Next Article in Journal
The Poverty of Farmers in a Main Grain-Producing Area in Northeast China
Previous Article in Journal
Mechanisms of Change in Urban Green Infrastructure—Evidence from Romania and Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of MODIS Surface Temperature Products of Greenland Ice Sheet Using In-Situ Measurements

1
College of Geography and Environment, Shandong Normal University, Jinan 250061, China
2
State Key Laboratory of Severe Weather, Chinese Academy of Meteorological Sciences, Beijing 100081, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Land 2022, 11(5), 593; https://doi.org/10.3390/land11050593
Submission received: 16 March 2022 / Revised: 13 April 2022 / Accepted: 16 April 2022 / Published: 19 April 2022

Abstract

:
Satellite-based data have promoted the research progress in polar regions under global climate change, meanwhile the uncertainties and limitations of satellite-derived surface temperatures are widely discussed over Greenland. This study validated the accuracy of ice surface temperature (IST) from the moderate-resolution imaging spectroradiometer (MODIS) over the Greenland ice sheet (GrIS). Daily MODIS IST was validated against the observational surface temperature from 24 automatic weather stations (AWSs) using the mean bias (MB), the root mean square (RMSE), and the correlation coefficient (R). The temporal and spatial variability over the GrIS spanning from March 2000 to December 2019 and the IST melt threshold (−1 °C) were analyzed. Generally, the MODIS IST was underestimated by an average of −2.68 °C compared to AWSs, with cold bias mainly occurring in winter. Spatially, the R and RMSE performed the better accuracy of MODIS IST on the northwest, northeast, and central part of the GrIS. Furthermore, the mean IST is mainly concentrated between −20 °C and −10 °C in summer while between −50 °C and −30 °C in winter. The largest positive IST anomalies (exceeds 3 °C) occurred in southwestern GrIS during 2010. IST shows the positive trends mainly in spring and summer and negative in autumn and winter.

1. Introduction

The Greenland ice sheet (GrIS) covers an area of 1.7 × 106 km2 and is the second largest ice sheet in the world after the Antarctic ice sheet [1]. The GrIS had been relatively stable during the 1970s, 1980s, and the early 1990s with the accumulation and loss staying relatively balanced [2,3]. In recent decades, there has been much attention paid to the GrIS due to its enhanced melting [4,5,6], particularly relevant to abnormal climatic conditions [7,8,9] and accelerated mass loss [10,11,12,13]. The enhanced ablation was measured using infrared and passive-microwave records [14,15,16,17,18]. The analysis of in-situ observations, satellite data, and regional climate models (RCMs) all revealed the accelerating GrIS mass loss [10,19]. The GrIS mass balance has decreased more than twice since the 21st century comparing with the last century [20,21]. The mass loss of GrIS over the past decade is about six times that of the 1980s and the GrIS has contributed 13.7 ± 1.1 mm to the global sea level rise since 1972 [2]. If the acceleration continues, several studies using numerical model strategies to predict the future evolution of GrIS suggested that GrIS could contribute to a rise in global sea level of 2.3 cm by 2050 [22], and 3.5 cm to 76.4 cm by 2300 [23,24]. Surface temperature is an important factor in controlling mass balance [25]. Recently, the increase in GrIS mass loss is mainly because of the increased surface meltwater runoff [26], and prolonged surface temperature elevate exacerbates melting [27]. Surface temperature affects the basal melt [28] as its rise is an important step in the melt process [29]. Thus, tracking IST can better account for surface melt, mass balance as well as ice sheet surface processes.
Currently, observational stations, regional climate models, and remote sensing are the three main sources to obtain IST for studying [30,31,32,33]. The AWSs on the ice sheet provide observations of the GrIS, for example, the Greenland Climate Network (GC-Net) [1,34] and the Greenland Ice Sheet Monitoring Program (PROMICE), which monitors surface and air temperature [35,36,37]. Meanwhile, existing coastal and low-density weather stations on the ice sheet make it hard to get accurate surface temperature observations of the entire GrIS. The difficulties of manipulating station devices in severe polar regions also make it hard to obtain continuous and complete observations record. RCMs can compensate for data recorded at certain periods and locations which could not be obtained from weather stations, especially in the polar regions [38]. RCMs are reliable in the entire GrIS context, however, in the regional and local context, climate variability can be missed occasionally [39,40]. Hence, the way to measure the surface temperature through satellite remote sensing is more outstanding than AWSs and RCMs with high temporal and spatial resolution over the GrIS. Nonetheless, the most practical way to get a wide range of surface temperature data is through the satellite remote sensing [41]. Accurate determination of satellite-derived IST will improve modeling of ice sheet melt and other ice sheet processes and changes.
The GrIS is sensitive to climate change and plays a vital role in the global climate [1]. Hall showed an increase IST of ~0.55 ± 0.44 °C/decade and the northeastern GrIS experienced the largest increase of ~0.95 ± 0.44 °C/decade [8]. Under these conditions, detailed and accurate monitoring of the surface temperature on the entire GrIS is of significance. Recently, a multilayer, daily IST-albedo-water vapor product of Greenland, spanning from March 2000 to December 2019, has been developed. This paper mainly validated the accuracy of the MODIS IST products based on 24 AWSs from PROMICE. A description of the data and methods used in this study were given in 2. In 3, the IST performance was first demonstrated, and then based on its favorable performance, the temporal and spatial variability of GrIS IST, IST anomalies, and trends were further reproduced. Finally, in 4, the MODIS IST melt threshold −1 °C using daily ablation data from 8 AWSs was also analyzed.

2. Materials and Methods

2.1. Observational Data

The 24 AWSs (Figure 1) data used in this study are from the PROMICE, which is operated by the Geological Survey of Denmark and Greenland (GEUS) in collaboration with the National Space Institute at the Technical University of Denmark [35]. The time series of the AWS has different observation periods and completeness (Table 1). In this study, the daily IST data was derived from measured downward and upward longwave irradiance and surface emissivity is set to 0.97 [42]. Mostly located in the coastal area of the GrIS, the spatial layout of the AWSs is uneven. Roughly 25% of the stations are located below 500 m and only 16.7% are above 1000 m (Table 1). Moreover, only EGP, KAN_U, and CEN are located at the accumulation area according to the ELA (equilibrium line altitude) of the corresponding basin in 2017 [43].

