Next Article in Journal
The Characterization of the Vertical Distribution of Surface Soil Moisture Using ISMN Multilayer In Situ Data and Their Comparison with SMOS and SMAP Soil Moisture Products
Previous Article in Journal
Joint Direction of Arrival-Polarization Parameter Tracking Algorithm Based on Multi-Target Multi-Bernoulli Filter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geophysics in Antarctic Research: A Bibliometric Analysis

1
School of Geophysics and Information Technology, China University of Geosciences, Beijing 100083, China
2
National Engineering Research Center of Offshore Oil and Gas Exploration, Beijing 100028, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(16), 3928; https://doi.org/10.3390/rs15163928
Submission received: 19 June 2023 / Revised: 28 July 2023 / Accepted: 5 August 2023 / Published: 8 August 2023

Abstract

:
Antarctica is of great importance in terms of global warming, the sustainability of resources, and the conservation of biodiversity. However, due to 99.66% of the continent being covered in ice and snow, geological research and geoscientific study in Antarctica face huge challenges. Geophysical surveys play a crucial role in enhancing comprehension of the fundamental structure of Antarctica. This study used bibliometric analysis to analyze citation data retrieved from the Web of Science for the period from 1982 to 2022 with geophysical research on Antarctica as the topic. According to the analysis results, the amount of Antarctic geophysical research has been steadily growing over the past four decades as related research countries/regions have become increasingly invested in issues pertaining to global warming and sustainability, and international cooperation is in sight. Moreover, based on keyword clustering and an analysis of highly cited papers, six popular research topics have been identified: Antarctic ice sheet instability and sea level change, Southern Ocean and Sea Ice, tectonic activity of the West Antarctic rift system, the paleocontinental rift and reorganization, magmatism and volcanism, and subglacial lakes and subglacial hydrology. This paper provides a detailed overview of these popular research topics and discusses the applications and advantages of the geophysical methods used in each field. Finally, based on keywords regarding abrupt changes, we identify and examine the thematic evolution of the nexus over three consecutive sub-periods (i.e., 1990–1995, 1996–2005, and 2006–2022). The relevance of using geophysics to support numerous and diverse scientific activities in Antarctica becomes very clear after analyzing this set of scientific publications, as is the importance of using multiple geophysical methods (satellite, airborne, surface, and borehole technology) to revolutionize the acquisition of new data in greater detail from inaccessible or hard-to-reach areas. Many of the advances that they have enabled be seen in the Antarctic terrestrial areas (detailed mapping of the geological structures of West and East Antarctica), ice, and snow (tracking glaciers and sea ice, along with the depth and features of ice sheets). These valuable results help identify potential future research opportunities in the field of Antarctic geophysical research and aid academic professionals in keeping up with recent advances.

1. Introduction

Antarctica is one of the cleanest places on Earth. Due to its unique geographical location and ecological features, it represents a matchless “natural laboratory” for vital scientific research to study the past, present, and future [1]. It has long been ranked as a priority research area for Earth-related and environmental science studies in many countries. However, due to the remoteness of the Antarctic continent, its difficulty of access, complex topography, and the fact that 99.66% of the continent is covered by ice and snow, “fieldwork” in Antarctica is challenging and expensive [2]. This means that geophysical methods, especially remote sensing techniques such as satellite imaging and airborne surveys, are the best and most effective options. Geophysics uses the principles and methods of physics to study the structure of the medium, composition, formation, and evolution of the Earth’s interior; in many ways, it offers unmatched advantages in terms of geological observations [3]. Conducting geophysical surveys help to improve understanding of the Antarctic continent and its ice sheet, revealing the geomorphology [4], tectonic evolution [5,6], ice-sheet stability [7,8], mineral resources [9,10,11], and climate change [12,13,14] in the Antarctic continent.
Antarctic geophysical surveys represent a key research area in Antarctic scientific study. The first long-term geomagnetic station was completed in the early 20th century, marking the beginning of Antarctic geophysical exploration. However, until the 1950s, little was known about the size and shape of the Antarctic continent, the volume of ice stored there, its previous changes, and its impact on the rest of the planet. With the advent of geophysical techniques such as seismic bathymetry, a young glaciologist named Gordon Robin began making the first measurements of ice thickness in 1952 and obtained reliable measurements of ice thicknesses and the bedrock beneath [15,16,17,18]. In the 1960s, Robin and his colleagues improved radio-echo-sounding (RES) techniques and utilized aircraft-mounted radar systems to study the internal structure and thickness of the ice sheet [19,20]. For example, during the International Geophysical Year of 1957–1958, scientists from multiple countries conducted extensive radar surveys, collecting crucial data that provided a better understanding of Antarctica’s ice sheet [21]. What followed was one of the key scientific expeditions in the history of Antarctic exploration—a joint United Kingdom–United States–Danish program comprising long-range airborne surveys of Antarctica over several seasons during the 1970s [22]. In the late 1980s and early 1990s, with the advent of satellite-based remote sensing, particularly from the European remote sensing satellite (ERS-1), technology once again propelled Antarctic research forward [23]. In the 21st century, the use of autonomous systems, including unmanned aerial vehicles (UAVs) [24,25] and underwater autonomous vehicles (AUVs) [26,27], has also facilitated data collection and the mapping of inaccessible or hazardous regions of Antarctica. To summarize, with advances in polar exploration techniques and novel detection methods, ranging from satellites to airborne and surface unmanned vehicles, to boreholes, there are currently more than 80 countries and regions conducting the international geophysical exploration of Antarctica. The continuing advancement of these technologies, combined with interdisciplinary approaches and international collaboration, has significantly expanded our understanding of Antarctica’s geophysical characteristics, glacial processes, and the implications of climate change on the continent.
Bibliometrics is a quantitative analysis method that is based on mathematical statistics. It takes the external characteristics of the scientific literature as the object of study and examines the distribution structure of and quantitative relationships among the literature, which was used to describe, assess, and predict the current status and trends of emerging topical areas in research [28,29]. Although many scholars have contributed to the field of Antarctic geophysical research, only a few related studies have been published that have undertaken a comprehensive review and investigation of this research topic. However, despite the quantity of Antarctica review articles, these studies are typically restricted to either one topic or one area, such as unmanned aerial vehicles (UVAs), whereas systematic geophysical reviews are relatively rare [14,30,31,32,33,34]. Therefore, to effectively explore this emerging trend, it is necessary to analyze the development, hot spots, and trends of this topic more systematically and comprehensively.
For this study, based on publications related to Antarctic geophysical research recorded in the Web of Science (WoS) database, the annual publication quantity of articles, national and institutional publication quantity and cooperation mapping, popular research areas, and the temporal evolution of studies published from 1982 to 2022 were systematically reviewed, using a bibliometric analysis method.

2. Data and Methods

In this study, we employed a unified methodological approach that was capable of unifying classical bibliometric analyses using qualitative descriptors. Our approach and procedures are shown in Figure 1. In accordance with the methodology used, our study was separated into two primary phases: Phase 1 and Phase 2. The first phase consisted of selecting a search database, identifying search terms relevant to the subject matter, and incorporating filters into the search engine. Next, we performed a manual screening by reading all the titles and abstracts of the articles returned in the search, identifying matches, and excluding possible articles that did not fit our objective (Phase 1). After these two steps were complete, we performed all the relevant analyses (Phase 2) on both data sets, including co-country, co-occurrence, cluster, and thematic evolution analysis. The results and data were also visualized and manipulated, which helped discover hidden patterns and trends in the data.

2.1. Data Collection

Literature databases are online repositories that provide access to articles, books, and other publications from a range of academic disciplines. These databases allow researchers to search for and access the relevant literature to support their research. Some popular literature databases include WoS, Scopus, and ScienceDirect. Compared to the WoS, Scopus has lower quality control standards than the WoS, and its coverage is limited compared to that of the WoS, which includes over 12,000 journals. In addition, search results from the WoS allow to be exported to other software programs for additional post-processing. For these reasons, we selected the WoS as the database of peer-reviewed literature used in this research.
We considered those documents with a publication year between 1982 and 2022 and studies regarding geophysical research in the Antarctic. To retrieve complete and suitable publications data from the WoS, we constructed search queries based on two main research pillars, treating Antarctica and geophysics as two broad concepts. We used the advanced feature and selected the keywords “Antarctic”, “Antarc”, or “geophysics”, which should appear in the title and/or abstracts. The search was supplemented by Antarctic-related geophysical and expedition station terms, such as “King George Island” and “Antarctic Peninsula”, also combined with specific geophysical terms, such as “seismic”, “gravity”, “magnetic”, “electrical”, “electromagnetic”, “radio”, “echo sounding”, and “RES”. The documents retrieved by this search then underwent manual screening to avoid word ambiguity. After the process was complete and any duplicates had been eliminated, a final total of 3606 published works had been identified.

2.2. Analysis Methods

With the rapid development of computers and information technology, there are now several software tools available for bibliometrics analysis and scientific mapping [35,36]. The most widely used analytical software programs include HistCite [37], VOSviewer [38], and CiteSpace [39]. The software programs used in this article were VOSviewer and CiteSpace. VOSviewer is a commonly used free text-mining tool that runs software invented by Van Eck and Walterman in the Netherlands [38]. VOSviewer is used to create bibliometric networks of different items (e.g., authors, organizations, keywords, etc.) using various network analysis methods, such as co-citation, term co-occurrence, and bibliographic coupling [38]. CiteSpace is a multi-dimensional, time-sharing, and dynamic citation visualization analysis software that focuses on the potential knowledge contained in analytical scientific analysis [40,41]. The software has the capability to draw a knowledge map of a specific field, Utilizing intuitive and visual formats, this tool proficiently translates complex information. As a result, it not only presents the structure of disciplines within the field in an accessible manner but also analyzes the laws and distribution of these disciplines. Through this software, we effectively understand the visual presentation of the developing trend of a discipline or knowledge field within a specific period of time [42,43]. The results obtained via VOSviewer and CiteSpace using different algorithms will be different. In this paper, we synthesize the conclusions of both methods and reference highly cited literature to help us to obtain more accurate conclusions.
The number of publications, countries, institutions, titles, keywords, research hotspots, keyword clustering, and keyword abrupt are some of the markers used in this work to measure the volume of literature. VOSviewer and CiteSpace were used to read the information into the software, using multiple algorithms, and then recombine it according to type and intensity. The study analyzed the co-authorship between countries and institutions, which is defined as a cooperative relationship if they both appear in the same paper. These tools create a map employing colored clusters (such as countries and keywords) and connecting lines to denote specific parameters (such as the degree of cooperation between countries, which is quantified and represented by the clustering circle size). In co-country studies, the strength of the links between countries and institutions is reflected in the number of publications that are co-authored by them. Co-occurrence analysis is then used to analyze the frequency of occurrence of a set of words in the research literature, that is, words that co-occur in a somewhat related body of literature. Cluster analysis is based on the similarity of objects, wherein collections of physical or abstract objects are grouped into multiple clusters that consist of similar objects for analysis [44]. In addition, CiteSpace offers citation burst analysis, a technique that was used to identify the number of suddenly changing citations within a certain period. Utilization of the mutation detection algorithm designed by Kleinberg identify emergencies in the frontiers and extract explosive nodes from big data [45,46].

3. Results and Discussion

3.1. Characteristics of Publication Output

To present an overview of the literature regarding geophysical research, the annual number of articles published from 1982 to 2022 (3606 articles in total) is shown in Figure 2. It should be noted that since the Web of Science core data set can only be searched regarding literature published from 1982 to the present day, only the development of Antarctic geophysics from 1982 to the present day can be analyzed; however, this does not imply that the study and development of Antarctic geophysics began in 1982.
Observing the growth trend depicted in Figure 2, it is evident thatthe literature on Antarctic geophysical research has shown an overall growth trend over a period of more than 40 years between 1982 and 2022. The increase from 3–4 articles in the 1980s to 174 articles in 2022 indicates that geophysical research and investment in surveys of the Antarctic continent is increasing year by year. Antarctic geophysical studies are gaining increasing attention for two main reasons. On the one hand, research in Antarctica is important for understanding global climate change [47,48], geological evolution [49], understanding global sea level changes [50], etc. On the other hand, more than 99.6 percent of the continent’s land mass is covered by an ice sheet that has accumulated over millions of years. The ice sheet is, on average, just over 2133.6 m thick, but in places, it is more than twice that thick. Geophysical and drilling research was used to understand the characteristics and evolution of subglacial geology [51] and provide important information on the englacial environment and bed conditions of the Antarctic ice sheet [52], as well as elucidating the histories of accumulation and ice flow [19,53,54].
In contrast, the development of these publications can be divided into three stages, the slow development stage (1980–1991), the steady development stage (1991–2005), and the rapid development stage (2006–2022). During the slow development stage, the annual output of Antarctic geophysics-related literature and its total volume was relatively small, with only a few studies published. The study area was mainly concentrated in East Antarctica [55,56] and near the marginal seas [57]. Dibble et al. studied the volcanic tectonics of Mount Erebus using seismic, infrasonic, and magnetic induction recordings [58]. During the steady development stage, the annual number of articles published shows an “N”-type distribution trend, with slight up- and downturned fluctuations. Compared with the previous stage, the annual number of publications in this period shows a rapid development trend, from 3–4 papers to more than 40 papers, indicating that international research on Antarctic geophysics was gradually gaining scientific importance. During this period, a total of 1003 papers on Antarctic geophysical research were published. The article “Geophysical studies of the West Antarctic rift system” was published in Tectonics in 1991 by Behrendt, LeMasurie, Cooper, et al. Elsewhere, the Bundesanstalt für Geowissenschaften and Rohstoffe (BGR), working alongside the United States Geological Survey (USGS), used aeromagnetic techniques to investigate the rift structure of the Ross Sea continental shelf. This article was cited 174 times [59]. During the rapid development stages, the number of articles issued shows a large fluctuation, with the annual average number of articles published rising to more than 100 papers. Since the Fourth International Polar Year (2007–2008), the annual number of publications has shown a rapid growth trend, indicating that international integrated geophysical programs have promoted the development of Antarctic geophysical research [60]. The number of publications peaked at 216 in 2021.

