Article

Properties of natural supraglacial debris in relation to modelling sub-debris ice ablation

Authors:
To read the full-text of this research, you can request a copy directly from the authors.

Abstract

As debris-covered glaciers become a more prominent feature of a shrinking mountain cryosphere, there is increasing need to successfully model the surface energy and mass balance of debris-covered glaciers, yet measurements of the processes operating in natural supraglacial debris covers are sparse. We report measurements of vertical temperature profiles in debris on the Ngozumpa glacier in Nepal, that show: (i) conductive processes dominate during the ablation season in matrix-supported diamict; (ii) ventilation may be possible in coarse surface layers; (iii) phase changes associated with seasonal change have a marked effect on the effective thermal diffusivity of the debris. Effective thermal conductivity determined from vertical temperature profiles in the debris is generally ~30% higher in summer than in winter, but values depend on the volume and phase of water in the debris. Surface albedo can vary widely over small spatial scales, as does the debris thickness. Measurements indicate that debris thickness is best represented as a probability density function with the peak debris thickness increasing down-glacier. The findings from Ngozumpa glacier indicate that the probability distribution of debris thickness changes from positively skewed in the upper glacier towards a more normal distribution nearer the terminus. Although many of these effects remain to be quantified, our observations highlight aspects of spatial and temporal variability in supraglacial debris that may require consideration in annual or multi-annual distributed modelling of debris-covered glacier surface energy and mass balance. Copyright © 2012 John Wiley & Sons, Ltd.

No full-text available

Request Full-text Paper PDF

To read the full-text of this research,
you can request a copy directly from the authors.

... Other debris characteristics, including moisture content, grain size and lithology can affect the relationship between debris thickness and sub-debris melt rate, for example, by altering the thermal conductivity of the supraglacial debris (e.g. Nicholson and Benn, 2012;Collier and others, 2014;Nakawo and Young, 1981). Therefore, it is also important to quantify the thermal conductivity of the debris layer in order to effectively simulate sub-debris glacial melt rates with a high degree of accuracy. ...
... where SW in is the incoming shortwave radiation (W m −2 ), α is the albedo (dimensionless), 1 is the emissivity (dimensionless), LW in is the incoming longwave radiation (W m −2 ), σ is the Stephan-Boltzmann constant (5.67 × 10 −8 W m −2 K −4 ) and T s is the surface temperature (K). Debris emissivity and albedo values of 0.94 and 0.3, respectively, were assumed (Salisbury and D'Aria, 1992;Nicholson and Benn, 2012). The sensible heat flux was calculated as: ...
... As shown by numerous previous studies (e.g. Conway and Rasmussen, 2000;Nicholson and Benn, 2012), our results further confirm that, during the daytime, the temperature of the supraglacial debris on Llaca Glacier is lowest at the debris ice interface and increases towards the surface of the debris layer (Fig. 4a). The increasingly lagged peaks in debris temperature from the surface to the base of the debris layer indicate that, with increasing depth within the debris layer, the delay time in the melt response to meteorological forcing also increases. ...
Article
Full-text available
Supraglacial debris cover regulates the melt rates of many glaciers in mountainous regions around the world, thereby modifying the availability and quality of downstream water resources. However, the influence of supraglacial debris is often poorly represented within glaciological models, due to the absence of a technique to provide high-precision, spatially continuous measurements of debris thickness. Here, we use high-resolution UAV-derived thermal imagery, in conjunction with local meteorological data, visible UAV imagery and vertically profiled debris temperature time series, to model the spatially distributed debris thickness across a portion of Llaca Glacier in the Cordillera Blanca of Peru. Based on our results, we simulate daily sub-debris melt rates over a 3-month period during 2019. We demonstrate that, by effectively calibrating the radiometric thermal imagery and accounting for temporal and spatial variations in meteorological variables during UAV surveys, thermal UAV data can be used to more precisely represent the highly heterogeneous patterns of debris thickness and sub-debris melt on debris-covered glaciers. Additionally, our results indicate a mean sub-debris melt rate nearly three times greater than the mean melt rate simulated from satellite-derived debris thicknesses, emphasising the importance of acquiring further high-precision debris thickness data for the purposes of investigating glacier-scale melt processes, calibrating regional melt models and improving the accuracy of runoff predictions.
... Surface debris typically has an albedo of 0.1-0.4 (Inoue & Yoshida, 1980;Kayastha et al., 2000;Lejeune et al., 2013;Nicholson & Benn, 2012;Rounce et al., 2018), compared to glacier ice albedo of 0.30-0.46 (Cuffey & Paterson, 2010); thus, Abstract Melt from debris-covered glaciers represents a regionally important freshwater source, especially in high-relief settings as found in central Asia, Alaska, and South America. ...
... Energy balance modeling of sub-debris melt involves evaluation of individual energy balance terms at the debris surface, imposing energy closure to constrain ground heat flux into the debris. The debris layer is then typically treated as a horizontally homogeneous medium (some vertical structure may be invoked) to calculate the conductive heat flux available for sub-debris melt, as pioneered by Reid and Brock (2010) and used further by Rounce and McKinney (2014), Nicholson and Benn (2012), Rounce et al. (2015Rounce et al. ( , 2021, and Reid et al. (2012). While this approach has proven useful in many studies, it ignores convective processes which may occur within the surface debris layer, and thus heat transfer by conduction may be overestimated. ...
... Convection can be significant in other contexts, such as within subarctic snowpack (Sturm & Benson, 1997;Sturm & Johnson, 1991) and in the permafrost active layer (Roth & Boike, 2001), which suggests they may be significant for debris-covered glaciers as well. Nicholson and Benn (2012) identified nonlinear behavior in mean temperature profiles for coarse supraglacial debris cover on the Ngozumpa Glacier, which they interpreted as evidence for convection cooling the uppermost debris layer. ...
Article
Full-text available
Melt from debris‐covered glaciers represents a regionally important freshwater source, especially in high‐relief settings as found in central Asia, Alaska, and South America. Sub‐debris melt is traditionally predicted from surface energy balance models that determine heat conduction through the supraglacial debris layer. Convection is rarely addressed, despite the porous nature of debris. Here we provide the first constraints on convection in supraglacial debris, through the development of a novel method to calculate individual conductive and nonconductive heat flux components from debris temperature profile data. This method was applied to data from Kennicott Glacier, Alaska, spanning two weeks in the summer of 2011 and two months in the summer of 2020. Both heat flux components exhibit diurnal cycles, the amplitude of which is coupled to atmospheric conditions. Mean diurnal nonconductive heat flux peaks at up to 43% the value of conductive heat flux, indicating that failure to account for it may lead to an incorrect representation of melt rates and their drivers. We interpret this heat flux to be dominated by latent heat as debris moisture content changes on a diurnal cycle. A sharp afternoon drop‐off in nonconductive heat flux is observed at shallow depths as debris dries. We expect these processes to be relevant for other debris‐covered glaciers. Debris properties such as porosity and tortuosity may play a large role in modulating it. Based on the present analysis, we recommend further study of convection in supraglacial debris for glaciers across the globe with different debris properties.
... The influence of supraglacial debris on ablation depends on its thickness and thermal properties (Mihalcea and others, 2006;Reid and others, 2012;Fyffe and others, 2020). While field observations of debris properties in HKKH are scarce (Conway and Rasmussen, 2000;Nicholson and Benn, 2013;Rounce and others, 2015;Chand and Kayastha, 2018;Rowan and others, 2021), they indicate that debris properties vary within and between glaciers at a range of scales, and through time (Mihalcea and others, 2008;Nicholson and others, 2018;Shah and others, 2019). Although the physical theory is understood, the actual variability of the debris layer properties is not currently well characterised, and the effects on ablation are poorly constrained (Nicholson and others, 2018). ...
... The alternate method of estimating sub-debris ablation from debris-temperature profiles recorded at auto-logging temperature sensors, which we focus on in this work, is much less labour-intensive. Only a couple of field visits per year are required to obtain both debris thermal properties and sub-debris ablation at up to daily temporal resolution (Nicholson and Benn, 2013). Conway and Rasmussen (2000) pioneered a method to compute the effective thermal diffusivity of a debris layer and estimate the subdebris ablation using observed vertical temperature profiles. ...
... Conway and Rasmussen (2000) pioneered a method to compute the effective thermal diffusivity of a debris layer and estimate the subdebris ablation using observed vertical temperature profiles. This approach, hereinafter referred to as the CR method (Conway and Rasmussen, 2000), has been adapted in several studies in HKKH (e.g., Haidong and others, 2006;Nicholson and Benn, 2013;Chand and Kayastha, 2018;Rowan and others, 2021) and elsewhere (e.g., Nicholson and Benn, 2006;Anderson and others, 2021). ...
Article
Full-text available
A supraglacial debris layer controls energy transfer to the ice surface and moderates ice ablation on debris-covered glaciers. Measurements of vertical temperature profiles within the debris enables the estimation of thermal diffusivities and sub-debris ablation rates. We have measured the debris-layer temperature profiles at 16 locations on Satopanth Glacier (central Himalaya) during the ablation seasons of 2016 and 2017. Debris temperature profile data are typically analysed using a finite-difference method, assuming that the debris layer is a homogeneous one-dimensional thermal conductor. We introduce three more methods for analysing such data that approximate the debris layer as either a single or a two-layered conductor. We analyse the performance of all four methods using synthetic experiments and by comparing the estimated ablation rates with in situ glaciological observations. Our analysis shows that the temperature measurements obtained at equispaced sensors and analysed with a two-layered model improve the accuracy of the estimated thermal diffusivity and sub-debris ablation rate. The accuracy of the ablation rate estimates is comparable to that of the in situ observations. We argue that measuring the temperature profile is a convenient and reliable method to estimate seasonal to sub-seasonal variations of ablation rates in the thickly debris-covered parts of glaciers.
... Field (e.g. Mattson et al., 1993;Nicholson and Benn, 2013;Rowan et al., 2021), laboratory (Reznichenko et al., 2010) and modeling (e.g. Nakawo and Young, 1981;Nicholson and Benn, 2006;Reid and Brock, 2010;Evatt et al., 2015) studies demonstrate that debris thickness is the primary determinant of how sub-debris ice ablation rate differs to that of clean ice, with other properties of the debris layer, such as lithology, porosity and moisture content playing only secondary roles (e.g. ...
... Nakawo and Young, 1981;Nicholson and Benn, 2006;Reid and Brock, 2010;Evatt et al., 2015) studies demonstrate that debris thickness is the primary determinant of how sub-debris ice ablation rate differs to that of clean ice, with other properties of the debris layer, such as lithology, porosity and moisture content playing only secondary roles (e.g. Reznichenko et al., 2010;Nicholson and Benn, 2013;Collier et al., 2014). ...
... There is a systematic tendency for supraglacial debris cover thickness to increase downglacier (e.g. Zhang et al., 2011;Nicholson and Benn, 2013). This is because melt out of englacial material can only occur in the ablation zone of the glacier, and debris is continually conveyed downglacier with ice flow (Kirkbride, 2000), and is concentrated towards the glacier terminus by ice flow deceleration towards the glacier front . ...
Article
Full-text available
Ongoing changes in mountain glaciers affect local water resources, hazard potential and global sea level. An increasing proportion of remaining mountain glaciers are affected by the presence of a surface cover of rock debris, and the response of these debris-covered glaciers to climate forcing is different to that of glaciers without a debris cover. Here we take a back-to-basics look at the fundamental terms that control the processes of debris evolution at the glacier surface, to illustrate how the trajectory of debris cover development is partially decoupled from prevailing climate conditions, and that the development of a debris cover over time should prevent the glacier from achieving steady state. We discuss the approaches and limitations of how this has been treated in existing modeling efforts and propose that “surrogate world” numerical representations of debris-covered glaciers would facilitate the development of well-validated parameterizations of surface debris cover that can be used in regional and global glacier models. Finally, we highlight some key research targets that would need to be addressed in order to enable a full representation of debris-covered glacier system response to climate forcing.
... Debris-covered glaciers (DCGs) respond differently to clean ice glaciers under the same climatic forcing (Nicholson and Benn, 2013). The empirical relationship between debris thickness and ablation rates is well established (Östrem, 1959;Nakawo and Young, 1981;Mattson et al., 1993;Nicholson and Benn, 2006). ...
... The combined surface temperature/debris thickness dataset from the six glaciers has a mean debris thickness of 2.02 m, a median of 1.64 m, a standard deviation of 1.33 m, and a range spanning 0-7.34 m. This would appear to be representative of the debris thickness distribution we might expect on DCGs in the HMA region (Nicholson and Benn, 2013;Juen et al., 2014;Rounce and McKinney, 2014;Rounce et al., 2015). For the collated dataset (n 151,821), the nonparametric Spearman's Rank Correlation Coefficient between surface temperature and debris thickness was 0.30 (99% confidence). ...
... Only a single debris thickness value can be derived for a pixel area represented by a single surface temperature value. However, debris thickness varies on a scale smaller than a 30 m × 30 m area (Nicholson and Benn, 2013). In the datasets used to derive the empirical relationships, there is often a range of debris thickness measurements associated with a single surface temperature (Figure 4). ...
Article
Full-text available
Mapping patterns of supraglacial debris thickness and understanding their controls are important for quantifying the energy balance and melt of debris-covered glaciers and building process understanding into predictive models. Here, we find empirical relationships between measured debris thickness and satellite-derived surface temperature in the form of a rational curve and a linear relationship consistently outperform two different exponential relationships, for five glaciers in High Mountain Asia (HMA). Across these five glaciers, we demonstrate the covariance of velocity and elevation, and of slope and aspect using principal component analysis, and we show that the former two variables provide stronger predictors of debris thickness distribution than the latter two. Although the relationship between debris thickness and slope/aspect varies between glaciers, thicker debris occurs at lower elevations, where ice flow is slower, in the majority of cases. We also find the first empirical evidence for a statistical correlation between curvature and debris thickness, with thicker debris on concave slopes in some settings and convex slopes in others. Finally, debris thickness and surface temperature data are collated for the five glaciers, and supplemented with data from one more, to produce an empirical relationship, which we apply to all glaciers across the entire HMA region. This rational curve: 1) for the six glaciers studied has a similar accuracy to but greater precision than that of an exponential relationship widely quoted in the literature; and 2) produces qualitatively similar debris thickness distributions to those that exist in the literature for three other glaciers. Despite the encouraging results, they should be treated with caution given our relationship is extrapolated using data from only six glaciers and validated only qualitatively. More (freely available) data on debris thickness distribution of HMA glaciers are required.
