Next Article in Journal
Validation of Analytical Solutions for Predicting Drilled Pile Behaviour under Bi-Directional Static Load Tests
Next Article in Special Issue
Drivers of Last Millennium Antarctic Climate Evolution in an Ensemble of Community Earth System Model Simulations
Previous Article in Journal
Collection, Standardization and Attribution of Robust Disaster Event Information—A Demonstrator of a National Event-Based Loss and Damage Database in Austria
Previous Article in Special Issue
Ice Core Chronologies from the Antarctic Peninsula: The Palmer, Jurassic, and Rendezvous Age-Scales
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Variety and Distribution of Diatom-Based Sea Ice Proxies in Antarctic Marine Sediments of the Past 2000 Years

1
British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 0ET, UK
2
School of Earth and Environmental Sciences, Cardiff University, Main Building, Park Place, Cardiff CF10 3AT, UK
*
Author to whom correspondence should be addressed.
Geosciences 2022, 12(8), 282; https://doi.org/10.3390/geosciences12080282
Submission received: 13 June 2022 / Revised: 1 July 2022 / Accepted: 8 July 2022 / Published: 22 July 2022

Abstract

:
Antarctic sea ice is an essential component of the global climate system. Reconstructions of Antarctic sea ice from marine sediment cores are a vital resource to improve the representation of Antarctic sea ice in climate models and to better understand natural variability in sea ice over centennial and sub-centennial timescales. The Thomas et al. (2019) review of Antarctic sea ice reconstructions from ice and marine cores highlighted the prominence of diatom-based proxies in this research. Here, focusing solely on the diatom-based proxy records in marine sediments, we review the composition of proxies, their advantages and limitations, as well as the spatial and temporal cover of the records over the past 2 ka in order to assess the scope for future assimilation and standardization. The archive comprises 112 records from 68 marine cores, with proxies based on more than 30 different combinations of diatom taxa as well as the relatively new, highly branched isoprenoid (HBI) biomarkers.

1. Introduction

The annual growth and retreat of sea ice is one of the most dramatic and dynamic seasonal patterns on Earth today. In the Southern Ocean, the seasonal expansion of sea ice effectively doubles the surface area of Antarctica and exerts a powerful influence over global climatic, oceanographic, and biological systems. Satellite observations provide detailed records of sea ice cover over the past 40 years and highlight the inter-annual variability in sea ice distribution [1]. Since 1979, whilst the Arctic region has seen a pronounced decline in sea ice cover, the total area of Antarctic sea ice has gradually increased, interrupted only with a precipitous decline between 2014 and 2017, resulting in an overall modest expansion [2]. The trend in the total area of Antarctic sea ice masks substantial inter-annual variation and strong regional divergence (e.g., marked reduction (increase) in the Bellingshausen (Ross) Sea [3]), contradicting the decreasing trajectory of sea ice cover predicted by the majority of climate simulations [2,4,5,6]. An improved understanding of the influence and response of Antarctic sea ice to climate change is of primary importance in order to improve the accuracy and predictive competence of climate models tasked with simulating future global change [7].
To resolve pre-satellite trends in sea ice cover and put recent changes into the context of natural variability, longer archives are needed to resolve decadal-to-centennial variations and establish the pre-industrial state of Antarctic sea ice. Thomas et al. [8] show that the majority of Antarctic sea ice reconstructions for the past 2 ka from marine sediment cores are derived from “diatom proxies”. Diatoms inhabit a diverse range of habitats within the Southern Ocean and its sea ice; individual species have unique requirements for light, temperature, salinity, and nutrient availability, and may bloom at different points in the year (Appendix A). The ecological preferences of these diatoms allow their fossils to be used as proxies for environmental changes. Biomarkers from a range of primary producers preserved in the sedimentary record have been used to investigate past environmental changes [9]. Over the past 10 years, biomarkers specific to diatoms have been developed as a sea ice proxy. The di-unsaturated C25 highly branched isoprenoid (HBI) alkene (known as diene or HBI-II) and the tri-unsaturated C25 HBI alkene (known as triene or HBI-III) are produced by a small number of Southern Ocean diatom species and are preserved in marine sediments [10]. The source-specific nature of these biomarkers allow for their use as effective proxies.
Diatom proxies are typically considered more robust than other proxies as they provide a direct link between the sea ice environment and the sediment archive. However, there is great variety in the composition and approaches grouped within the “diatom proxies”. This diversity reflects the wide range of oceanographic settings of the core sites and the sensitivity of diatom assemblages to environmental gradients. Providing a detailed review of the proxies and the most commonly employed taxa, followed by mapping out the distribution and resolution of existing records, will readily show areas where regional syntheses may be possible; highlight gaps in the existing coverage of sea ice records; and provide valuable information on the suitability, application, and comparison of diatom-based proxies for different regions of Antarctica.
The structure of the paper is as follows:
In the first section, we collate all the published marine sediment records spanning all, or part, of the past 2000 years with sea ice reconstructions derived from diatom-based proxies. We review the composition and variety of proxies used to reconstruct sea ice and map their distribution.
We then consider the advantages and limitations of the proxies and discuss the potential for improving consistency/standardization between records and the scope for regional syntheses.