2.2. MODIS IST Products

MODIS is a polar-orbiting, 36-channel, and across-track scanning spectro-radiometer whose images cover the whole region of the planet every one to two days [44]. MODIS has been flying aboard NASA Earth Observation System (EOS) Terra satellites since 2000 and Aqua since 2002 [45], so the GrIS swath-based daily gridded products began to be available at the beginning of 2000. Terra’s orbit around the Earth is timed so that it passes across Greenland in the afternoon (14:30–16:30 UTC) [41]. MODIS Collection 6.1 IST used in this study is from multilayer, daily IST-albedo-water vapor products developed using standard MODIS datasets from the Terra satellite [28].
Its algorithm was derived from the algorithm developed for the MODIS sea ice product MOD29 or MYD29 (MOD and MYD stand for Terra and Aqua products, respectively) [45,46,47]. The “MOD29” IST products during March 2000~December 2019, provide daily and monthly mean ISTs with polar-stereographic grids at 0.78 × 0.78 km resolution [28]. As the input product of the MOD29 algorithm, MOD35 determines the cloud obscuration of the MODIS IST products. The cloudy grid cells without IST value are excluded calculation process.

2.3. MODIS Products Preprocessing

The cloud coverage affects the accuracy of MODIS IST products. The internal cloud cover incorrectly identified instances of the ice surface as cloudless on a lot of days in JJA (June, July, and August) [8]. Therefore, the ISTs are actually the cloud top temperatures, far below the clear surface temperature (up to 30 °C) [8]. Therefore, we checked the MODSI IST of each grid over two standard deviations to reduce outliers caused by cloud masks or other reasons. The percentage of the removed pixels was shown in Table 2.
As the ice/snow surface cannot exist in the form of ice/snow above 0 °C [8], the fake pixels values (≥0 °C) found on the ice sheet are removed. The most likely reason for a few pixels with the wrong IST is incorrect cloud masking caused by MODIS cloud mask [48]. Additionally, some pixel values representing no data, cloud cover, and fill value, are also removed.

2.4. Statistical Indexes for MODIS IST Assessment

The daily MODIS ISTs used to make a comparison with the observations were extracted respectively from the nearest gridded MODIS IST at the ice sheet. Meanwhile, three classical statistical indexes were used to estimate the performance of the MODIS product, including the bias, root mean square (RMSE), and the correlation coefficient (R). The calculation formulas are as follows:
Bias = i = 1 N ( M i O i ) N
RMSE = i = 1 N ( O i M i ) 2 N
r = i = 1 N ( O i O ¯ ) ( M i M ¯ ) i = 1 N ( O i O ¯ ) 2 i = 1 N ( M i M ¯ ) 2
where N is the available number of the sample, O is the observed IST, and M is the corresponding MODIS IST.

2.5. Anomaly and Trend Analysis

As the MODIS IST dataset of the GrIS used in this study is available since March 2000, the annual anomaly of 2000 was not calculated. For each grid, the annual IST anomalies were calculated using the monthly data, relative to the multi-year average annual IST from 2001 to 2019. The summer IST includes the year 2000. When seasonal and annual IST trends of each grid were calculated, only the grids with 20-year data are used to calculate.

3. Results

3.1. MODIS IST Performance

To initially obtain the performance of MODIS IST, 12 AWSs are selected according to their location and relative data volume. Figure 2 shows that the MODIS IST agrees well with the observations, with the R > 0.84 (TAS_ L and QAS_L). A significant characteristic is MODIS IST has a cold deviation of 2–3 °C. MODIS IST shows the best agreement with KPC_U AWS observation, with a low RMSE (RMSE = 3.71 °C) and a high correlation (R = 0.97), whereas, generally poorer agreement with TAS_L and QAS_L.
Figure 3 also suggests that the bias indicates the IST tends to underestimate surface temperature (2–4 °C). The R and RMSE indicate that MODIS IST is highly consistent with ground measurements. Spatially, the performance of MODIS IST in the north is better than that in the south, with a higher average R (0.96) and lower average RMSE (4.19 °C). Compared with other regions, the R and RMSE show a better accuracy of MODIS IST on the northwest, northeast, and central GrIS. In these areas, the R is close to 1 and the range of RMSE is concentrated at 3.7–4.9 °C, indicating a high correlation. The MODIS IST shows the best agreement with observations in the northeast, with the highest average R (0.96) and the lowest average RMSE (3.93 °C), however, generally poorer agreement with AWSs observations in the southeast.
Figure 4 compares MODIS IST time series with AWSs to analyze the typical temporal variation and (in) consistency of time series derived from MODIS IST. KPC_U, KAN_U, and EGP were selected for their high-continuity data. The MODIS IST performance is good, and the time series are consistent well with observations, only with a slight underestimation in summer. Notably, the cold bias of MODIS IST mainly occurs in winter. The average cold deviations in summer and winter are 2.04 °C and 4.35 °C, respectively.
Overall, the performance of MODIS products is good, and the range of the cold bias is concentrated at 2–4 °C. A cold deviation of 0.98 °C was found when MODIS IST was compared with the observation data measured by Summit Station of Greenland [28]. The results of thermochron data on the Summit also showed a cold deviation of 3.14 °C of MODIS IST [45]. Given the limited coverage of AWSs on GrIS, we suggest that ISTs obtained from the MODIS IST products are characterized by high accuracy and could be reliably used as an alternative or supplement to Greenland temperature monitoring, especially for IST in summer. Generally, fine space-based climate information obtained by analyzing MODIS IST can effectively evaluate the temporal and spatial variability of the GrIS and be applied in many fields.