3.2. Research Influence and Cooperation Analysis

3.2.1. Major Countries and Institutions for Antarctic Geophysical Research

Antarctic geophysical research involves a total of 82 countries (regions), mainly in Europe, North America, and Asia. Figure 3 lists the top 15 countries with the highest levels of academic productivity and summarizes their total number of articles and centrality. Here, country centrality refers to key nodes with the greatest influence in the co-occurring network, or key nodes with larger intermediary bridging roles [61]. The higher the value of country centrality, the greater the importance of that country in the research area. The formula for centrality is as follows [62]:
C B = s v t V σ s t ( v ) σ s t
where  σ s t ( v )  is the number of shortest path entries from s to t, passing through node v; and  σ s t  is the number of all the shortest paths from node s to node t.
Among these fifteen countries, eight were European, three were Asian, two were Oceanic, and two were North American. Except for Canada, the other 14 countries are either original signatories to the Antarctic Treaty or consultative parties. The United States was responsible for the highest number of publications (1220), accounting for 25.5% of the total number of papers. Their centrality is also the highest, indicating that the United States is a world leader in Antarctic geophysical research. Eight European countries are among the top 15 countries/regions in terms of published Antarctic geophysical research. The United Kingdom and Germany are also among the top three publishers, indicating that Europe is also in a key position in the field of Antarctic geophysics. Other countries/regions are also at the top of the list of publications, such as Australia, China, and Japan, and have made significant contributions to Antarctic geophysical research. China has been conducting comprehensive geophysical surveys in Antarctica since its introduction to the field in 1984 [63,64]. Although China is a latecomer in the field of Antarctic research, it has made rapid progress. In less than 40 years, China has established four Antarctic research stations in Antarctica. In addition, the “Snow Eagle 601” airborne platform serves China’s Antarctic expeditions [65,66]. Snow Eagle 601 was instrumented with ice-penetrating radar (IPR) along with a gravimeter, magnetometer, laser altimeter, and optical camera, and was employed in an investigation of Princess Elizabeth Land (PEL) in 2015 [66].
To date, 497 institutions have contributed to this area of research. The top 20 institutions with the highest number of publications in the field of ocean-based remote sensing research for the entire study period are shown in Figure 4. The Helmholtz Association tops the list with the largest number of publications (278) and a share of 7.71%. The Natural Environment Research Council published 274 papers (7.52%), while the British Antarctic Survey published 263 papers (7.63%). There is little difference between the productivity of the top institutions. Eight of these institutions are from the United States, three are from the United Kingdom, and two are from Germany, thus demonstrating the importance of the United States, the United Kingdom, and Germany in Antarctic geophysical research and the breadth of institutional involvement. Antarctic geophysical research in the United States developed early, with numerous and decentralized forces, showing a clear structure of systemic innovation. The National Science Foundation (NSF) operates the Antarctic basic equipment and organizes university and research institute involvement in Antarctic geophysics-related matters. The University of California and the University of Texas have their own characteristics, maintaining a leading position in Antarctic geophysical research. The University of California is characterized by its leadership in the field of ice-sheet research. In 2019, an Irvine-led team of glaciologists from the university unveiled the most accurate portrait yet created of the contours of the land beneath Antarctica’s ice sheet, the results of which were published in the journal Nature Geoscience [6]. The University of Texas employed space-based, airborne, land-based, and marine geophysical methods to better understand ice sheet evolution, climate, and geologic processes in the polar regions. These efforts provided valuable insights into the dynamics of Antarctica’s unique environment and contributed to the global understanding of climate change impacts [67]. The NSF invests USD 70 million annually into scientific research on Antarctica and the Southern Ocean, along with USD 255 million in the provision of related facilities [68]. The Helmholtz Association is the most prolific publisher of academic papers and is the largest research institution in Germany. It operates five ice-breakers and two polar planes, making it possible to obtain geophysical measurements in Antarctica [69]. As the largest investment institution in the UK, the Nature Environment Research Council (NERC) invested GBP 67.029 billion for polar research facilities in 2022 [70]. The British Antarctic Survey (BAS) has also carried out extensive scientific work and is a world leader in terms of Antarctic research [71]. Other institutions, such as the Centre National de la Recherche Scientifique (CNRS), the National Polar Research Institute, and the Chinese Academy of Sciences also play important roles in Antarctic geophysical research.

3.2.2. Influence Distribution and Cooperation among Core Countries/Institutions

By promoting scientific cooperation among scientists in different disciplines, international cooperation has improved research productivity [72]. In particular, the Scientific Committee on Antarctic Research (SCAR), which was established in 1985, has played a key role in Antarctic cooperation [73,74]. SCAR, being a major international Antarctic research coordination body, supports and promotes various initiatives and projects, including the International Collaborative Exploration of the Cryosphere through Airborne Profiling program (ICECAP) [75,76,77], Instability and Thresholds in Antarctica (INSTANT) [78,79], RINGS [80], and AntArchitecture [81,82,83]. ICECAP is a significant and collaborative international research project involving the United States, the United Kingdom, Australia, France, and China. ICECAP fieldwork is led by the Polar Research Institute of China, using an equipment suite that includes radar and gravity measurements developed by the United States [84]. All these nations, along with the various institutions that are also linked to the program, collaborate on data analysis by offering scientific expertise in geophysics, geology, glacial history, and glaciology. This collaboration aims to gather data to help predict whether and how the East Antarctic Ice Sheet might collapse in the future [85]. These international cooperation projects operate under the umbrella of SCAR, furthering our understanding of Antarctica’s role in the global environment. International cooperation has also made a range of geological and geophysical data sets available to the wider scientific community. Antarctic geophysical data sets, such as ADMAP (The Antarctic Digital Magnetic Anomaly Project), Bedmap (bedrock topography of the Antarctic), AntGG (Gravity and Geoid in Antarctica), and other data sets, are extremely useful for understanding the geological structure and evolution of the Antarctic continent and accelerating scientific discovery related to the Antarctic region [86,87,88,89].
To visualize this cooperation between countries and institutions, national and institutional cooperation networks were obtained via CiteSpace and VOSviewer, from which images were drawn using Scimago Graphic. The top 15 countries, listed in terms of their cooperation strength, are shown in Figure 5, where the nodes represent countries, the links join the various countries at their ends, and the size of the nodes indicates the centrality of the country. The more that the various countries cooperate, the larger their number of nodes. Figure 5 shows that those countries contributing to the field of Antarctic geophysical research have the following characteristics. (1) The largest node in Figure 5 is the United States. As can be seen from our mapping of the country’s cooperation network, it has the largest nodes, representing cooperation links with many more countries compared to others on the map. The United States has led numerous international cooperation programs, such as Polar Earth Observing Network (POLENET) [90,91] and Antarctica’s Gamburtsev Province (AGAP) [92], while Operation IceBridge (OIB) [93,94] is a partnership between NASA and its international partners, which include Canada, Germany, and the United Kingdom [93]. This mission has been ongoing since 2009 and has covered large areas of the polar regions, including Antarctica and the Arctic [95]. The OIB has been designed to bridge the gap between satellite observations and ground-based measurements, providing a more comprehensive and accurate picture of the polar regions [93,96,97]. (2) Due to the strong level of cooperation between European countries in terms of their geographical location and language exchange, the United Kingdom, Germany, Italy, Norway, France, and Russia have developed cooperative relationships and close ties. Russia has not been ranked among the top countries in terms of published literature but maintains close cooperative ties with several countries. The United Kingdom maintains its ongoing track record of successful exploration in remote and hostile frontier regions through major international collaborations, such as the AGAP project in the Gamburtsev Subglacial Mountains. (3) Asian countries are relatively less cooperative than their European counterparts. Japan, China, and India were more active and were closely linked to the United States, Norway, Russia, and others. China also actively participated in international cooperation in the field of Antarctic geophysics. In the fourth IPY, the Prydz Bay–Amery Ice Shelf–Dome A (the PANDA program) is led by China and also represents a core research plan [98].

3.3. Focus on Antarctic Geophysical Research

Keywords are at the heart of academic papers and can highly refine their academic content [44]. By analyzing the co-occurrence patterns of keywords in scholarly articles and applying clustering algorithms, software programs such as CiteSpace can effectively identify the various research groups or clusters, based on their shared interests or thematic areas [99]. The main steps are as follows. First, the CiteSpace program collects data sets that have been obtained from the WoS, comprising scholarly articles or publications related to Antarctic geophysics. Second, CiteSpace performs a co-occurrence analysis of the keywords extracted from the data set; those keywords that frequently appear together are considered relevant and are likely to belong to the same research group. Third, CiteSpace uses a clustering algorithm to organize the co-occurrence network data, to identify the six most popular research groups [41]. These popular research groups comprise Antarctic ice sheet instability and sea level change, the Southern Ocean and sea ice, tectonic activity of the West Antarctic rift system, the paleocontinental rift and reorganization, magmatism and volcanism, and finally, subglacial lakes and subglacial hydrology. To ensure the accuracy of the groups exported from CiteSpace, these groups were critically evaluated and validated according to the most highly cited article and expert validation. By using these methods, we can assess whether the identified groups align with the existing Antarctic geophysical research areas and themes. Details of the six groups are as follows.

3.3.1. Antarctic Ice Sheet Instability and Sea Level Change

The Antarctic ice sheet is the largest in the world and, if it melts completely, the global mean sea level (GMSL) will rise by 60 m [100,101,102]. The Intergovernmental Panel on Climate Change (IPCC) published the Special Report on the Ocean and Cryosphere in a Changing Climate (SROCC), which specifically identified the greatest challenge to accurately quantifying the rate and magnitude of sea level rise as being our limited knowledge of the mechanisms of ice sheet change [103].
Geophysical techniques can be used to accurately measure ice sheet characteristics, as well as their temporal and spatial variability [104,105], which has greatly promoted research on the progress of Antarctic ice sheet movement and research on its instability [106,107,108]. Based on the techniques of satellite altimetry and gravimetry, among others, the Antarctic ice-sheet mass balance can be estimated [109,110,111]. Since 2002, the measurements reported by the Gravity Recovery and Climate Experiment (GRACE) satellite program provide new and key observations for detecting, monitoring, and understanding ice-sheet mass balance [112,113,114,115]. However, the satellite has low resolution, and continuous time series detection is not possible in high-accumulation areas. The results yielded by the different analytical technologies are somewhat different [116,117]. Airborne and ground-based geophysical surveys provide critical data, more detailed information, and observations that can be used to calibrate and validate satellite retrieval results that cannot be measured from space. Ice thickness and near-surface density were measured via a reflection and refraction survey [17], while bedrock elevation was mapped using ground-based and airborne ice-penetrating radar [88]. These data are useful for fine-tuning information on the instability of the Antarctic ice sheet. Figure 6 shows the main geophysical methods used to probe the Antarctic ice sheet.
The recorded geophysical measurements indicate that the mass of the Antarctic ice sheet has been decreasing [100,118,119]. Since extensive measurement-taking began in 1992, the West Antarctic ice sheet (WAIS) and the Antarctic Peninsula have been losing mass [120,121,122,123,124]. Recent mass loss and an increase in the WAIS ice sheet are concentrated in the Amundsen Sea (ASE) and along the coast of the Bellingshausen Sea. These changes are consistent with the recorded observations of grounding line retreat [125] and a decline in the Pine Island Glacier [126,127]. From 2008 to 2015, Gardner et al. used Landsat 7 and 8 imagery, spanning the period from 2013 to 2015, to compare the physical observations to earlier estimates, and calculated a mass loss of −214 ± 51 Gt yr−1 for the WAIS [123]. In contrast, it is generally considered that the East Antarctic ice sheet (EAIS), which encompasses a larger ice mass, has been in a state of mass equilibrium or has had a slightly positive mass balance over the past two decades [100,128,129]. However, recent observations have detected thinning in some of the glaciers on the EAIS, resulting in a negative mass balance in the ice catchments [115,130]. Increased ice velocity and glacier terminal retreat in the EAIS outflow have been observed at Wilkes Land, which indicates mass loss [131]. However, increased EAIS mass on the Siple Coast and Dronning Maud Land have been reported [132]. All these findings suggest that the EAIS is highly dynamic in nature. However, the mass balance of the terrestrially dominated EAIS is still less clearly understood due to the lack of observational evidence.

3.3.2. Southern Ocean and Sea Ice

The Southern Ocean plays an extremely important role in the effective functioning of the Earth’s systems and is a major regulator of planetary climate, acting as an important carbon sink for anthropogenic carbon dioxide from the atmosphere [133]. Antarctic Sea ice was formed when the surface of the Southern Ocean froze; it interacts in various ways with adjacent ice shelves and stranded ice [134]. The small-scale spatial properties of Antarctic Sea ice help to assess the stability and variability of global sea ice cover. Sea ice in the Southern Ocean affects these important global functions; therefore, understanding how ocean–ice interactions occur is a high-priority scientific issue.
The factors of the sea ice itself make it difficult to obtain consistent and continuous data in field measurements; therefore, satellite remote sensing is required to obtain regional thickness distribution data on the circumpolar Antarctic Sea ice [135]. Satellite radar and laser altimetry are currently the most widely used and effective tools for the inversion of sea ice thickness [136]. These tools establish the range [137,138,139,140], thickness [141,142,143], and drift map [144,145] of sea ice, in order to deeply understand the impact of sea ice on the climate [146] and its impact on the ecosystem. Since the late 1970s, variations in the extent of Antarctic Sea ice have been charted regularly using passive microwave satellite imagery [137]. Until the mid-1990s, there was no significant trend in the annual mean total Antarctic sea ice extent or the extent at the annual minimum [147]. However, after reaching a record-high annual mean Antarctic Sea ice extent in 2014, the extent experienced a dramatic decline. In 2017, both the annual daily maximum and minimum extent, as well as the annual mean extent, reached record or near-record lows [148]) (Figure 7a). In February 2022, the annual minimum Antarctic Sea ice extent reached a historic low in the satellite era, with less than two million square kilometers recorded. This stands in stark contrast to the slightly positive trend in the Antarctic SIE observed before 2014 [149]. Based on satellite observations, from 1979 to 2015, the Antarctic sea ice area experienced a statistically significant growth rate that was roughly one-third the rate of the retreat observed in Arctic sea ice [150] (Figure 7b). Satellite Radar and Laser Altimetry are currently the most widely used and effective tools for inversion of sea ice thickness [136]. Kurtz et al. analyzed the basin-wide trends in Antarctic Sea ice thickness and volume over five years from 2003 to 2008, utilizing passive microwave observations and satellite laser altimetry data drawn from NASA’s ice, cloud, and land elevation satellite (ICESat), shown in Figure 7c [142]. Satellite Radar and Laser Altimetry are currently the most widely used and effective tools for inversion of sea ice thickness [136]. Kurtz et al. analyzed basin-wide trends in Antarctic Sea ice thickness and volume over five years from 2003 to 2008 utilizing passive microwave observations and satellite laser altimetry data from NASA’s Ice, Cloud, and Land Elevation Satellite (ICESat) [142]. Through analysis, it has been observed that there is a slight negative trend in sea ice thickness during the summer, but balanced losses in thickness, leading to small overall volume changes. However, estimating the thickness of Antarctic Sea ice from remote sensing data is still challenging. These measurements have linked airborne projects (e.g., NASA’s Operation IceBridge, etc.), providing a vital component for understanding long-term changes in Antarctic Sea ice and its impact on the climate [94,95].