2. Results

In order to maximize the number of records included in this study, broad selection criteria were adopted: (1) include at least two data points to infer sea ice conditions during some or all of the past 2 ka, (2) a dated horizon to establish the surface/near-surface age (e.g., lead-210; radiocarbon) and at least one other dated horizon to determine the depth of the 2 ka interval. Records that were published up to and including 2021 were all considered, but duplicated records with matching age models and resolution were excluded. These criteria were met by 112 proxy records from 68 marine core sites, comprising 8 sites from the deep Southern Ocean and 60 from the Antarctic continental shelf (Figure 1 and Table 1 [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68]). Duplicate and lower resolution records in other publications were also excluded. All dates were based on the published age models and presented in the same units as in the original publications. Radiocarbon dates were reported as either corrected radiocarbon (14C ka BP) or calendar (ka BP) ages. The mean offset between calendar and 14C ages for the 2 ka time interval is +/−0.092 ka [69].

2.1. Proxy Records

The 112 records comprise more than 30 individually described diatom-based sea ice proxies across 68 marine sites. Of the 112 records, 30 are based on a single species, 16 on two species, 10 on transfer functions, 14 on diatom-specific HBI biomarkers and the remaining 42 are based on groups of more than two species.
Fragilariopsis curta is included in 77 of the 112 proxy records and is the most common diatom taxa used in reconstructions of Antarctic sea ice (Table 2). Of these 77 records, 23 are based solely on F. curta% and 54 comprising various groups or ratios that include F. curta (Table 3). F. cylindrus is the second most common taxa and is included in the composition of 54 proxies. Thalassiosira antarctica, Fragilariopsis sublinearis, F. obliquecostata, F. vanheurckii, F. rhombica, F. kerguelensis, Actinocyclus actinochilus, Porosira glacialis, F. ritscheri and F. separanda are each included in at least 10 of the 112 proxy records (Table 2). The habitat preferences of these species are available in Appendix A.
The selection of taxa comprising the various groups is mostly based on statistical and/or ecological associations, but seasonal timing within varved sediments is also used to determine the composition of some groups [32,38]. Statistical techniques used to define groups include the modern analogue technique (MAT), the Imbrie Kipp method (IKM), the generalized additive model (GAM) method, principal component analysis (PCA), and cluster analysis [26,27,38,61,62,63,70]. Most proxies are used to infer qualitative changes in seasonal sea ice cover, whilst the MAT and GAM methods produce a quantitative estimate of sea ice conditions (e.g., months per year of sea ice cover or sea ice concentrations). In nearshore locations and sites close to ice shelves, some taxa and/or taxa groups are associated with more niche sea ice conditions or types e.g., the fast ice index [22,23,26].
In order to review the prevalence and distribution of proxies used to reconstruct sea ice in the marine 2 ka records considered here, the proxies were divided into eight categories according to their composition. As F. curta was the most commonly used diatom (Table 2), the diatom groups used as sea ice proxies were classified based on whether they included F. curta or not (Table 3).
Categories 1 to 4 are based on the relative abundances of 1–2 diatom species or on ratios of 2–3 diatom species that are used by more than two authors; categories 5 and 6 include proxies based on groups of >3 taxa that either include or exclude F. curta, respectively; category 7 contains all other diatom proxies that do not fit in categories 1–6, most of which are presented in a single publication or by a specific author; and category 8 comprises all HBI diatom-specific biomarker records. Further details regarding the habitat preferences associated with the diatoms used in these proxy categories are provided in Appendix A.