3.2. Temporal and Spatial Variability of GrIS IST

Figure 5 are clear-sky IST maps illustrating the MODIS IST on the GrIS at seasonal and annual scales. The MODIS IST revealed distinct spatial gradients varying from outer (low elevation) to inner (high elevation), showing a colder inside with higher elevation. Apparently, the IST in the northern GrIS is much lower than that in the southern region while there is little difference between the eastern and western GrIS. The mean autumn IST is a little colder than the spring and annual mainly reflected in the inner of the GrIS. The mean summer IST is mainly concentrated between −20 °C and −10 °C with the highest IST (−1.33 °C) in the southwest margin of the GrIS and the lowest IST (−17.98 °C) in the inner. Conversely, the mean winter IST is mainly concentrated between −50 °C and −30 °C with the highest and lowest IST values of −9.85 °C and −50.13 °C, respectively. The minimum IST values in winter are found at high elevations just as in summer but tend to migrate farther north, indicating that the coldest ISTs are not always located at the highest elevations, consistent with the previous study [49].

3.3. GrIS IST Anomalies

Figure 6 shows the temporal and spatial annual IST anomaly from 2001 to 2019 respectively. The IST anomalies between −1 °C and 1 °C are dominated in most of the study period except for some abnormally warm and cold years. The largest positive IST anomalies (exceeds 3 °C) occurred in the southwest in 2010 when almost the entire GrIS experienced positive IST anomalies consistent with [8]. In that year, Greenland experienced melting up to another 60 days compared with the 1960–2010 average (1980–2010 were simulated data), with the largest differences occurring at the southwestern and western margins of the GrIS [6]. The previous study shows that the unusual melt event in 2010 was trigged by large positive air temperature anomalies during May, accelerating snowpack metamorphism, and premature bare ice exposure [50]. The GrIS IST anomaly in May 2010 may be relevant to the eruption of Eyjafjallajőkull volcano on March 2010 which once had a profound influence on the environment and climatic conditions of neighboring regions [51]. With the positive albedo feedback and the fact that the wet snow could absorb up to three times more incident solar energy than dry snow, the further melt was fostered [52].
In addition to 2010, relatively large positive IST anomalies (1–3 °C) were mainly found in 2003, 2012, 2016, and 2019. The high positive IST anomalies were mainly found in central and southern GrIS in 2003, southeastern GrIS in 2012, and almost the entire GrIS in 2016 while northern and western GrIS in 2019. By comparison, a relatively large region of negative IST anomalies that between –1 °C and –3 °C were found in 2001, 2011, 2013 and 2015. Especially, the largest negative IST anomalies (below −3 °C) was found in 2015 in southwestern GrIS.
Figure 7 shows the temporal and spatial characteristics of the mean summer IST anomaly from 2000 to 2019, respectively. As reported by Hall et al. [8], the summer of 2012 is the warmest summer, with the southeast experiencing the largest positive IST anomalies (exceeding 3 °C), related to the strongest North American heat wave since 1895 [53]. In 2012, almost the entire GrIS (98.6%) experienced an extreme melting event., even at Summit [7]. This extreme melt event was associated with an abnormal ridge of warm air that was stagnant over Greenland and the radiative effects of low-level liquid layer clouds [7,54]. Previous studies had shown that on 11–13 July, the GrIS experienced the most extensive melt extent based on satellite-derived data [7,9]. Figure 8 are the MODIS IST on 11–13 July of the GrIS showing that the IST is concentrated between 0 °C and −3 °C, especially the western GrIS experienced the largest area of highest IST.
In addition, a relatively large area of positive IST anomalies (0–2 °C) in summer was mainly found in 2007, 2010, 2016 and 2019. The recent warm summer found in our study period is the summer of 2019, agreeing with the previous study [55,56]. The increasing IST is the basic factor causing the increasing GrIS melting [57]. The warm, moist air intrusions could potentially drive melt events [54,58]. At the end of July 2019, the GrIS melting extent achieved more than 62%, even at Summit [55] and the ablation zones continued through August and September, entering accumulation conditions in early October [59]. Contrary to the above-average warmth conditions during the summers mentioned above, relatively large areas of negative IST anomalies were found in 2000, 2001, 2002, 2006, 2009, 2013, 2017 and 2018.

3.4. IST Trend Analysis of GrIS

Figure 9 shows the positive IST trends mainly in spring and summer and negative trends in fall and winter. The annual IST trends are much more moderate than the seasonal IST trends and the most IST trends are concentrated ~+−0.5 °C/decade, showing a lower spatial variability. On an annual scale, the northern GrIS mainly experienced positive IST trends (0~0.5 °C/decade) while the southern GrIS mainly experienced negative IST trends (−1~0 °C/decade). Autumn cooling was common and Hall et al. [8] also found a decrease in IST (~−1.49 ± 1.20 °C/decade) in autumn during the period 2000–2012. Similarly, this work also showed the northwestern GrIS experienced positive IST trends except autumn during the study period, especially in spring with the positive IST trends reaching up to 1.5 °C/decade. Furthermore, the minimum negative IST trends (below −1.5 °C/decade) were found in the small area of southern and eastern GrIS in winter and a small area of southern GrIS in autumn.

