]
  148. Hawkings, J.R.; Wadham, J.L.; Tranter, M.; Lawson, E.; Sole, A.; Cowton, T.; Tedstone, A.J.; Bartholomew, I.; Nienow, P.; Chandler, D.; et al. The Effect of Warming Climate on Nutrient and Solute Export from the Greenland Ice Sheet. Geochem. Persp. Let. 2015, 94–104. [Google Scholar] [CrossRef]
  149. Musilova, M.; Tranter, M.; Bennett, S.A.; Wadham, J.; Anesio, A.M. Stable Microbial Community Composition on the Greenland Ice Sheet. Front. Microbiol. 2015, 6, 193. [Google Scholar] [CrossRef] [PubMed]
  150. Varliero, G.; Holland, A.; Barker, G.L.A.; Yallop, M.L.; Fountain, A.G.; Anesio, A.M. Glacier Clear Ice Bands Indicate Englacial Channel Microbial Distribution. J. Glaciol. 2021, 67, 811–823. [Google Scholar] [CrossRef]
  151. Martinez-Alonso, E.; Pena-Perez, S.; Serrano, S.; Garcia-Lopez, E.; Alcazar, A.; Cid, C. Taxonomic and Functional Characterization of a Microbial Community from a Volcanic Englacial Ecosystem in Deception Island, Antarctica. Sci. Rep. 2019, 9, 12158. [Google Scholar] [CrossRef]
  152. Bakermans, C.; Skidmore, M.L. Microbial Metabolism in Ice and Brine at −5°C: Microbial Metabolism in Ice and Brine at −5°C. Environ. Microbiol. 2011, 13, 2269–2278. [Google Scholar] [CrossRef]
  153. Dani, K.G.S.; Mader, H.M.; Wolff, E.W.; Wadham, J.L. Modelling the Liquid-Water Vein System within Polar Ice Sheets as a Potential Microbial Habitat. Earth Planet. Sci. Lett. 2012, 333–334, 238–249. [Google Scholar] [CrossRef]
  154. Mader, H.M.; Pettitt, M.E.; Wadham, J.L.; Wolff, E.W.; Parkes, R.J. Subsurface Ice as a Microbial Habitat. Geology 2006, 34, 169. [Google Scholar] [CrossRef]
  155. Rohde, R.A.; Price, P.B. Diffusion-Controlled Metabolism for Long-Term Survival of Single Isolated Microorganisms Trapped within Ice Crystals. Proc. Natl. Acad. Sci. USA 2007, 104, 16592–16597. [Google Scholar] [CrossRef]
  156. Li, Y.; Cha, Q.-Q.; Dang, Y.-R.; Chen, X.-L.; Wang, M.; McMinn, A.; Espina, G.; Zhang, Y.-Z.; Blamey, J.M.; Qin, Q.-L. Reconstruction of the Functional Ecosystem in the High Light, Low Temperature Union Glacier Region, Antarctica. Front. Microbiol. 2019, 10, 2408. [Google Scholar] [CrossRef]
  157. Mitchell, A.C.; Lafrenière, M.J.; Skidmore, M.L.; Boyd, E.S. Influence of Bedrock Mineral Composition on Microbial Diversity in a Subglacial Environment. Geology 2013, 41, 855–858. [Google Scholar] [CrossRef]
  158. Skidmore, M.; Anderson, S.P.; Sharp, M.; Foght, J.; Lanoil, B.D. Comparison of Microbial Community Compositions of Two Subglacial Environments Reveals a Possible Role for Microbes in Chemical Weathering Processes. Appl. Environ. Microbiol. 2005, 71, 6986–6997. [Google Scholar] [CrossRef]
  159. Tranter, M.; Sharp, M.J.; Lamb, H.R.; Brown, G.H.; Hubbard, B.P.; Willis, I.C. Geochemical Weathering at the Bed of Haut Glacier d’Arolla, Switzerland?A New Model. Hydrol. Process. 2002, 16, 959–993. [Google Scholar] [CrossRef]
  160. Anderson, S.P. Glaciers Show Direct Linkage between Erosion Rate and Chemical Weathering Fluxes. Geomorphology 2005, 67, 147–157. [Google Scholar] [CrossRef]
  161. Macdonald, M.L.; Wadham, J.L.; Telling, J.; Skidmore, M.L. Glacial Erosion Liberates Lithologic Energy Sources for Microbes and Acidity for Chemical Weathering Beneath Glaciers and Ice Sheets. Front. Earth Sci. 2018, 6, 212. [Google Scholar] [CrossRef]
  162. Thór Marteinsson, V.; Rúnarsson, Á.; Stefánsson, A.; Thorsteinsson, T.; Jóhannesson, T.; Magnússon, S.H.; Reynisson, E.; Einarsson, B.; Wade, N.; Morrison, H.G.; et al. Microbial Communities in the Subglacial Waters of the Vatnajökull Ice Cap, Iceland. ISME J. 2013, 7, 427–437. [Google Scholar] [CrossRef]