3.3.3. Tectonic Activity of the West Antarctic Rift System

Understanding the West Antarctic lithospheric structure is important for elucidating the dynamics and mass balance of the Antarctic ice sheet [151,152]. The West Antarctic rift system (WARS) is one of the largest continental rift systems on Earth [153]. Its size is comparable to that of the North American basin and the East African Rift Valley [59,154,155]. Unlike the stable East Antarctic Craton, the WARS has a complex tectonic history. From the Cenozoic era to the present day, West Antarctic tectonic activity has continued from time to time, and the area is prone to instability and potential collapse [156]. The West Antarctic ice sheet poses the most immediate threat of a large sea level rise [157]. Therefore, an in-depth understanding of the configuration and development of the West Antarctic rift system is essential for accurate mass balance [158,159] and glacial isostatic adjustment [160].
Direct knowledge of the rift system comes primarily from geological and geophysical surveys and drilling. Geophysical data have played a key role in defining the structure of West Antarctica [161]. Since 1980, such data have been collected in the form of about 35,000 km of marine common depth point (CDP) reflection profiles [162]. In 2006/2010, airborne magnetic and gravity measurements taken by the Polarstern helicopter affiliated with the Alfred Wegener Institute (AWI), which took readings in the Amundsen Sea Embayment, revealed important stages in the tectonic evolution of the region [163]. In 2012, the BAS used airborne ice-penetrating radar and magnetic and gravity measurements to identify a mile-deep rift under the ice in West Antarctica, which is assumed to be part of the WARS [164]. The 32,000-line km of aeromagnetic data collected near the Pine Island (PIG) catchment region [165] was used by researchers in the UK to interpret the dynamic partial stability of the West Antarctic ice sheet [86,166]. The seismic and geodetic networks deployed by the Polar Earth-Observing Network (POLENET-ANET) project, working in the Transantarctic Mountains (TAM), have played an important role in understanding the WARS [167,168]. As part of the Polar Earth Observing Network, Lloyd et al. employed the data from 13 temporary broadband seismic stations that were deployed from January 2010 to January 2012 to identify reduced seismic wave velocities in the uppermost mantle beneath the WARS, which are thought to be a residual thermal signature of Neoproterozoic rifting [169].

3.3.4. Paleocontinental Rift and Reorganization

Antarctica has played a crucial role in the study of the geological history and reconstruction of the supercontinents Rodinia and Gondwana [170,171,172,173]. The study of the lithosphere and the identification of geological boundaries in Antarctica and the neighboring continents are essential to an understanding of the geodynamic evolution of the planet, as revealed in the processes that led to the dispersal of Gondwana’s constituent land masses [174]. However, due to the mostly ice-covered nature of Antarctica, the geometries of these blocks, the outlines of the moving bands, and the sutures between them are poorly known [175]. Airborne geophysical surveys, especially magnetic and gravity data, are the only tools that reveal subglacial geology in a regional sense [174].
The regional tectonic zones of Antarctica have become better understood as a result of the collection of 40 years of extensive airborne geophysical data from across the continent. Eastern Dronning Maud Land (DML) in Antarctica represents a crucial area for enhancing our comprehension of the crustal fragments involved in the merging, amalgamation, and breaking up of Rodinia and Gondwana [175]. The BAS acquired 15,500-line km of data, based on high-resolution airborne magnetic measurements from the “Magmatism as a Monitor of Gondwana Break-up” (MAMOG) project conducted on the Jutulstraumen ice stream in western Dronning Maud Land. This provides new constraints on the magmatic and structural context of the DML’s margins and aims to address the mantle processes that led to the initial breaking-up of Gondwana [176]. The AWI has also conducted several airborne geophysical surveys at the site. Early offshore surveys taken in an area north of the DML by the EMAGE project focused on the breaking-up history of Gondwana, specifically, the opening of the Weddell Sea and, thus, the dispersal of Antarctica and Africa/South America [177,178]. The VISA project, run by the AWI in collaboration with the Dresden University of Technology, focused on the East Antarctic Shield and the major structures on its continental margin to better understand the reconstruction of the paleocontinental region [179]. Airborne magnetic, gravity, and geological data from the Antarctic Peninsula have contributed to the scientific interpretation of Mesozoic arc magmatism and terrane accretion at the paleo-Pacific margin of Gondwana [180,181,182]. Overall, airborne gravity and magnetic data provide crucial evidence for revealing Antarctic geological and tectonic information, understanding the regional tectonic provinces of the continent, and providing insight into the breakup and assembly of the paleocontinental region.

3.3.5. Magmatism and Volcanism

Magmatism has played an important role in the geological evolution of Antarctica [183]. Magmatism in Antarctica has occurred in a variety of tectonic settings, resulting in diverse magma types and eruptive styles. Some of the volcanism behavior mirrors the type of magma that occurs [184]. Additionally, the enormous volume of ice cover in Antarctica and its interaction with volcanism have provided rare opportunities for scientists to use the volcanic record to understand past environmental conditions [183,185,186].
Geophysics researchers have inferred the presence of past active magma activity under the Antarctic ice sheet [187]. Lough et al. analyzed seismic data that were recorded by 37 seismic stations deployed in Marie Byrd Land. The observed seismic activity was interpreted as a sign of magma movement within an active subglacial magmatic system, demonstrating that volcanism continues to migrate southwards along the Executive Committee Range [188]. The aeromagnetic method has been the most useful geophysical tool for the identification of subglacial volcanic rocks. Short-wavelength, shallow-source, high-amplitude (100 to >1000 nT) magnetic anomalies were observed, ranging from 5 to 20 km in width at half-amplitude, located about 1 km above the 2–3 km thick moving ice. These magnetic anomalies are thought to be subglacial volcanic rocks [187,189,190]. Van et al. used aeromagnetic, gravity, and satellite imagery to identify 138 subglacial volcanoes in West Antarctica that are concentrated along the West Antarctic rift system [191]. High-resolution airborne magnetic surveys have successfully mapped the extent of Cenozoic magnetism along the arc/pre-arc boundary of the West Antarctic Peninsula, which straddles Adelaide Island [182,192]. Magnetotelluric investigations have provided significant evidence of volcanic activity. Using magnetotelluric data from the Erebus volcano, Hill et al. argued that the steep, melt-related low resistivity of the upper mantle defines the underlying magma system [193]. It has been suggested that the dynamical intersection of the Terror rift and the accommodation zone fracture structure may control the geometry and style of magma transport and storage in the Erebus volcanic system [193]. Gupta et al. also applied tomographic methods to detailed three-dimensional P-wave velocity imaging of the apocalyptic travel times that were recorded at 42 portable seismic stations in and around Mount Erebus, with a focus on understanding the deep structure and evolution of Mount Erebus [194].

3.3.6. Subglacial Lakes and Subglacial Hydrology

Ice-sheet hydrology has long been recognized as a crucial component in the understanding of ice sheets, their behavior, and their evolution [195]. The presence and distribution of water at the ice sheet bed are widely considered to control ice motion by facilitating basal sliding and substrate deformation [101,196]. In addition, as an enclosed special body of water under the ice, the subglacial lake may preserve primitive life forms and unique geological features, recording the evolution of life and changes in the climate and environment [197,198]. Therefore, the study of Antarctic subglacial hydrology has become a hot topic in Antarctic geophysical research.
Most of the identification and characterization of Antarctic subglacial hydrology has relied on remote geophysical observations (Figure 8a). Satellite data (e.g., ICESat, CryoSat-2 data sets) reveal the hydrological conditions present in the substrate. Satellite altimetry techniques have been used on numerous occasions to identify the presence of “active” subglacial lakes, along with their connectivity and their impact on ice dynamics (Figure 8b) [199,200,201,202]. At the same time, various geophysical methods can also be used to reconstruct the subglacial hydrological environment [203]. Radio-echo sounding (RES) is the most effective technique for detecting and characterizing subglacial water bodies (Figure 8c) [204,205]. The utilization of the scattering characteristics of returned bed echoes, such as the specularity content, trailing bed echoes, the bed-echo coherent index, and bed-echo variability, has advanced the quantitative identification of subglacial water and our understanding of subglacial drainage systems [206]. Subglacial lake drilling provides direct access to subglacial lake water and sediment samples, in a process that can extract valuable information on the paleoclimate and paleoenvironment [207,208]. Other geophysical methods, such as active seismic surveys (Figure 8d) and electromagnetic (EM) approaches, play important roles in revealing the geological and hydrological conditions prevalent in subglacial lakes [209,210,211,212].
Geophysical studies have greatly expanded our scientific understanding of subglacial hydrology. The first detection of Antarctic subglacial lakes using RES was achieved in the late 1960s [216]. In the 1970s, Oswald et al. analyzed RES data drawn from the “SPRI-NSF-TUD” database to find the largest subglacial lake ever discovered, Lake Vostok in Antarctica [217]. Subsequently, RES was widely applied in subglacial lake exploration in Antarctica. High-precision ERS-1 satellite radar altimetry data from 1997 to 1999 highlighted a drop in elevation of more than 2 m in the East Central Antarctic ice sheet, a change that was well above instrumental errors, representing evidence of a subglacial lake outburst [201]. This finding shed light on the highly active underlying hydrology of Antarctica and suggests that some subglacial lakes are not completely isolated. Continuous observations of ice surface changes made by the CryoSat [218] and ICESat/IceBridge projects [200] enable effective data capture, which was targeted to predict subglacial lakes and flow paths [219] and can be calibrated via the in situ measurement of these inferred lakes [220,221]. With the increasingly abundant satellite and RES data, the number, location, size, and depth of the Antarctic ice lake are constantly being corrected and updated [222]. Up until 2022, a total of 675 lakes have been discovered under the Antarctic ice sheet by utilizing airborne ice-penetrating radar and satellite altimetry [206] (Figure 9). Active seismic surveys, conducted in conjunction with radar surveys [223,224], can confirm the presence of liquid water beneath the ice and can be used to characterize the lake floor properties (i.e., hard bedrock, sediment, and till porosity) [215,225]. After conducting a thorough seismic field investigation of the lake near South Pole Station, it has been determined that the lake contains a sediment floor and reaches a maximum water depth of 32 m [215]. The “Vibroseis” technique employs snow flow receivers [226] to quantify the permeability, porosity, and groundwater content of rock formations within sedimentary basins at a high resolution [227,228]. Airborne electromagnetic techniques have also been used to image shallow groundwater areas that are 100 to 200 m below some of the thin glaciers and permafrost found in the McMurdo Dry Valleys. However, these techniques can only penetrate about 350 m of ice [229,230,231].

3.4. Theme Evolution in Antarctic Geophysical Research

An analysis of keyword bursts can identify the frontiers and hotspots of the different research periods seen during the evolution of Antarctic geophysical research, thereby revealing changes in the popularity of certain research directions. Statistical measurement functions available from the CiteSpace software were employed to create keyword intensity maps [232]. Table 1 shows 24 mutated keywords in the frontier region of Antarctic geophysics. The citation bursts represent the intensity of the sudden change calculated by CiteSpace. The year variable shows the duration of the mutation. The rise in research hotspots will lead to a burst of keywords occurring within a short time, where the mutation intensity is much greater than that for the common keywords. However, the number of articles published in the period from 1983 to 1990 was very small, with only 3–5 publications. The articles at this stage in the literature are insufficient in number to obtain keywords for strong mutations and clear research hotspots. Thus, the selection from the literature for the period 1991–2022 is not only representative but also provides valuable insights into the theme evolution found in Antarctic geophysical research. Between 1991 and 2022, there have been different studies at the edges of each period, with different focuses. These can be divided into three main phases.
In 1990–1995, the keywords that appeared include “Margin”, “Antarctic Peninsula”, “Anomaly”, “Marine sediments”, “Weddell Sea”, “Gondwanaland”, “Reconstruction”, “Plate”, “Boundary”, and “Seismic waves”. This shows that the short-term popular areas for Antarctic scientific research were mainly concentrated in the areas of “Antarctic Peninsula”, “Weddell Sea”, “Marine”, etc. The reconstruction of Gondwana and plate tectonics are also popular research topics. “Margin” is the keyword with the greatest abrupt intensity of 10.51, indicating that geophysical research interest in the Antarctic continental margins and paleocontinental margins was high from 1991 until 2007. In addition to keywords about the object of study, there are also keywords regarding various geophysical detection techniques, such as “Seismic waves”, indicating an increasing interest in seismic surveys in the period from 1999 to 2007.
In 1996–2005, the most popular keywords that appeared include “Magnetic anomalies”, “Fracture zone”, “Magmatism”, “Plate tectonics”, “Magnetic field”, “Crustal”, and “Prydz Bay”. The keywords “Fracture zone” and “Magmatism” present high intensity and long duration in the literature from 1998 to 2012, indicating that Antarctic geophysical studies focused on rift tectonics and magmatism during this period. The bursting of “Crustal” indicates the increasing interest of scientists in the study of deep structures, such as the Antarctic mantle and the crust. The keyword “Magnetic field” was highlighted in this period, which lasted from 2003 to 2011, indicating that the magnetic method became a popular research direction in Antarctic geophysics at this time.
In 2006–2022, the most significant keywords that appear include “Pacific margin”, “West Antarctica”, “Pine Island Glacier”, “Ice shelves”, “Amundsen Sea Embayment”, “Antarctic glaciology”, and “Rayleigh wave”. The keyword “West Antarctica” had the highest burst intensity of 7.95 and lasted from 2009 to 2017, indicating a strong focus on West Antarctica during this period. The abrupt mutation of “Pine Island glacier” also verifies this phenomenon. These keywords relate to recent research frontiers in Antarctic geophysics, such as the ice sheet mass balance, the Pacific margin, etc.