4. Discussion

GrIS melt and runoff have increased rapidly since the early 1990s [26]. The low density and discontinuity of AWS measurement are the inherent limitation in evaluating the melt on the entire ice sheet. Surface melting conditions that cannot be measured due to the AWSs constraints can only be obtained by remote sensing [60]. Given the significance of remotely sensed melt products for monitoring GrIS melt, an assessment of the MODIS IST-derived melt data is essential.
The MODIS IST-derived melt data is determined by the melt threshold (−1 °C) and a non-cloud-obscured IST grid that is ≥−1 °C is defined as “melt”. Therefore, the melt threshold is the key factor to be validated. The IST products have an accuracy of ±1 °C and melt may occur when temperatures are slightly below freezing with a strong solar radiation [28]. Given the completeness and amount of the AWSs daily ablation data, the daily ablation data from 5 AWSs at low elevations and 3 AWSs at high elevations were selected to validate the melt threshold. The daily ablation data was set to 25 mm based on the accuracy of the pressure transducer [37]. Only the data that were both available for MODIS and AWS were used in this study. The days that ware detected melt or no melt both from MODIS and AWS were recorded as “BB” and these days can be considered well-matched days. If the days were detected melt only from MODIS or AWS, then they were recorded as “MM” or “SM” and these days can be considered failed-matched days.
Figure 10 shows the day-to-day melt agreement conditions between MODIS and AWSs. Obviously, AWSs at low elevations reproduced more melt occasions than at high elevations. In the strong ablation year of 2012, except for the KPC_L, MODIS IST reproduced fewer melt conditions than AWSs. In general, MODIS IST performed an average agreement of 55% with the highest agreement in KPC_L (61%), located at the northeastern GrIS. NUK_L at southwestern GrIS had the lowest consistency of 41% with much fewer melt days than that detected by MODIS IST. Other AWSs show an agreement between 50% and 60%. Overall, the −1 °C melt threshold suggested a 41–61% consistency of melt conditions (BB) between MODIS and AWSs, and generally, fewer MODIS melt days occurred than AWSs. Given the fewer MODIS melt days based on −1 °C and the cold bias found in this work, lower melt thresholds were used to possibly achieve higher consistencies (Figure 11). The results showed that consistencies were 46–66% and 51–70% based on melt thresholds respectively −1.5 °C and −2 °C, respectively. With the melt threshold decreasing, the consistency was generally becoming higher especially for the stations QAS_L, NUK_L, KAN_L, and TAS_U while KPC_L and QAS_U had lower consistencies.

5. Conclusions

In this study, based on PROMICE AWSs observations, three statistical indexes, bias, RMSE, and R are used to comprehensively evaluate the performance of the MODIS IST over the GrIS. Generally, MODIS IST has a widespread cold bias over the GrIS, ranging from −2 to −4 °C and this cold bias mainly occurred in winter. The consistency between MODIS IST and in-situ IST is very high on the GrIS, with an average of R = 0.92 and RMSE = 4.36 °C. Spatially, the R and RMSE agree on a better accuracy of MODIS IST on the northwestern, northeastern, and central GrIS.
The MODIS IST can well demonstrate the temporal and spatial variability of GrIS. The mean summer IST is mainly concentrated between −20 °C and −10 °C with the highest and lowest IST values of −1.33 °C and −17.98 °C, respectively. Besides, the largest positive IST anomalies (exceeding 3 °C) occurred in southwestern GrIS in 2010. Summer 2012 was the warmest summer with the largest positive IST anomalies (exceeding 3 °C) in the southeastern GrIS. The positive IST trends are mainly in spring and summer and the negative IST trends are in autumn and winter. Importantly, most of the annual IST trends are concentrated between ~±0.5 °C/decade.
This work also verifies the rationality of the MODIS IST melt threshold of −1 °C using the daily ablation data from AWSs. The results show that the −1 °C melt threshold suggested a 41–61% consistency of melt conditions between MODIS and AWSs. Furthermore, with the melt threshold decreasing (−1.5 °C, −2 °C), the consistency is generally becoming higher.

Author Contributions

X.Y.: Formal analysis, visualization, writing—original draft. T.W.: formal analysis, visualization, writing—original draft. M.D.: formal analysis, writing—original draft. Y.W.: formal analysis, methodology. W.S.: formal analysis, methodology. Q.Z.: methodology, software, validation. B.H.: funding acquisition, supervision, conceptualization, methodology, validation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China grant number 42171121, 41701059. And The APC was funded by 42171121.

Data Availability Statement

All data and the script of the whole processes are available through an email request to the authors.

Acknowledgments

This work was funded by the Natural Science Foundation of China (42171121, 41701059). The authors gratefully acknowledge data availability from the Programme for Monitoring of the Greenland Ice Sheet (PROMICE) and the Greenland Analogue Project (GAP) provided by the Geological Survey of Denmark and Greenland (GEUS) at http://www.promice.dk (accessed on 6 October 2021).