... Field measurements of debris thickness are scarce and challenging to collect: manual excavations (e.g. others, 2008a, 2008b;Reid and others, 2012;Rounce and McKinney, 2014), extrapolations from ice-cliff surveying (Nicholson and Benn, 2013;Nicholson and Mertes, 2017) and ground-penetrating radar measurements (McCarthy and others, 2017;Nicholson and Mertes, 2017) are the most commonly used methods for measuring or estimating debris thicknesses. However, time and labour constraints limit the spatial distribution of measurements and these methods rely on the assumption that the area sampled is representative of the entire debris-covered area (Rounce and McKinney, 2014). ...
... However, time and labour constraints limit the spatial distribution of measurements and these methods rely on the assumption that the area sampled is representative of the entire debris-covered area (Rounce and McKinney, 2014). The thickness of a surface debris layer influences its surface temperature, such that thicker debris layers generally have higher surface temperatures than thin debris layers under the same meteorological forcing, because of longer distances for conduction that slowdown transport of the heat absorbed at the surface into the underlying ice (Nicholson and Benn, 2013). ...
... In contrast, it showed a minimal sensitivity to albedo, incoming shortwave radiation and finally surface temperature. As the model is relatively insensitive to changes in albedo, following the method of Rounce and McKinney (2014), we used an albedo value of 0.3 (Nicholson and Benn, 2013). The model was moderately sensitive to variations in the non-linear temperature gradient factor -G ratio (for a ±14% change in G ratio , the debris thickness varies by +0.08 m; Table 3). ...
Article
Full-text available
Surface energy-balance models are commonly used in conjunction with satellite thermal imagery to estimate supraglacial debris thickness. Removing the need for local meteorological data in the debris thickness estimation workflow could improve the versatility and spatiotemporal application of debris thickness estimation. We evaluate the use of regional reanalysis data to derive debris thickness for two mountain glaciers using a surface energy-balance model. Results forced using ERA-5 agree with AWS-derived estimates to within 0.01 ± 0.05 m for Miage Glacier, Italy, and 0.01 ± 0.02 m for Khumbu Glacier, Nepal. ERA-5 data were then used to estimate spatiotemporal changes in debris thickness over a ~20-year period for Miage Glacier, Khumbu Glacier and Haut Glacier d'Arolla, Switzerland. We observe significant increases in debris thickness at the terminus for Haut Glacier d'Arolla and at the margins of the expanding debris cover at all glaciers. While simulated debris thickness was underestimated compared to point measurements in areas of thick debris, our approach can reconstruct glacier-scale debris thickness distribution and its temporal evolution over multiple decades. We find significant changes in debris thickness over areas of thin debris, areas susceptible to high ablation rates, where current knowledge of debris evolution is limited.
... The magnitude of melt of the underlying ice interface is strongly modulated by debris thickness because the gradient of the temperature profile at the debris-ice interface determines the conductive heat flux to or from the ice surface (Østrem, 1959;Nicholson and Benn, 2006;Reid and Brock, 2010). Supraglacial debris in the Central Himalaya is typically tens of centimetres thick and commonly exceeds 2.0 m in thickness (Nicholson and Benn, 2013;McCarthy and others, 2017;Nicholson and Mertes, 2017). Sub-debris ablation is therefore expected to be strongly reduced compared to that of a clean-ice surface (Mihalcea and others, 2006;Nicholson and Benn, 2006;Brock and others, 2010;Reid and others, 2012). ...
... Debris layers thicker than a few centimetres heat up rapidly due to warm daytime air temperatures and incoming shortwave radiation, but due to thermal inertia of the debris, a portion of the energy absorbed during the day is reemitted to the atmosphere instead of being transmitted to the underlying ice (Reznichenko and others, 2010). The diurnal cycling of surface energy receipts ensures that the debris cover can rarely reach instantaneous thermal equilibrium with air temperature (Nicholson and Benn, 2013). Vertical profiles of debris temperature demonstrate thermally unstable behaviour at sub-diurnal timescales (Conway and Rasmussen, 2000) and therefore the temperature change within the debris with depth from the surface, hereafter the debris temperature gradient, is likely to be non-linear due to variable meteorological conditions (Reid and Brock, 2010;Foster and others, 2012;Rounce and McKinney, 2014;Schauwecker and others, 2015). ...
... To date, studies have yet to (conclusively) demonstrate how the relationship between debris temperature and thickness varies through the year in the monsoon-influenced Himalaya, or if it can be approximated at a suitable timescale for calculating annual ablation. Previous studies have suggested that, in the absence of snowfall and phase changes within the debris, quasi-linear temperature profiles can be expected over the core ablation season in the Himalaya (Nicholson and Benn, 2013), which potentially simplifies the calculation of annual ablation from debris-covered glaciers. Furthermore, sub-debris melt appears to be more strongly controlled by debris thickness rather than glacier elevation (Shah and others, 2019). ...
Article
Full-text available
Rock debris covers about 30% of glacier ablation areas in the Central Himalaya and modifies the impact of atmospheric conditions on mass balance. The thermal properties of supraglacial debris are diurnally variable but remain poorly constrained for monsoon-influenced glaciers over the timescale of the ablation season. We measured vertical debris profile temperatures at 12 sites on four glaciers in the Everest region with debris thickness ranging from 0.08–2.8 m. Typically, the length of the ice ablation season beneath supraglacial debris was 160 days (15 May to 22 October)—a month longer than the monsoon season. Debris temperature gradients were approximately linear (r2 > 0.83), measured as –40°C m–1 where debris was up to 0.1 m thick, –20°C m–1 for debris 0.1–0.5 m thick, and –4°C m–1 for debris greater than 0.5 m thick. Our results demonstrate that the influence of supraglacial debris on the temperature of the underlying ice surface, and therefore melt, is stable at a seasonal timescale and can be estimated from near-surface temperature. These results have the potential to greatly improve the representation of ablation in calculations of debris-covered glacier mass balance and projections of their response to climate change.
... Table 1 contrasts bulk thermal conductivity, heat capacity, and density of dry debris with debris of the same porosity (φ = 0.39) that has water-filled and ice-filled interstitial spaces. A number of studies (e.g., Conway and Rasmussen, 2000;Reznichenko et al., 2010;Nicholson and Benn, 2012; have emphasized the importance of moisture to the thermal properties of debris, particularly in transition seasons. Rounce and McKinney (2014) found a dramatic increase in conductivity from the top 10 cm of debris on Khumbu region glaciers to the deeper depths; they attribute this difference to water content, noting that Nicholson and Benn (2006) found the conductivity of fully saturated debris to be a factor of 2-3 greater than that of dry debris. ...
... Observations suggest that moisture transport in glacier debris is neither completely reservoir-like (as parameterized in nor fully governed by Darcy's law (as in the original ISBA for soil) but rather some of both simultaneously. A number of studies (e.g., Nicholson and Benn, 2012) mention a saturated basal layer of debris, and Rounce and McKinney (2014) discuss deeper, wet debris overlain by dry debris; our own field observations are consistent. The concentration of wetness at the debris base is due both to the fact that debris coarsens upward (Reid and Brock, 2010) and to the fact that the permeability of the overlying debris is greater than that of the debris at the interface (precipitation quickly moves through the debris until it reaches the impermeable ice surface). ...
... Since it takes a saturated sandy soil 24-48 h to drain to its field capacity, 48 h for τ max is consistent with measurements of the kinds of particles at the base of a debris layer. A shape factor of 30 is consistent with observations of wetted debris right at the debris-ice interface (Nakawo and Young, 1981;Conway and Rasmussen, 2000;Nicholson and Benn, 2012). ...
Article
Full-text available
Few surface energy balance models for debris-covered glaciers account for the presence of moisture in the debris, which invariably affects the debris layer's thermal properties and, in turn, the surface energy balance and sub-debris melt of a debris-covered glacier. We adapted the interactions between soil, biosphere, and atmosphere (ISBA) land surface model within the SURFace EXternalisée (SURFEX) platform to represent glacier debris rather than soil (referred to hereafter as ISBA-DEB). The new ISBA-DEB model includes the varying content, transport, and state of moisture in debris with depth and through time. It robustly simulates not only the thermal evolution of the glacier–debris–snow column but also moisture transport and phase changes within the debris – and how these, in turn, affect conductive and latent heat fluxes. We discuss the key developments in the adapted ISBA-DEB and demonstrate the capabilities of the model, including how the time- and depth-varying thermal conductivity and specific heat capacity depend on evolving temperature and moisture. Sensitivity tests emphasize the importance of accurately constraining the roughness lengths and surface slope. Emissivity, in comparison to other tested parameters, has less of an effect on melt. ISBA-DEB builds on existing work to represent the energy balance of a supraglacial debris layer through time in its novel application of a land surface model to debris-covered glaciers. Comparison of measured and simulated debris temperatures suggests that ISBA-DEB includes some – but not all – processes relevant to melt under highly permeable debris. Future work, informed by further observations, should explore the importance of advection and vapor transfer in the energy balance.
... This phenomenon might be related to the progressive melting of permafrost at these altitudes and increasing periglacial pro cesses ( Figure 4). It is acknowledged that some centimeters of debris cover isolate the ic and reduce the melting process [42][43][44][45][46]. Despite this, these types of glaciers are becomin thinner, their velocity is modified, and it has been observed that there is compression o the ice through the weight of the debris cover [47]. ...
... This phenomenon might be related to the progressive melting of permafrost at these altitudes and increasing periglacial processes ( Figure 4). It is acknowledged that some centimeters of debris cover isolate the ice and reduce the melting process [42][43][44][45][46]. Despite this, these types of glaciers are becoming thinner, their velocity is modified, and it has been observed that there is compression of the ice through the weight of the debris cover [47]. ...
Article
Full-text available
The Infierno Glacier is located in Aragon (Spain), Pyrenees Mountain range, the only one in this country that still preserves white glaciers. These are the southernmost glaciers in Europe and are currently in rapid decline. The work analyzes the evolution of the glacier between 2016 and 2022 and provides data, for this period, which lacked this information, in an area bordering the glacial ice survival. In addition to the observations on the glacier itself, the variables (precipitation, temperature, snow volume and thickness) that allow an understanding of this evolution are studied. The results show a setback of the glacier (thickness losses: 4.6 m; front retreat; 14.9 m). The evolution has frequent trend changes, linked to the interannual climatic irregularity characteristic of the Pyrenees. The main explanatory factor is the thermal increase. The thermal anomalies with respect to the average reference values have increased, in this period, by +0.55 °C. The year 2022 has been particularly warm and has recorded the greatest losses for this glacier. With respect to precipitation, it has an irregular behavior and shows a tendency to decrease (−9% in the same period). This work has the additional interest of analyzing a glacier in the terminal phase, which if current trends continue, evolves into dead ice.
... Nevertheless, observations of debris thickness and thermal properties of the debris layer are available for very few glaciers in the HPKH region (e.g., Mihalcea et al., 2008;Rounce and McKinney, 2014;Ashraf andKhan, 2016, 2017;McCarthy et al., 2017;Nicholson et al., 2018;Shah et al., 2019;Steiner et al., 2021). In situ debris thickness measurements conducted on several glaciers in the HPKH region through manual excavation (Rounce and McKinney, 2014;Patel et al., 2016), surveying debris above exposed ice cliffs (Nicholson and Benn, 2013;Nicholson and Mertes, 2017) and ground penetrating radar (McCarthy et al., 2017) indicate considerable variation in debris thickness over short horizontal distances. Recent studies suggest that melt beneath the debris layer is particularly sensitive to the thermal conductivity and thickness of debris cover, and the variability of the thermal conductivity in space is high on a single glacier (Nicholson and Benn, 2013;Rounce and McKinney, 2014;Steiner et al., 2021). ...
... In situ debris thickness measurements conducted on several glaciers in the HPKH region through manual excavation (Rounce and McKinney, 2014;Patel et al., 2016), surveying debris above exposed ice cliffs (Nicholson and Benn, 2013;Nicholson and Mertes, 2017) and ground penetrating radar (McCarthy et al., 2017) indicate considerable variation in debris thickness over short horizontal distances. Recent studies suggest that melt beneath the debris layer is particularly sensitive to the thermal conductivity and thickness of debris cover, and the variability of the thermal conductivity in space is high on a single glacier (Nicholson and Benn, 2013;Rounce and McKinney, 2014;Steiner et al., 2021). Therefore, knowledge gaps regarding physical and thermal properties of the debris layer remain for large-scale glacier mass-balance and runoff simulations in the HPKH region. ...
Article
Supraglacial debris is widespread on the Hindu Kush-Pamir-Karakoram-Himalaya (HPKH) glaciers, which influences ice melt rates and glacier response to climate change, with important consequences for regional water resources. Supraglacial debris has shown an expanding trend with glacier shrinkage and mass loss, but knowledge about regional spatial patterns of debris cover and associated impacts on HPKH glaciers is still incomplete, which markedly affects the assessment accuracy of regional debris-covered glacier status and hydrological impacts. Here, we address these issues based on ASTER imagery, glacier inventory, and a physically-based ice melt model. We find that about 14.3 % of the regional glacier area is covered by supraglacial debris, and its thickness decreases gradually from northwest to southeast of the region overall. Except for the glaciers of western Himalaya and Karakoram, where the acceleration effect of debris cover is dominant, the insulation effect of debris cover on the glaciers of Pamir, Hindu-Kush, central and eastern Himalaya is particularly significant. The heterogeneous distribution of debris thickness and the resulting melt hotspots (ice cliffs and supraglacial ponds) enhance debris-covered surface thinning in the HPKH region, but it is not the dominant cause of surface thinning. Overall, the net effect of considering spatially distributed debris thickness can reduce total mass loss of debris-covered glaciers by 0.44 ± 0.04 m water equivalent (w. e.) yr⁻¹. As a result, the presence of debris cover retards accelerated glacier melting caused by current climate warming, with important implications for slowing down regional water shortage and ecological environment risk in the HPKH region. Our findings highlight the importance of including the debris-cover effect in regional glacier models for assessing future glacier change and associated hydrological impacts in the HPKH region.
... The observed glacier loss in the Himalaya would adversely impact the water supplies for irrigation and other dependent services and livelihoods in the region (Abdullah et al., 2020;Immerzeel et al., 2020;Nie et al., 2021). These concerns have led to a number of glacier studies in the region (Scherler et al., 2011;Nicholson and Benn, 2013;Kääb et al., 2015;Racoviteanu et al., 2015;Garg et al., 2017;Murtaza and Romshoo, 2017;Bhushan et al., 2018;Vijay and Braun., 2018;Muhammad et al., 2019;Kumar et al., 2019;Abdullah et al., 2020;Romshoo et al., 2020), which indicated that glaciers in the Himalaya are showing a significant area and mass loss and that the glacier behavior observed across the Himalaya is not uniform due to the varying influence of topographic, climatic and other driving factors. ...