  163. Achberger, A.M.; Michaud, A.B.; Vick-Majors, T.J.; Christner, B.C.; Skidmore, M.L.; Priscu, J.C.; Tranter, M. Microbiology of Subglacial Environments. In Psychrophiles: From Biodiversity to Biotechnology; Margesin, R., Ed.; Springer International Publishing: Cham, Switzerland, 2017; pp. 83–110. ISBN 978-3-319-57057-0. [Google Scholar]
  164. Elberling, B. Temperature and Oxygen Control on Pyrite Oxidation in Frozen Mine Tailings. Cold Reg. Sci. Technol. 2005, 41, 121–133. [Google Scholar] [CrossRef]
  165. Nixon, S.L.; Telling, J.P.; Wadham, J.L.; Cockell, C.S. Viable Cold-Tolerant Iron-Reducing Microorganisms in Geographically Diverse Subglacial Environments. Biogeosciences 2017, 14, 1445–1455. [Google Scholar] [CrossRef]
  166. Hamilton, T.L.; Peters, J.W.; Skidmore, M.L.; Boyd, E.S. Molecular Evidence for an Active Endogenous Microbiome beneath Glacial Ice. ISME J. 2013, 7, 1402–1412. [Google Scholar] [CrossRef]
  167. Wynn, P.M.; Hodson, A.; Heaton, T. Chemical and Isotopic Switching within the Subglacial Environment of a High Arctic Glacier. Biogeochemistry 2006, 78, 173–193. [Google Scholar] [CrossRef]
  168. Wynn, P.M.; Hodson, A.J.; Heaton, T.H.E.; Chenery, S.R. Nitrate Production beneath a High Arctic Glacier, Svalbard. Chem. Geol. 2007, 244, 88–102. [Google Scholar] [CrossRef]
  169. Liermann, L.J.; Hausrath, E.M.; Anbar, A.D.; Brantley, S.L. Assimilatory and Dissimilatory Processes of Microorganisms Affecting Metals in the Environment. J. Anal. At. Spectrom. 2007, 22, 867. [Google Scholar] [CrossRef]
  170. Muyzer, G.; Stams, A.J.M. The Ecology and Biotechnology of Sulphate-Reducing Bacteria. Nat. Rev. Microbiol. 2008, 6, 441–454. [Google Scholar] [CrossRef] [PubMed]
  171. Stibal, M.; Wadham, J.L.; Lis, G.P.; Telling, J.; Pancost, R.D.; Dubnick, A.; Sharp, M.J.; Lawson, E.C.; Butler, C.E.H.; Hasan, F.; et al. Methanogenic Potential of Arctic and Antarctic Subglacial Environments with Contrasting Organic Carbon Sources. Glob. Change Biol. 2012, 18, 3332–3345. [Google Scholar] [CrossRef]
  172. Anesio, A.M.; Hodson, A.J.; Fritz, A.; Psenner, R.; Sattler, B. High Microbial Activity on Glaciers: Importance to the Global Carbon Cycle. Glob. Chang. Biol. 2009, 15, 955–960. [Google Scholar] [CrossRef]
  173. Price, P.B. A Habitat for Psychrophiles in Deep Antarctic Ice. Proc. Natl. Acad. Sci. USA 2000, 97, 1247–1251. [Google Scholar] [CrossRef]
  174. Mikucki, J.A.; Pearson, A.; Johnston, D.T.; Turchyn, A.V.; Farquhar, J.; Schrag, D.P.; Anbar, A.D.; Priscu, J.C.; Lee, P.A. A Contemporary Microbially Maintained Subglacial Ferrous "Ocean". ” Science 2009, 324, 397–400. [Google Scholar] [CrossRef]
  175. Lyons, W.B.; Mikucki, J.A.; German, L.A.; Welch, K.A.; Welch, S.A.; Gardner, C.B.; Tulaczyk, S.M.; Pettit, E.C.; Kowalski, J.; Dachwald, B. The Geochemistry of Englacial Brine From Taylor Glacier, Antarctica. J. Geophys. Res. Biogeosci. 2019, 124, 633–648. [Google Scholar] [CrossRef]
  176. Vick-Majors, T.J.; Achberger, A.M.; Michaud, A.B.; Priscu, J.C. Metabolic and Taxonomic Diversity in Antarctic Subglacial Environments. In Life in Extreme Environments; di Prisco, G., Edwards, H.G.M., Elster, J., Huiskes, A.H.L., Eds.; Cambridge University Press: Cambridge, UK, 2020; pp. 279–296. ISBN 978-1-108-68331-9. [Google Scholar]
  177. Telling, J.; Boyd, E.S.; Bone, N.; Jones, E.L.; Tranter, M.; MacFarlane, J.W.; Martin, P.G.; Wadham, J.L.; Lamarche-Gagnon, G.; Skidmore, M.L.; et al. Rock Comminution as a Source of Hydrogen for Subglacial Ecosystems. Nat. Geosci. 2015, 8, 851–855. [Google Scholar] [CrossRef]