Conflicts of Interest

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

References

  1. Steffen, K.; Box, J. Surface climatology of the Greenland ice sheet: Greenland climate network 1995–1999. J. Geophys. Res.-Atmos. 2001, 106, 33951–33964. [Google Scholar] [CrossRef]
  2. Mouginot, J.; Rignot, E.; Bjork, A.A.; van den Broeke, M.; Millan, R.; Morlighem, M.; Noel, B.; Scheuchl, B.; Wood, M. Forty-six years of Greenland Ice Sheet mass balance from 1972 to 2018. Proc. Natl. Acad. Sci. USA 2019, 116, 9239–9244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Blunden, J.; Arndt, D. State of the Climate in 2019. Bull. Am. Meteorol. Soc. 2020, 101, S1–S429. [Google Scholar] [CrossRef]
  4. Hanna, E.; Huybrechts, P.; Steffen, K.; Cappelen, J.; Huff, R.; Shuman, C.; Irvine-Fynn, T.; Wise, S.; Griffiths, M. Increased runoff from melt from the Greenland Ice Sheet: A response to global warming. J. Clim. 2008, 21, 331–341. [Google Scholar] [CrossRef] [Green Version]
  5. Tedesco, M.; Serreze, M.; Fettweis, X. Diagnosing the extreme surface melt event over southwestern Greenland in 2007. Cryosphere 2008, 2, 159–166. [Google Scholar] [CrossRef] [Green Version]
  6. Mernild, S.H.; Mote, T.L.; Liston, G.E. Greenland ice sheet surface melt extent and trends: 1960–2010. J. Glaciol. 2011, 57, 621–628. [Google Scholar] [CrossRef] [Green Version]
  7. Nghiem, S.V.; Hall, D.K.; Mote, T.L.; Tedesco, M.; Albert, M.R.; Keegan, K.; Shuman, C.A.; DiGirolamo, N.E.; Neumann, G. The extreme melt across the Greenland ice sheet in 2012. Geophys. Res. Lett. 2012, 39, L20502. [Google Scholar] [CrossRef]
  8. Hall, D.K.; Comiso, J.C.; DiGirolamo, N.E.; Shuman, C.A.; Box, J.E.; Koenig, L.S. Variability in the surface temperature and melt extent of the Greenland ice sheet from MODIS. Geophys. Res. Lett. 2013, 40, 2114–2120. [Google Scholar] [CrossRef]
  9. Hanna, E.; Fettweis, X.; Mernild, S.H.; Cappelen, J.; Ribergaard, M.H.; Shuman, C.A.; Steffen, K.; Wood, L.; Mote, T.L. Atmospheric and oceanic climate forcing of the exceptional Greenland ice sheet surface melt in summer 2012. Int. J. Climatol. 2014, 34, 1022–1037. [Google Scholar] [CrossRef] [Green Version]
  10. Rignot, E.; Velicogna, I.; van den Broeke, M.R.; Monaghan, A.; Lenaerts, J. Acceleration of the contribution of the Greenland and Antarctic ice sheets to sea level rise. Geophys. Res. Lett. 2011, 38, L05503. [Google Scholar] [CrossRef] [Green Version]
  11. Enderlin, E.M.; Howat, I.M. Submarine melt rate estimates for floating termini of Greenland outlet glaciers (2000–2010). J. Glaciol. 2013, 59, 67–75. [Google Scholar] [CrossRef] [Green Version]
  12. Fettweis, X.; Hanna, E.; Lang, C.; Belleflamme, A.; Erpicum, M.; Gallee, H. Important role of the mid-tropospheric atmospheric circulation in the recent surface melt increase over the Greenland ice sheet. Cryosphere 2013, 7, 241–248. [Google Scholar] [CrossRef] [Green Version]
  13. Wouters, B.; Bamber, J.L.; van den Broeke, M.R.; Lenaerts, J.T.M.; Sasgen, I. Limits in detecting acceleration of ice sheet mass loss due to climate variability. Nat. Geosci. 2013, 6, 613–616. [Google Scholar] [CrossRef]
  14. Comiso, J.C.; Yang, J.; Honjo, S.; Krishfield, R.A. Detection of change in the Arctic using satellite and in situ data. J. Geophys. Res. Ocean. 2003, 108, C12. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, X.J.; Key, J.R. Recent trends in arctic surface, cloud, and radiation properties from space. Science 2003, 299, 1725–1728. [Google Scholar] [CrossRef]
  16. Wang, X.J.; Key, J.R. Arctic surface, cloud, and radiation properties based on the AVHRR Polar Pathfinder dataset. Part I: Spatial and temporal characteristics. J. Clim. 2005, 18, 2558–2574. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, X.J.; Key, J.R. Arctic surface, cloud, and radiation properties based on the AVHRR Polar Pathfinder dataset. Part II: Recent trends. J. Clim. 2005, 18, 2575–2593. [Google Scholar] [CrossRef]
  18. Mote, T.L. Greenland surface melt trends 1973–2007: Evidence of a large increase in 2007. Geophys. Res. Lett. 2007, 34, L22507. [Google Scholar] [CrossRef]
  19. van den Broeke, M.; Bamber, J.; Ettema, J.; Rignot, E.; Schrama, E.; van de Berg, W.J.; van Meijgaard, E.; Velicogna, I.; Wouters, B. Partitioning Recent Greenland Mass Loss. Science 2009, 326, 984–986. [Google Scholar] [CrossRef] [Green Version]
  20. Reddy, S. Comments on IPCC’s 24th September 2019 Report on “The Ocean and Cryosphere in a Changing Climate: Summary for Policy Makers”. Acta Sci. Agric 2019, 3, 16–19. [Google Scholar] [CrossRef]
  21. Pörtner, H.-O.; Roberts, D.C.; Masson-Delmotte, V.; Zhai, P.; Tignor, M.; Poloczanska, E.; Weyer, N. The ocean and cryosphere in a changing climate. IPCC Spec. Rep. Ocean Cryosphere A Chang. Clim. 2019. Available online: https://www.ipcc.ch/site/assets/uploads/sites/3/2019/12/02_SROCC_FM_FINAL.pdf (accessed on 15 March 2022).
  22. Muntjewerf, L.; Petrini, M.; Vizcaino, M.; da Silva, C.E.; Sellevold, R.; Scherrenberg, M.D.W.; Thayer-Calder, K.; Bradley, S.L.; Lenaerts, J.T.M.; Lipscomb, W.H.; et al. Greenland Ice Sheet Contribution to 21st Century Sea Level Rise as Simulated by the Coupled CESM2.1-CISM2.1. Geophys. Res. Lett. 2020, 47, e2019GL086836. [Google Scholar] [CrossRef]