4. Conclusions

Since the early 1970s, it has been recognized that the field of geophysics has significant implications for Antarctic science. Therefore, a vast amount of research has been published on this topic. The main purpose of this paper was to provide a bibliometrics analysis of this vast body of literature using text-mining techniques and science mapping tools. The current study complements previously published reviews by mapping the existing science and providing a performance analysis. In addition, the overall conceptual evolution of the field is also explored. The specific conclusions are as follows.
(1) Regarding the publication trends, Antarctic geophysical research has consistently increased in terms of the number of papers that are published per year, suggesting a rise in global scientific interest in the subject. The trends can also be divided into three stages of development, based on the number of articles, the focuses of research directions, and major events (the IPY, etc.).
(2) Among the 82 participating countries/regions, the substantial investment and well-developed infrastructure developed by the United States, Germany, and the United Kingdom in Antarctica is a further indication of their leadership in Antarctic geophysical research. In terms of international cooperation, SCAR is a major international Antarctic research coordination body. It advocated for and led the development of the ICECAP, INSTANT, and RINGS programs, driving forward our understanding of Antarctica’s role in the global environment. Research collaboration among the top 15 countries was quite frequent. ADMAP, OIB, AGAP, and other projects have contributed to the development and thematic evolution of geophysics.
(3) The science map reveals six popular research areas through a network analysis of keyword co-occurrences. By illustrating these six popular areas of research, we found that the wider use of satellite remote sensing to investigate sea ice, ice sheets, and subglacial lakes has greatly improved the identification and characterization of cryospheric features. Airborne geophysical methods are the most useful geophysical tools for identifying large-scale paleocontinental reconstructions, magma, and volcanism. Seismic and electromagnetic applications have played a crucial role in studying the subsurface structure and the evolution of the Antarctic ice sheet. Furthermore, radar surveys of the ice sheet, englacial, and basal environments have become increasingly important for evaluating the impact of climate change on Antarctica. In particular, ice radar techniques play a key role in identifying subglacial lakes.
(4) Evolutionary trends in Antarctic geophysical research can be obtained by an examination of the abrupt changes in keywords. It can be seen that the research area gradually shifts from the Antarctic continental margin to the deep interior of the continent. This phenomenon is also inseparable from the progress of technology. In addition, because of global warming and rising sea levels, the mass balance of the Antarctic ice cap has become a hot topic among scientists in many fields. As research in Antarctic geophysics continues to evolve, it will be essential to integrate the latest technologies to further our understanding of this critical region.
Overall, Antarctica is a unique geographic focus for scientific endeavors, with a community of scientists from a variety of countries. The study of Antarctic geophysics is dedicated to advancing our understanding of the Antarctic continent and its ice sheet. The bibliometric study presented in this article used an essential research instrument to obtain a global perspective on developments and trends in the field of Antarctic geophysics research. This study will help researchers to realize the current state of Antarctic geophysics and provide promising directions for future research.