  178. Greening, C.; Biswas, A.; Carere, C.R.; Jackson, C.J.; Taylor, M.C.; Stott, M.B.; Cook, G.M.; Morales, S.E. Genomic and Metagenomic Surveys of Hydrogenase Distribution Indicate H2 Is a Widely Utilised Energy Source for Microbial Growth and Survival. ISME J. 2016, 10, 761–777. [Google Scholar] [CrossRef]
  179. Yang, Z.; Zhang, Y.; Lv, Y.; Yan, W.; Xiao, X.; Sun, B.; Ma, H. H2 Metabolism Revealed by Metagenomic Analysis of Subglacial Sediment from East Antarctica. J. Microbiol. 2019, 57, 1095–1104. [Google Scholar] [CrossRef]
  180. Holland, A.T.; Bergk Pinto, B.; Layton, R.; Williamson, C.J.; Anesio, A.M.; Vogel, T.M.; Larose, C.; Tranter, M. Over Winter Microbial Processes in a Svalbard Snow Pack: An Experimental Approach. Front. Microbiol. 2020, 11, 1029. [Google Scholar] [CrossRef]
  181. Kranzler, C.; Kessler, N.; Keren, N.; Shaked, Y. Enhanced Ferrihydrite Dissolution by a Unicellular, Planktonic Cyanobacterium: A Biological Contribution to Particulate Iron Bioavailability: Bio-Dissolution of Ferrihydrite by Cyanobacteria. Environ. Microbiol. 2016, 18, 5101–5111. [Google Scholar] [CrossRef]
  182. Koffman, B.G.; Yoder, M.F.; Methven, T.; Hanschka, L.; Sears, H.B.; Saylor, P.L.; Wallace, K.L. Glacial Dust Surpasses Both Volcanic Ash and Desert Dust in Its Iron Fertilization Potential. Glob. Biogeochem. Cycles 2021, 35. [Google Scholar] [CrossRef]
  183. Hood, E.; Battin, T.J.; Fellman, J.; O’Neel, S.; Spencer, R.G.M. Storage and Release of Organic Carbon from Glaciers and Ice Sheets. Nat. Geosci. 2015, 8, 91–96. [Google Scholar] [CrossRef]
  184. Kellerman, A.M.; Hawkings, J.R.; Wadham, J.L.; Kohler, T.J.; Stibal, M.; Grater, E.; Marshall, M.; Hatton, J.E.; Beaton, A.; Spencer, R.G.M. Glacier Outflow Dissolved Organic Matter as a Window Into Seasonally Changing Carbon Sources: Leverett Glacier, Greenland. J. Geophys. Res. Biogeosci. 2020, 125, e2019JG005161. [Google Scholar] [CrossRef]
  185. Feng, L.; An, Y.; Xu, J.; Kang, S. Characteristics and Sources of Dissolved Organic Matter in a Glacier in the Northern Tibetan Plateau: Differences between Different Snow Categories. Ann. Glaciol. 2018, 59, 31–40. [Google Scholar] [CrossRef]
  186. Säwström, C.; Mumford, P.; Marshall, W.; Hodson, A.; Laybourn-Parry, J. The Microbial Communities and Primary Productivity of Cryoconite Holes in an Arctic Glacier (Svalbard 79°N). Polar Biol. 2002, 25, 591–596. [Google Scholar] [CrossRef]
  187. Mindl, B.; Anesio, A.M.; Meirer, K.; Hodson, A.J.; Laybourn-Parry, J.; Sommaruga, R.; Sattler, B. Factors Influencing Bacterial Dynamics along a Transect from Supraglacial Runoff to Proglacial Lakes of a High Arctic Glacieri: Bacterial Dynamics in a Glacial Environment. FEMS Microbiol. Ecol. 2007, 59, 307–317. [Google Scholar] [CrossRef]
  188. Irvine-Fynn, T.D.L.; Edwards, A.; Newton, S.; Langford, H.; Rassner, S.M.; Telling, J.; Anesio, A.M.; Hodson, A.J. Microbial Cell Budgets of an Arctic Glacier Surface Quantified Using Flow Cytometry. Environ. Microbiol. 2012, 14, 2998–3012. [Google Scholar] [CrossRef] [PubMed]
  189. Gaidos, E.; Marteinsson, V.; Thorsteinsson, T.; Jóhannesson, T.; Rúnarsson, Á.R.; Stefansson, A.; Glazer, B.; Lanoil, B.; Skidmore, M.; Han, S.; et al. An Oligarchic Microbial Assemblage in the Anoxic Bottom Waters of a Volcanic Subglacial Lake. ISME J. 2009, 3, 486–497. [Google Scholar] [CrossRef] [PubMed]
  190. Christner, B.C.; Priscu, J.C.; Achberger, A.M.; Barbante, C.; Carter, S.P.; Christianson, K.; Michaud, A.B.; Mikucki, J.A.; Mitchell, A.C.; Skidmore, M.L.; et al. A Microbial Ecosystem beneath the West Antarctic Ice Sheet. Nature 2014, 512, 310–313. [Google Scholar] [CrossRef] [PubMed]