  23. Calov, R.; Beyer, S.; Greve, R.; Beckmann, J.; Willeit, M.; Kleiner, T.; Rueckamp, M.; Humbert, A.; Ganopolski, A. Simulation of the future sea level contribution of Greenland with a new glacial system model. Cryosphere 2018, 12, 3097–3121. [Google Scholar] [CrossRef] [Green Version]
  24. Rueckamp, M.; Greve, R.; Humbert, A. Comparative simulations of the evolution of the Greenland ice sheet under simplified Paris Agreement scenarios with the models SICOPOLIS and ISSM. Polar Sci. 2019, 21, 14–25. [Google Scholar] [CrossRef]
  25. Bell, R.E. The role of subglacial water in ice-sheet mass balance. Nat. Geosci. 2008, 1, 297–304. [Google Scholar] [CrossRef]
  26. van den Broeke, M.R.; Enderlin, E.M.; Howat, I.M.; Munneke, P.K.; Noel, B.P.Y.; van de Berg, W.J.; van Meijgaard, E.; Wouters, B. On the recent contribution of the Greenland ice sheet to sea level change. Cryosphere 2016, 10, 1933–1946. [Google Scholar] [CrossRef] [Green Version]
  27. Smith, L.C.; Chu, V.W.; Yang, K.; Gleason, C.J.; Pitcher, L.H.; Rennermalm, A.K.; Legleiter, C.J.; Behar, A.E.; Overstreet, B.T.; Moustafa, S.E.; et al. Efficient meltwater drainage through supraglacial streams and rivers on the southwest Greenland ice sheet. Proc. Natl. Acad. Sci. USA 2015, 112, 1001–1006. [Google Scholar] [CrossRef] [Green Version]
  28. Hall, D.K.; Cullather, R.I.; DiGirolamo, N.E.; Comiso, J.C.; Medley, B.C.; Nowicki, S.M. A Multilayer Surface Temperature, Surface Albedo, and Water Vapor Product of Greenland from MODIS. Remote Sens. 2018, 10, 555. [Google Scholar] [CrossRef] [Green Version]
  29. Miller, N.B.; Shupe, M.D.; Cox, C.J.; Noone, D.; Persson, P.O.G.; Steffen, K. Surface energy budget responses to radiative forcing at Summit, Greenland. Cryosphere 2017, 11, 497–516. [Google Scholar] [CrossRef] [Green Version]
  30. Noel, B.; van de Berg, W.J.; van Wessem, J.M.; van Meijgaard, E.; van As, D.; Lenaerts, J.T.M.; Lhermitte, S.; Munneke, P.K.; Smeets, C.J.P.P.; van Ulft, L.H.; et al. Modelling the climate and surface mass balance of polar ice sheets using RACMO2-Part 1: Greenland (1958–2016). Cryosphere 2018, 12, 811–831. [Google Scholar] [CrossRef] [Green Version]
  31. Noel, B.; van de Berg, W.J.; Lhermitte, S.; van den Broeke, M.R. Rapid ablation zone expansion amplifies north Greenland mass loss. Sci. Adv. 2019, 5, eaaw0123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Huai, B.; van den Broeke, M.R.; Reijmer, C.H. Long-term surface energy balance of the western Greenland Ice Sheet and the role of large-scale circulation variability. Cryosphere 2020, 14, 4181–4199. [Google Scholar] [CrossRef]
  33. Nielsen-Englyst, P.; Hoyer, J.L.; Madsen, K.S.; Tonboe, R.T.; Dybkjaer, G.; Skarpalezos, S. Deriving Arctic 2 m air temperatures over snow and ice from satellite surface temperature measurements. Cryosphere 2021, 15, 3035–3057. [Google Scholar] [CrossRef]
  34. Shuman, C.A.; Steffen, K.; Box, J.E.; Stearns, C.R. A dozen years of temperature observations at the summit: Central Greenland automatic weather stations 1987–1999. J. Appl. Meteorol. 2001, 40, 741–752. [Google Scholar] [CrossRef]
  35. Ahlstrom, A.P.; Gravesen, P.; Andersen, S.B.; van As, D.; Citterio, M.; Fausto, R.S.; Nielsen, S.; Jepsen, H.F.; Kristensen, S.S.; Christensen, E.L.; et al. A new programme for monitoring the mass loss of the Greenland ice sheet. Geol. Surv. Den. Greenl. Bull. 2008, 15, 61–64. [Google Scholar] [CrossRef]
  36. van As, D.; Fausto, R.S.; Team, P.P. Programme for Monitoring of the Greenland Ice Sheet (PROMICE): First temperature and ablation records. Geol. Surv. Den. Greenl. Bull. 2011, 23, 73–76. [Google Scholar] [CrossRef] [Green Version]
  37. Fausto, R.S.; Van As, D.; Ahlstrom, A.P.; Citterio, M. Assessing the accuracy of Greenland ice sheet ice ablation measurements by pressure transducer. J. Glaciol. 2012, 58, 1144–1150. [Google Scholar] [CrossRef] [Green Version]
  38. Box, J.E.; Rinke, A. Evaluation of Greenland ice sheet surface climate in the HIRHAM regional climate model using automatic weather station data. J. Clim. 2003, 16, 1302–1319. [Google Scholar] [CrossRef] [Green Version]
  39. Leeson, A.A.; Van Wessem, J.M.; Ligtenberg, S.R.M.; Shepherd, A.; Van Den Broeke, M.R.; Killick, R.; Skvarca, P.; Marinsek, S.; Colwell, S. Regional climate of the Larsen B embayment 1980–2014. J. Glaciol. 2017, 63, 683–690. [Google Scholar] [CrossRef] [Green Version]
  40. Medley, B.; Joughin, I.; Das, S.B.; Steig, E.J.; Conway, H.; Gogineni, S.; Criscitiello, A.S.; McConnell, J.R.; Smith, B.E.; van den Broeke, M.R.; et al. Airborne-radar and ice-core observations of annual snow accumulation over Thwaites Glacier, West Antarctica confirm the spatiotemporal variability of global and regional atmospheric models. Geophys. Res. Lett. 2013, 40, 3649–3654. [Google Scholar] [CrossRef] [Green Version]