... The slight deviation in the shrinkage rates between these studies and our observations could be because of the variation in the climatic conditions (Shrestha et al., 2012) and tectonic setting of the basins (DiPietro and Pogue, 2004). However, the physical, thermal and lithological properties of the debris-cover like thickness, composition, thermal conductivity, moisture content, porosity (Conway et al., 2000;Nicholson and Benn, 2013) depth to the critical thickness of debris (Nakawo and Young, 1981) and the regional variation in the distribution of debris-cover influence the recession rates (Scherler et al., 2011). ...
Article
Out of 4530 glaciers covering ∼4770 km² in the Upper Indus Basin, viz., Chenab, Zanaskar and Suru sub-basins, we selected topographically and morphologically homogeneous set of 98 glaciers; 56 debris-covered and 42 clean to assess the impact of debris-cover on glacier melting. Satellite data analysis from 2000 to 2017 showed that the debris-covered glaciers have melted slower (2.49 ± 0.05%; 0.15%a⁻¹) than the clean glaciers (4.07 ± 0.03%; 0.24%a⁻¹). The intra-basin analyses revealed that clean and debris-covered glaciers in the Chenab have melted 2.34 ± 0.17% and 1.1 ± 0.30% (0.14%a⁻¹ and 0.06%a⁻¹) and retreated 8.82 ± 2.67ma⁻¹ and 5.7 ± 2.67ma⁻¹ respectively. Similarly, clean and debris-covered glaciers in the Zanaskar have melted at similar rates of 4.58 ± 0.07% and 4.55 ± 0.21% (0.27%a⁻¹ and 0.25%a⁻¹) and retreated 10.0 ± 2.67ma⁻¹ and 8.4 ± 2.67ma⁻¹ respectively. Clean and debris-covered glaciers in the Suru have melted 5.56 ± 0.09% and 4.53 ± 0.14% (0.33%a⁻¹ and 0.27%a⁻¹) and retreated 9.6 ± 2.67ma⁻¹ and 7.0 ± 2.67ma⁻¹ respectively. Extensive debris-cover of ∼32.3% modulates glacier loss in the Chenab and sparse debris-cover of 11.4% and 6.4% in the Suru and Zanaskar enhances the melting of the debris-covered glaciers. Overall, debris-covered glaciers melted slower than the clean glaciers, with inter- and intra-basin variations, suggesting that debris-cover plays a complex but significant role in modulating glacier melting under climate change.
... However, field studies (e.g. Fujii 1977;Mattson et al. 1993), laboratory experiments (Reznichenko et al. 2010) and modelling studies (e.g., Nakawo and Young 1982;Nicholson and Benn 2006;Reid and Brock 2010) demonstrate that debris thickness is the primary determinant of how sub-debris ice ablation rates differ to clean-ice melt rates, with the properties of the debris layer playing secondary roles (Reznichenko et al. 2010;Nicholson and Benn 2012;Collier et al. 2014). As surface debris is continuously conveyed downglacier with ice flow, debris cover thickness increases towards the glacier terminus (Rowan et al. 2015;Anderson and Anderson 2018). ...
... These range from simple band thresholding (Ranzi et al. 2004) to exponential curve fitting based on the empirical relationship between surface temperature and thickness (Juen et al. 2014;Kraaijenbrink et al. 2018) or energy-balance inversion, often requiring model spin-up (Mihalcea et al. 2008;Zhang et al. 2011;Foster et al. 2012;Rounce and McKinney 2014;Schauwecker et al. 2015;Rounce et al. 2021;Stewart et al. 2021). An intercomparison of these methods is needed and has been identified as a research target for the IACS Debris-covered Glaciers Working Group (https://cryosphericsciences.org/activities/wgdebris/); o Elevation change/surface mass balance: Since the thickness of debris moderates energy transfer to the ice, it also controls ice melt rates (Østrem 1959;Nicholson and Benn 2012). Surface mass balance data can thus be used to invert an energy mass balance model for debris thickness (Ragettli et al. 2015;Rounce et al. 2018), although this often requires careful consideration of ice dynamics to estimate surface mass balance from elevation change, and the long-duration melt modelling is computationally expensive (Rounce et al. 2021); o Polarimetric SAR: As certain wavelengths of radar can penetrate into the debris surface, the attenuation of radar signals is indicative of debris thickness. ...
Article
Full-text available
Glaciers respond sensitively to climate variability and change, with associated impacts on meltwater production, sea-level rise and geomorphological hazards. There is a strong societal interest to understand the current response of all types of glacier systems to climate change and how they will continue to evolve in the context of the whole glacierized landscape. In particular, understanding the current and future behaviour of debris-covered glaciers is a ‘hot topic’ in glaciological research because of concerns for eater resources and glacier-related hazards. The state of these glaciers is closely related to various hazardous geomorphological processes which are relatively poorly understood. Understanding the implications of debris-covered glacier evolution requires a systems approach. This includes the interplay of various factors such as local geomorphology, ice ablation patterns, debris characteristics, glacier lake growth and development. Such a broader, contextualized understanding is prerequisite to identifying and monitoring the geohazards and hydrologic implications associated with changes in the debris-covered glacier system under future climate scenarios. This paper presents a comprehensive review of current knowledge of the debris-covered glacier landsystem. Specifically, we review state-of-the-art field and remote sensing-based methods for monitoring debris-covered glacier characteristics and lakes and their evolution under future climate change. We advocate a holistic process-based framework for assessing hazards associated with moraine-dammed glacio-terminal lakes that are a projected end-member state for many debris-covered glaciers under a warming climate.
... Higher surface temperatures, and a larger amplitude in the diurnal cycle of near-surface temperature have been found over debris-covered glaciers, compared to clean-ice glaciers (Steiner and Pellicciotti, 2016;Yang et al., 2017;Nicholson and Stiperski, 2020). It has been shown that after heating during the day, debris cover greater than 0.05 m returns the energy to the atmosphere at night, rather than heating the ice below (Reznichenko et al., 2010;Nicholson and Benn, 2013). Differences have also been found in near-surface wind speed and incoming short wave radiation between debris-covered and clean-ice glaciers (Yang et al., 2017). ...
... For this simple representation of debris cover in the WRF model, the debris is assumed to occupy the full soil depth in the WRF model, which extends to 255 cm below the surface (i.e., there is no representation of the ice underneath). As such, this method best represents thick debris cover, as in general thick debris cover is thought to provide insulation between the air and the glacier ice below the debris cover (Reznichenko et al., 2010;Nicholson and Benn, 2013). ...
Article
Full-text available
Abstract Many of the glaciers in the Nepalese Himalaya are partially covered in a layer of loose rock known as debris cover. In the Dudh Koshi River Basin, Nepal, approximately 25% of glaciers are debris‐covered. Debris‐covered glaciers have been shown to have a substantial impact on near‐surface meteorological variables and the surface energy balance, in comparison to clean‐ice glaciers. The Weather Research and Forecasting (WRF) model is often used for high‐resolution weather and climate modelling, however representation of debris‐covered glaciers is not included in the standard land cover and soil categories. Here we include a simple representation of thick debris‐covered glaciers in the WRF model, and investigate the impact on the near‐surface atmosphere over the Dudh Koshi River Basin for July 2013. Inclusion of this new category is found to improve the model representation of near‐surface temperature and relative humidity, in comparison with a simulation using the default category of clean‐ice glaciers, when compared to observations. The addition of the new debris‐cover category in the model warms the near‐surface air over the debris‐covered portion of the glacier, and the wind continues further up the valley, compared to the simulation using clean‐ice. This has consequent effects on water vapour and column‐integrated total water path, over both the portions of the glacier with and without debris cover. Correctly simulating meteorological variables such as these is vital for accurate precipitation forecasts over glacierized regions, and therefore estimating future glacier melt and river runoff in the Himalaya. These results highlight the need for debris cover to be included in high‐resolution regional climate models over debris‐covered glaciers.
... Other properties of supraglacial debris, such as moisture content, rock type, and grain size, can alter the thermal conductivity of the debris layer, consequently modifying the relationship between debris thickness and surface melt rates (e.g. [15][16][17]). The ablation rates on heavily debris-covered glaciers are extremely difficult to measure, due to the challenges that are associated with drilling stakes through the debris layer, as well as the large heterogeneity of local ablation rates (e.g. ...
... Differences in debris-thickness distribution, as well as local debris properties, such as lithology, grain size, and moisture content, may contribute towards the contrasting breakpoint elevations and magnitudes of reversed altitudinal SMB gradients that were observed within these regions (e.g. [15][16][17]. Supraglacial ice cliffs and ponds can also influence the SMB of debris-covered glaciers by creating localised areas of enhanced melting [24][25][26][27]. Therefore, it is possible that these features could also partially explain the heterogeneity that was observed between debris-covered glaciers. ...
Article
Full-text available
Meltwater from the glaciers in High Mountain Asia plays a critical role in water availability and food security in central and southern Asia. However, observations of glacier ablation and accumulation rates are limited in spatial and temporal scale due to the challenges that are associated with fieldwork at the remote, high-altitude settings of these glaciers. Here, using a remote-sensing-based mass-continuity approach, we compute regional-scale surface mass balance of glaciers in five key regions across High Mountain Asia. After accounting for the role of ice flow, we find distinctively different altitudinal surface-mass-balance gradients between heavily debris-covered and relatively debris-free areas. In the region surrounding Mount Everest, where debris coverage is the most extensive, our results show a reversed mean surface-mass-balance gradient of -0.21 ± 0.18 m w.e. a−1 (100m)−1 on the low-elevation portions of glaciers, switching to a positive mean gradient of 1.21 ± 0.41 m w.e. a−1 (100m)−1 above an average elevation of 5520 ± 50 m. Meanwhile, in West Nepal, where the debris coverage is minimal, we find a continuously positive mean gradient of 1.18 ± 0.40 m w.e. a−1 (100m)−1. Equilibrium line altitude estimates, which are derived from our surface-mass-balance gradients, display a strong regional gradient, increasing from northwest (4490 ± 140 m) to southeast (5690 ± 130 m). Overall, our findings emphasise the importance of separating signals of surface mass balance and ice dynamics, in order to constrain better their contribution towards the ice thinning that is being observed across High Mountain Asia.
... In our model, we assumed a linear gradient of debris temperature with depth, and demonstrated that the assumption holds true for the daily scale (see Fig. 5), but it is not verified for sub-daily time periods. This is in agreement with Nicholson and Benn (Nicholson and Benn, 2013;Nicholson and Benn, 2006), who demonstrated this behaviour for debris of different lithology and grain size. To compute melt at an hourly scale, a different approach would therefore be needed. ...
... For clean-surface glaciers, the englacial debris is entrained by ice flow and exposed at the glacier margins. For debris-covered glaciers, supraglacial debris will affect ablation substantially depending on the debris characteristics, particularly for its thickness [9][10][11][12][13][14]. Theoretically, these two types of glacier states might be mutually transformed for the same glacier as it expands and shrinks with varying ice influx [15][16][17]. ...
Article
Full-text available
The Tibetan Plateau contains a large number of mountain glaciers with clean surfaces, where englacial debris is generally entrained by the ice flow and exposed at the glacier margins. The long-term observation on one of the typical clean surface glaciers (the Qiyi Glacier, northern Tibetan Plateau) suggests an early emergence of englacial debris on its transport pathway, with accelerated surface melting from the mid-2000s onwards. Given that the englacial debris layers of the tongue part of Qiyi Glacier are approximately parallel to the glacier surface, the continuing melting might be expected to result in the rapid expansion of exposed debris. Compared with the clean surface ice, debris cover at the same elevation reduced glacier mass loss by ~25.4% during a hydrological year (2020–2021), indicating that the early emergence of englacial debris can protect the glacier from climate warming with prolonged life expectance. As such, future glacial runoff will then reach its peak earlier and be followed by a gentler decreasing trend than model projections with constant clean surface ice. These findings imply that the emerging debris on clean surface glacier may mitigate the glacial-runoff risk, which has so far been neglected in projections of future water supplies.
... The debris on these glaciers is mainly from the slope sliding on both sides of the glacier and the upwelling of inside moraine by ice flow [3]. Debris cover alters the glacier dynamics by changing surface albedo, thermal conductivity, and surface roughness [4][5][6]. The thin debris cover can enhance the surface melting by absorbing more energy. ...
Article
Full-text available
Debris-covered glaciers have contrasting melting mechanisms and climate response patterns if compared with debris-free glaciers and thus show a unique influence on the hydrological process. Based on high-resolution satellite images and unpiloted aerial vehicle surveys, this study investigated the dynamic changes of Zhuxi Glacier, a thick debris-covered glacier in the southeastern Tibetan Plateau. Our result shows that the whole glacier can be divided into the active regime and stagnant regime along the elevation of 3400 m a.s.l. The mean surface velocity of the active regime was 13.1 m yr−1, which was five times higher than that of the stagnant regime. The surface-lowing rate of this debris-covered glacier reaches more than 1 m yr−1 and displays an accelerating trend. The majority of ice loss concentrates around ice cliffs and supraglacial ponds, the ablation hotspots. These hotspots can be roughly classified into three types, including persistent, expanding, and shrinking patterns, at different dynamic regimes on the Zhuxi Glacier. With the evolution of these hotpots and glacier dynamic changes, the supraglacial ponds showed significant change, with the total number fluctuating from 15 to 38 and the total area increasing from 1128 m2 to 95790 m2 during the past decade. The recent exponential expansion of the proglacial lake and the significant downwasting of stagnant ice inside the dammed terminus moraine possibly trigger the glacial lake outburst flood and thus threaten the security of livelihoods and infrastructure downstream.
... This phenomenon might be related to the progressive melting of permafrost at these altitudes and to increasing periglaciar dynamics (Figure 4). It is acknowledged that the upper covers of washout located at some centimeters isolate the ice and reduce the melting process [40][41][42][43][44]. In spite of it, these glaciers are becoming thinner, their velocity is modified, and it has been observed a compression of the ice through the weight of the melt [45]. ...