  191. Mikucki, J.A.; Lee, P.A.; Ghosh, D.; Purcell, A.M.; Mitchell, A.C.; Mankoff, K.D.; Fisher, A.T.; Tulaczyk, S.; Carter, S.; Siegfried, M.R.; et al. Subglacial Lake Whillans Microbial Biogeochemistry: A Synthesis of Current Knowledge. Philos. Trans. R. Soc. A 2016, 374, 20140290. [Google Scholar] [CrossRef]
  192. Hood, E.; Fellman, J.; Spencer, R.G.M.; Hernes, P.J.; Edwards, R.; D’Amore, D.; Scott, D. Glaciers as a Source of Ancient and Labile Organic Matter to the Marine Environment. Nature 2009, 462, 1044–1047. [Google Scholar] [CrossRef]
  193. Bhatia, M.P.; Kujawinski, E.B.; Das, S.B.; Breier, C.F.; Henderson, P.B.; Charette, M.A. Greenland Meltwater as a Significant and Potentially Bioavailable Source of Iron to the Ocean. Nat. Geosci. 2013, 6, 274–278. [Google Scholar] [CrossRef]
  194. Hawkings, J.R.; Wadham, J.L.; Tranter, M.; Raiswell, R.; Benning, L.G.; Statham, P.J.; Tedstone, A.; Nienow, P.; Lee, K.; Telling, J. Ice Sheets as a Significant Source of Highly Reactive Nanoparticulate Iron to the Oceans. Nat. Commun. 2014, 5, 3929. [Google Scholar] [CrossRef]
  195. Hawkings, J.R.; Benning, L.G.; Raiswell, R.; Kaulich, B.; Araki, T.; Abyaneh, M.; Stockdale, A.; Koch-Müller, M.; Wadham, J.L.; Tranter, M. Biolabile Ferrous Iron Bearing Nanoparticles in Glacial Sediments. Earth Planet. Sci. Lett. 2018, 493, 92–101. [Google Scholar] [CrossRef]
  196. Hopwood, M.J.; Statham, P.J.; Tranter, M.; Wadham, J.L. Glacial Flours as a Potential Source of Fe(II) and Fe(III) to Polar Waters. Biogeochemistry 2014, 118, 443–452. [Google Scholar] [CrossRef]
  197. Kim, I.; Kim, G.; Choy, E.J. The Significant Inputs of Trace Elements and Rare Earth Elements from Melting Glaciers in Antarctic Coastal Waters. Polar Res. 2015, 34, 24289. [Google Scholar] [CrossRef]
  198. Tockner, K.; Malard, F.; Uehlinger, U.; Ward, J.V. Nutrients and Organic Matter in a Glacial River-Floodplain System (Val Roseg, Switzerland). Limnol. Oceanogr. 2002, 47, 266–277. [Google Scholar] [CrossRef]
  199. Warner, K.A.; Saros, J.E.; Simon, K.S. Nitrogen Subsidies in Glacial Meltwater: Implications for High Elevation Aquatic Chains: Nitrogen Subsidies in Glacial Meltwater. Water Resour. Res. 2017, 53, 9791–9806. [Google Scholar] [CrossRef]
  200. Martin, J.B.; Pain, A.J.; Martin, E.E.; Rahman, S.; Ackerman, P. Comparisons of Nutrients Exported From Greenlandic Glacial and Deglaciated Watersheds. Glob. Biogeochem. Cycles 2020, 34, e2020GB006661. [Google Scholar] [CrossRef]
  201. Miller, J.B.; Frisbee, M.D.; Hamilton, T.L.; Murugapiran, S.K. Recharge from Glacial Meltwater Is Critical for Alpine Springs and Their Microbiomes. Environ. Res. Lett. 2021, 16, 064012. [Google Scholar] [CrossRef]
  202. Paulsen, M.L.; Robson, B.A. Glaciers and Land-to-Ocean Flux of Carbon. In Encyclopedia of Water; Maurice, P., Ed.; Wiley: Hoboken, NJ, USA, 2019; pp. 1–12. ISBN 978-1-119-30075-5. [Google Scholar]
  203. Zumsteg, A.; Luster, J.; Göransson, H.; Smittenberg, R.H.; Brunner, I.; Bernasconi, S.M.; Zeyer, J.; Frey, B. Bacterial, Archaeal and Fungal Succession in the Forefield of a Receding Glacier. Microb. Ecol. 2012, 63, 552–564. [Google Scholar] [CrossRef]
  204. Alfaro, F.D.; Salazar-Burrows, A.; Bañales-Seguel, C.; García, J.-L.; Manzano, M.; Marquet, P.A.; Ruz, K.; Gaxiola, A. Soil Microbial Abundance and Activity across Forefield Glacier Chronosequence in the Northern Patagonian Ice Field, Chile. Arct. Antarct. Alp. Res. 2020, 52, 553–562. [Google Scholar] [CrossRef]
  205. Wojcik, R.; Eichel, J.; Bradley, J.A.; Benning, L.G. How Allogenic Factors Affect Succession in Glacier Forefields. Earth-Sci. Rev. 2021, 218, 103642. [Google Scholar] [CrossRef]