  41. Hall, D.K.; Box, J.E.; Casey, K.A.; Hook, S.J.; Shuman, C.A.; Steffen, K. Comparison of satellite-derived and in-situ observations of ice and snow surface temperatures over Greenland. Remote Sens. Environ. 2008, 112, 3739–3749. [Google Scholar] [CrossRef] [Green Version]
  42. Fausto, R.S.; van As, D.; Mankoff, K.D.; Vandecrux, B.; Citterio, M.; Ahlstrom, A.P.; Andersen, S.B.; Colgan, W.; Karlsson, N.B.; Kjeldsen, K.K.; et al. Programme for Monitoring of the Greenland Ice Sheet (PROMICE) automatic weather station data. Earth Syst. Sci. Data 2021, 13, 3819–3845. [Google Scholar] [CrossRef]
  43. Ryan, J.; Smith, L.; Van As, D.; Cooley, S.; Cooper, M.; Pitcher, L.; Hubbard, A. Greenland Ice Sheet surface melt amplified by snowline migration and bare ice exposure. Sci. Adv. 2019, 5, eaav3738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Barnes, W.L.; Pagano, T.S.; Salomonson, V.V. Prelaunch characteristics of the Moderate Resolution Imaging Spectroradiometer (MODIS) on EOS-AM1. IEEE Trans. Geosci. Remote Sens. 1998, 36, 1088–1100. [Google Scholar] [CrossRef] [Green Version]
  45. Hall, D.K.; Comiso, J.C.; DiGirolamo, N.E.; Shuman, C.A.; Key, J.R.; Koenig, L.S. A Satellite-Derived Climate-Quality Data Record of the Clear-Sky Surface Temperature of the Greenland Ice Sheet. J. Clim. 2012, 25, 4785–4798. [Google Scholar] [CrossRef]
  46. Key, J.; Haefliger, M. Arctic ice surface temperature retrieval from AVHRR thermal channels. J. Geophys. Res. Atmos. 1992, 97, 5885–5893. [Google Scholar] [CrossRef]
  47. Key, J.R.; Collins, J.B.; Fowler, C.; Stone, R.S. High-latitude surface temperature estimates from thermal satellite data. Remote Sens. Environ. 1997, 61, 302–309. [Google Scholar] [CrossRef]
  48. Hall, D.K.; Williams, R.S., Jr.; Luthcke, S.B.; Digirolamo, N.E. Greenland ice sheet surface temperature, melt and mass loss: 2000–2006. J. Glaciol. 2008, 54, 81–93. [Google Scholar] [CrossRef] [Green Version]
  49. Stroeve, J.; Steffen, K. Variability of AVHRR-derived clear-sky surface temperature over the Greenland ice sheet. J. Appl. Meteorol. 1998, 37, 23–31. [Google Scholar] [CrossRef]
  50. Tedesco, M.; Fettweis, X.; van den Broeke, M.R.; van de Wal, R.S.W.; Smeets, C.J.P.P.; van de Berg, W.J.; Serreze, M.C.; Box, J.E. The role of albedo and accumulation in the 2010 melting record in Greenland. Environ. Res. Lett. 2011, 6, 014005. [Google Scholar] [CrossRef]
  51. Gałaś, A. Impact of volcanic eruptions on the environment and climatic conditions in the area of Poland (Central Europe). Earth-Sci. Rev. 2016, 162, 58–64. [Google Scholar] [CrossRef]
  52. Abdalati, W.; Steffen, K. Greenland ice sheet melt extent: 1979–1999. J. Geophys. Res. -Atmos. 2001, 106, 33983–33988. [Google Scholar] [CrossRef]
  53. Neff, W.; Compo, G.P.; Ralph, F.M.; Shupe, M.D. Continental heat anomalies and the extreme melting of the Greenland ice surface in 2012 and 1889. J. Geophys. Res. -Atmos. 2014, 119, 6520–6536. [Google Scholar] [CrossRef]
  54. Bennartz, R.; Shupe, M.D.; Turner, D.D.; Walden, V.P.; Steffen, K.; Cox, C.J.; Kulie, M.S.; Miller, N.B.; Pettersen, C. July 2012 Greenland melt extent enhanced by low-level liquid clouds. Nature 2013, 496, 83–86. [Google Scholar] [CrossRef] [PubMed]
  55. Cullather, R.I.; Andrews, L.C.; Croteau, M.J.; Digirolamo, N.E.; Hall, D.K.; Lim, Y.-K.; Loomis, B.D.; Shuman, C.A.; Nowicki, S.M.J. Anomalous Circulation in July 2019 Resulting in Mass Loss on the Greenland Ice Sheet. Geophys. Res. Lett. 2020, 47, e2020GL087263. [Google Scholar] [CrossRef]
  56. Tedesco, M.; Fettweis, X. Unprecedented atmospheric conditions (1948–2019) drive the 2019 exceptional melting season over the Greenland ice sheet. Cryosphere 2020, 14, 1209–1223. [Google Scholar] [CrossRef] [Green Version]
  57. Hearty, T.J., III; Lee, J.N.; Wu, D.L.; Cullather, R.; Blaisdell, J.M.; Susskind, J.; Nowicki, S.M.J. Intercomparison of Surface Temperatures from AIRS, MERRA, and MERRA-2 with NOAA and GC-Net Weather Stations at Summit, Greenland. J. Appl. Meteorol. Climatol. 2018, 57, 1231–1245. [Google Scholar] [CrossRef]
  58. Cullather, R.I.; Nowicki, S.M.J. Greenland Ice Sheet Surface Melt and Its Relation to Daily Atmospheric Conditions. J. Clim. 2018, 31, 1897–1919. [Google Scholar] [CrossRef]
  59. Sasgen, I.; Wouters, B.; Gardner, A.S.; King, M.D.; Tedesco, M.; Landerer, F.W.; Dahle, C.; Save, H.; Fettweis, X. Return to rapid ice loss in Greenland and record loss in 2019 detected by the GRACE-FO satellites. Commun. Earth Environ. 2020, 1, 8. [Google Scholar] [CrossRef]
  60. Kimball, J.S.; Du, J.; Meierbachtol, T.W.; Kim, Y.; Johnson, J.V. Comparing Greenland Ice Sheet Melt Variability from Different Satellite Passive Microwave Remote Sensing Products Over a Common 5-year Record. Front. Earth Sci. 2021, 9, 654220. [Google Scholar] [CrossRef]
Figure 1. Map of Greenland showing the elevation and locations of the PROMICE AWS. Six transects are magnified in subplots.
Figure 1. Map of Greenland showing the elevation and locations of the PROMICE AWS. Six transects are magnified in subplots.
Land 11 00593 g001
Figure 2. MODIS and observed daily IST, the red diagonal represents the 1:1 line. Statistics show Correlation coefficient (R), root mean squared error (RMSE), and bias.