Author Contributions

Conceptualization, Y.Z. and X.L.; Data curation, C.P., Y.Z. and H.Z.; Investigation, Y.Z. and H.Z.; Supervision, C.Z. and C.P.; Writing—original draft, Y.Z.; Writing—review and editing, C.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China, (No. 42242402); Consulting Project of the Chinese Academy of Engineering (2022-XY-87).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stokes, C.R.; Abram, N.J.; Bentley, M.J.; Edwards, T.L.; England, M.H.; Foppert, A.; Jamieson, S.S.; Jones, R.S.; King, M.A.; Lenaerts, J.T. Response of the East Antarctic Ice Sheet to past and future climate change. Nature 2022, 608, 275–286. [Google Scholar] [CrossRef] [PubMed]
  2. Jankowski, E.J.; Drewry, D. The structure of West Antarctica from geophysical studies. Nature 1981, 291, 17–21. [Google Scholar] [CrossRef]
  3. Claerbout, J.F. Fundamentals of Geophysical Data Processing; McGraw-Hill: New York, NY, USA, 1976; Volume 274. [Google Scholar]
  4. Franke, S.; Eisermann, H.; Jokat, W.; Eagles, G.; Asseng, J.; Miller, H.; Steinhage, D.; Helm, V.; Eisen, O.; Jansen, D. Preserved landscapes underneath the Antarctic Ice Sheet reveal the geomorphological history of Jutulstraumen Basin. Earth Surf. Process. Landf. 2021, 46, 2728–2745. [Google Scholar] [CrossRef]
  5. Bo, S.; Siegert, M.J.; Mudd, S.M.; Sugden, D.; Fujita, S.; Xiangbin, C.; Yunyun, J.; Xueyuan, T.; Yuansheng, L. The Gamburtsev mountains and the origin and early evolution of the Antarctic Ice Sheet. Nature 2009, 459, 690–693. [Google Scholar] [CrossRef] [PubMed]
  6. Morlighem, M.; Rignot, E.; Binder, T.; Blankenship, D.; Drews, R.; Eagles, G.; Eisen, O.; Ferraccioli, F.; Forsberg, R.; Fretwell, P. Deep glacial troughs and stabilizing ridges unveiled beneath the margins of the Antarctic ice sheet. Nat. Geosci. 2020, 13, 132–137. [Google Scholar] [CrossRef] [Green Version]
  7. Gulick, S.P.; Shevenell, A.E.; Montelli, A.; Fernandez, R.; Smith, C.; Warny, S.; Bohaty, S.M.; Sjunneskog, C.; Leventer, A.; Frederick, B. Initiation and long-term instability of the East Antarctic Ice Sheet. Nature 2017, 552, 225–229. [Google Scholar] [CrossRef] [Green Version]
  8. Joughin, I.; Alley, R.B. Stability of the West Antarctic ice sheet in a warming world. Nat. Geosci. 2011, 4, 506–513. [Google Scholar] [CrossRef]
  9. Behrendt, J.C. Petroleum and Mineral Resources of Antarctica; Geological Survey: Washington, DC, USA, 1983.
  10. Nan, Z.; Grigor’evich, T.P. Antarctic oil and mineral resources: A subject off limits or future reality? Authorea Prepr. 2022. [Google Scholar] [CrossRef]
  11. Jin, Y.; Lee, M.; Kim, Y.; Nam, S.; Kim, K. Gas hydrate volume estimations on the South Shetland continental margin, Antarctic Peninsula. Antarct. Sci. 2003, 15, 271–282. [Google Scholar] [CrossRef]
  12. Easterbrook, D. Evidence-Based Climate Science: Data Opposing CO2 Emissions as the Primary Source of Global Warming; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar]
  13. Rohling, E.J.; Grant, K.; Bolshaw, M.; Roberts, A.; Siddall, M.; Hemleben, C.; Kucera, M. Antarctic temperature and global sea level closely coupled over the past five glacial cycles. Nat. Geosci. 2009, 2, 500–504. [Google Scholar] [CrossRef]
  14. Turner, J.; Bindschadler, R.; Convey, P.; Di Prisco, G.; Fahrbach, E.; Gutt, J.; Hodgson, D.; Mayewski, P.; Summerhayes, C. Antarctic Climate Change and the Environment; SCAR: Cambridgeshire, UK, 2009. [Google Scholar]
  15. Naylor, S.; Dean, K.; Siegert, M. The IGY and the ice sheet: Surveying Antarctica. J. Hist. Geogr. 2008, 34, 574–595. [Google Scholar] [CrossRef]
  16. Robin, G.d.Q. Norwegian-British-Swedish Antarctic Expedition, 1949–1952. Polar Rec. 1953, 6, 608–616. [Google Scholar] [CrossRef]
  17. Robin, G.d.Q. Seismic Shooting and Related Investigations; Norsk Polarinstitut: Tromsø, Norway, 1958. [Google Scholar]
  18. Siegert, M.J.; Jamieson, S.S.; White, D. Exploration of Subsurface Antarctica: Uncovering Past Changes and Modern Processes; Geological Society: London, UK, 2018; Volume 461, pp. 1–6. [Google Scholar]
  19. Schroeder, D.M.; Bingham, R.G.; Blankenship, D.D.; Christianson, K.; Eisen, O.; Flowers, G.E.; Karlsson, N.B.; Koutnik, M.R.; Paden, J.D.; Siegert, M.J. Five decades of radioglaciology. Ann. Glaciol. 2020, 61, 1–13. [Google Scholar] [CrossRef] [Green Version]
  20. Dean, K.; Naylor, S.; Turchetti, S.; Siegert, M. Data in Antarctic science and politics. Soc. Stud. Sci. 2008, 38, 571–604. [Google Scholar] [CrossRef]
  21. Winter, A.; Steinhage, D.; Creyts, T.T.; Kleiner, T.; Eisen, O. Age stratigraphy in the East Antarctic Ice Sheet inferred from radio-echo sounding horizons. Earth Syst. Sci. Data 2019, 11, 1069–1081. [Google Scholar] [CrossRef] [Green Version]
  22. Frémand, A.C.; Bodart, J.A.; Jordan, T.A.; Ferraccioli, F.; Robinson, C.; Corr, H.F.; Peat, H.J.; Bingham, R.G.; Vaughan, D.G. British Antarctic Survey’s aerogeophysical data: Releasing 25 years of airborne gravity, magnetic, and radar datasets over Antarctica. Earth Syst. Sci. Data 2022, 14, 3379–3410. [Google Scholar] [CrossRef]
  23. Huybrechts, P. A 3-D model for the Antarctic ice sheet: A sensitivity study on the glacial-interglacial contrast. Clim. Dyn. 1990, 5, 79–92. [Google Scholar] [CrossRef] [Green Version]
  24. Funaki, M.; Higashino, S.-I.; Sakanaka, S.; Iwata, N.; Nakamura, N.; Hirasawa, N.; Obara, N.; Kuwabara, M. Small unmanned aerial vehicles for aeromagnetic surveys and their flights in the South Shetland Islands, Antarctica. Polar Sci. 2014, 8, 342–356. [Google Scholar] [CrossRef] [Green Version]
  25. Lucieer, A.; Robinson, S.; Turner, D.; Harwin, S.; Kelcey, J. Using a micro-UAV for ultra-high resolution multi-sensor observations of Antarctic moss beds. ISPRS Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2012, XXXIX-B1, 429–433. [Google Scholar] [CrossRef] [Green Version]
  26. McPhail, S.; Templeton, R.; Pebody, M.; Roper, D.; Morrison, R. Autosub long range AUV missions under the Filchner and Ronne ice shelves in the Weddell sea, Antarctica-an engineering perspective. In Proceedings of the OCEANS 2019-Marseille, Marseille, France, 17–20 June 2019; pp. 1–8. [Google Scholar]
  27. King, P.; Williams, G.; Coleman, R.; Zürcher, K.; Bowden-Floyd, I.; Ronan, A.; Kaminski, C.; Laframboise, J.-M.; McPhail, S.; Wilkinson, J. Deploying an AUV beneath the Sørsdal Ice Shelf: Recommendations from an expert-panel workshop. In Proceedings of the 2018 IEEE/OES Autonomous Underwater Vehicle Workshop (AUV), Porto, Portugal, 6–9 November 2018; pp. 1–6. [Google Scholar]
  28. Nederhof, A.J. Bibliometric monitoring of research performance in the social sciences and the humanities: A review. Scientometrics 2006, 66, 81–100. [Google Scholar] [CrossRef]
  29. Donthu, N.; Kumar, S.; Mukherjee, D.; Pandey, N.; Lim, W.M. How to conduct a bibliometric analysis: An overview and guidelines. J. Bus. Res. 2021, 133, 285–296. [Google Scholar] [CrossRef]
  30. Adie, R. Earth sciences: The geology of Antarctica: A review. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1977, 279, 123–130. [Google Scholar]
  31. Fogg, G.E. A History of Antarctic Science; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  32. Dastidar, P.; Ramachandran, S. Intellectual structure of Antarctic science: A 25-years analysis. Scientometrics 2008, 77, 389–414. [Google Scholar] [CrossRef]
  33. Ji, Q.; Pang, X.; Zhao, X. A bibliometric analysis of research on Antarctica during 1993–2012. Scientometrics 2014, 101, 1925–1939. [Google Scholar] [CrossRef]
  34. Pina, P.; Vieira, G. UAVs for science in Antarctica. Remote Sens. 2022, 14, 1610. [Google Scholar] [CrossRef]
  35. Cobo, M.J.; López-Herrera, A.G.; Herrera-Viedma, E.; Herrera, F. An approach for detecting, quantifying, and visualizing the evolution of a research field: A practical application to the Fuzzy Sets Theory field. J. Informetr. 2011, 5, 146–166. [Google Scholar] [CrossRef]
  36. Cobo, M.J.; López-Herrera, A.G.; Herrera-Viedma, E.; Herrera, F. Science mapping software tools: Review, analysis, and cooperative study among tools. J. Am. Soc. Inf. Sci. Technol. 2011, 62, 1382–1402. [Google Scholar] [CrossRef]
  37. Garfield, E. From the science of science to Scientometrics visualizing the history of science with HistCite software. J. Informetr. 2009, 3, 173–179. [Google Scholar] [CrossRef] [Green Version]
  38. Van Eck, N.; Waltman, L. Software survey: VOSviewer, a computer program for bibliometric mapping. Scientometrics 2010, 84, 523–538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Chen, C. Searching for intellectual turning points: Progressive knowledge domain visualization. Proc. Natl. Acad. Sci. USA 2004, 101, 5303–5310. [Google Scholar] [CrossRef]
  40. Guo, P.; Tian, W.; Li, H.; Zhang, G.; Li, J. Global characteristics and trends of research on construction dust: Based on bibliometric and visualized analysis. Environ. Sci. Pollut. Res. 2020, 27, 37773–37789. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, C. CiteSpace II: Detecting and visualizing emerging trends and transient patterns in scientific literature. J. Am. Soc. Inf. Sci. Technol. 2006, 57, 359–377. [Google Scholar] [CrossRef] [Green Version]
  42. Chen, C.; Hu, Z.; Liu, S.; Tseng, H. Emerging trends in regenerative medicine: A scientometric analysis in CiteSpace. Expert Opin. Biol. Ther. 2012, 12, 593–608. [Google Scholar] [CrossRef] [PubMed]
  43. Li, J.; Chen, C. CiteSpace: Text Mining and Visualization in Scientific Literature; Capital University of Economics and Business Press: Beijing, China, 2016; pp. 149–152. [Google Scholar]
  44. Duan, P.; Wang, Y.; Yin, P. Remote sensing applications in monitoring of protected areas: A bibliometric analysis. Remote Sens. 2020, 12, 772. [Google Scholar] [CrossRef] [Green Version]
  45. Kleinberg, J. Bursty and hierarchical structure in streams. In Proceedings of the Eighth ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, Edmonton, AB, Canada, 23–36 July 2002; pp. 91–101. [Google Scholar]
  46. Xu, T.; Wang, F.; Yi, Q.; Xie, L.; Yao, X. A Bibliometric and Visualized Analysis of Research Progress and Trends in Rice Remote Sensing over the Past 42 Years (1980–2021). Remote Sens. 2022, 14, 3607. [Google Scholar] [CrossRef]
  47. Brook, E.J.; Buizert, C. Antarctic and global climate history viewed from ice cores. Nature 2018, 558, 200–208. [Google Scholar] [CrossRef]
  48. Rintoul, S.R.; Chown, S.L.; DeConto, R.M.; England, M.H.; Fricker, H.A.; Masson-Delmotte, V.; Naish, T.R.; Siegert, M.J.; Xavier, J.C. Choosing the future of Antarctica. Nature 2018, 558, 233–241. [Google Scholar] [CrossRef]
  49. Klages, J.P.; Salzmann, U.; Bickert, T.; Hillenbrand, C.-D.; Gohl, K.; Kuhn, G.; Bohaty, S.M.; Titschack, J.; Müller, J.; Frederichs, T. Temperate rainforests near the South Pole during peak Cretaceous warmth. Nature 2020, 580, 81–86. [Google Scholar] [CrossRef]
  50. Whitehouse, P.L.; Gomez, N.; King, M.A.; Wiens, D.A. Solid Earth change and the evolution of the Antarctic Ice Sheet. Nat. Commun. 2019, 10, 503. [Google Scholar] [CrossRef] [Green Version]
  51. Kennicutt, M.C.; Bromwich, D.; Liggett, D.; Njåstad, B.; Peck, L.; Rintoul, S.R.; Ritz, C.; Siegert, M.J.; Aitken, A.; Brooks, C.M. Sustained Antarctic research: A 21st century imperative. One Earth 2019, 1, 95–113. [Google Scholar] [CrossRef] [Green Version]
  52. Cui, X.; Jeofry, H.; Greenbaum, J.S.; Guo, J.; Li, L.; Lindzey, L.E.; Habbal, F.A.; Wei, W.; Young, D.A.; Ross, N. Bed topography of princess Elizabeth land in east Antarctica. Earth Syst. Sci. Data 2020, 12, 2765–2774. [Google Scholar] [CrossRef]
  53. Koutnik, M.R.; Fudge, T.; Conway, H.; Waddington, E.D.; Neumann, T.A.; Cuffey, K.M.; Buizert, C.; Taylor, K.C. Holocene accumulation and ice flow near the West Antarctic Ice Sheet Divide ice core site. J. Geophys. Res. Earth Surf. 2016, 121, 907–924. [Google Scholar] [CrossRef] [Green Version]
  54. Smith, A.; Anker, P.; Nicholls, K.; Makinson, K.; Murray, T.; Rios-Costas, S.; Brisbourne, A.; Hodgson, D.; Schlegel, R.; Anandakrishnan, S. Ice stream subglacial access for ice-sheet history and fast ice flow: The BEAMISH Project on Rutford Ice Stream, West Antarctica and initial results on basal conditions. Ann. Glaciol. 2021, 62, 203–211. [Google Scholar] [CrossRef]
  55. Ito, K.; Ikami, A. Crustal structure of the Mizuho Plateau, East Antarctica from geophysical data. J. Geodyn. 1986, 6, 285–296. [Google Scholar] [CrossRef]
  56. Martin, A.; Hartnady, C. Plate tectonic development of the South West Indian Ocean: A revised reconstruction of East Antarctica and Africa. J. Geophys. Res. Solid Earth 1986, 91, 4767–4786. [Google Scholar] [CrossRef]
  57. Cook, R.A.; Davey, F. The hydrocarbon exploration of the basins of the Ross sea, Antarctica, from modelling of the geophysical data. J. Pet. Geol. 1984, 7, 213–225. [Google Scholar] [CrossRef]
  58. Dibble, R.; Kienle, J.; Kyle, P.; Shibuya, K. Geophysical studies of Erebus volcano, Antarctica, from 1974 December to 1982 January. N. Z. J. Geol. Geophys. 1984, 27, 425–455. [Google Scholar] [CrossRef]
  59. Behrendt, J.C.; LeMasurier, W.; Cooper, A.; Tessensohn, F.; Trehu, A.; Damaske, D. Geophysical studies of the West Antarctic rift system. Tectonics 1991, 10, 1257–1273. [Google Scholar] [CrossRef]
  60. Barr, S.; Lüdecke, C. The History of the International Polar Years (IPYs); Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  61. Li, B.; Si, G.; Ding, J.; Wang, F. A faster algorithm to calculate centrality based on Shortest Path Layer. In Proceedings of the 2017 29th Chinese Control and Decision Conference (CCDC), Chongqing, China, 28–30 May 2017; pp. 6283–6290. [Google Scholar]
  62. Puzis, R.; Zilberman, P.; Elovici, Y.; Dolev, S.; Brandes, U. Heuristics for speeding up betweenness centrality computation. In Proceedings of the 2012 International Conference on Privacy, Security, Risk and Trust and 2012 International Confernece on Social Computing, Amsterdam, The Netherlands, 3–5 September 2012; pp. 302–311. [Google Scholar]
  63. Chen, L.; Liu, X.; Bian, L.; Chen, B.; Huang, H.; Hu, H.; Luo, W.; Shi, G.; Shi, J.; Xu, C. Overview of China’s Antarctic research progress 1984–2016. Adv. Polar Sci. 2017, 28, 151–160. [Google Scholar] [CrossRef]
  64. Gao, J.; Shen, Z.; Yang, C.; Wang, W.; Ji, F.; Wu, Z.; Niu, X.; Ding, W.; Li, D.; Zhang, Q. Progress in Antarctic marine geophysical research by the Chinese Polar Program. Adv. Polar Sci. 2017, 28, 256–267. [Google Scholar]
  65. Cui, X.; Greenbaum, J.S.; Lang, S.; Zhao, X.; Li, L.; Guo, J.; Sun, B. The scientific operations of Snow Eagle 601 in Antarctica in the past five austral seasons. Remote Sens. 2020, 12, 2994. [Google Scholar] [CrossRef]