Preprint
The Infierno glacier is located in Aragon (Spain), Pyrenees mountain range, southern slope, the only one in this country that still preserves white glaciers. These are the southernmost glaciers in Europe and are currently in rapid regression. The work analyzes the evolution of the glacier between 2016 and 2022 (taking as "year zero" or starting point, 2015). In addition to the observations on the glacier itself, the variables (precipitation, temperatures, snow volumes and thicknesses) that allow understanding this evolution are studied. The results offer strong regression, with thickness losses in that period of 4.6 m and retreat of its front of 14.9 m. The evolution has frequent trend changes, linked to the interannual climatic irregularity characteristic of this mountain range. The main explanatory factor is the thermal increase. The thermal anomalies with respect to the average reference values have increased, in this period, +0.55° C. The year 2022 has been particularly warm and has recorded the greatest regression of the glacier (between May and August, the thermal anomalies were between +4° C and +2° C). Regarding precipitations, they have irregular tendencies and show a decreasing trend (-9% in the same period of time).
... When sediment supply to a glacier exceeds the rate of sediment evacuation to moraines, a supraglacial debris layer may develop from the glacier terminus and extend across the ablation area (Anderson & Anderson, 2018;Benn et al., 2012;Kirkbride, 2000). Supraglacial debris thicker than a few centimeters reduces ablation, allowing the glacier tongue to persist at a lower elevation than a climatically equivalent clean-ice surface (Anderson et al., 2021;Nicholson & Benn, 2013;Rounce et al., 2018). Glaciers in the Khumbu Valley are currently mantled by extensive debris layers that formed in response to recent mass loss (Herreid & Pellicciotti, 2020;Rowan et al., 2015). ...
Article
Full-text available
The dynamic response of large mountain glaciers to climatic forcing operates over timescales of several centuries and therefore understanding how these glaciers change requires observations of their behavior through the Holocene. We used Be‐10 exposure‐age dating and geomorphological mapping to constrain the evolution of glaciers in the Khumbu Valley in the Everest region of Nepal. Khumbu and Lobuche Glaciers are surrounded by high‐relief lateral and terminal moraines from which seven glacial stages were identified and dated to 7.4 ± 0.2, 5.0 ± 0.3, 3.9 ± 0.1, 2.8 ± 0.2, 1.3 ± 0.1, 0.9 ± 0.02, and 0.6 ± 0.16 ka. These stages correlate to each of the seven latest Holocene regional glacial stages identified across the monsoon‐influenced Himalaya, demonstrating that a coherent record of high elevation terrestrial palaeoclimate change can be extracted from dynamic mountain landscapes. The time‐constrained moraine complex represents a catchment‐wide denudation rate of 0.8–1.4 mm a⁻¹ over the last 8 kyr. The geometry of the ablation area of Khumbu Glacier changed around 4 ka from a broad, shallow ice tongue to become narrower and thicker as restricted by the topographic barrier of the terminal moraine complex.
... Generally, debris thickness increases toward a glacier terminus ( Anderson and Anderson, 2018), but exhibits a strong small-scale variability, caused by a variety of factors (Nicholson and others, 2018; Shah and others, 2019). Manual excavations (e.g., Reid and others, 2012), observations of debris thickness above exposed ice cliffs (e.g., Nicholson and Benn, 2013) or ground-penetrating radar surveys (e.g. McCarthy and others, 2017; Giese and others, 2021) at a high enough spatial resolution to capture this small-scale variability are both time and labor-intensive. ...
Article
Full-text available
Debris-covered glaciers are an important component of the mountain cryosphere and influence the hydrological contribution of glacierized basins to downstream rivers. This study examines the potential to make estimates of debris thickness, a critical variable to calculate the sub-debris melt, using ground-based thermal infrared radiometry (TIR) images. Over four days in August 2019, a ground-based, time-lapse TIR digital imaging radiometer recorded sequential thermal imagery of a debris-covered region of Peyto Glacier, Canadian Rockies, in conjunction with 44 manual excavations of debris thickness ranging from 10 to 110 cm, and concurrent meteorological observations. Inferring the correlation between measured debris thickness and TIR surface temperature as a base, the effectiveness of linear and exponential regression models for debris thickness estimation from surface temperature was explored. Optimal model performance ( R ² of 0.7, RMSE of 10.3 cm) was obtained with a linear model applied to measurements taken on clear nights just before sunrise, but strong model performances were also obtained under complete cloud cover during daytime or nighttime with an exponential model. This work presents insights into the use of surface temperature and TIR observations to estimate debris thickness and gain knowledge of the state of debris-covered glacial ice and its potential hydrological contribution.
... The Himalayan glaciers are characterized by the presence of debris in their lower ablation zones and the thickness of debris varies from a few centimetres to meters (Benn et al., 2012;Chand et al., 2015;Chand and Kayastha, 2018;Fujii and Higuchi, 1977;Mattson and E., 1993;Östrem, 1959). The source of debris for these glaciers is due to the presence of large valleys and steep headwalls (Kraaijenbrink et al., 2017), which alter the surface energy balance and work as a barrier between the atmosphere and ice (Nicholson and Benn, 2013). The melt rate of the clean ice and debris-covered ice is different due to the difference in interaction between atmosphere and ice. ...
... Supra-glacial debris thickness can also play a key role in changes of mass balance for individual glaciers. This is broadly confirmed that supra-glacial debris cover affects surface melt with increasing ablation in cases of thin debris cover (less than a few cm) or decreasing ablation under continuous thick debris cover [43][44][45]. It is also possible that ice cliffs and supra-glacial ponds play a key role by enhancing the total ablation of debris-covered glaciers in the Greater Caucasus. ...
Article
Full-text available
Glaciers and snow in the Caucasus are major sources of runoff for populated places in many parts of this mountain region. These glaciers have shown a continuous area decrease; however, the magnitude of mass balance changes at the regional scale need to be further investigated. Here, we analyzed regional changes in surface elevation (or thickness) and geodetic mass balance for 1861 glaciers (1186.1 ± 53.3 km2) between 2000 and 2019 from existing dataset and outlines of the Caucasus glacier inventory. We used a debris-covered glacier dataset to compare the changes between debris-free and debris-covered glaciers. We also used 30 m resolution ASTER GDEM (2011) to determine topographic details, such as aspect, slope, and elevation distribution of glaciers. Results indicate that the mean rate of glacier mass loss has accelerated from 0.42 ± 0.61 m of water equivalent per year (m w.e. a−1) over 2000–2010, to 0.64 ± 0.66 m w.e. a−1 over 2010–2019. This was 0.53 ± 0.38 m w.e. a−1 in 2000–2019. Mass loss rates differ between the western, central, and eastern Greater Caucasus, indicating the highest mean annual mass loss in the western section (0.65 ± 0.43 m w.e. a−1) in 2000–2019 and much lower in the central (0.48 ± 0.35 m w.e. a−1) and eastern (0.38 ± 0.37 m w.e. a−1) sections. No difference was found between the northern and southern slopes over the last twenty years corresponding 0.53 ± 0.38 m w.e. a−1. The observed decrease in mean annual geodetic mass balance is higher on debris-covered glaciers (0.66 ± 0.17 m w.e. a−1) than those on debris-free glaciers (0.49 ± 0.15 m w.e. a−1) between 2000 and 2019. Thickness change values in 2010–2019 were 1.5 times more negative (0.75 ± 0.70 m a−1) than those in 2000–2010 (0.50 ± 0.67 m a−1) in the entire region, suggesting an acceleration of ice thinning starting in 2010. A significant positive trend of May-September air temperatures at two selected meteorological stations (Terskol and Mestia) along with a negative trend of October-April precipitation might be responsible for the negative mass balances and thinning for all Caucasus glaciers over the study period. These results provide insight into the change processes of regional glaciers, which is key information to improve glaciological and hydrological projections in the Caucasus region.
... 川表碛厚度的精确测量或反演在冰川消融率估算、物质平衡模拟、区域水资源评估乃至 冰川灾害演化研究等方面具有重要意义 [12] 。 确定冰川表碛厚度有两类方式:① 传统的现场实地测量,这种方法工作量大且无法 覆盖整个表碛范围,对于无法到达的冰川,获得表碛厚度值更为困难 [13] ;② 利用遥感技 术反演区域冰川表碛厚度 [14] 。随着遥感科学的快速发展,利用遥感影像反演冰川表碛厚 度已经较为成熟,但对于如何提高反演精度及结果验证仍面临较大困难。在具体应用 中,地表温度-表碛厚度关系法、雷达影像反演表碛厚度、基于能量平衡和表碛热阻系 数估算表碛厚度等方法运用最为广泛 [6] 。Zhang 等 [15] 实地测量了海螺沟冰川消融区的表碛 厚度,发现表碛厚度随海拔上升而降低且与基于热红外影像提取的表碛层热阻系数相关 性很强。Mihalcea 等 [16] 基于地表能量平衡方程反演 Baltoro 冰川表碛厚度并估算了不同表 碛厚度下的消融率,这种方法的优点在于不需要太多观测数据就能对表碛厚度分布进行 制图。Mihalcea 等 [17] 通过建立冰川表面温度与表碛厚度之间的关系模拟了 Miage 冰川表碛 的空间分布,发现地表测量和遥感反演均能很好地估算消融区的表碛厚度。此外,也有 学者利用合成孔径雷达 (SAR) 影像的 L 波段估算了天山托木尔峰地区冰川表碛厚度 [18] 。 冰川表碛分布与冰面湖、冰崖等冰面地貌演化过程密切相关。总结国内外相关文 献,利用遥感影像反演冰川表碛厚度已经相对较为成熟,但通过多时相分析表碛厚度变 化研究较少。在气候变暖冰川消融加剧背景下,冰川表碛覆盖存在分布范围向上游扩张 和厚度向下游区增厚的趋势。贡嘎山位于青藏高原东南缘,是典型地海洋型冰川作用 区,这种趋势在贡嘎山地区表现得更为明显。目前,针对贡嘎山地区冰川表碛的研究还 只停留在实地测量阶段,缺乏全面系统的时空对比研究。本文将利用 Landsat TM/TIRS 影像的热红外波段反演贡嘎山地区冰川表碛厚度,揭示其典型冰川在 1990-2019 年间表 碛覆盖范围和厚度变化特征以及贡嘎山东西坡差异,并在此基础上进行分析讨论。 2 贡嘎山位于青藏高原东南缘,是中国重要的海洋型冰川作用区 [19] 。针对贡嘎山地区 冰川的研究最早起源于 20 世纪 30 年代,之后国内外学者先后多次对该区域内的冰川进行 考察和监测。1990 年中苏联合贡嘎山冰川考察,为后来山地冰川研究积累了大量宝贵研 究资料。中国科学院四川贡嘎山森林生态系统野外科学观测研究站于 1987 年建站,为该 区域内冰川的监测与研究提供了重要平台 [20] 。 贡嘎山最高峰达 7514 m,区内有现代冰川 74 条,冰川总面积 255.1 km 2 [21] 。基于自 动气象站连续观测表明,贡嘎山东坡海螺沟冰川区附近 (3000 m) 年平均气温 (5 ℃) 高于西坡贡巴冰川区附近 (3700 m) 的贡嘎寺 (3 ℃) ,对应站点东坡多年平均年降水量 约为 1700 mm,然而西坡仅约 900 mm。受东亚夏季风的影响,贡嘎山东坡冰川较为发 育,典型的较大型山岳冰川有海螺沟冰川、磨子沟冰川、燕子沟冰川和南门关沟冰川; 西坡发育冰川数量多,但规模小,其中典型冰川有大贡巴冰川和小贡巴冰川 (图 1) [22] 。 以上 6 条冰川除小贡巴冰川外长度均超过了 10 km,面积占据了整个贡嘎山冰川的 47.6%,冰川消融区均有大范围表碛覆盖 [23] 。Zhang 等 [15] [32] ...
Article
Full-text available
表碛覆盖型冰川是中国西部较为常见的冰川类型。表碛层存在于大气—冰川冰界面, 强烈影响大气圈与冰冻圈之间的热交换。表碛厚度的空间异质性可极大地改变冰川的消融率 和物质平衡过程,进而影响冰川径流过程和下游水资源。基于Landsat TM/TIRS数据,运用能 量平衡方程反演了贡嘎山地区冰川表碛厚度,研究了贡嘎山地区冰川在1990—2019年间表碛 覆盖范围及厚度变化情况,同时对比了东西坡差异。结果表明:①贡嘎山地区冰川表碛扩张总 面积达43.824 km2。其中,海螺沟冰川扩张2.606 km2、磨子沟冰川1.959 km2、燕子沟冰川1.243 km2、大贡巴冰川0.896 km2、小贡巴冰川0.509 km2、南门关沟冰川2.264 km2,年均扩张率分别为 3.2%、11.1%、1.5%、0.9%、1.0%和6.5%;②海螺沟冰川、磨子沟冰川、燕子沟冰川、大贡巴冰川、 小贡巴冰川、南门关沟冰川表碛平均增厚分别为5.2 cm、3.1 cm、3.7 cm、6.8 cm、7.3 cm和13.1 cm;③西坡冰川表碛覆盖度高,表碛覆盖年均扩张率低,冰川末端退缩量小;东坡冰川表碛覆 盖年均扩张率高,但表碛覆盖度总体低于西坡,冰川末端退缩量大。
... They were attached with bamboo stakes, and they continuously logged temperature data at 30 min intervals between September 2016 and October 2017 ( Table 1). Many studies have used the same thermistor probes and dataloggers for temperature profiling for similar investigations of debris-covered glaciers in the Nepal Himalaya (Mihalcea et al., 2006;Brock et al., 2010;Nicholson and Benn, 2013;Rowan et al., 2021). The Tinytag datalogger (TGP 4520) has an accuracy of ± 0.4°C at 0°C, and two thermal probes can be connected into the logger unit. ...
Article
Full-text available
A large number of glaciers in the Hindu-Kush Himalaya are covered with debris in the lower part of the ablation zone, which is continuously expanding due to enhanced glacier mass loss. The supraglacial debris transported over the melting glacier surface acts as an insulating barrier between the ice and atmospheric conditions and has a strong influence on the spatial distribution of surface ice melt. We conducted in-situ field measurements of point-wise ablation rate, supraglacial debris thickness, and debris temperature to examine the thermal resistivity of the debris pack and its influence on ablation over three glaciers (Bara Shigri, Batal, and Kunzam) in Chandra Basin of Western Himalaya during 2016-2017. Satellite-based supraglacial debris cover assessment shows an overall debris covered area of 15% for Chandra basin. The field data revealed that the debris thickness varied between 0.5 and 326 cm, following a spatially distributed pattern in the Chandra basin. The studied glaciers have up to 90% debris cover within the ablation area, and together represent ∼33.5% of the total debris-covered area in the basin. The supraglacial debris surface temperature and near-surface air temperature shows a significant correlation (r > 0.88, p < 0.05), which reflects the effective control of energy balance over the debris surface. The thermal resistivity measurements revealed low resistance (0.009 ± 0.01 m 2°C W −1) under thin debris pack and high resistance (0.55 ± 0.09 m 2°C W −1) under thick debris. Our study revealed that the increased thickness of supraglacial debris significantly retards the glacier ablation due to its high thermal resistivity.