Figure 2. MODIS and observed daily IST, the red diagonal represents the 1:1 line. Statistics show Correlation coefficient (R), root mean squared error (RMSE), and bias.
Land 11 00593 g002
Figure 3. Correlation coefficient (R), bias, and root mean squared error (RMSE) between daily MODIS IST and observations from all AWSs (AWSs shown in bold are above the ELA).
Figure 3. Correlation coefficient (R), bias, and root mean squared error (RMSE) between daily MODIS IST and observations from all AWSs (AWSs shown in bold are above the ELA).
Land 11 00593 g003
Figure 4. Time series of MODIS IST (red line) and surface temperature (green line) from EGP (a), KAN_U (b), and KPC_U (c).
Figure 4. Time series of MODIS IST (red line) and surface temperature (green line) from EGP (a), KAN_U (b), and KPC_U (c).
Land 11 00593 g004
Figure 5. Seasonal MODIS IST over the GrIS in spring (a), summer (b), autumn (c), and winter (d) and the mean annual MODIS IST (e) from March 2000 to December 2019.
Figure 5. Seasonal MODIS IST over the GrIS in spring (a), summer (b), autumn (c), and winter (d) and the mean annual MODIS IST (e) from March 2000 to December 2019.
Land 11 00593 g005
Figure 6. Annual IST anomalies (°C) calculated from monthly-derived MODIS IST.
Figure 6. Annual IST anomalies (°C) calculated from monthly-derived MODIS IST.
Land 11 00593 g006
Figure 7. Summer IST anomalies (°C) calculated from monthly IST (2000 to 2019).
Figure 7. Summer IST anomalies (°C) calculated from monthly IST (2000 to 2019).
Land 11 00593 g007
Figure 8. MODIS IST maps for (a): 11 July, (b): 12 July, and (c): 13 July 2012. The IST grid that is under cloud cover or has no data is considered as no data.
Figure 8. MODIS IST maps for (a): 11 July, (b): 12 July, and (c): 13 July 2012. The IST grid that is under cloud cover or has no data is considered as no data.
Land 11 00593 g008
Figure 9. Seasonal MODIS IST trends over the GrIS in spring (a), summer (b), autumn (c) and winter (d) from March 2000 to December 2019 and the annual MODIS IST (e).
Figure 9. Seasonal MODIS IST trends over the GrIS in spring (a), summer (b), autumn (c) and winter (d) from March 2000 to December 2019 and the annual MODIS IST (e).
Land 11 00593 g009
Figure 10. The consistency between MODIS and AWSs melt condition.
Figure 10. The consistency between MODIS and AWSs melt condition.
Land 11 00593 g010
Figure 11. Consistency of melt conditions (BB) between MODIS and station based on the melt threshold −1 °C, −1.5 °C, and −2 °C.
Figure 11. Consistency of melt conditions (BB) between MODIS and station based on the melt threshold −1 °C, −1.5 °C, and −2 °C.
Land 11 00593 g011
Table 1. Geographical characteristics of the meteorological stations used in this study.
Table 1. Geographical characteristics of the meteorological stations used in this study.
Station NameLatitude (°N)Longitude (°W)Elevation (m)Elevation-ELA (m)Start Date
KPC_L *79.9124.08370−70717 July 2008
KPC_U79.8325.17870−20717 July 2008
EGP75.6235.97266015831 May 2016
SCO_L72.2226.82460−119221 July 2008
SCO_U72.3927.23970−68221 July 2008
MIT65.6937.83440−9703 May 2009
TAS_L65.6438.90250−116028 August 2007
TAS_U65.7038.87570−84015 August 2007
TAS_A65.7838.90 890−52023 August 2013
QAS_L61.0346.85280−119024 August 2007
QAS_M61.1046.83630−84011 August 2016
QAS_U61.1846.82900−5707 August 2008
QAS_A61.2446.731000−47020 August 2012
NUK_L64.4849.54530−88520 August 2007
NUK_U64.5149.271120−29520 August 2007
NUK_N64.9549.89920−49525 July 2010
KAN_L67.1049.95670−7451 September 2008
KAN_M67.0748.841270−1452 September 2008
KAN_U67.0047.0318404254 April 2009
UPE_L72.8954.30220−84317 August 2009
UPE_U72.8953.58940−12317 August 2009
THU_L76.4068.27570−4939 August 2010
THU_U76.4268.15760−3039 August 2010
CEN77.1761.11188081723 May 2017
* L: Lower station, M: Middle station, U: Upper station.
Table 2. The percentage of the MODIS IST pixels that had been removed.
Table 2. The percentage of the MODIS IST pixels that had been removed.
YearPercentage (%)YearPercentage (%)
20000.5620100.46
20010.5520110.49
20020.5220120.48
20030.5220130.50
20040.5220140.50
20050.4920150.50
20060.4820160.47
20070.5020170.48
20080.5020180.48
20090.4820190.45
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, X.; Wang, T.; Ding, M.; Wang, Y.; Sun, W.; Zhang, Q.; Huai, B. Assessment of MODIS Surface Temperature Products of Greenland Ice Sheet Using In-Situ Measurements. Land 2022, 11, 593. https://doi.org/10.3390/land11050593

AMA Style

Yu X, Wang T, Ding M, Wang Y, Sun W, Zhang Q, Huai B. Assessment of MODIS Surface Temperature Products of Greenland Ice Sheet Using In-Situ Measurements. Land. 2022; 11(5):593. https://doi.org/10.3390/land11050593

Chicago/Turabian Style

Yu, Xiaoge, Tingting Wang, Minghu Ding, Yetang Wang, Weijun Sun, Qinglin Zhang, and Baojuan Huai. 2022. "Assessment of MODIS Surface Temperature Products of Greenland Ice Sheet Using In-Situ Measurements" Land 11, no. 5: 593. https://doi.org/10.3390/land11050593

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

Article Metrics

Back to TopTop