  206. Heckmann, T.; McColl, S.; Morche, D. Retreating Ice: Research in pro-Glacial Areas Matters: Research in pro-Glacial Areas. Earth Surf. Process. Landf. 2016, 41, 271–276. [Google Scholar] [CrossRef]
  207. Saros, J.E.; Rose, K.C.; Clow, D.W.; Stephens, V.C.; Nurse, A.B.; Arnett, H.A.; Stone, J.R.; Williamson, C.E.; Wolfe, A.P. Melting Alpine Glaciers Enrich High-Elevation Lakes with Reactive Nitrogen. Environ. Sci. Technol. 2010, 44, 4891–4896. [Google Scholar] [CrossRef]
  208. Fegel, T.S.; Baron, J.S.; Fountain, A.G.; Johnson, G.F.; Hall, E.K. The Differing Biogeochemical and Microbial Signatures of Glaciers and Rock Glaciers: Defining Glacier-Type Signatures. J. Geophys. Res. Biogeosci. 2016, 121, 919–932. [Google Scholar] [CrossRef]
  209. Slemmons, K.E.H.; Rodgers, M.L.; Stone, J.R.; Saros, J.E. Nitrogen Subsidies in Glacial Meltwaters Have Altered Planktonic Diatom Communities in Lakes of the US Rocky Mountains for at Least a Century. Hydrobiologia 2017, 800, 129–144. [Google Scholar] [CrossRef]
  210. Peter, H.; Sommaruga, R. Shifts in Diversity and Function of Lake Bacterial Communities upon Glacier Retreat. ISME J. 2016, 10, 1545–1554. [Google Scholar] [CrossRef] [PubMed]
  211. Zhang, R.; John, S.G.; Zhang, J.; Ren, J.; Wu, Y.; Zhu, Z.; Liu, S.; Zhu, X.; Marsay, C.M.; Wenger, F. Transport and Reaction of Iron and Iron Stable Isotopes in Glacial Meltwaters on Svalbard near Kongsfjorden: From Rivers to Estuary to Ocean. Earth Planet. Sci. Lett. 2015, 424, 201–211. [Google Scholar] [CrossRef]
  212. Chifflard, P.; Fasching, C.; Reiss, M.; Ditzel, L.; Boodoo, K.S. Dissolved and Particulate Organic Carbon in Icelandic Proglacial Streams: A First Estimate. Water 2019, 11, 748. [Google Scholar] [CrossRef]
  213. Hodson, A.; Nowak, A.; Sabacka, M.; Jungblut, A.; Navarro, F.; Pearce, D.; Ávila-Jiménez, M.L.; Convey, P.; Vieira, G. Climatically Sensitive Transfer of Iron to Maritime Antarctic Ecosystems by Surface Runoff. Nat. Commun. 2017, 8, 14499. [Google Scholar] [CrossRef]
  214. Meire, L.; Mortensen, J.; Meire, P.; Juul-Pedersen, T.; Sejr, M.K.; Rysgaard, S.; Nygaard, R.; Huybrechts, P.; Meysman, F.J.R. Marine-Terminating Glaciers Sustain High Productivity in Greenland Fjords. Glob. Chang. Biol. 2017, 23, 5344–5357. [Google Scholar] [CrossRef]
  215. Torsvik, T.; Albretsen, J.; Sundfjord, A.; Kohler, J.; Sandvik, A.D.; Skarðhamar, J.; Lindbäck, K.; Everett, A. Impact of Tidewater Glacier Retreat on the Fjord System: Modeling Present and Future Circulation in Kongsfjorden, Svalbard. Estuar. Coast. Shelf Sci. 2019, 220, 152–165. [Google Scholar] [CrossRef]
  216. Hopwood, M.J.; Carroll, D.; Dunse, T.; Hodson, A.; Holding, J.M.; Iriarte, J.L.; Ribeiro, S.; Achterberg, E.P.; Cantoni, C.; Carlson, D.F.; et al. Review Article: How Does Glacier Discharge Affect Marine Biogeochemistry and Primary Production in the Arctic? Cryosphere 2020, 14, 1347–1383. [Google Scholar] [CrossRef]
  217. Stuart-Lee, A.E.; Mortensen, J.; van der Kaaden, A.-S.; Meire, L. Seasonal Hydrography of Ameralik: A Southwest Greenland Fjord Impacted by a Land-Terminating Glacier. J. Geophys. Res. Ocean. 2021, 126, e2021JC017552. [Google Scholar] [CrossRef]
  218. Lyons, W.B.; Dailey, K.R.; Welch, K.A.; Deuerling, K.M.; Welch, S.A.; McKnight, D.M. Antarctic Streams as a Potential Source of Iron for the Southern Ocean. Geology 2015, 43, 1003–1006. [Google Scholar] [CrossRef]
  219. Bhatia, M.P.; Waterman, S.; Burgess, D.O.; Williams, P.L.; Bundy, R.M.; Mellett, T.; Roberts, M.; Bertrand, E.M. Glaciers and Nutrients in the Canadian Arctic Archipelago Marine System. Glob. Biogeochem. Cycles 2021, 35, e2021GB006976. [Google Scholar] [CrossRef]