... We calculated the sensible and latent heat fluxes following e.g. [82]: ...
Preprint
Supraglacial debris strongly modulates glacier melt rates and can be decisive for ice dynamics and mountain hydrology. It is ubiquitous in High-Mountain Asia (HMA), yet because its thickness and supply rate from local topography are poorly known, our ability to forecast regional glacier change and streamflow is limited. Here we resolved the spatial distribution of supraglacial debris thickness (SDT) for 4401 glaciers in HMA for 2000-2016, via an inverse approach using a new dataset of glacier mass balance. We then determined debris-supply rate (DSR) to 3843 of those glaciers using a debris mass-balance model. Our results reveal high spatial variability in both SDT and DSR, with supraglacial debris most concentrated around Everest, and DSR highest in the Pamir-Alai. We demonstrate that DSR and, by extension, SDT increase with the temperature and slope of debris-supply slopes regionally and that SDT increases as ice flow decreases locally. Our centennial-scale estimates of DSR are an order of magnitude lower than millennial-scale estimates of headwall-erosion rate from 10Be cosmogenic nuclides, indicating that debris supply to the region's glaciers is highly episodic. We anticipate that our datasets will enable improved representation of the complex response of HMA's glaciers to climatic warming in future modelling efforts.
... Measurements of near-surface ice temperatures are important for modelling the surface energy balance and projecting the future mass-balance response of glaciers to anticipated climate change. This is especially the case for glaciers with a thick supraglacial debris layer that insulates the underlying ice (according to debris layer thickness and lithology), reducing ablation and potentially extending glacier longevity (Nicholson and Benn, 2006;Nicholson and Benn, 2013;Anderson and Anderson, 2016). ...
Article
Full-text available
Meltwater from high-elevation debris-covered glaciers—particularly those located in the greater Himalaya and Andes—shapes the water supply of major rivers and nourishes substantial terrestrial, estuarine, and marine habitats (Kraaijenbrink et al., 2017; Immerzeel et al., 2020). However, the relative inaccessibility and high elevation of such glaciers results in a paucity of data relating to their fundamental physical properties and processes, limiting the information available to constrain and evaluate numerical models of their behaviour and project future change. Knowledge of the subsurface properties of such glaciers is particularly deficient because it is largely obscured to satellite and airborne remote sensing; englacial investigations therefore commonly require direct access (Miles et al., 2020). Of the physical properties of glaciers, ice temperature exerts an important control over glaciological processes, such as glacier motion, and their modelled behaviour. For example, ice viscosity is sensitive to temperature such that, under the same stress, ice approaching the melting point deforms 5–10 times more rapidly than it would at −10°C (Deeley and Woodward, 1908; Cuffey and Paterson, 2010). Basal motion depends on lubrication facilitated by the presence of meltwater at the ice-bed interface and/or within the pore space of a subglacial sediment layer. Measurements of near-surface ice temperatures are important for modelling the surface energy balance and projecting the future mass-balance response of glaciers to anticipated climate change. This is especially the case for glaciers with a thick supraglacial debris layer that insulates the underlying ice (according to debris layer thickness and lithology), reducing ablation and potentially extending glacier longevity (Nicholson and Benn, 2006; Nicholson and Benn, 2013; Anderson and Anderson, 2016). Here, we present a one-year time series of near-surface ice temperatures, measured between 1.5 and 7.0 m below the ice surface, in a borehole drilled by hot water into the debris-covered tongue of Khumbu Glacier, Nepal.
... For the case presented in the main text we assumed that C is uniform and does not vary in time, that porosity is 0.3 (Nicholson and Benn, 2013), and that the rock density is 2,200 kg m −3 (see MacKevett and Smith, 1972;Miles et al., 2021). Because we have no a priori knowledge of the pattern or magnitude of C, porosity, and rock density we ran 13 additional inversions in which 1) C increases downglacier linearly through the swath profile by factors of 2, 10, 100, and 1,000; 2) C decreases linearly downglacier through the swath profile by factors of 0.5, 0.1, 0.01, and 0.001; 3) porosity varies between 0.1 and 0.4; 4) rock density varies between 2000 and 2,700 kg m −3 ; and 5) use the remotely sensed debris pattern of Rounce et al. (2021) instead of the in situ debris thicknesses as the 2011 evaluating dataset (Supplementary Table S5). ...
Article
Full-text available
The cause of debris-covered glacier thinning remains controversial. One hypothesis asserts that melt hotspots (ice cliffs, ponds, or thin debris) increase thinning, while the other posits that declining ice flow leads to dynamic thinning under thick debris. Alaska’s Kennicott Glacier is ideal for testing these hypotheses, as ice cliffs within the debris-covered tongue are abundant and surface velocities decline rapidly downglacier. To explore the cause of patterns in melt hotspots, ice flow, and thinning, we consider their evolution over several decades. We compile a wide range of ice dynamical and mass balance datasets which we cross-correlate and analyze in a step-by-step fashion. We show that an undulating bed that deepens upglacier controls ice flow in the lower 8.5 km of Kennicott Glacier. The imposed velocity pattern strongly affects debris thickness, which in turn leads to annual melt rates that decline towards the terminus. Ice cliff abundance correlates highly with the rate of surface compression, while pond occurrence is strongly negatively correlated with driving stress. A new positive feedback is identified between ice cliffs, streams and surface topography that leads to chaotic topography. As the glacier thinned between 1991 and 2015, surface melt in the study area decreased, despite generally rising air temperatures. Four additional feedbacks relating glacier thinning to melt changes are evident: the debris feedback (negative), the ice cliff feedback (negative), the pond feedback (positive), and the relief feedback (positive). The debris and ice cliff feedbacks, which are tied to the change in surface velocity in time, likely reduced melt rates in time. We show this using a new method to invert for debris thickness change and englacial debris content (∼0.017% by volume) while also revealing that declining speeds and compressive flow led to debris thickening. The expansion of debris on the glacier surface follows changes in flow direction. Ultimately, glacier thinning upvalley from the continuously debris-covered portion of Kennicott Glacier, caused by mass balance changes, led to the reduction of flow into the study area. This caused ice emergence rates to decline rapidly leading to the occurrence of maximum, glacier-wide thinning under thick, insulating debris.
... Debris-covered ice represents 30% of the glacier mass in ablation areas in High Mountain Asia (Kraaijenbrink et al., 2017). Supraglacial debris in the Everest region is typically sufficiently thick to reduce ablation by insulating the underlying ice surface (Nicholson & Benn, 2013). As a result, these debris-covered glaciers have experienced lower sensitivity to atmospheric warming than would be expected for climatically equivalent clean-ice surfaces (Benn et al., 2012). ...
Article
Full-text available
Sustained mass loss from Himalayan glaciers is causing supraglacial debris to expand and thicken, with the expectation that thicker debris will suppress ablation and extend glacier longevity. However, debris-covered glaciers are losing mass at similar rates to clean-ice glaciers in High Mountain Asia. This rapid mass loss is attributed to the combined effects of; (a) low or reversed mass balance gradients across debris-covered glacier tongues, (b) differential ablation processes that locally enhance ablation within the debris-covered section of the glacier, for example, at ice cliffs and supraglacial ponds, and (c) a decrease in ice flux from the accumulation area in response to climatic warming. Adding meter-scale spatial variations in supraglacial debris thickness to an ice-flow model of Khumbu Glacier, Nepal, increased mass loss by 47% relative to simulations assuming a continuous debris layer over a 31-year period (1984–2015 CE) but overestimated the reduction in ice flux. Therefore, we investigated if simulating the effects of dynamic detachment of the upper active glacier from the debris-covered tongue would give a better representation of glacier behavior, as suggested by observations of change in glacier dynamics and structure indicating that this process occurred during the last 100 years. Observed glacier change was reproduced more reliably in simulations of the active, rather than entire, glacier extent, indicating that Khumbu Glacier has passed a dynamic tipping point by dynamically detaching from the heavily debris-covered tongue that contains 20% of the former ice volume.
... The median conductivity of 1.29 J m −1 K −1 matches well with the few available direct measurements from our field site. Reported conductivity values are rare for specific seasons (Nicholson and Benn 2006;Nicholson and Benn 2013), but do exist for unspecified conditions (Conway and Rasmussen 2000;Reid and Brock 2010;Juen et al., 2013;Rounce et al., 2015). These reported values correspond well with the range in conductivity values that we have derived ( Figure 5B). ...
Article
Full-text available
Debris-covered glaciers, especially in high-mountain Asia, have received increased attention in recent years. So far, few field-based observations of distributed mass loss exist and both the properties of the debris layer as well as the atmospheric drivers of melt below debris remain poorly understood. Using multi-year observations of on-glacier atmospheric data, debris properties and spatial surface elevation changes from repeat flights with an unmanned aerial vehicle (UAV), we quantify the necessary variables to compute melt for the Lirung Glacier in the Himalaya. By applying an energy balance model we reproduce observed mass loss during one monsoon season in 2013. We show that melt is especially sensitive to thermal conductivity and thickness of debris. Our observations show that previously used values in literature for the thermal conductivity through debris are valid but variability in space on a single glacier remains high. We also present a simple melt model, which is calibrated based on the results of energy balance model, that is only dependent on air temperature and debris thickness and is therefore applicable for larger scale studies. This simple melt model reproduces melt under thin debris (<0.5 m) well at an hourly resolution, but fails to represent melt under thicker debris accurately at this high temporal resolution. On the glacier scale and using only off-glacier forcing data we however are able to reproduce the total melt volume of a debris-covered tongue. This is a promising result for catchment scale studies, where quantifying melt from debris covered glaciers remains a challenge.
... Zhang and others, 2011). Other approaches include geometrical scaling estimations of exposed debris (Nicholson and Benn, 2012;Nicholson and Mertes, 2017) and calculations from energy-balance models in concert with remote thermal imagery (e.g. Foster and others, 2012;Rounce and McKinney, 2014;Schauwecker and others, 2015) or surface height changes (Ragettli and others, 2015). ...
Article
Full-text available
The thickness of a supraglacial layer is critical to the magnitude and time frame of glacier melt. Field-based, short pulse, ground-penetrating radar (GPR) has successfully measured debris thickness during a glacier's melt season, when there is a strong return from the ice–debris interface, but profiling with GPR in the absence of a highly reflective ice interface has not been explored. We investigated the performance of 960 MHz signals over 2 km of transects on Changri Nup Glacier, Nepal, during the post-monsoon. We also performed laboratory experiments to interpret the field data and investigate electromagnetic wave propagation into dry rocky debris. Laboratory tests confirmed wave penetration into the glacier ice and suggest that the ice–debris interface return was missing in field data because of a weak dielectric contrast between solid ice and porous dry debris. We developed a new method to estimate debris thicknesses by applying a statistical approach to volumetric backscatter, and our backscatter-based calculated thickness retrievals gave reasonable agreement with debris depths measured manually in the field (10–40 cm). We conclude that, when melt season profiling is not an option, a remote system near 1 GHz could allow dry debris thickness to be estimated based on volumetric backscatter.
... a Thickness of the supraglacial debris layer that would result from melting the borehole length or full ice column (as indicated), assuming no debris redistribution at the surface and a bulk effective porosity of 0.33 (ref. 31 ). Calculations are provided the 'Methods' section. ...
Article
Full-text available
Surface melting of High Mountain Asian debris-covered glaciers shapes the seasonal water supply to millions of people. This melt is strongly influenced by the spatially variable thickness of the supraglacial debris layer, which is itself partially controlled by englacial debris concentration and melt-out. Here, we present measurements of deep englacial debris concentrations from debris-covered Khumbu Glacier, Nepal, based on four borehole optical televiewer logs, each up to 150 m long. The mean borehole englacial debris content is ≤ 0.7% by volume in the glacier’s mid-to-upper ablation area, and increases to 6.4% by volume near the terminus. These concentrations are higher than those reported for other valley glaciers, although those measurements relate to discrete samples while our approach yields a continuous depth profile. The vertical distribution of englacial debris increases with depth, but is also highly variable, which will complicate predictions of future rates of surface melt and debris exhumation at such glaciers.
... A seasonal cycle of albedo has been demonstrated in previous observational studies and modelling efforts of broadband albedo, highlighting the importance of continuous measurements (e.g. Hoinkes and Wendler, 1968;Nicholson and Benn, 2013;Möller and Möller, 2017). ...
Article
Full-text available
As Alpine glaciers become snow-free in summer, more dark, bare ice is exposed, decreasing local albedo and increasing surface melting. To include this feedback mechanism in models of future deglaciation, it is important to understand the processes governing broadband and spectral albedo at a local scale. However, few in situ reflectance data have been measured in the ablation zones of mountain glaciers. As a contribution to this knowledge gap, we present spectral reflectance data (hemispherical–conical–reflectance factor) from 325 to 1075 nm collected along several profile lines in the ablation zone of Jamtalferner, Austria. Measurements were timed to closely coincide with a Sentinel-2 and Landsat 8 overpass and are compared to the respective ground reflectance (bottom-of-atmosphere) products. The brightest spectra have a maximum reflectance of up to 0.7 and consist of clean, dry ice. In contrast, reflectance does not exceed 0.2 for dark spectra where liquid water and/or fine-grained debris are present. Spectra can roughly be grouped into dry ice, wet ice, and dirt or rocks, although gradations between these groups occur. Neither satellite captures the full range of in situ reflectance values. The difference between ground and satellite data is not uniform across satellite bands, between Landsat and Sentinel, and to some extent between ice surface types (underestimation of reflectance for bright surfaces, overestimation for dark surfaces). We highlight the need for further, systematic measurements of in situ spectral reflectance properties, their variability in time and space, and in-depth analysis of time-synchronous satellite data.
... However, the subtle weathering rind (Fig. 6) and rounded margins ( Fig. 1) do suggest that, at some time, the boulder was exposed to cold-climate, physical weathering and that stream-related processes may have transported the boulder onto the glacier. Because supraglacial debris tends to become more abundant and smaller in particle size down-glacier (Nakawo, 1979;Nicholson and Benn, 2012), the apparent lack of any other debris in the Cleveland Shale and the apparently isolated nature of the large boulder imply that the boulder may have originated in the shallow reaches of the glacier away from terminal areas. ...