2.1.1. Fragilariopsis curta + F. cylindrus %

Based on the diatom assemblage composition of sediment trap material and surface sediments in the Atlantic sector of the Southern Ocean, Gersonde and Zielinski [71] demonstrated a close correlation between the combined relative abundance of F. curta and F. cylindrus (F. c + cy %) and the distribution of Antarctic winter sea ice. This study determined that the relative abundances of F. c + cy > 3% can be applied as a proxy for areas south of the mean winter sea ice extent in the Southern Ocean.
The F. c + cy % proxy is used in 10 records from nine sites across the deep South Atlantic (Sites 58, 59, 60, and 61), Palmer Deep (Sites 37 and 38), and the Dumont d’Urville Trough (Sites 33, 35, and 68; Figure 2). Nine of the F. c + cy % records extend from 2.0 ka BP to between 1.0 and 0.1 ka BP with maximum sample intervals between 190 and 600 years at the deep ocean sites and between 180 and 4 years for the continental shelf sites (Table 4). The highest resolution F. c + cy % record, with sampling intervals of 4 years or less, is from core DTGC2011 (Site 68) in the Dumont D’Urville Trough and spans from 0.4 to ~0 ka BP [68].

2.1.2. Fragilariopsis curta %

Leventer et al. [11] assigned F. curta as the most diagnostic diatom species for sea ice reconstructions in the southern Ross Sea (Sites 1, 3, and 4) with relative abundances of >60% in areas where sea ice cover persists for 10–11 m/y [12]. In contrast with other regions of the Antarctic continental shelf, F. cylindrus is considered more indicative of open water conditions in the Ross Sea, with the transition from open water to dense pack ice reflected in a shift in the dominant diatom species from F. cylindrus to F. curta [11]. Subsequent studies in the Ross Sea (Sites 2, 5–9, and 12) have continued to use F. curta % as the prevalent species to infer seasonal sea ice conditions [12,13,14]. Beyond the Ross Sea, F. curta % has been used as the principal taxa to reconstruct sea ice conditions in Prydz Bay (Site 13) [16], the Ingrid Christensen coast (Sites 18 and 21) [22], the MacRobertson Shelf (Sites 26 and 27) [28,29], and the Antarctic Peninsula (Sites 25, 44, 52–54, and 65–67) [27,51,58,59,60]. Maximum sample intervals of the F. curta % records range between 17 and 400 years. In the northern Antarctic Peninsula and Ross Sea, there are six records with sample resolution of 50 years or less (Sites 8, 9, 12, and 52–54). Except for the study of discrete laminations [32], sites on the MacRobertson Shelf and in Prydz Bay have records with maximum sample intervals of 200 to 400 years. The highest resolution record of F. curta % (sampling intervals of ≤17 years) is from Marian Cove in the South Shetland Islands (Site 52) and covers the past 1.7 ka BP [58]. The other F. curta % records cover 0.5 to 2.0 ka of the past 2.0 ka BP with basal ages ranging from 2.0 ka BP to 0.5 ka (Figure 3 and Table 4).

2.1.3. Fragilariopsis curta + F. cylindrus/T. antarctica

The F. c + cy/T. antarctica ratio was first applied by Leventer et al. [45] to characterize the overall changes in diatom assemblage and to limit the impact of large relative abundance shifts in other diatom taxa. In the Palmer Deep (Site 38), an increase (decrease) in the ratio suggests greater (lesser) importance of sea ice melt on surface ocean conditions [44,45].
The F. c + cy/T. antarctica proxy has only been used at two other sites (Sites 41 and 48), both in the Antarctic Peninsula [48]. Located on the Anvers Shelf, these sites are immediately offshore from the Palmer Deep and comprise records covering 1.1 and 1.9 ka from 2.0 ka BP with sample intervals between 99 and 220 years (Figure 4 and Table 4). The Palmer Deep record (Site 38) covers from 2.0 ka BP to 0.3 ka BP with maximum sample intervals of 85 years [44,45].