  220. Williams, P.L.; Burgess, D.O.; Waterman, S.; Roberts, M.; Bertrand, E.M.; Bhatia, M.P. Nutrient and Carbon Export From a Tidewater Glacier to the Coastal Ocean in the Canadian Arctic Archipelago. JGR Biogeosci. 2021, 126, e2021JG006289. [Google Scholar] [CrossRef]
  221. Hopwood, M.J.; Bacon, S.; Arendt, K.; Connelly, D.P.; Statham, P.J. Glacial Meltwater from Greenland Is Not Likely to Be an Important Source of Fe to the North Atlantic. Biogeochemistry 2015, 124, 1–11. [Google Scholar] [CrossRef]
  222. Cape, M.R.; Straneo, F.; Beaird, N.; Bundy, R.M.; Charette, M.A. Nutrient Release to Oceans from Buoyancy-Driven Upwelling at Greenland Tidewater Glaciers. Nat. Geosci. 2019, 12, 34–39. [Google Scholar] [CrossRef]
  223. Halbach, L.; Vihtakari, M.; Duarte, P.; Everett, A.; Granskog, M.A.; Hop, H.; Kauko, H.M.; Kristiansen, S.; Myhre, P.I.; Pavlov, A.K.; et al. Tidewater Glaciers and Bedrock Characteristics Control the Phytoplankton Growth Environment in a Fjord in the Arctic. Front. Mar. Sci. 2019, 6, 254. [Google Scholar] [CrossRef]
  224. Tremblay, J.-É.; Anderson, L.G.; Matrai, P.; Coupel, P.; Bélanger, S.; Michel, C.; Reigstad, M. Global and Regional Drivers of Nutrient Supply, Primary Production and CO2 Drawdown in the Changing Arctic Ocean. Progress. Oceanogr. 2015, 139, 171–196. [Google Scholar] [CrossRef]
  225. Hopwood, M.J.; Carroll, D.; Browning, T.J.; Meire, L.; Mortensen, J.; Krisch, S.; Achterberg, E.P. Non-Linear Response of Summertime Marine Productivity to Increased Meltwater Discharge around Greenland. Nat. Commun. 2018, 9, 3256. [Google Scholar] [CrossRef]
  226. Lam, P.J.; Bishop, J.K.B. The Continental Margin Is a Key Source of Iron to the HNLC North Pacific Ocean: Continental Margin and the North Pacific. Geophys. Res. Lett. 2008, 35, L07608. [Google Scholar] [CrossRef]
  227. Laufkötter, C.; Stern, A.A.; John, J.G.; Stock, C.A.; Dunne, J.P. Glacial Iron Sources Stimulate the Southern Ocean Carbon Cycle. Geophys. Res. Lett. 2018, 45, 13377–13385. [Google Scholar] [CrossRef]
  228. Arrigo, K.R.; Dijken, G.L.; Castelao, R.M.; Luo, H.; Rennermalm, Å.K.; Tedesco, M.; Mote, T.L.; Oliver, H.; Yager, P.L. Melting Glaciers Stimulate Large Summer Phytoplankton Blooms in Southwest Greenland Waters. Geophys. Res. Lett. 2017, 44, 6278–6285. [Google Scholar] [CrossRef]
  229. Cauvy-Fraunié, S.; Dangles, O. A Global Synthesis of Biodiversity Responses to Glacier Retreat. Nat. Ecol. Evol. 2019, 3, 1675–1685. [Google Scholar] [CrossRef] [PubMed]
  230. Hopwood, M.J.; Connelly, D.P.; Arendt, K.E.; Juul-Pedersen, T.; Stinchcombe, M.C.; Meire, L.; Esposito, M.; Krishna, R. Seasonal Changes in Fe along a Glaciated Greenlandic Fjord. Front. Earth Sci. 2016, 4. [Google Scholar] [CrossRef]
  231. Lewis, S.M.; Smith, L.C. Hydrologic Drainage of the Greenland Ice Sheet. Hydrol. Process. 2009, 23, 2004–2011. [Google Scholar] [CrossRef]
  232. Ding, Y.; Zhang, S.; Zhao, L.; Li, Z.; Kang, S. Global Warming Weakening the Inherent Stability of Glaciers and Permafrost. Sci. Bull. 2019, 64, 245–253. [Google Scholar] [CrossRef]
  233. Fransson, A.; Chierici, M.; Nomura, D.; Granskog, M.A.; Kristiansen, S.; Martma, T.; Nehrke, G. Effect of Glacial Drainage Water on the CO2 System and Ocean Acidification State in an Arctic Rctic Tidewater-glacier Fjord during Two Contrasting Years. J. Geophys. Res. Ocean. 2015, 120, 2413–2429. [Google Scholar] [CrossRef]
  234. Murray, C.; Markager, S.; Stedmon, C.A.; Juul-Pedersen, T.; Sejr, M.K.; Bruhn, A. The Influence of Glacial Melt Water on Bio-Optical Properties in Two Contrasting Greenlandic Fjords. Estuar. Coast. Shelf Sci. 2015, 163, 72–83. [Google Scholar] [CrossRef]