... In 2010-2012, the values of k d increased significantly from June to a maximum in September, which is reflected in Figure 5b and the KM d 2 coefficients for month in Table 4. Short-term observations have found that heat storage in supraglacial debris is close to zero on the order of days (Brock and others, 2007) but may be higher at seasonal scale (Nicholson and Benn, 2013). The increase of values of k d between June and September at Dokriani Glacier suggests that heat accumulated in the debris layer after melting had begun, so that melt rates were increasingly sensitive to surface energy inputs. ...
Article
Full-text available
Supraglacial debris is significant in many regions and complicates modeling of glacier melt, which is required for predicting glacier change and its influences on hydrology and sea-level rise. Temperature-index models are a popular alternative to energy-balance models when forcing data are limited, but their transferability among glaciers and inherent uncertainty have not been documented in application to debris-covered glaciers. Here, melt factors were compiled directly from published studies or computed from reported melt and MERRA-2 air temperature for 27 debris-covered glaciers around the world. Linear mixed-effects models were fit to predict melt factors from debris thickness and variables including debris lithology and MERRA-2 radiative exchange. The models were tested by leave-one-site-out cross-validation based on predicted melt rates. The best model included debris thickness (fixed effect) and glacier and year (random effects). Predictions were more accurate using MERRA-2 than on-site air temperature data, and pooling MERRA-2-derived and reported melt factors improved cross-validation accuracy more than including additional predictors such as shortwave or longwave radiation. At one glacier where monthly ablation was measured over 4 years, seasonal variation of melt factors suggested that heat storage significantly affected the relation between melt and energy exchange at the debris surface.
... The Astore and Hunza sub-basins range from 1504 to 8069 and 1420 to 7809 m.a.s.l. with around 14% and 28% of the total area, respectively covered by glaciers and permanent ice cover [13,30]. Importantly, it may also be necessary to consider that glacier cover melt-rates vary depending on whether a glacier consideration is debris-covered or clean-ice [31][32][33]. Similarly, the most recent studies based on snow cover assessment and hydrological modelling were carried out in the Gilgit River Basin (GRB) and neighboring river basins [27][28][29]34]. ...
Article
Full-text available
In contrast to widespread glacier retreat evidenced globally, glaciers in the Karakoram region have exhibited positive mass balances and general glacier stability over the past decade. Snow and glacier meltwater from the Karakoram and the western Himalayas, which supplies the Indus River Basin, provide an essential source of water to more than 215 million people, either directly, as potable water, or indirectly, through hydroelectric generation and irrigation for crops. This study focuses on water resources in the Upper Indus Basin (UIB) which combines the ranges of the Hindukush, Karakoram and Himalaya (HKH). Specifically, we focus on the Gilgit River Basin (GRB) to inform more sustainable water use policy at the sub-basin scale. We employ two degree-day approaches, the Spatial Processes in Hydrology (SPHY) and Snowmelt Runoff Model (SRM), to simulate runoff in the GRB during 2001-2012. The performance of SRM was poor during July and August, the period when glacier melt contribution typically dominates runoff. Consequently, SPHY outperformed SRM, likely attributable to SPHY's ability to discriminate between glacier, snow, and rainfall contributions to runoff during the ablation period. The average simulated runoff revealed the prevalent snowmelt contribution as 62%, followed by the glacier melt 28% and rainfall 10% in GRB. We also assessed the potential impact of climate change on future water resources, based on two Representative Concentration Pathways (RCP) (RCP 4.5 and RCP 8.5). We estimate that summer flows are projected to increase by between 5.6% and 19.8% due to increased temperatures of between 0.7 and 2.6 °C over the period 2039-2070. If realized, increased summer flows in the region could prove beneficial for a range of sectors, but only over the short to medium term and if not associated with extreme events. Long-term projections indicate declining water resources in the region in terms of snow and glacier melt.
... Also, Collier et al. (2015) linearly distributed the debris thickness from terminus to the upper debris-covered zone to model and compare melt rates between debris cover and debris free ice. The findings are useful for the impact assessment of debris cover on glacier-melt but for accurate model parametrization, estimation of field-based melt rate and spatial distribution of debris is necessary (Nicholson and Benn, 2013). ...
... Also, Collier et al. (2015) linearly distributed the debris thickness from terminus to the upper debris-covered zone to model and compare melt rates between debris cover and debris free ice. The findings are useful for the impact assessment of debris cover on glacier-melt but for accurate model parametrization, estimation of field-based melt rate and spatial distribution of debris is necessary (Nicholson and Benn, 2013). ...
Article
Full-text available
The assessment of meltwater sourcing from the clean and debris-covered glaciers is scarce in High Mountain Asia (HMA). The melting rate varies with the debris cover thickness and glacier orientation. The present study quantifies glacier melting rate attributed to varying thickness of debris cover in the Karakoram. We observed daily melting rates by installing ablation stakes over debris-free and debris-covered ice during a field expedition. The stakes were installed on glacier surface with debris cover thickness ranges between 0.5 and 40 cm at selected experimental sites during the ablation period (September and October 2018) and (July to August 2019). We selected three glaciers including Ghulkin, Hinarchi, and Hoper facing east, south, and north, respectively to assess the role of glacier orientation on melting rates. We observed that the debris-free ice melts faster than the debris-covered ice. Intriguingly, a thin debris layer of 0.5 cm does not enhance melting compared to the clean ice which is inconsistent with the earlier studies. The melting rate decreases as the thickness of debris cover increases at all the three selected glaciers. Furthermore, south-facing glacier featured the highest melting (on average ~ 25% more). However, the north and eastfacing glaciers revealed almost same melting rates. We observed that the average degree-day factors (DDF) slightly varies within a range of 0.58–0.73 and 0.55–0.68 cm °C−1 day−1 for debris-free and 0.5 cm debris-covered ice, respectively, however, DDF largely reduces to 0.13–0.25 cm °C−1 day−1 for 40 cm
... In turn, a debris cover may develop and expand. Ice ablation is retarded once the debris reaches a critical thickness (Østrem, 1959;Mattson et al., 1993;Kayastha et al., 2000;Kirkbride & Dugmore, 2003;Nicholson & Benn, 2013), triggering an increase in relief relative to adjacent debris-free surfaces. Thereafter, gravity-driven secondary dispersal mechanisms redistribute debris across the glacier surface to form a continuous debris mantle. ...
Article
Full-text available
There exists a need to advance our understanding of debris‐covered glacier surfaces over relatively short timescales due to rapid, climatically induced areal expansion of debris cover at the global scale, and the impact debris has on mass balance. We applied UAV‐SfM and DEM differencing with debris thickness and debris stability modelling to unravel the evolution of a 0.15 km2 region of the debris‐covered Miage Glacier, Italy, between June 2015 and July 2018. DEM differencing revealed widespread surface lowering (mean 4.1 ± 1.0 m a‐1; max 13.3 m a‐1). We combined elevation change data with local meteorological data and a sub‐debris melt model, and used these relationships to produce high resolution, spatially distributed maps of debris thickness. These maps were differenced to explore patterns and mechanisms of debris redistribution. Median debris thicknesses ranged from 0.12 – 0.17 m and were spatially variable. We observed localised debris thinning across ice cliff faces, except those which were decaying, where debris thickened. We observed pervasive debris thinning across larger, backwasting slopes, including those bordered by supraglacial streams, as well as ingestion of debris by a newly exposed englacial conduit. Debris stability mapping showed that 18.2‐26.4% of the survey area was theoretically subject to debris remobilisation. By linking changes in stability to changes in debris thickness, we observed that slopes that remain stable, stabilise, or remain unstable between periods almost exclusively show net debris thickening (mean 0.07 m a‐1) whilst those which become newly unstable exhibit both debris thinning and thickening. We observe a systematic downslope increase in the rate at which debris cover thickens which can be described as a function of the topographic position index and slope gradient. Our data provide quantifiable insights into mechanisms of debris remobilisation on glacier surfaces over sub‐decadal timescales, and open avenues for future research to explore glacier‐scale spatiotemporal patterns of debris remobilisation.
... Also, Collier et al. (2015) linearly distributed the debris thickness from terminus to the upper debris-covered zone to model and compare melt rates between debris cover and debris free ice. The findings are useful for the impact assessment of debris cover on glacier-melt but for accurate model parametrization, estimation of field-based melt rate and spatial distribution of debris is necessary (Nicholson and Benn, 2013). ...
Article
Full-text available
The assessment of meltwater sourcing from the clean and debris-covered glaciers is scarce in High Mountain Asia (HMA). The melting rate varies with the debris cover thickness and glacier orientation. The present study quantifies glacier melting rate attributed to varying thickness of debris cover in the Karakoram. We observed daily melting rates by installing ablation stakes over debris-free and debris-covered ice during a field expedition. The stakes were installed on glacier surface with debris cover thickness ranges between 0.5 and 40 cm at selected experimental sites during the ablation period (September and October 2018) and (July to August 2019). We selected three glaciers including Ghulkin, Hinarchi, and Hoper facing east, south, and north, respectively to assess the role of glacier orientation on melting rates. We observed that the debris-free ice melts faster than the debris-covered ice. Intriguingly, a thin debris layer of 0.5 cm does not enhance melting compared to the clean ice which is inconsistent with the earlier studies. The melting rate decreases as the thickness of debris cover increases at all the three selected glaciers. Furthermore, south-facing glacier featured the highest melting (on average ~ 25% more). However, the north and east-facing glaciers revealed almost same melting rates. We observed that the average degree-day factors (DDF) slightly varies within a range of 0.58–0.73 and 0.55–0.68 cm °C−1 day−1 for debris-free and 0.5 cm debris-covered ice, respectively, however, DDF largely reduces to 0.13–0.25 cm °C−1 day−1 for 40 cm debris-covered ice. We suggest continuous physical glacier ablation observations for various debris cover throughout the ablation zone to better understand the role of debris on melting.
... However, the subtle weathering rind (Fig. 6) and rounded margins ( Fig. 1) do suggest that, at some time, the boulder was exposed to cold-climate, physical weathering and that stream-related processes may have transported the boulder onto the glacier. Because supraglacial debris tends to become more abundant and smaller in particle size down-glacier (Nakawo, 1979;Nicholson and Benn, 2012), the apparent lack of any other debris in the Cleveland Shale and the apparently isolated nature of the large boulder imply that the boulder may have originated in the shallow reaches of the glacier away from terminal areas. ...
Preprint
Full-text available
Debris-covered glaciers exist in many mountain ranges and play an important role in the regional water cycle. However, modelling the surface mass balance, runoff contribution and future evolution of debris-covered glaciers is fraught with uncertainty as accurate information on small-scale variations in debris thickness and sub-debris ice melt rates is only available for a few locations worldwide. Here we present a customised low-cost UAV for high-resolution thermal imaging of mountain glaciers and a complete open-source pipeline that facilitates the generation of accurate surface temperature and debris thickness maps from radiometric images. First, a thermal orthophoto is computed from individual radiometric UAV images using structure-from-motion and multi-view-stereo techniques. User-specific calibration and correction procedures can then be applied to the raw thermal orthophoto to account for atmospheric and environmental influences that affect the radiometric measurement. The corrected thermal orhthophoto reflects spatial variations in surface temperature across the surveyed debris-covered area. Finally, a high-resolution debris thickness map is derived from the corrected thermal orthophoto using in-situ measurements in conjuction with an empirical or inverse surface energy balance model that relates surface temperature to debris thickness. Our results from a small-scale experiment on the Kanderfirn in the Swiss Alps show that the surface temperature and thickness of a relatively thin debris layer (ca. 0–15 cm) can be mapped with high accuracy. On snow and ice surfaces, the mean deviation of the mapped surface temperature from the melting point (∼0 °C) was 0.4 ±1.0 °C. The root-mean-square error of the modelled debris thickness was 1.2 cm. Through the detailed mapping, typical small-scale debris features and debris thickness patterns become visible, which are not spatially resolved by the thermal infrared sensors of current-generation satellites. The presented approach paves the way for glacier-wide high-resolution debris thickness mapping and opens up new opportunities for more accurate monitoring and modelling of debris-covered glaciers.
Article
The ablation model for debris-covered glaciers that uses thermal resistance, which includes both debris thickness and thermal conductivity of the debris layer and requires only surface temperature and meteorological data to be acquired simultaneously. However, this method overestimates the ablation amounts for thick debris layers, since it neglects changes in heat storage in the debris layer. Conversely, recent studies on debris-covered glaciers have concentrated on developing the physical model of ablation of debris covered ice. However, physical model requires several parameters such as thermal conductivity and thickness of debris layers, which are difficult to obtain without digging debris layers. Therefore, it is difficult to employ this model over a wide area. By considering the advantages and disadvantages of each physical model, we could improve the ablation model of debris-covered glaciers.
Book
Full-text available
Considerable parts of the ablation zone of many high mountain glaciers are covered with supraglacial debris. The presence of supraglacial debris complicates the interplay of processes linking climate change, topography, and glacier dynamics. Debris-covered glaciers (DCGs) thus differ significantly from their clean-ice counterparts, as they form a more complex system of forcing factors, couplings, and feedback mechanisms that are yet to be fully understood. Resolving the uncertain response and evolution of debris-covered glaciers is vital for devising sustainable management strategies for freshwater availability, glacier-related hazards, hydro-power generation, and also for more precise estimation of their contribution to eustatic sea level changes. The articles in this Research Topic cover conceptual, modelling, and observational approaches to study DCGs at point to glacier and regional scales.
Article
Full-text available
Supraglacial debris covers 7% of mountain glacier area globally and generally reduces glacier surface melt. Enhanced energy absorption at ice cliffs and supraglacial ponds scattered across the debris surface leads these features to contribute disproportionately to glacier-wide ablation. However, the degree to which cliffs and ponds actually increase melt rates remains unclear, as these features have only been studied in a detailed manner for selected locations, almost exclusively in High Mountain Asia. In this study we model the surface energy balance for debris-covered ice, ice cliffs, and supraglacial ponds at a set of automatic weather station records representing the global prevalence of debris-covered glacier ice. We generate 5000 random sets of values for physical parameters using probability distributions derived from literature, which we use to investigate relative melt rates and to isolate the melt responses of debris, cliffs and ponds to the site-specific meteorological forcing. Modelled sub-debris melt rates are primarily controlled by debris thickness and thermal conductivity. At a reference thickness of 0.1 m, sub-debris melt rates vary considerably, differing by up to a factor of four between sites, mainly attributable to air temperature differences. We find that melt rates for ice cliffs are consistently 2-3x the melt rate for clean glacier ice, but this melt enhancement decays with increasing clean ice melt rates. Energy absorption at supraglacial ponds is dominated by latent heat exchange and is therefore highly sensitive to wind speed and relative humidity, but is generally less than for clean ice. Our results provide reference melt enhancement factors for melt modelling of debris-covered glacier sites, globally, while highlighting the need for direct measurement of debris-covered glacier surface characteristics, physical parameters, and local meteorological conditions at a variety of sites around the world.