  66. Cui, X.; Greenbaum, J.S.; Beem, L.H.; Guo, J.; Ng, G.; Li, L.; Blankenship, D.; Sun, B. The first fixed-wing aircraft for Chinese Antarctic expeditions: Airframe, modifications, scientific instrumentation and applications. J. Environ. Eng. Geophys. 2018, 23, 1–13. [Google Scholar] [CrossRef]
  67. The University of Texas. Available online: https://ig.utexas.edu/polar-and-planetary/ (accessed on 18 June 2023).
  68. Niu, Y.; Zhang, S.; Zhao, J.; Xiong, Y.; Wu, X. Comprehensive evaluation and analysis of the international polar research from 2010 to 2016. J. Glaciol. Geocryol. 2017, 39, 1039–1046. [Google Scholar]
  69. Helmholtz. Available online: https://www.helmholtz.de/en/about-us/helmholtz-centers/centers-a-z/centre/alfred-wegener-institute-helmholtz-centre-for-polar-and-marine-research-awi/ (accessed on 16 June 2023).
  70. UK Research and Innovation. Available online: https://www.ukri.org/news/uk-invests-to-modernise-polar-science/ (accessed on 16 June 2023).
  71. UK Research and Innovation. Available online: https://www.ukri.org/about-us/nerc/who-we-are/nerc-research-centres/ (accessed on 7 December 2022).
  72. Erb, K.A. International collaboration in the Antarctic for global science. Sci. Dipl. Antarct. Sci. Gov. Int. Spaces 2011, 77, 265–270. [Google Scholar]
  73. Lüdecke, C. Scientific collaboration in Antarctica (1901–04): A challenge in times of political rivalry. Polar Rec. 2003, 39, 35–48. [Google Scholar] [CrossRef]
  74. Walton, D.W.; Kennicutt, M.C.; Summerhayes, C.P. Antarctic Scientific Collaboration: The Role of the SCAR. In Exploring the Last Continent: An Introduction to Antarctica; Springer: Berlin/Heidelberg, Germany, 2015; pp. 573–588. [Google Scholar]
  75. Jamieson, S.S.; Ross, N.; Greenbaum, J.S.; Young, D.A.; Aitken, A.R.; Roberts, J.L.; Blankenship, D.D.; Bo, S.; Siegert, M.J. An extensive subglacial lake and canyon system in Princess Elizabeth Land, East Antarctica. Geology 2016, 44, 87–90. [Google Scholar] [CrossRef] [Green Version]
  76. Tang, X.; Sun, B. Towards an integrated study of subglacial conditions in Princess Elizabeth Land, East Antarctica. Adv. Polar Sci 2021, 32, 75–77. [Google Scholar]
  77. Greenbaum, J.; Blankenship, D.; Young, D.; Richter, T.; Roberts, J.; Aitken, A.; Legresy, B.; Schroeder, D.; Warner, R.; Van Ommen, T. Ocean access to a cavity beneath Totten Glacier in East Antarctica. Nat. Geosci. 2015, 8, 294–298. [Google Scholar] [CrossRef]
  78. Tang, X. 2021 international symposium on “Instability and Thresholds in Antarctica (INSTANT)”, Scientific Committee on Antarctic Research (SCAR). Chin. J. Polar Res. 2021, 33, 309. [Google Scholar]
  79. Siegert, M.; Florindo, F.; De Santis, L.; Naish, T.R. The future evolution of Antarctic climate: Conclusions and upcoming programmes. In Antarctic Climate Evolution; Elsevier: Amsterdam, The Netherlands, 2022; pp. 769–775. [Google Scholar]
  80. Brooks, S.M.; Becker, T.M.; Baillie, K.; Becker, H.N.; Bradley, E.T.; Colwell, J.E.; Cuzzi, J.; De Pater, I.; Eckert, S.; Elmoutamid, M. Frontiers in Planetary Rings Science. Bull. Am. Astron. Soc. 2021, 53, 258. [Google Scholar]
  81. Cavitte, M.G.; Young, D.A.; Mulvaney, R.; Ritz, C.; Greenbaum, J.S.; Ng, G.; Kempf, S.D.; Quartini, E.; Muldoon, G.R.; Paden, J. A detailed radiostratigraphic data set for the central East Antarctic Plateau spanning from the Holocene to the mid-Pleistocene. Earth Syst. Sci. Data 2021, 13, 4759–4777. [Google Scholar] [CrossRef]
  82. Sutter, J.; Fischer, H.; Eisen, O. Investigating the internal structure of the Antarctic ice sheet: The utility of isochrones for spatiotemporal ice-sheet model calibration. Cryosphere 2021, 15, 3839–3860. [Google Scholar] [CrossRef]
  83. Delf, R.; Schroeder, D.M.; Curtis, A.; Giannopoulos, A.; Bingham, R.G. A comparison of automated approaches to extracting englacial-layer geometry from radar data across ice sheets. Ann. Glaciol. 2020, 61, 234–241. [Google Scholar] [CrossRef]
  84. Young, D.A.; Greenbaum, J.S.; Blankenship, D.D.; Siegert, M.J.; Roberts, J.L.; van Ommen, T.D.; Aitken, A.; Grima, C.; Le Meur, E. ICECAP’s contribution to NASA’s Operation IceBridge in East Antarctica. In Proceedings of the AGU Fall Meeting Abstracts, San Francisco, CA, USA, 9–13 December 2019; p. C31C-1562. [Google Scholar]
  85. Wei, W.; Greenbaum, J.S.; Gourmelen, N.; Dow, C.F.; Bo, S.; Guo, J.; van Ommen, T.D.; Roberts, J.L.; Young, D.A.; Blankenship, D.D. The bathymetric and subglacial hydrological context for basal melting of the West Ice Shelf in East Antarctica. In Proceedings of the AGU Fall Meeting Abstracts, Washington, DC, USA, 10–14 December 2018; p. C21C-1355. [Google Scholar]
  86. Golynsky, A.; Bell, R.; Blankenship, D.; Damaske, D.; Ferraccioli, F.; Finn, C.; Golynsky, D.; Ivanov, S.; Jokat, W.; Masolov, V. Air and shipborne magnetic surveys of the Antarctic into the 21st century. Tectonophysics 2013, 585, 3–12. [Google Scholar] [CrossRef] [Green Version]
  87. Kim, H.; Golynsky, A.; Golynsky, D.; Yu, H.; von Frese, R.; Hong, J. New magnetic anomaly constraints on the Antarctic crust. J. Geophys. Res. Solid Earth 2022, 127, e2021JB023329. [Google Scholar] [CrossRef]
  88. Fretwell, P.; Pritchard, H.D.; Vaughan, D.G.; Bamber, J.L.; Barrand, N.E.; Bell, R.; Bianchi, C.; Bingham, R.; Blankenship, D.D.; Casassa, G. Bedmap2: Improved ice bed, surface and thickness datasets for Antarctica. Cryosphere 2013, 7, 375–393. [Google Scholar] [CrossRef] [Green Version]
  89. Scheinert, M.; Ferraccioli, F.; Schwabe, J.; Bell, R.; Studinger, M.; Damaske, D.; Jokat, W.; Aleshkova, N.; Jordan, T.; Leitchenkov, G. New Antarctic gravity anomaly grid for enhanced geodetic and geophysical studies in Antarctica. Geophys. Res. Lett. 2016, 43, 600–610. [Google Scholar] [CrossRef] [Green Version]
  90. Wiens, D.; Wilson, T.; Dietrich, R. POLENET: Polar Earth Observing Network for the International Polar Year. In Proceedings of the AGU Fall Meeting Abstracts, San Francisco, CA, USA, 10–14 December 2007; p. U24D-03. [Google Scholar]
  91. Wilson, T.; Wiens, D.; Smalley, B.; Raymond, C.; Nyblade, A.; Huerta, A.; Dalziel, I.; Bevis, M.; Aster, R.; Anandakrishnan, S. POLENET seismic and GPS network in West Antarctica. In Proceedings of the AGU Fall Meeting Abstracts, San Francisco, CA, USA, 15–19 December 2008; p. V11F-02. [Google Scholar]
  92. Ferraccioli, F.; Finn, C.A.; Jordan, T.A.; Bell, R.E.; Anderson, L.M.; Damaske, D. East Antarctic rifting triggers uplift of the Gamburtsev Mountains. Nature 2011, 479, 388–392. [Google Scholar] [CrossRef]
  93. MacGregor, J.A.; Boisvert, L.N.; Medley, B.; Petty, A.A.; Harbeck, J.P.; Bell, R.E.; Blair, J.B.; Blanchard-Wrigglesworth, E.; Buckley, E.M.; Christoffersen, M.S. The scientific legacy of NASA’s Operation IceBridge. Rev. Geophys. 2021, 59, e2020RG000712. [Google Scholar] [CrossRef]
  94. Kurtz, N.; Farrell, S.; Studinger, M.; Galin, N.; Harbeck, J.; Lindsay, R.; Onana, V.; Panzer, B.; Sonntag, J. Sea ice thickness, freeboard, and snow depth products from Operation IceBridge airborne data. Cryosphere 2013, 7, 1035–1056. [Google Scholar] [CrossRef] [Green Version]
  95. Farrell, S.L.; Kurtz, N.; Connor, L.N.; Elder, B.C.; Leuschen, C.; Markus, T.; McAdoo, D.C.; Panzer, B.; Richter-Menge, J.; Sonntag, J.G. A first assessment of IceBridge snow and ice thickness data over Arctic sea ice. IEEE Trans. Geosci. Remote Sens. 2011, 50, 2098–2111. [Google Scholar] [CrossRef]
  96. Studinger, M.; Koenig, L.; Martin, S.; Sonntag, J. Operation icebridge: Using instrumented aircraft to bridge the observational gap between Icesat and Icesat-2. In Proceedings of the 2010 IEEE International Geoscience and Remote Sensing Symposium, Honolulu, HI, USA, 25–30 July 2010; pp. 1918–1919. [Google Scholar]
  97. Kurtz, N.T.; Farrell, S.L. Large-scale surveys of snow depth on Arctic sea ice from Operation IceBridge. Geophys. Res. Lett. 2011, 38, L20505. [Google Scholar] [CrossRef]
  98. Zhang, Q.; Wang, Y. Zhongguo nanji kaocha 28 nian lai de jinzhan [Antarctic research progress of China in the past 28 years]. Ziran Zazhi 2008, 30, 252–258. [Google Scholar]
  99. Montemurro, M.A.; Zanette, D.H. Keywords and co-occurrence patterns in the Voynich manuscript: An information-theoretic analysis. PLoS ONE 2013, 8, e66344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Rignot, E.; Mouginot, J.; Scheuchl, B.; Van Den Broeke, M.; Van Wessem, M.J.; Morlighem, M. Four decades of Antarctic Ice Sheet mass balance from 1979–2017. Proc. Natl. Acad. Sci. USA 2019, 116, 1095–1103. [Google Scholar] [CrossRef] [Green Version]
  101. Alley, R.B.; Blankenship, D.D.; Bentley, C.R.; Rooney, S. Deformation of till beneath ice stream B, West Antarctica. Nature 1986, 322, 57–59. [Google Scholar] [CrossRef]
  102. Kennicutt, M.C.; Chown, S.L.; Cassano, J.J.; Liggett, D.; Massom, R.; Peck, L.S.; Rintoul, S.R.; Storey, J.W.; Vaughan, D.G.; Wilson, T.J. Polar research: Six priorities for Antarctic science. Nature 2014, 512, 23–25. [Google Scholar] [CrossRef] [Green Version]
  103. 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: Geneva, Switzerland, 2019; Volume 1155. [Google Scholar]
  104. Allison, I.; Alley, R.; Fricker, H.; Thomas, R.; Warner, R. Ice sheet mass balance and sea level. Antarct. Sci. 2009, 21, 413–426. [Google Scholar] [CrossRef] [Green Version]
  105. Kennicutt, M.C.; Chown, S.L.; Cassano, J.J.; Liggett, D.; Peck, L.S.; Massom, R.; Rintoul, S.; Storey, J.; Vaughan, D.; Wilson, T. A roadmap for Antarctic and Southern Ocean science for the next two decades and beyond. Antarct. Sci. 2015, 27, 3–18. [Google Scholar] [CrossRef] [Green Version]
  106. Treuer, G.A. The psychology of Miami’s struggle to adapt to sea-level rise. Bull. At. Sci. 2018, 74, 155–159. [Google Scholar] [CrossRef]
  107. Siegert, M.J.; Kulessa, B.; Bougamont, M.; Christoffersen, P.; Key, K.; Andersen, K.R.; Booth, A.D.; Smith, A.M. Antarctic subglacial groundwater: A concept paper on its measurement and potential influence on ice flow. Geol. Soc. Lond. Spec. Publ. 2018, 461, 197–213. [Google Scholar] [CrossRef]
  108. Yang, T.; Liang, Q.; Zheng, L.; Li, T.; Chen, Z.; Hui, F.; Cheng, X. Mass Balance of the Antarctic Ice Sheet in the Early 21st Century. Remote Sens. 2023, 15, 1677. [Google Scholar] [CrossRef]
  109. Van den Broeke, M.R.; Bamber, J.; Lenaerts, J.; Rignot, E. Ice sheets and sea level: Thinking outside the box. Surv. Geophys. 2011, 32, 495–505. [Google Scholar] [CrossRef]
  110. Rémy, F.; Parouty, S. Antarctic ice sheet and radar altimetry: A review. Remote Sens. 2009, 1, 1212–1239. [Google Scholar] [CrossRef] [Green Version]
  111. Yu, A.; Shi, H.; Wang, Y.; Yang, J.; Gao, C.; Lu, Y. A Bibliometric and Visualized Analysis of Remote Sensing Methods for Glacier Mass Balance Research. Remote Sens. 2023, 15, 1425. [Google Scholar] [CrossRef]
  112. Velicogna, I. Increasing rates of ice mass loss from the Greenland and Antarctic ice sheets revealed by GRACE. Geophys. Res. Lett. 2009, 36, L19503. [Google Scholar] [CrossRef] [Green Version]
  113. Velicogna, I.; Wahr, J. Acceleration of Greenland ice mass loss in spring 2004. Nature 2006, 443, 329–331. [Google Scholar] [CrossRef] [Green Version]
  114. Velicogna, I.; Wahr, J. Measurements of time-variable gravity show mass loss in Antarctica. Science 2006, 311, 1754–1756. [Google Scholar] [CrossRef] [Green Version]
  115. Chen, J.; Wilson, C.; Blankenship, D.; Tapley, B. Accelerated Antarctic ice loss from satellite gravity measurements. Nat. Geosci. 2009, 2, 859–862. [Google Scholar] [CrossRef]
  116. Pritchard, H.D.; Arthern, R.J.; Vaughan, D.G.; Edwards, L.A. Extensive dynamic thinning on the margins of the Greenland and Antarctic ice sheets. Nature 2009, 461, 971–975. [Google Scholar] [CrossRef] [PubMed]
  117. Riva, R.E.; Gunter, B.C.; Urban, T.J.; Vermeersen, B.L.; Lindenbergh, R.C.; Helsen, M.M.; Bamber, J.L.; van de Wal, R.S.; van den Broeke, M.R.; Schutz, B.E. Glacial isostatic adjustment over Antarctica from combined ICESat and GRACE satellite data. Earth Planet. Sci. Lett. 2009, 288, 516–523. [Google Scholar] [CrossRef]
  118. Noble, T.; Rohling, E.; Aitken, A.; Bostock, H.; Chase, Z.; Gomez, N.; Jong, L.; King, M.A.; Mackintosh, A.; McCormack, F. The sensitivity of the Antarctic ice sheet to a changing climate: Past, present, and future. Rev. Geophys. 2020, 58, e2019RG000663. [Google Scholar] [CrossRef]
  119. Jawak, S.D.; Luis, A.J.; Pandit, P.H.; Wankhede, S.F.; Convey, P.; Fretwell, P.T. Exploratory Mapping of Blue Ice Regions in Antarctica Using Very High-Resolution Satellite Remote Sensing Data. Remote Sens. 2023, 15, 1287. [Google Scholar] [CrossRef]
  120. Martin-Español, A.; Bamber, J.L.; Zammit-Mangion, A. Constraining the mass balance of East Antarctica. Geophys. Res. Lett. 2017, 44, 4168–4175. [Google Scholar] [CrossRef] [Green Version]
  121. Zwally, H.J.; Li, J.; Robbins, J.W.; Saba, J.L.; Yi, D.; Brenner, A.C. Mass gains of the Antarctic ice sheet exceed losses. J. Glaciol. 2015, 61, 1019–1036. [Google Scholar] [CrossRef] [Green Version]
  122. Bamber, J.L.; Westaway, R.M.; Marzeion, B.; Wouters, B. The land ice contribution to sea level during the satellite era. Environ. Res. Lett. 2018, 13, 063008. [Google Scholar] [CrossRef]
  123. Gardner, A.S.; Moholdt, G.; Scambos, T.; Fahnstock, M.; Ligtenberg, S.; Van Den Broeke, M.; Nilsson, J. Increased West Antarctic and unchanged East Antarctic ice discharge over the last 7 years. Cryosphere 2018, 12, 521–547. [Google Scholar] [CrossRef] [Green Version]
  124. Luthcke, S.B.; Sabaka, T.; Loomis, B.; Arendt, A.; McCarthy, J.; Camp, J. Antarctica, Greenland and Gulf of Alaska land-ice evolution from an iterated GRACE global mascon solution. J. Glaciol. 2013, 59, 613–631. [Google Scholar] [CrossRef]
  125. Rignot, E.; Mouginot, J.; Morlighem, M.; Seroussi, H.; Scheuchl, B. Widespread, rapid grounding line retreat of Pine Island, Thwaites, Smith, and Kohler glaciers, West Antarctica, from 1992 to 2011. Geophys. Res. Lett. 2014, 41, 3502–3509. [Google Scholar] [CrossRef] [Green Version]
  126. Pritchard, H.; Ligtenberg, S.R.; Fricker, H.A.; Vaughan, D.G.; van den Broeke, M.R.; Padman, L. Antarctic ice-sheet loss driven by basal melting of ice shelves. Nature 2012, 484, 502–505. [Google Scholar] [CrossRef]
  127. Turner, J.; Orr, A.; Gudmundsson, G.H.; Jenkins, A.; Bingham, R.G.; Hillenbrand, C.D.; Bracegirdle, T.J. Atmosphere-ocean-ice interactions in the Amundsen Sea embayment, West Antarctica. Rev. Geophys. 2017, 55, 235–276. [Google Scholar] [CrossRef] [Green Version]
  128. Shepherd, A.; Ivins, E.R.; Barletta, V.R.; Bentley, M.J.; Bettadpur, S.; Briggs, K.H.; Bromwich, D.H.; Forsberg, R.; Galin, N.; Horwath, M. A reconciled estimate of ice-sheet mass balance. Science 2012, 338, 1183–1189. [Google Scholar] [CrossRef] [Green Version]