  235. Buongiorno, J.; Herbert, L.C.; Wehrmann, L.M.; Michaud, A.B.; Laufer, K.; Røy, H.; Jørgensen, B.B.; Szynkiewicz, A.; Faiia, A.; Yeager, K.M.; et al. Complex Microbial Communities Drive Iron and Sulfur Cycling in Arctic Fjord Sediments. Appl. Environ. Microbiol. 2019, 85, e00949-19. [Google Scholar] [CrossRef]
  236. Cantoni, C.; Hopwood, M.J.; Clarke, J.S.; Chiggiato, J.; Achterberg, E.P.; Cozzi, S. Glacial Drivers of Marine Biogeochemistry Indicate a Future Shift to More Corrosive Conditions in an Arctic Fjord. J. Geophys. Res. Biogeosci. 2020, 125. [Google Scholar] [CrossRef]
  237. Jørgensen, B.B.; Laufer, K.; Michaud, A.B.; Wehrmann, L.M. Biogeochemistry and Microbiology of High Arctic Marine Sediment Ecosystems—Case Study of Svalbard Fjords. Limnol. Oceanogr. 2021, 66, S273–S292. [Google Scholar] [CrossRef]
  238. Szeligowska, M.; Trudnowska, E.; Boehnke, R.; Dąbrowska, A.M.; Dragańska-Deja, K.; Deja, K.; Darecki, M.; Błachowiak-Samołyk, K. The Interplay between Plankton and Particles in the Isfjorden Waters Influenced by Marine- and Land-Terminating Glaciers. Sci. Total. Environ. 2021, 780, 146491. [Google Scholar] [CrossRef]
Figure 1. Scheme of hydrological, nutrient, and microbial metabolism shifts during ablation and accumulation season. Width and height of the colored shapes represent the occurrence and the intensity of each specific process. X-axis represents time (accumulation and ablation seasons), whereas the y-axis represents different glacial realms (supraglacial, englacial, and subglacial environments). Water release from weathering crust, cryoconite holes, and especially firn aquifers is not linear as it also goes through daily and irregular cycles. The proposed trends of hydrological [1,36,55,64], nutrient [13,65,66,67], and microbial metabolism [11,17,18,68] shifts are broadly based on published peer-reviewed studies.
Figure 1. Scheme of hydrological, nutrient, and microbial metabolism shifts during ablation and accumulation season. Width and height of the colored shapes represent the occurrence and the intensity of each specific process. X-axis represents time (accumulation and ablation seasons), whereas the y-axis represents different glacial realms (supraglacial, englacial, and subglacial environments). Water release from weathering crust, cryoconite holes, and especially firn aquifers is not linear as it also goes through daily and irregular cycles. The proposed trends of hydrological [1,36,55,64], nutrient [13,65,66,67], and microbial metabolism [11,17,18,68] shifts are broadly based on published peer-reviewed studies.
Microorganisms 11 01153 g001
Figure 2. Water fluxes during the ablation season and main hotspots for microbial metabolisms. Subglacial lakes (not represented in this figure) can also represent hotspots for microbial activity.
Figure 2. Water fluxes during the ablation season and main hotspots for microbial metabolisms. Subglacial lakes (not represented in this figure) can also represent hotspots for microbial activity.
Microorganisms 11 01153 g002
Figure 3. Land-terminating glaciers (A), marine-terminating glaciers (B), and main consequences of ice retreat of proglacial systems (C).
Figure 3. Land-terminating glaciers (A), marine-terminating glaciers (B), and main consequences of ice retreat of proglacial systems (C).
Microorganisms 11 01153 g003
Table 1. Water availability and residence times in glacial/ice sheet compartments.
Table 1. Water availability and residence times in glacial/ice sheet compartments.