2.1.4. Fragilariopsis curta/Fragilariopsis kerguelensis

Similar to the application of the F. c + cy/T. antarctica ratio, the F. curta/F. kerguelensis ratio was introduced by Denis et al. [39] to reflect the relative dominance of sea ice or sea ice-free conditions in the growing season. Used to interpret the high accumulation sediment record from the Adélie margin (Site 35), F. curta and F. kerguelensis were chosen based on their elevated abundance in the diatom assemblages and their respective affinities for sea ice and open water conditions.
The F. curta/F. kerguelensis ratio has also been applied to infer past sea ice conditions at three other sites (Sites 23, 36, and 63) in the Antarctic Peninsula [25,40,46]. All four F. curta/F. kerguelensis records span from 2.0 ka BP (Figure 5 and Table 4), extending to 1.0 ka BP (Site 35) [39], 1.2 ka BP (Site 36) [40], and 1.9 ka BP (Sites 23 and 63) [25,46]. The highest resolution F. curta/F. kerguelensis record is presented for the Adélie margin record (Site 35) [39] with sample intervals between 10 and 25 years. The sites from the Antarctic Peninsula (Sites 23, 36, and 63) all have maximum sample intervals between 100 and 230 years (Figure 5 and Table 4).

2.1.5. Groups including Fragilariopsis curta

A summary of the F. curta groups used as sea ice proxies in the records considered here (Table 5) show that there are up to 17 different compositions with at least three species in each. The most common taxa accompanying F. curta in the various groups are: F. cylindrus, F. vanheurckii, F. obliquecostata, F. sublinearis, F. rhombica, F. ritscheri and Porosira glacialis. Of the other diatom taxa included in the various F. curta groups, several are only used at one site within a single study, including Berkeleya rutilans, Cocconeis spp., Corethron spp., Entomoneis kufferathii, Eucampia antarctica and Synedra spp. The remaining taxa (e.g. Actinocyclus actinochilus, Chaetoceros resting spores (CRS), Pseudonitzschia turgiduloides, Fragilariopsis kerguelensis, Pentalamina corona, Thalassionema spp., Thalassiosira antarctica, T. lentiginosa, T. tumida and Thalassiothrix antarctica etc.) are included within the ‘sea ice groups’ of at least two sites within this category.
The F. curta groups have been applied as proxies for sea ice throughout the Antarctic, including the deep Southern Ocean (Sites 56–62), the Antarctic Peninsula (Sites 22, 25, 37, 40, 41, 43, 46, 47, 48, 49, 50, 52, and 55), and the EA margin (Sites 14, 15, 16, 17, 24, 28, 32, 34, and 35).
The majority (n = 27) of the 36 records using a F. curta group as a sea ice proxy date back to at least 2 ka BP, of which 11 cover the entire 2 ka period. The 9 records that start after 2 ka BP include 3 that span more than 1.5 ka and 6 records that cover between 0.8 and 0.08 ka BP to the present (Figure 6 and Table 4). One F. curta group record (Site 32) from the Dumont D’Urville Trough characterizes the seasonal sequence of sea ice and open water production from specific intervals of Late Holocene laminated sediments [32], whilst all other F. curta group records have maximum and minimum sample intervals ranging from 1000 to 2 years and 500 to 1 year, respectively (Figure 6 and Table 4).

2.1.6. Groups Excluding Fragilariopsis curta

There are seven proxy records based on three groups of diatoms excluding F. curta (Table 6). Each of the groups excluding F. curta are comprised of a unique set of taxa (Table 5), with only N. glaceii incorporated in groups both including and excluding F. curta (Table 5 and Table 6). The habitat preferences of these species are available in Appendix A.
The seven records utilizing the three ‘non-F. curta groups’ to infer sea ice conditions are presented for coastal sites of the Vestfold Hills (Sites 18–21), Wilkes Land (Sites 30 and 31), and James Ross Island (Site 42) In many of these coastal settings, F. curta is considered more indicative of ‘open water’ conditions rather than extensive pack or fast ice [17,22]. All records date back to at least 2 ka BP, with three records that span the complete 2.0 ka interval (Sites 42, 31, and 18), two that are continuous through to 0.8 ka BP (Sites 20 and 21), one that covers up to 0.3 ka BP (Site 30), and one that covers up to 0.2 ka BP (Site 19). Across the seven sites, maximum sample intervals range between 200 and 30 years and minimum sample intervals range from 105 to 12 years (Figure 7 and Table 4).