  129. Shen, Q.; Wang, H.; Shum, C.; Jiang, L.; Hsu, H.T.; Dong, J. Recent high-resolution Antarctic ice velocity maps reveal increased mass loss in Wilkes Land, East Antarctica. Sci. Rep. 2018, 8, 4477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. McMillan, M.; Shepherd, A.; Sundal, A.; Briggs, K.; Muir, A.; Ridout, A.; Hogg, A.; Wingham, D. Increased ice losses from Antarctica detected by CryoSat-2. Geophys. Res. Lett. 2014, 41, 3899–3905. [Google Scholar] [CrossRef]
  131. Li, X.; Rignot, E.; Mouginot, J.; Scheuchl, B. Ice flow dynamics and mass loss of Totten Glacier, East Antarctica, from 1989 to 2015. Geophys. Res. Lett. 2016, 43, 6366–6373. [Google Scholar] [CrossRef]
  132. Velicogna, I.; Sutterley, T.C.; Van Den Broeke, M.R. Regional acceleration in ice mass loss from Greenland and Antarctica using GRACE time-variable gravity data. Geophys. Res. Lett. 2014, 41, 8130–8137. [Google Scholar] [CrossRef] [Green Version]
  133. Meredith, M.P.; Brandon, M.A. Oceanography and Sea Ice in the Southern Ocean; John Wiley & Sons, Ltd.: Chichester, UK, 2017. [Google Scholar]
  134. Shepherd, A.; Fricker, H.A.; Farrell, S.L. Trends and connections across the Antarctic cryosphere. Nature 2018, 558, 223–232. [Google Scholar] [CrossRef]
  135. Ozsoy-Cicek, B.; Ackley, S.; Xie, H.; Yi, D.; Zwally, J. Sea ice thickness retrieval algorithms based on in situ surface elevation and thickness values for application to altimetry. J. Geophys. Res. Ocean. 2013, 118, 3807–3822. [Google Scholar] [CrossRef]
  136. Kern, S.; Ozsoy-Çiçek, B.; Worby, A.P. Antarctic sea-ice thickness retrieval from ICESat: Inter-comparison of different approaches. Remote Sens. 2016, 8, 538. [Google Scholar] [CrossRef] [Green Version]
  137. Zwally, H.J.; Parkinson, C.L.; Comiso, J.C. Variability of Antarctic Sea ice: And changes in carbon dioxide. Science 1983, 220, 1005–1012. [Google Scholar] [CrossRef]
  138. Simmonds, I.; Li, M. Trends and variability in polar sea ice, global atmospheric circulations, and baroclinicity. Ann. N. Y. Acad. Sci. 2021, 1504, 167–186. [Google Scholar] [CrossRef] [PubMed]
  139. Turner, J.; Guarino, M.V.; Arnatt, J.; Jena, B.; Marshall, G.J.; Phillips, T.; Bajish, C.; Clem, K.; Wang, Z.; Andersson, T. Recent decrease of summer sea ice in the Weddell Sea, Antarctica. Geophys. Res. Lett. 2020, 47, e2020GL087127. [Google Scholar] [CrossRef]
  140. Serreze, M.C.; Meier, W.N. The Arctic’s sea ice cover: Trends, variability, predictability, and comparisons to the Antarctic. Ann. N. Y. Acad. Sci. 2019, 1436, 36–53. [Google Scholar] [CrossRef] [PubMed]
  141. Zwally, H.J.; Yi, D.; Kwok, R.; Zhao, Y. ICESat measurements of sea ice freeboard and estimates of sea ice thickness in the Weddell Sea. J. Geophys. Res. Ocean. 2008, 113, C02S15. [Google Scholar] [CrossRef] [Green Version]
  142. Kurtz, N.T.; Markus, T. Satellite observations of Antarctic sea ice thickness and volume. J. Geophys. Res. Ocean. 2012, 117, C08025. [Google Scholar] [CrossRef] [Green Version]
  143. Gloersen, P.; Campbell, W.J.; Cavalieri, D.J.; Comiso, J.C.; Parkinson, C.L.; Zwally, H.J. Satellite passive microwave observations and analysis of Arctic and Antarctic sea ice, 1978–1987. Ann. Glaciol. 1993, 17, 149–154. [Google Scholar] [CrossRef] [Green Version]
  144. Heil, P.; Fowler, C.; Lake, S. Antarctic sea-ice velocity as derived from SSM/I imagery. Ann. Glaciol. 2006, 44, 361–366. [Google Scholar] [CrossRef] [Green Version]
  145. Holland, P.R.; Kwok, R. Wind-driven trends in Antarctic sea-ice drift. Nat. Geosci. 2012, 5, 872–875. [Google Scholar] [CrossRef]
  146. Massom, R.A.; Stammerjohn, S.E. Antarctic sea ice change and variability–physical and ecological implications. Polar Sci. 2010, 4, 149–186. [Google Scholar] [CrossRef] [Green Version]
  147. Turner, J.; Holmes, C.; Caton Harrison, T.; Phillips, T.; Jena, B.; Reeves-Francois, T.; Fogt, R.; Thomas, E.R.; Bajish, C. Record low Antarctic sea ice cover in February 2022. Geophys. Res. Lett. 2022, 49, e2022GL098904. [Google Scholar] [CrossRef]
  148. Fogt, R.L.; Sleinkofer, A.M.; Raphael, M.N.; Handcock, M.S. A regime shift in seasonal total Antarctic sea ice extent in the twentieth century. Nat. Clim. Change 2022, 12, 54–62. [Google Scholar] [CrossRef]
  149. Wang, S.; Liu, J.; Cheng, X.; Yang, D.; Kerzenmacher, T.; Li, X.; Hu, Y.; Braesicke, P. Contribution of the deepened Amundsen sea low to the record low Antarctic sea ice extent in February 2022. Environ. Res. Lett. 2023, 18, 054002. [Google Scholar] [CrossRef]
  150. Sun, S.; Eisenman, I. Observed Antarctic sea ice expansion reproduced in a climate model after correcting biases in sea ice drift velocity. Nat. Commun. 2021, 12, 1060. [Google Scholar] [CrossRef] [PubMed]
  151. Dalziel, I.; Lawver, L. The lithospheric setting of the West Antarctic ice sheet. West Antarct. Ice Sheet Behav. Environ. 2001, 77, 29–44. [Google Scholar]
  152. Jordan, T.A.; Riley, T.R.; Siddoway, C.S. The geological history and evolution of West Antarctica. Nat. Rev. Earth Environ. 2020, 1, 117–133. [Google Scholar] [CrossRef] [Green Version]
  153. Cande, S.C.; Stock, J.M.; Müller, R.D.; Ishihara, T. Cenozoic motion between east and west Antarctica. Nature 2000, 404, 145–150. [Google Scholar] [CrossRef]
  154. Behrendt, J.C. Crustal and lithospheric structure of the West Antarctic Rift System from geophysical investigations—A review. Glob. Planet. Change 1999, 23, 25–44. [Google Scholar] [CrossRef]
  155. Behrendt, J.; LeMasurier, W.; Cooper, A.; Tessensohn, F.; Trehu, A.; Damaske, D. The West Antarctic rift system: A review of geophysical investigations. Contrib. Antarct. Res. II 1991, 53, 67–112. [Google Scholar]
  156. Joughin, I.; Smith, B.E.; Medley, B. Marine ice sheet collapse potentially under way for the Thwaites Glacier Basin, West Antarctica. Science 2014, 344, 735–738. [Google Scholar] [CrossRef]
  157. Oppenheimer, M. Global warming and the stability of the West Antarctic Ice Sheet. Nature 1998, 393, 325–332. [Google Scholar] [CrossRef]
  158. Afonso, J.C.; Fullea, J.; Yang, Y.; Connolly, J.; Jones, A. 3-D multi-observable probabilistic inversion for the compositional and thermal structure of the lithosphere and upper mantle. II: General methodology and resolution analysis. J. Geophys. Res. Solid Earth 2013, 118, 1650–1676. [Google Scholar] [CrossRef] [Green Version]
  159. Lösing, M.; Ebbing, J.; Szwillus, W. Geothermal heat flux in Antarctica: Assessing models and observations by Bayesian inversion. Front. Earth Sci. 2020, 8, 105. [Google Scholar] [CrossRef] [Green Version]
  160. Whitehouse, P.L.; Bentley, M.J.; Milne, G.A.; King, M.A.; Thomas, I.D. A new glacial isostatic adjustment model for Antarctica: Calibrated and tested using observations of relative sea-level change and present-day uplift rates. Geophys. J. Int. 2012, 190, 1464–1482. [Google Scholar] [CrossRef] [Green Version]
  161. Winberry, J.P.; Anandakrishnan, S. Crustal structure of the West Antarctic rift system and Marie Byrd Land hotspot. Geology 2004, 32, 977–980. [Google Scholar] [CrossRef]
  162. Cooper, A.K.; Barker, P.F.; Brancolini, G.; Hambrey, M. Geology and Seismic Stratigraphy of the Antarctic Margin; American Geophysical Union: Washington, DC, USA, 1995; Volume 68. [Google Scholar]
  163. Gohl, K.; Teterin, D.; Eagles, G.; Netzeband, G.; Grobys, J.; Parsiegla, N.; Schlüter, P.; Leinweber, V.T.; Larter, R.D.; Uenzelmann-Neben, G. Geophysical Survey Reveals Tectonic Structures in the Amundsen Sea Embayment. In Proceedings of the 10th International Symposium on Antarctic Earth Sciences (USGS of-2007-1047), Washington, DC, USA, 26 August–1 September 2007. Short Research Paper 047. [Google Scholar]
  164. Bingham, R.G.; Ferraccioli, F.; King, E.C.; Larter, R.D.; Pritchard, H.D.; Smith, A.M.; Vaughan, D.G. Inland thinning of West Antarctic Ice Sheet steered along subglacial rifts. Nature 2012, 487, 468–471. [Google Scholar] [CrossRef] [PubMed]
  165. Ferraccioli, F.; Jordan, T.A.; Vaughan, D.G.; Holt, J.; James, M.; Corr, H.; Blankenship, D.D.; Fairhead, J.; Diehl, T.M. New aerogeophysical survey targets the extent of the West Antarctic Rift System over Ellsworth Land. In Proceedings of the 10th International Symposium on Antarctic Earth Sciences (USGS of-2007-1047), Washington, DC, USA, 26 August–1 September 2007. Open-File Report 1047. [Google Scholar]
  166. Vaughan, D.G.; Corr, H.F.; Ferraccioli, F.; Frearson, N.; O’Hare, A.; Mach, D.; Holt, J.W.; Blankenship, D.D.; Morse, D.L.; Young, D.A. New boundary conditions for the West Antarctic ice sheet: Subglacial topography beneath Pine Island Glacier. Geophys. Res. Lett. 2006, 33, L09501. [Google Scholar] [CrossRef] [Green Version]
  167. An, M.; Wiens, D.A.; Zhao, Y.; Feng, M.; Nyblade, A.A.; Kanao, M.; Li, Y.; Maggi, A.; Lévêque, J.J. S-velocity model and inferred Moho topography beneath the Antarctic Plate from Rayleigh waves. J. Geophys. Res. Solid Earth 2015, 120, 359–383. [Google Scholar] [CrossRef]
  168. Dylan Mikesell, T.; Mordret, A.; Xu, Z.; Frank, W.B. Crustal structure across the West Antarctic rift system from multicomponent ambient noise surface wave tomography. Seismol. Soc. Am. 2022, 93, 2201–2217. [Google Scholar] [CrossRef]
  169. Lloyd, A.J.; Wiens, D.A.; Nyblade, A.A.; Anandakrishnan, S.; Aster, R.C.; Huerta, A.D.; Wilson, T.J.; Dalziel, I.W.; Shore, P.J.; Zhao, D. A seismic transect across West Antarctica: Evidence for mantle thermal anomalies beneath the Bentley Subglacial Trench and the Marie Byrd Land Dome. J. Geophys. Res. Solid Earth 2015, 120, 8439–8460. [Google Scholar] [CrossRef]
  170. Golynsky, A.; Ivanov, S.; Kazankov, A.J.; Jokat, W.; Masolov, V.; Von Frese, R.; Group, A.W. New continental margin magnetic anomalies of East Antarctica. Tectonophysics 2013, 585, 172–184. [Google Scholar] [CrossRef]
  171. Harley, S.L.; Fitzsimons, I.C.; Zhao, Y. Antarctica and supercontinent evolution: Historical perspectives, recent advances and unresolved issues. Geol. Soc. Lond. Spec. Publ. 2013, 383, 1–34. [Google Scholar] [CrossRef] [Green Version]
  172. Jordan, T.A.; Ferraccioli, F.; Forsberg, R. An embayment in the East Antarctic basement constrains the shape of the Rodinian continental margin. Commun. Earth Environ. 2022, 3, 52. [Google Scholar] [CrossRef]
  173. Boger, S.D. Antarctica—Before and after Gondwana. Gondwana Res. 2011, 19, 335–371. [Google Scholar] [CrossRef]
  174. Riedel, S.; Jokat, W.; Steinhage, D. Mapping tectonic provinces with airborne gravity and radar data in Dronning Maud Land, East Antarctica. Geophys. J. Int. 2012, 189, 414–427. [Google Scholar] [CrossRef] [Green Version]
  175. Ruppel, A.; Jacobs, J.; Eagles, G.; Läufer, A.; Jokat, W. New geophysical data from a key region in East Antarctica: Estimates for the spatial extent of the Tonian Oceanic Arc Super Terrane (TOAST). Gondwana Res. 2018, 59, 97–107. [Google Scholar] [CrossRef]
  176. Ferraccioli, F.; PC, J.; ML, C.; PT, L.; TR, R. Tectonic and magmatic patterns in the Jutulstraumen rift (?) region, East Antarctica, as imaged by high-resolution aeromagnetic data. Earth Planets Space 2005, 57, 767–780. [Google Scholar] [CrossRef] [Green Version]
  177. Jokat, W.; Boebel, T.; König, M.; Meyer, U. Timing and geometry of early Gondwana breakup. J. Geophys. Res. Solid Earth 2003, 108(B9), 2428. [Google Scholar] [CrossRef]
  178. König, M.; Jokat, W. The mesozoic breakup of the weddell sea. J. Geophys. Res. Solid Earth 2006, 111, B12102. [Google Scholar] [CrossRef]
  179. Riedel, S.; Jokat, W. A Compilation of New Airborne Magnetic and Gravity Data across Dronning Maud Land, Antarctica. In Proceedings of the 10th International Symposium on Antarctic Earth Sciences (USGS of-2007-1047), Washington, DC, USA, 26–31 August 2007. Extended Abstract 149. [Google Scholar]
  180. Ferraccioli, F.; Jones, P.; Vaughan, A.; Leat, P. New aerogeophysical view of the Antarctic Peninsula: More pieces, less puzzle. Geophys. Res. Lett. 2006, 33, L05310. [Google Scholar] [CrossRef]
  181. Ghidella, M.; Zambrano, O.; Ferraccioli, F.; Lirio, J.; Zakrajsek, A.; Ferris, J.; Jordan, T. Analysis of James Ross Island volcanic complex and sedimentary basin based on high-resolution aeromagnetic data. Tectonophysics 2013, 585, 90–101. [Google Scholar] [CrossRef]
  182. Jordan, T.; Neale, R.; Leat, P.; Vaughan, A.; Flowerdew, M.; Riley, T.; Whitehouse, M.J.; Ferraccioli, F. Structure and evolution of Cenozoic arc magmatism on the Antarctic Peninsula: A high resolution aeromagnetic perspective. Geophys. J. Int. 2014, 198, 1758–1774. [Google Scholar] [CrossRef] [Green Version]
  183. Smellie, J.L.; Martin, A.P.; Panter, K.S.; Kyle, P.R.; Geyer, A. Magmatism in Antarctica and its relation to Zealandia. N. Z. J. Geol. Geophys. 2020, 63, 578–588. [Google Scholar] [CrossRef]
  184. Quartini, E.; Blankenship, D.D.; Young, D.A. Chapter 7.5 Active subglacial volcanism in West Antarctica. Geol. Soc. Lond. Mem. 2021, 55, 785–803. [Google Scholar] [CrossRef]
  185. Antoniades, D.; Giralt, S.; Geyer, A.; Álvarez-Valero, A.M.; Pla-Rabes, S.; Granados, I.; Liu, E.J.; Toro, M.; Smellie, J.L.; Oliva, M. The timing and widespread effects of the largest Holocene volcanic eruption in Antarctica. Sci. Rep. 2018, 8, 17279. [Google Scholar] [CrossRef] [Green Version]
  186. Ellis, M.; King, G. Structural control of flank volcanism in continental rifts. Science 1991, 254, 839–842. [Google Scholar] [CrossRef]
  187. Behrendt, J.C. Mobile magma under Antarctic ice. Nat. Geosci. 2013, 6, 990–991. [Google Scholar] [CrossRef]
  188. Lough, A.C.; Wiens, D.A.; Grace Barcheck, C.; Anandakrishnan, S.; Aster, R.C.; Blankenship, D.D.; Huerta, A.D.; Nyblade, A.; Young, D.A.; Wilson, T.J. Seismic detection of an active subglacial magmatic complex in Marie Byrd Land, Antarctica. Nat. Geosci. 2013, 6, 1031–1035. [Google Scholar] [CrossRef]
  189. Blankenship, D.D.; Bell, R.E.; Hodge, S.M.; Brozena, J.M.; Behrendt, J.C.; Finn, C.A. Active volcanism beneath the West Antarctic ice sheet and implications for ice-sheet stability. Nature 1993, 361, 526–529. [Google Scholar] [CrossRef]
  190. Behrendt, J.C.; Wold, R.J. Depth to magnetic ‘basement’in west Antarctica. J. Geophys. Res. 1963, 68, 1145–1153. [Google Scholar] [CrossRef]
  191. Van Wyk de Vries, M.; Bingham, R.G.; Hein, A.S. A new volcanic province: An inventory of subglacial volcanoes in West Antarctica. Geol. Soc. Lond. Spec. Publ. 2018, 461, 231–248. [Google Scholar] [CrossRef]
  192. Maslanyj, M.; Storey, B. Regional aeromagnetic anomalies in Ellsworth Land: Crustal structure and Mesozoic microplate boundaries within West Antarctica. Tectonics 1990, 9, 1515–1532. [Google Scholar] [CrossRef]