Article
Full-text available
Known for their important role in locally enhancing surface melt, supraglacial ponds and ice cliffs are common features on debris-covered glaciers. We use high resolution satellite imagery to describe pond-cliff systems and surface velocity on Verde debris-covered glacier, Monte Tronador, and Southern Chile. Ponds and ice cliffs represent up to 0.4 and 2.7% of the glacier debris-covered area, respectively. Through the analyzed period and the available data, we found a seasonality in the number of detected ponds, with larger number of ponds at the beginning of the ablation season and less at the end of it. Using feature tracking, we determined glacier surface velocity, finding values up to 55 m/yr on the upper part of the debris-covered area, and decreasing almost to stagnation in the terminus. We found that larger ponds develop in glacier zones of low velocity, while zones of high velocity only contain smaller features. Meanwhile, ice cliffs appeared to be less controlled by surface velocity and gradient. Persistent ice cliffs were detected between 2009 and 2019 and backwasting up to 24 m/yr was measured, highlighting significant local glacier wastage.
Article
Full-text available
Supraglacial debris affects glacier mass balance as a thin layer enhances surface melting, while a thick layer reduces it. While many glaciers are debris-covered, global glacier models do not account for debris because its thickness is unknown. We provide the first globally distributed debris thickness estimates using a novel approach combining sub-debris melt and surface temperature inversion methods. Results are evaluated against observations from 22 glaciers. We find the median global debris thickness is ∼0.15 ± 0.06 m. In all regions, the net effect of accounting for debris is a reduction in sub-debris melt, on average, by 37%, which can impact regional mass balance by up to 0.40 m water equivalent (w.e.) yr-1. We also find recent observations of similar thinning rates over debris-covered and clean ice glacier tongues is primarily due to differences in ice dynamics. Our results demonstrate the importance of accounting for debris in glacier modeling efforts.
Article
Full-text available
During the last few decades, the lake-terminating glaciers in the Himalaya have receded faster than the land-terminating glaciers as proglacial lakes have exacerbated the mass loss of their host glaciers. Monitoring the impacts of glacier recession and dynamics on lake extent and water volume provides an approach to assess the mass interplay between glaciers and proglacial lakes. We describe the recession of Longbasaba Glacier and estimate the mass wastage and its contribution to the water volume of its proglacial lake. The results show that the glacier area has decreased by 3% during 1988–2018, with a more variable recession prior to 2008 than in the last decade. Longbasaba Lake has expanded by 164% in area and 237% in water volume, primarily as a result of meltwater inflow produced from surface lowering of the glacier. Over the periods 1988–2000 and 2000–18, the mass loss contributed by glacier thinning has decreased from 81 to 61% of the total mass loss, accompanied by a nearly doubled contribution from terminus retreat. With the current rate of retreat, Longbasaba glacier is expected to terminate in its proglacial lake for another four decades. The hazard risk of this lake is expected to continue to increase in the near future because of the projected continued glacier mass loss and related lake expansion.
Article
Full-text available
Thaw slumps in ice‐rich permafrost can retreat tens of metres per summer, driven by the melt of subaerially exposed ground ice. However, some slumps retain an ice‐veneering debris cover as they retreat. A quantitative understanding of the thermal regime and geomorphic evolution of debris‐covered slumps in a warming climate is largely lacking. To characterize the thermal regime, we instrumented four debris‐covered slumps in the Canadian Low Arctic and developed a numerical conduction‐based model. The observed surface temperatures 20°C and steep thermal gradients indicate that debris insulates the ice by shifting the energy balance towards radiative and turbulent losses. After the model was calibrated and validated with field observations, it predicted sub‐debris ice melt to decrease four‐fold from 1.9 to 0.5 m as the thickness of the fine‐grained debris quadruples from 0.1 to 0.4 m. With warming temperatures, melt is predicted to increase most rapidly, in relative terms, for thick (~0.5‐1.0 m) debris covers. The morphology and evolution of the debris‐covered slumps were characterized using field and remote sensing observations, which revealed differences in association with morphology and debris composition. Two low‐angle slumps retreated continually despite their persistent fine‐grained debris covers. The observed elevation losses decreased from ~1.0 m/yr where debris thickness ~.2 m to 0.1 m/yr where thickness ~1.0 m. Conversely, a steep slump with a coarse‐grained debris veneer underwent short‐lived bursts of retreat, hinting at a complex interplay of positive and negative feedback processes. The insulative protection and behaviour of debris vary significantly with factors such as thickness, grain size and climate: debris thus exerts a fundamental, spatially variable influence on slump trajectories in a warming climate.
Article
Full-text available
Debris-covered glaciers account for almost one-fifth of the total glacier ice volume in High Mountain Asia; however, their contribution to the total glacier melt remains uncertain, and the drivers controlling this melt are still largely unknown. Debris influences the properties (e.g. albedo, thermal conductivity, roughness) of the glacier surface and thus the surface energy balance and glacier melt. In this study we have used sensitivity tests to assess the effect of surface properties of debris on the spatial distribution of micrometeorological variables such as wind fields, moisture and temperature. Subsequently we investigated how those surface properties drive the turbulent fluxes and eventually the conductive heat flux of a debris-covered glacier. We simulated a debris-covered glacier (Lirung Glacier, Nepal) at a 1 m resolution with the MicroHH model, with boundary conditions retrieved from an automatic weather station (temperature, wind and specific humidity) and unmanned aerial vehicle flights (digital elevation map and surface temperature). The model was validated using eddy covariance data. A sensitivity analysis was then performed to provide insight into how heterogeneous surface variables control the glacier microclimate. Additionally, we show that ice cliffs are local melt hot spots and that turbulent fluxes and local heat advection amplify spatial heterogeneity on the surface. The high spatial variability of small-scale meteorological variables suggests that point-based station observations cannot be simply extrapolated to an entire glacier. These outcomes should be considered in future studies for a better estimation of glacier melt in High Mountain Asia.
Article
The hydrological characteristics of debris-covered glaciers are known to be fundamentally different from those of clean-ice glaciers, even within the same climatological, geological and geomorphological setting. Understanding how these characteristics influence the timing and magnitude of meltwater discharge is particularly important for regions like High Mountain Asia, where downstream communities rely on this resource for sanitation, irrigation and hydropower. The hydrology of debris-covered glaciers is relatively complex: rugged surface topographies typically route meltwater through compound supraglacial-englacial systems involving both channels and ponds, as well as pathways that remain unknown. Low-gradient tongues that extend several kilometres retard water conveyance and promote englacial storage. Englacial channels are frequently abandoned and reactivated as water supply changes, new lines of permeability are exploited, and drainage is captured due to high rates of surface and subsurface change. Seasonal influences, such as the monsoon, are superimposed on these distinctive characteristics, reorganising surface and subsurface drainage rapidly from one season to the next. Recent advances in understanding have mostly come from studies aimed at quantifying and describing supraglacial processes; little is known about the subsurface hydrology, particularly the nature (or even existence) of subglacial drainage. In this review, we consider in turn the supraglacial, englacial, subglacial, and proglacial hydrological domains of debris-covered glaciers in High Mountain Asia. We summarise different lines of evidence to establish the current state of knowledge and, in doing so, identify major knowledge gaps. Finally, we use this information to suggest priorities for future hydrological research at High Mountain Asian debris-covered glaciers, and how they may influence our ability to be able to make long-term predictions of changes in the water they supply.
Article
Full-text available
Temperatures from a bore hole through an active rock glacier in the eastern Swiss Alps are presented and thermal conditions within the slowly creeping permafrost are analyzed. Present mean annual temperature in the uppermost part of the permafrost is −3°C. Permafrost is 52 m thick and reaches heavily fissured bedrock. Thermal conductivity as determined in situ from seasonal temperature variations and measured in a cold laboratory using frozen samples is close to 2.5–3.0 W m−1 °C−1. Vertical heat flow is anomalously high (around 150 mW m-2), probably due to heat advection from circulating ground water or air within the fissured bedrock zone. Beneath this zone, which could in fact represent a non-frozen intra-permafrost layer or “talik”, relic permafrost from past centuries may possibly exist as indicated by a corresponding heat-flow inversion. Given the current temperature condition at the surface of the rock glacier and the fact that the twentieth century is among the warmest in post-glacial time, permafrost conditions may be assumed to have existed during the whole of the Holocene and, hence, during the entire time of rock-glacier formation.
Article
Full-text available
Quantitative investigations have been made of ice-cored dirt cones on Bersaerkerbræ in north-east Greenland. Experiments were also undertaken to evaluate field observations. Measurements included: maximum cone dimensions, sediment thickness and particle size, cone growth rates, slope angles and the temperature distribution within the debris layer and ice core. Particle size, which has not been stressed in previous studies, and related liquid consistency limits, appear as the dominant controls in cone formation, independent of debris thickness within the observed range of 10 mm to 125 mm. A threshold grain-size for dirt-cone inception was found, between 0.2 mm and 0.6 mm. The growth of cones was usually not more than 50% of the ablation over “clean” ice. Temperature measurements within dirt cones has enabled heat-flow studies to be made, evaluating the thermal conductivity of a sediment layer and the heat transfer involved in melting the ice core. A simple model of dirt-cone dynamics is proposed, characterized by negative feedbacks and describing a steady-state system.
Article
Full-text available
Temperatures from a bore hole through an active rock glacier in the eastern Swiss Alps are presented and thermal conditions within the slowly creeping permafrost are analyzed. Vertical heat flow is anomalously high (around 150 MW m-2), probably due to heat advection from circulating ground water on air within the fissured bedrock zone. Beneath this zone, which could in fact represent a non-frozen intra-permafrost layer or "talik', relic permafrost from past centuries may possibly exist as indicated by a corresponding heat-flow inversion. Given the current temperature conditions at the surface of the rock glacier and the fact that the twentieth century is among the warmest in post-glacial time, permafrost conditions may be assumed to have existed during the whole of the Holocene and, hence, during the entire time of rock-glacier formation. -from Authors
Article
Full-text available
Temperature measurements made during summer within supraglacial debris on Khumbu Glacier, Nepal show a sttong diurnal signal that diffused downward into the debris with decreasing amplitude and increasing lag. Surface temperatures during the day were up to 35°C higher than the air temperature; energy transfer into the debris was dominated by the solar radiative flux. Temperature profiles through the debris indicate that heat flow deeper than about 0.2 m was primarily by conduction. The thermal conductivity k of the debris, estimated from a calculated thermal diffusivity and a representative volumetric heat capacity, was 0.85 ± 0.20 W rn 1 K" 1 at one site and 1.28 ± 0.15 W m" 1 K" 1 at another. At the first site the debris was 0.40 m thick and the average temperature gradient dT/dz = 19 K m' 1 ; the average flux of energy through the debris was sufficient to melt 4—6 mm of ice per day. The debris was thicker (estimated to be 2.5 m) and the temperature gradient lower (4.5 K m' 1) at the second site, and the calculated ice-melt was less than 2 mm day" 1 .
Article
Full-text available
Temperatures from a bore hole through an active rock glacier in the eastern Swiss Alps are presented and thermal conditions within the slowly creeping permafrost are analyzed. Present mean annual temperature in the uppermost part of the permafrost is -3· C. Permafrost is 52 m thick and reaches heavily fissured bedrock. Thermal conductivity as determined ill silu from seasonal temperature variations and measured in a cold laboratory using frozen samples is close to 2.5-3.0 W m-I ·C-l . Vertical heat flow is anomalously high (around 150 mW m-2), probably due to heat advection from circulating ground water or air within the fissured bedrock zone. Beneath this zone, which could in fact represent a non-frozen intra-permafrost layer or "talik", relic permafrost from past centuries may possibly exist as indicated by a corresponding heat-flow inversion. Given the current temperature condition at the surface of the rock glacier and the fact that the twentieth century is among the warmest in post-glacial time, permafrost conditions may be assumed to have existed during the whole of the Holocene and, hence, during the entire time of rock-glacier formation.
Article
Full-text available
There are many supraglacial ponds on debris-covered glaciers in the Nepal Himalayas. The heat absorbed at the surface of a pond was estimated from heat budget observations on the Lirung Glacier in Langtang Valley, Nepal. The results indicated an average heat absorption of 170 W m-2 during the summer monsoon season. This rate is about 7 times the average for the whole debris-covered zone. Analysis of the heat budget for a pond suggests that at least half of the heat absorbed at a pond surface is released with the water outflow from the pond, indicating that the water warmed in the pond enlarges the englacial conduit that drains water from the pond and produces internal ablation. Furthermore, the roof of the conduit could collapse, leading to the formation of ice cliffs and new ponds, which would accelerate the ablation of the debris-covered glacier.
Conference Paper
Full-text available
Coarse blocks are a widespread ground cover in cold mountain areas. They have been recognized to exert a cooling influence on subsurface temperatures in comparison with other types of surface material and are employed in man- made structures for ground cooling and permafrost protection. The contrast in heat transfer between the atmosphere and the ground caused by thermally driven convection in winter and stable stratification of interstitial air during summer is usually invoked to explain this “thermal diode” effect. Based on measurements and model calculations, we propose an additional cooling mechanism, which is independent of convection, and solely functions based on the interplay of a winter snow cover and a layer of coarse blocks with low thermal conductivity. The thermal conductivity of a block layer with a porosity of 0.4 is reduced by about an order of magnitude compared to solid rock. We use a simple and purely conductive model experiment to demonstrate that low-conductivity layers reduce the temperature below the winter snow cover as well as mean annual ground temperatures by comparison with other ground materials. Coarse block layers reduce the warming effect of the snow cover and can result in cooling of blocky surfaces in comparison with surrounding areas in the order of one or several degrees. The characteristics of this mechanism correspond to existing measurements.