LocationWater Presence in Accumulation SeasonResidence Time and Water FlowImportant References
Ice-lidded cryoconite holesYes, but some can completely freeze during accumulation seasonYears, but occasionally connected to the rest of the system during accumulation season. Estimates say that ~50% of the cryoconite holes hydrologically connect to the supraglacial system every accumulation season.[61,91]
Open cryoconite holesNoDuring accumulation season, they have a higher connectivity to the glacial system than ice-lidded cryoconite holes, showing lower permanence times.[39,63]
Weathering crustYes, if insulated by a snowpack.Several days, water is released to the glacial when the system is saturated. Water flow is slow in the interstitial space.[55,58,60]
Supraglacial streamsNoDepending on the ablation stage of the system, water can flow at different rates.[64]
Supraglacial ponds and lakesYes. Lakes can form ice lids and maintain water during accumulation season.Years; water is released to the glacial when the system is saturated. Smaller ponds can be drained by meltwater streams.[39,50,61,64]
Firn aquiferYesPerennial; water is released to the glacial when the system is saturated.[52,92,93,94,95]
Englacial conduitsYes, when the presence of solutes and particles lowers water freezing temperature.Fast-flowing systems with permanence of hours up to a day. However, water can be present all year round and even for multiple years if water pockets are formed by collapsed conduits. Depending on the ablation stage of the system, water can flow at different rates.[88,96,97]
Ice veinsYesResidence times in the ice veins are unknown. However, due to the low water flowing rate, we assume it to be in the order of years (at the very least).[98]
Subglacial cavitiesYesDays to months, and water is usually present all year round.[99,100,101]
Subglacial channelsNoHours, water is usually present only during peak ablation season.[99,100,101]
Subglacial tillYesPotentially all-year round.[36,78]
Subglacial lakesYesYears; some systems have estimated water residence times of millions of years.[79,80,102]
Table 2. Cell concentration in different water environments. VLP: Viral-Like Particles.
Table 2. Cell concentration in different water environments. VLP: Viral-Like Particles.
Cell ConcentrationSourceSampling Ablation SeasonReference
1.0–4.5 × 104 cells mL−1 (3.97–12.7 × 104 VLP mL−1)Cryoconite holes, Midtre Lovénbreen2000 and 2001[186]
1.38 × 104–4.84 × 104 cells mL−1Supraglacial meltwater runoff and cryoconite holes, Midtre Lovénbreen2004[187]
5.4 ± 1.6 × 104 cells mL−1Cryoconite holes, Austre Broggerbreen2005[31]
3.4 ± 1.2 × 104 cells mL−1Cryoconite holes, Midtre Lovénbreen2005[31]
4.1 ± 3.8 × 104 cells mL−1Cryoconite holes, Rotmoosferner2004[31]
3.7 ± 1.4 × 104 cells mL−1Cryoconite holes, Stubacher Sonnblickkees2007[31]
1.3 ± 8.2 × 104 cells mL−1Cryoconite holes, blue ice close to Patriot Hills2002[31]
4.4 ± 2.4 × 104 cells mL−1Cryoconite holes, Canada, Commonwealth, and Taylor glaciers2005[31]
2 × 104 cells mL−1Supraglacial meltwater runoff, Midtre Lovénbreen2010[188]
8.38 × 103 ± 9.85 × 103 cells mL−1Supraglacial meltwater runoff, Russell glacier2012[6]
2.2 × 104 ± 5.5 × 104 cells mL−1Weathering crust, Northern Hemisphere glaciers2014, 2015 and 2016[29]
6 × 104 cells ml−1Subglacial brine, Blood Falls2004[174]
106–108 cells mL−1Ice vein water/[154]
4.7–5.7 × 105 cells mL−1Subglacial water, Skaftá subglacial lake2006[189]
4.4 ± 2.2 × 105 cells mL−1Subglacial water, East Skaftárkatlar subglacial lakes2007[162]
1.3 × 105 cells mL−1Subglacial water, subglacial lake Whillans2013[190]
1.3 × 105 cells mL−1Subglacial water, subglacial lake Whillans2013[191]
1.15 × 105 ± 1.38 × 105 cells mL−1Subglacial meltwater runoff, Leverett glacier2012[6]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Varliero, G.; Lebre, P.H.; Frey, B.; Fountain, A.G.; Anesio, A.M.; Cowan, D.A. Glacial Water: A Dynamic Microbial Medium. Microorganisms 2023, 11, 1153. https://doi.org/10.3390/microorganisms11051153

AMA Style

Varliero G, Lebre PH, Frey B, Fountain AG, Anesio AM, Cowan DA. Glacial Water: A Dynamic Microbial Medium. Microorganisms. 2023; 11(5):1153. https://doi.org/10.3390/microorganisms11051153

Chicago/Turabian Style

Varliero, Gilda, Pedro H. Lebre, Beat Frey, Andrew G. Fountain, Alexandre M. Anesio, and Don A. Cowan. 2023. "Glacial Water: A Dynamic Microbial Medium" Microorganisms 11, no. 5: 1153. https://doi.org/10.3390/microorganisms11051153

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

Article Metrics

Back to TopTop