  193. Hill, G.; Wannamaker, P.; Maris, V.; Stodt, J.; Kordy, M.; Unsworth, M.; Bedrosian, P.; Wallin, E.; Uhlmann, D.; Ogawa, Y. Trans-crustal structural control of CO2-rich extensional magmatic systems revealed at Mount Erebus Antarctica. Nat. Commun. 2022, 13, 2989. [Google Scholar] [CrossRef]
  194. Gupta, S.; Zhao, D.; Rai, S. Seismic imaging of the upper mantle under the Erebus hotspot in Antarctica. Gondwana Res. 2009, 16, 109–118. [Google Scholar] [CrossRef]
  195. Greenwood, S.L.; Clason, C.C.; Helanow, C.; Margold, M. Theoretical, contemporary observational and palaeo-perspectives on ice sheet hydrology: Processes and products. Earth-Sci. Rev. 2016, 155, 1–27. [Google Scholar] [CrossRef] [Green Version]
  196. Engelhardt, H.; Kamb, B. Basal hydraulic system of a West Antarctic ice stream: Constraints from borehole observations. J. Glaciol. 1997, 43, 207–230. [Google Scholar] [CrossRef] [Green Version]
  197. Siegert, M.J.; Kennicutt, M.C.; Bindschadler, R.A. Antarctic Subglacial Aquatic Environments; John Wiley & Sons: Hoboken, NJ, USA, 2013. [Google Scholar]
  198. Yan, S.; Blankenship, D.D.; Greenbaum, J.S.; Young, D.A.; Li, L.; Rutishauser, A.; Guo, J.; Roberts, J.L.; van Ommen, T.D.; Siegert, M.J. A newly discovered subglacial lake in East Antarctica likely hosts a valuable sedimentary record of ice and climate change. Geology 2022, 50, 949–953. [Google Scholar] [CrossRef]
  199. Pattyn, F.; Carter, S.P.; Thoma, M. Advances in modelling subglacial lakes and their interaction with the Antarctic ice sheet. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20140296. [Google Scholar] [CrossRef] [PubMed]
  200. Fricker, H.A.; Scambos, T.; Carter, S.; Davis, C.; Haran, T.; Joughin, I. Synthesizing multiple remote-sensing techniques for subglacial hydrologic mapping: Application to a lake system beneath MacAyeal Ice Stream, West Antarctica. J. Glaciol. 2010, 56, 187–199. [Google Scholar] [CrossRef] [Green Version]
  201. Wingham, D.J.; Siegert, M.J.; Shepherd, A.; Muir, A.S. Rapid discharge connects Antarctic subglacial lakes. Nature 2006, 440, 1033–1036. [Google Scholar] [CrossRef]
  202. Stearns, L.A.; Smith, B.E.; Hamilton, G.S. Increased flow speed on a large East Antarctic outlet glacier caused by subglacial floods. Nat. Geosci. 2008, 1, 827–831. [Google Scholar] [CrossRef]
  203. Siegfried, M.R.; Fricker, H.A. Thirteen years of subglacial lake activity in Antarctica from multi-mission satellite altimetry. Ann. Glaciol. 2018, 59, 42–55. [Google Scholar] [CrossRef] [Green Version]
  204. Hao, T.; Jing, L.; Liu, J.; Wang, D.; Feng, T.; Zhao, A.; Li, R. Automatic Detection of Subglacial Water Bodies in the AGAP Region, East Antarctica, Based on Short-Time Fourier Transform. Remote Sens. 2023, 15, 363. [Google Scholar] [CrossRef]
  205. Sun, Y.; Li, B.; Fan, X.; Li, Y.; Li, G.; Yu, H.; Li, H.; Wang, D.; Zhang, N.; Gong, D. Brief communication: New sonde to unravel the mystery of polar subglacial lakes. Cryosphere 2023, 17, 1089–1095. [Google Scholar] [CrossRef]
  206. Livingstone, S.J.; Li, Y.; Rutishauser, A.; Sanderson, R.J.; Winter, K.; Mikucki, J.A.; Björnsson, H.; Bowling, J.S.; Chu, W.; Dow, C.F. Subglacial lakes and their changing role in a warming climate. Nat. Rev. Earth Environ. 2022, 3, 106–124. [Google Scholar] [CrossRef]
  207. Gaidos, E.; Lanoil, B.; Thorsteinsson, T.; Graham, A.; Skidmore, M.; Han, S.-K.; Rust, T.; Popp, B. A viable microbial community in a subglacial volcanic crater lake, Iceland. Astrobiology 2004, 4, 327–344. [Google Scholar] [CrossRef] [PubMed]
  208. Bulat, S.A. Microbiology of the subglacial Lake Vostok: First results of borehole-frozen lake water analysis and prospects for searching for lake inhabitants. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2016, 374, 20140292. [Google Scholar] [CrossRef] [Green Version]
  209. Key, K.; Siegfried, M.R. The feasibility of imaging subglacial hydrology beneath ice streams with ground-based electromagnetics. J. Glaciol. 2017, 63, 755–771. [Google Scholar] [CrossRef] [Green Version]
  210. Kulessa, B.; Hubbard, B.; Brown, G.H. Time-lapse imaging of subglacial drainage conditions using three-dimensional inversion of borehole electrical resistivity data. J. Glaciol. 2006, 52, 49–57. [Google Scholar] [CrossRef] [Green Version]
  211. Kulessa, B.; Murray, T.; Rippin, D. Active seismoelectric exploration of glaciers. Geophys. Res. Lett. 2006, 33. [Google Scholar] [CrossRef]
  212. Xiao, E.; Jiang, F.; Guo, J.; Latif, K.; Fu, L.; Sun, B. 3D Interpretation of a Broadband Magnetotelluric Data Set Collected in the South of the Chinese Zhongshan Station at Prydz Bay, East Antarctica. Remote Sens. 2022, 14, 496. [Google Scholar] [CrossRef]
  213. Siegfried, M.; Fricker, H. Illuminating active subglacial lake processes with ICESat-2 laser altimetry. Geophys. Res. Lett. 2021, 48, e2020GL091089. [Google Scholar] [CrossRef]
  214. Ashmore, D.W.; Bingham, R.G. Antarctic subglacial hydrology: Current knowledge and future challenges. Antarct. Sci. 2014, 26, 758–773. [Google Scholar] [CrossRef] [Green Version]
  215. Peters, L.; Anandakrishnan, S.; Holland, C.; Horgan, H.; Blankenship, D.; Voigt, D. Seismic detection of a subglacial lake near the South Pole, Antarctica. Geophys. Res. Lett. 2008, 35, L23501. [Google Scholar] [CrossRef]
  216. Robin, G.d.Q.; Drewry, D.; Meldrum, D. International studies of ice sheet and bedrock. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1977, 279, 185–196. [Google Scholar]
  217. Oswald, G.; Robin, G.d.Q. Lakes beneath the Antarctic ice sheet. Nature 1973, 245, 251–254. [Google Scholar] [CrossRef]
  218. McMillan, M.; Corr, H.; Shepherd, A.; Ridout, A.; Laxon, S.; Cullen, R. Three-dimensional mapping by CryoSat-2 of subglacial lake volume changes. Geophys. Res. Lett. 2013, 40, 4321–4327. [Google Scholar] [CrossRef] [Green Version]
  219. Livingstone, S.; Clark, C.; Woodward, J.; Kingslake, J. Potential subglacial lake locations and meltwater drainage pathways beneath the Antarctic and Greenland ice sheets. Cryosphere 2013, 7, 1721–1740. [Google Scholar] [CrossRef] [Green Version]
  220. Christianson, K.; Jacobel, R.W.; Horgan, H.J.; Anandakrishnan, S.; Alley, R.B. Subglacial Lake Whillans—Ice-penetrating radar and GPS observations of a shallow active reservoir beneath a West Antarctic ice stream. Earth Planet. Sci. Lett. 2012, 331, 237–245. [Google Scholar] [CrossRef]
  221. Siegert, M.J.; Ross, N.; Corr, H.; Smith, B.; Jordan, T.; Bingham, R.G.; Ferraccioli, F.; Rippin, D.M.; Le Brocq, A. Boundary conditions of an active West Antarctic subglacial lake: Implications for storage of water beneath the ice sheet. Cryosphere 2014, 8, 15–24. [Google Scholar] [CrossRef] [Green Version]
  222. Dirscherl, M.; Dietz, A.J.; Kneisel, C.; Kuenzer, C. Automated mapping of Antarctic supraglacial lakes using a machine learning approach. Remote Sens. 2020, 12, 1203. [Google Scholar] [CrossRef] [Green Version]
  223. Murray, T.; Corr, H.; Forieri, A.; Smith, A. Contrasts in hydrology between regions of basal deformation and sliding beneath Rutford Ice Stream, West Antarctica, mapped using radar and seismic data. Geophys. Res. Lett. 2008, 35, L12504. [Google Scholar] [CrossRef] [Green Version]
  224. Smith, A.M.; Jordan, T.A.; Ferraccioli, F.; Bingham, R.G. Influence of subglacial conditions on ice stream dynamics: Seismic and potential field data from Pine Island Glacier, West Antarctica. J. Geophys. Res. Solid Earth 2013, 118, 1471–1482. [Google Scholar] [CrossRef] [Green Version]
  225. Horgan, H.J.; Anandakrishnan, S.; Jacobel, R.W.; Christianson, K.; Alley, R.B.; Heeszel, D.S.; Picotti, S.; Walter, J.I. Subglacial Lake Whillans—Seismic observations of a shallow active reservoir beneath a West Antarctic ice stream. Earth Planet. Sci. Lett. 2012, 331, 201–209. [Google Scholar] [CrossRef]
  226. Hofstede, C.; Eisen, O.; Diez, A.; Jansen, D.; Kristoffersen, Y.; Lambrecht, A.; Mayer, C. Investigating englacial reflections with vibro-and explosive-seismic surveys at Halvfarryggen ice dome, Antarctica. Ann. Glaciol. 2013, 54, 189–200. [Google Scholar] [CrossRef]
  227. Eisen, O.; Hofstede, C.; Diez, A.; Kristoffersen, Y.; Lambrecht, A.; Mayer, C.; Blenkner, R.; Hilmarsson, S. On-ice vibroseis and snowstreamer systems for geoscientific research. Polar Sci. 2015, 9, 51–65. [Google Scholar] [CrossRef] [Green Version]
  228. Eisen, O.; Hofstede, C.; Miller, H.; Kristoffersen, Y.; Blenkner, R.; Lambrecht, A.; Mayer, C. A new approach for exploring ice sheets and sub-ice geology. EOS Trans. Am. Geophys. Union 2010, 91, 429–430. [Google Scholar] [CrossRef] [Green Version]
  229. Dugan, H.; Doran, P.; Tulaczyk, S.; Mikucki, J.; Arcone, S.A.; Auken, E.; Schamper, C.; Virginia, R. Subsurface imaging reveals a confined aquifer beneath an ice-sealed Antarctic lake. Geophys. Res. Lett. 2015, 42, 96–103. [Google Scholar] [CrossRef]
  230. Mikucki, J.A.; Auken, E.; Tulaczyk, S.; Virginia, R.; Schamper, C.; Sørensen, K.; Doran, P.; Dugan, H.; Foley, N. Deep groundwater and potential subsurface habitats beneath an Antarctic dry valley. Nat. Commun. 2015, 6, 6831. [Google Scholar] [CrossRef] [Green Version]
  231. Foley, N.; Tulaczyk, S.; Auken, E.; Schamper, C.; Dugan, H.; Mikucki, J.; Virginia, R.; Doran, P. Helicopter-borne transient electromagnetics in high-latitude environments: An application in the McMurdo Dry Valleys, AntarcticaAEM resistivity in the Dry Valleys. Geophysics 2016, 81, WA87–WA99. [Google Scholar] [CrossRef] [Green Version]
  232. Zhong, D.; Li, Y.; Huang, Y.; Hong, X.; Li, J.; Jin, R. Molecular mechanisms of exercise on cancer: A bibliometrics study and visualization analysis via CiteSpace. Front. Mol. Biosci. 2022, 8, 1360. [Google Scholar] [CrossRef]
Figure 1. Bibliometric lines of research for Phases 1 (light blue) and 2 (dark blue).
Figure 1. Bibliometric lines of research for Phases 1 (light blue) and 2 (dark blue).
Remotesensing 15 03928 g001
Figure 2. Time evolution of annual publications and the accumulated literature from 1982 to 2022.
Figure 2. Time evolution of annual publications and the accumulated literature from 1982 to 2022.
Remotesensing 15 03928 g002
Figure 3. Distribution of the top 15 countries with publications in Antarctic geophysical research.
Figure 3. Distribution of the top 15 countries with publications in Antarctic geophysical research.
Remotesensing 15 03928 g003
Figure 4. Distribution of major countries/institutions for Antarctic geophysical research.
Figure 4. Distribution of major countries/institutions for Antarctic geophysical research.
Remotesensing 15 03928 g004
Figure 5. The cooperation network of the most productive countries, in terms of Antarctic geophysical research.
Figure 5. The cooperation network of the most productive countries, in terms of Antarctic geophysical research.
Remotesensing 15 03928 g005
Figure 6. Geophysical observation methods applied to the Antarctic ice sheet (modified from [118]).
Figure 6. Geophysical observation methods applied to the Antarctic ice sheet (modified from [118]).
Remotesensing 15 03928 g006
Figure 7. (a) The minimum daily sea ice extent (SIE) (106 km2) for the Southern Ocean along with the daily maximum and annual mean for 1979–2022 [147]; (b) Antarctic sea ice drift velocity trend during 1992–2015 [150]; (c) Maps of the average sea ice thickness data for each season from 2003 to 2008 [142].
Figure 7. (a) The minimum daily sea ice extent (SIE) (106 km2) for the Southern Ocean along with the daily maximum and annual mean for 1979–2022 [147]; (b) Antarctic sea ice drift velocity trend during 1992–2015 [150]; (c) Maps of the average sea ice thickness data for each season from 2003 to 2008 [142].
Remotesensing 15 03928 g007
Figure 8. (a) Main geophysical methods for detecting subglacial hydrology [206]; (b) ICESat-2 altimetry coverage of active subglacial lakes in Antarctica [213]; (c) Radargrams of the Antarctic subglacial hydrological environment [214]; (d) The subglacial lake was imaged using seismic data [215].
Figure 8. (a) Main geophysical methods for detecting subglacial hydrology [206]; (b) ICESat-2 altimetry coverage of active subglacial lakes in Antarctica [213]; (c) Radargrams of the Antarctic subglacial hydrological environment [214]; (d) The subglacial lake was imaged using seismic data [215].
Remotesensing 15 03928 g008
Figure 9. (a) Antarctic Inventory of subglacial lakes, Red circles represent stable lakes and blue triangles represent active lakes [206]; (b) subglacial water system of the Siple Coast in West Antarctica Major water-flow pathways connect subglacial lakes (in black) within a large-scale distributed system of subglacial till layers and linked cavities [107].
Figure 9. (a) Antarctic Inventory of subglacial lakes, Red circles represent stable lakes and blue triangles represent active lakes [206]; (b) subglacial water system of the Siple Coast in West Antarctica Major water-flow pathways connect subglacial lakes (in black) within a large-scale distributed system of subglacial till layers and linked cavities [107].
Remotesensing 15 03928 g009
Table 1. The top 24 keywords with the strongest citation bursts, from 1991 to 2022.
Table 1. The top 24 keywords with the strongest citation bursts, from 1991 to 2022.
RankKeywordsCitation BurstsYear
1Margin10.511991–2007
Antarctic Peninsula8.351991–2007
Anomaly5.171993–2017
Marine sediments5.311993–2002
Weddell Sea7.201994–2007
Gondwanaland5.061994–2002
Reconstruction4.971994–2007
Plate7.961994–2012
Boundary9.551995–2007
2Seismic waves6.311995–2007
Magnetic anomalies5.941996–2002
Fracture zone7.311998–2012
Magmatism6.171999–2012
Plate tectonics4.951999–2007
Magnetic field6.882003–2012
Crustal5.272003–2012
Prydz Bay5.182003–2012
3Pacific margin6.532008–2017
West Antarctica7.952009–2017
Pine Island Glacier5.752013–2022
Ice shelves5.142015–2022
Amundsen Sea Embayment6.442018–2022
Antarctic glaciology5.302018–2022
Rayleigh wave5.572019–2022
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Zou, C.; Peng, C.; Lan, X.; Zhang, H. Geophysics in Antarctic Research: A Bibliometric Analysis. Remote Sens. 2023, 15, 3928. https://doi.org/10.3390/rs15163928

AMA Style

Zhang Y, Zou C, Peng C, Lan X, Zhang H. Geophysics in Antarctic Research: A Bibliometric Analysis. Remote Sensing. 2023; 15(16):3928. https://doi.org/10.3390/rs15163928

Chicago/Turabian Style

Zhang, Yuanyuan, Changchun Zou, Cheng Peng, Xixi Lan, and Hongjie Zhang. 2023. "Geophysics in Antarctic Research: A Bibliometric Analysis" Remote Sensing 15, no. 16: 3928. https://doi.org/10.3390/rs15163928

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