Article
Full-text available
Surface glacier debris samples and field spectra were collected from the ablation zones of Nepal Himalaya Ngozumpa and Khumbu glaciers in November and December 2009. Geochemical and mineral compositions of supraglacial debris were determined by X-ray diffraction and X-ray fluorescence spectroscopy. This composition data was used as ground truth in evaluating field spectra and satellite supraglacial debris composition and mapping methods. Satellite remote sensing methods for characterizing glacial surface debris include visible to thermal infrared hyper- and multispectral reflectance and emission signature identification, semi-quantitative mineral abundance indicies and spectral image composites. Satellite derived supraglacial debris mineral maps displayed the predominance of layered silicates, hydroxyl-bearing and calcite minerals on Khumbu Himalayan glaciers. Supraglacial mineral maps compared with satellite thermal data revealed correlations between glacier surface composition and glacier surface temperature. Glacier velocity displacement fields and shortwave, thermal infrared false color composites indicated the magnitude of mass flux at glacier confluences. The supraglacial debris mapping methods presented in this study can be used on a broader scale to improve, supplement and potentially reduce errors associated with glacier debris radiative property, composition, areal extent and mass flux quantifications.
Article
Full-text available
In areas of high relief, many glaciers have extensive covers of supraglacial debris in their ablation zones, which alters both rates and spatial patterns of melting, with important consequences for glacier response to climate change. Wastage of debris-covered glaciers can be associated with the formation of large moraine-dammed lakes, posing risk of glacier lake outburst floods (GLOFs). In this paper, we use observations of glaciers in the Mount Everest region to present an integrated view of debris-covered glacier response to climate change, which helps provide a long-term perspective on evolving GLOF risks.
Article
Full-text available
During the 2005-2007 June-September ablation seasons, meteorological conditions were recorded on the lower and upper parts of the debris-covered ablation zone of Miage Glacier, Italy. In 2005, debris temperature and subdebris ice melt were also monitored at 25 points with debris thickness of 0.04-0.55 m, spread over 5 km2 of the glacier. The radiative fluxes were directly measured, and near-closure of the surface energy balance is achieved, providing support for the bulk aerodynamic calculation of the turbulent fluxes. Surface-layer meteorology and energy fluxes are dominated by the pattern of incoming solar radiation which heats the debris, driving strong convection. Mean measured subdebris ice melt rates are 6-33 mm d-1, and mean debris thermal conductivity is 0.96 W m-1 K-1, displaying a weak positive relationship with debris thickness. Mean seasonal values of the net shortwave, net longwave, and debris heat fluxes show little variation between years, despite contrasting meteorological conditions, while the turbulent latent (evaporative) heat flux was more than twice as large in the wet summer of 2007 compared with 2005. The increase in energy output from the debris surface in response to increasing surface temperature means that subdebris ice melt rates are fairly insensitive to atmospheric temperature variations in contrast to debris-free glaciers. Improved knowledge of spatial patterns of debris thickness distribution and 2 m air temperature, and the controls on evaporation of rainwater from the surface, are needed for distributed physically based melt modeling of debris-covered glaciers.
Article
Full-text available
Monitoring of air and ground temperature at Plateau Mountain (South-Western Alberta) at short intervals (20 minutes) for two years shows vastly different thermal regimes in and beneath coarse blocky materials as opposed to mineral soils and rocks lacking substantial interconnecting voids. The dominant process of heat transfer in the upper layers is by rapid air movement through the voids to at least 50 cm depth as compared with slow conduction through the individual grains. Thermal response to a change in air temperature (positive or negative) is immediate and substantial, so it is not merely the result of the Balch effect. Rain and snow can also penetrate more deeply. These blocky materials are called kurums in Russia. Mean annual ground temperatures are 4-7 °C cooler in the blocky materials than in the adjacent mineral soils in cold climates, but this would be different in warmer climates. The ground temperature envelope is cone-shaped rather than bell-shaped, and this difference also occurs in mineral soils under a thin cover of blocks. There is also a smaller geothermal gradient within the zone affected directly by cooling/heating due to air movement. These processes appear to explain the occurrence of permafrost and substantial ice bodies in block fields such as rock glaciers below the limit of regional continuous permafrost in adjacent rocks and mineral soils. They also affect permafrost mapping and heat flow modelling, but offer a means of cooling near-surface soils.
Article
Full-text available
Debris-covered glaciers respond differently to any given set of climatic conditions than clean glaciers. This difference stems from change in ablation rate caused by a debris cover, approaching zero ablation with sufficiently thick protective cover, a mechanism not yet considered in the context of ice sheet growth. Critical to applying the mechanism to ice sheets is supply of debris. We postulate that periods after major interglacials offer the best conditions for dirty advances. This is because the volume of debris, previously deposited and weathered in interglacial time, the latter a function of exposure length, should be at a maximum. Advances of dirty ice sheets generate landforms and in positions deviating from clean-ice advances under similar climatic conditions. Hence, inferences of both ice sheet properties inferred from such dirty ice advances and the climate conditions related to such advances must be cautious. Furthermore, modeling of past ice sheets must consider the effect of debris cover whenever indications exist for such a cover. r 2002 Elsevier Science Ltd and INQUA. All rights reserved.
Article
Full-text available
Generalized numerical models of sub-debris ice ablation are preferable to empirical approaches for predicting runoff and glacier response to climate change, as empirical methods are site-specific and strongly dependent upon the conditions prevailing during the measurement period. We present a modified surface energy-balance model to calculate melt beneath a surface debris layer from daily mean meteorological variables. Despite numerous simplifications, the model performs well and modelled melt rates give a good match to observed melt rates, suggesting that this model can produce reliable estimates of ablation rate beneath debris layers several decimetres thick. This is a useful improvement on previous models which are inappropriate for thick debris cover.
Article
Full-text available
This paper presents a simple model to estimate ice ablation under a thick supraglacial debris cover. The key method employed in the model is to establish a link between the debris heat flux and the debris temperature at a certain depth when the heat transfer in the debris is described by a diffusion process. Given surface temperature, debris thermal properties and relevant boundary conditions, the proposed model can estimate mean debris temperature at interfaces of different debris layers using an iterative procedure, and then the heat flux for ice ablation. The advantage of the proposed model is that it only requires a few parameters to conduct the modeling, which is simpler and more applicable than others. The case study on Koxkar glacier, west Tien Shan, China, shows, in general, that the proposed model gives good results for the prediction of debris temperatures, except for an apparent phase shift between modeled and observed values. We suggest that this error is mainly due to complex phase relations between debris temperature and debris heat flux. The modeled ablation rates at three experimental sites also show good results, using a direct comparison with observed data and an indirect comparison with a commonly used energy-balance model.
Article
Full-text available
Permafrost is known to respond to changes in atmospheric temperatures but our knowledge of the scale and the processes involved are still not fully understood, especially not when the surface is other than bedrock. In order to gain a better understanding of the influence of the surface and ground characteristics on the thermal regime of the active layer, boreholes were drilled in the high alpine discontinuous permafrost in the Murtel-Chastelets area, Eastern Swiss Alps, and instrumented with chains of thermistor arrays. Five shallow boreholes were drilled during the summer of 2002, adding to the network of the existing four boreholes. Data from the first year of this multi-annual study demonstrate a pronounced thermal offset for loose material composed of gravel or boulders. Also, the thermal diffusivity was calculated over the summer period, adding to the understanding of the spatial discontinuity in the Alpine permafrost.
Article
Full-text available
Here we report a laboratory study of the effects of debris thickness, diurnally cyclic radiation and rainfall on melt rates beneath rock-avalanche debris and sand (representing typical highly permeable supraglacial debris). Under continuous, steady-state radiation, sand cover >50mm thick delays the onset of ice-surface melting by >12 hours, but subsequent melting matches melt rates of a bare ice surface. Only when diurnal cycles of radiation are imposed does the debris reduce the longterm rate of ice melt beneath it. This is because debris >50 mm thick never reaches a steady-state heat flux, and heat acquired during the light part of the cycle is partially dissipated to the atmosphere during the nocturnal part of the cycle, thereby continuously reducing total heat flux to the ice surface underneath. The thicker the debris, the greater this effect. Rain advects heat from high-permeability supraglacial debris to the ice surface, thereby increasing ablation where thin, highly porous material covers the ice. In contrast, low-permeability rock-avalanche material slows water percolation, and heat transfer through the debris can cease when interstitial water freezes during the cold/night part of the cycle. This frozen interstitial water blocks heat advection to the ice-debris contact during the warm/day part of the cycle, thereby reducing overall ablation. The presence of metre-deep rock-avalanche debris over much of the ablation zone of a glacier can significantly affect the mass balance, and thus the motion, of a glacier. The length and thermal intensity of the diurnal cycle are important controls on ablation, and thus both geographical location and altitude significantly affect the impact of debris on glacial melting rates; the effect of debris cover is magnified at high altitude and in lower latitudes.
Book
This book describes the effects of cold climates on the surface of the earth. Using scientific principles, the authors describe the evolution of ground thermal conditions and the origin of natural features such as frost heave, solifluction, slope instabilities, patterned ground, pingos and ice wedges. The thermodynamic conditions accompanying the freezing of water in porous materials are examined and their fundamental role in the ice segregation and frost heave processes is demonstrated in a clear and simple manner. This book concentrates on the analysis of the causes and effects of frozen ground phenomena, rather than on the description of the natural features characteristic of freezing or thawing ground. Its scientific approach provides a basis for geotechnical analyses such as those essential to resource development.
Article
A simple model suggests that the ablation under a debris layer could be estimated from meteorological variables if the surface temperature data of the layer are available. This method was tested by analyzing the data obtained from experiments with artificial debris layers. Fairly good agreement was obtained between the estimated and the experimental data.
Article
The annually thawing active layer of permafrost is central to considerations of climate change consequences in arctic areas and interpretations of deep permafrost temperatures that constitute and exceptional archive of past climate change. Moreover, a sound understanding of the thermal regime of the active layer is of great interest, because all chemical, biological and physical processes are concentrated there. The author studied this layer by examining the soil physical properties and heat transfer processes that dictate soil temperatures for an arctic desert site in northwestern Spitsbergen. A wide array of soil physical properties based on field observations and laboratory measurements were defined. These include mineralogy, grain size distribution, local regolith thickness, porosity, density, typical soil moisture profile, heat capacity and thermal conductivity. Heat transfer processes were studied through modeling of soil temperatures. The heat transfer model accounted for much of the observed soil thermal regime. It was found that thermal conduction, phase change of soil water at 0°C, and changes in unfrozen water content are the primary thermal processes that explain the observed soil temperatures in this field site. Melt-water infiltration, which is often overlooked in the energy budget, causes abrupt warming events and delivers considerable energy to the soil in late spring. An increase in frequency or magnitude of infiltration events could mimic simple spring time surface warming. Advection of ground water and soil internal evaporation were found to be generally unimportant at the site studied.
Article
Glaciers and Glaciation is the classic textbook for all students of glaciation. Stimulating and accessible, it has established a reputation as a comprehensive and essential resource. In this new edition, the text, references, and illustrations have been thoroughly updated to give today's reader an up-to-the minute overview of the nature, origin, and behavior of glaciers and the geological and geomorphological evidence for their past history on earth. The first part of the book investigates the processes involved in forming glacier ice, the nature of glacier/climate relationships, the mechanisms of glacier flow, and the interactions of glaciers with other natural systems such as rivers, lakes, and oceans. In the second part, the emphasis moves to landforms and sediment, the interpretation of the earth's glacial legacy, and the reconstruction of glacial depositional environments and palaeoglaciology.
Article
Reports on ablation research carried out on the Rakhiot Glacier, Punjab, Himalaya. Specifically, detailed measurements of ablation rates on debris covered and debris free surfaces allow specification of relationships between ablation and debris cover thickness. Direct ablation measurements indicate a sharp increase in ablation with debris cover thickness increasing from 0.0 to 10 mm followed by a decrease in ablation with debris cover thickness increasing beyond 10 mm. Field observations reveal a critical thickness of 30 mm indicating that at any greater debris thickness ablation is suppressed from that expected on debris-free ice. A comparison with previous research indicates similar hyperbolic trends in the relationship between debris cover thickness and ablation, however, the intensity of these trends differ with global location. -Authors
Article
The effect of a thin layer of fine debris on the melting of snow and ice was investigated over Dokriani Glacier at an altitude of about 4000 m in the Garhwal Himalayas (31°49-31 52 N, 78 47-78 51 E). Such investigations were made during summer 1995, 1997 and 1998. The average melt rate with respect to unit temperature or degree-day factor for clean and dusted snow was computed to be 5.8 and 6.4 mm °C-1 day-1, whereas for clean ice and debris-covered ice the value of this factor was 7.3 and 8.0 mm °C-1 day-1, respectively. Melt rate of clean ice was about 1.26 times greater than that for clean snow under similar weather conditions. The effect of debris layer on the melt rate of snow was more prominent than that for ice. The presence of debris on the ice increased the melt rate by about 8.5%, whereas for snow it increased by about 11.6%. Information on such factors is useful for runoff modelling of glacierized drainage basins.
Article
In order to examine the role of supraglacial debris during ablation, a field experiment was carried out on a snowpatch beside the Rikha Samba Glacier, Hidden Valley. Black phyllite sand was scattered on nine test fields varying in thickness from about 0.5 cm to 8 cm. The thickness of the debris cover and the amount of surface ablation by melting of snow were measured carefully at 5 points in each test field. The results are as follows : (1) The ablation rate was accelerated under a thin debris layer and was retarded under a thick one as compared with that of a natural snow surface. The critical thickness of the debris layer was 1.6 cm. (2) The acceleration of ablation was greatest under a debris cover 0.5 cm thick.
Article
A simple model suggests that the ablation under a debris layer could be estimated from meteorological variables if the surface temperature data of the layer are available. This method was tested by analyzing the data obtained from experiments with artificial debris layers.-from Authors
Article
Quantitative investigations have been made of ice-cored dirt cones on Bersaerkerbræ in north-east Greenland. Experiments were also undertaken to evaluate field observations. Measurements included: maximum cone dimensions, sediment thickness and particle size, cone growth rates, slope angles and the temperature distribution within the debris layer and ice core. Particle size, which has not been stressed in previous studies, and related liquid consistency limits, appear as the dominant controls in cone formation, independent of debris thickness within the observed range of 10 mm to 125 mm. A threshold grain-size for dirt-cone inception was found, between 0.2 mm and 0.6 mm. The growth of cones was usually not more than 50% of the ablation over “clean” ice. Temperature measurements within dirt cones has enabled heat-flow studies to be made, evaluating the thermal conductivity of a sediment layer and the heat transfer involved in melting the ice core. A simple model of dirt-cone dynamics is proposed, characterized by negative feedbacks and describing a steady-state system.