2.1.7. Other

Records in this category are based on generic features of the diatom assemblage; distinct qualities of a single taxa; or reconfigured versions of the proxies described in proxy categories 1–4 (Table 7). The habitat preferences of these species are presented in Appendix A.
Entomoneis kjellmannii % (previously Amphiprora kjellmannii) is primarily used as an indicator of perennial sea ice, especially fast ice [17,22,23]. E. kjellmannii is reported as a common, sometimes dominant, member of the sea ice diatom community in coastal areas of the Ross Sea and is also abundant in the sea ice and sea ice adjacent waters of the Ingrid Christensen coast and in parts of the Victoria Land coast [74,75,76]. E. kjellmannii is a component of the ‘fast ice index’ (one of the non-F. curta groups—Section 2.1.6 and Table 6) presented by McMinn [22] and McMinn et al. [23] and used by Berg et al. [17] as evidence for the presence of fast ice on the Rauer Group coast, Prydz Bay.
Eucampia index is introduced by Kaczmarska et al. [77] and based on the distinct morphology of the ‘pointed’ or ‘horned’ terminal valves and ‘flat’ intercalary valves of both the polar E. antarctica var. recta and the sub-polar E. antarctic var. antarctica morphotypes [78]. The Eucampia index refers to the ratio of terminal to intercalary (‘pointy to flat’) valves. E. antarctica colonies growing in colder waters with greater sea ice cover are characterized by shorter chains with a higher relative proportion of terminal valves [79,80]. Milliken et al. [57] use the Eucampia index to infer changes in sea ice cover over Maxwell Bay in the South Shetland Islands (Site 51). Resolution of this record ranges between 475 and 950 years and spans between 2.0 and 0.1 ka BP (Figure 8 and Table 4).
Fragilariopsis cylindrus % is used as the primary indicator of prolonged sea ice cover (>8.5 months) by Campagne et al. [30,33], owing to F. cylindrus’s high statistical significance in explaining variability in the diatom assemblages of the past 40 years in the Adélie Margin (Sites 29 and 33) where F. curta was not significant. Yoon et al. [58,60] present separate down-core abundance plots for F. curta and F. cylindrus% (Sites 52 and 54), showing that peaks are often off-set and highlighting the sensitive habitat preferences of the two species.
The Fragilariopsis group/Thalassiosira antarctica (T2) is similar to the more commonly applied F. curta + F. cylindrus/T. antarctica ratio. The Fragilariopsis group/T. antarctica (T2) ratio also aims to characterize the major shifts in surface ocean conditions with elevated (reduced) contributions of the Fragilariopsis group (T. antarctica) indicative of more persistent (ephemeral) seasonal sea ice. As presented in Kamanidou et al. [55], the Fragilariopsis group comprises of the combined abundance of F. curta, F. cylindrus, F. sublinearis, and F. vanheurckii. Applied to the sediment record from the Perseverance Drift north of Joinville Island in the northern-most Antarctic Peninsula (Site 47), the record covers from 0.8 to 0.0 ka BP at a resolution between 40 and 21 years [55].
Porosira glacialis/Thalassiosira antarctica are both common components of the diatom assemblage in Antarctic continental shelf sediments with P. glacialis preferring slightly cooler ocean climate conditions than T. antarctica [81,82]. As established by Pike et al. [20], the ratio of P. glacialis/T. antarctica reflects the subtle difference in environmental preferences of the two species, with ratios of >0.1 indicating annual sea ice cover greater than 7.5 months/year [20]. Applied to sediment records from the Svenner Channel, Prydz Bay, and the Dumont d’Urville Trough (Sites 16 and 35), the P. glacialis/T. antarctica ratio is used to reconstruct changes in the duration of sea ice cover [20]. The record from the Svenner Channel (Site 16) covers between 2.0 and 0.6 ka BP with a maximum and minimum resolution of 36 and 70 years, respectively [20]. The Dumont d’Urville record (Site 35) spans the period from 2.0 to 1.0 ka BP with sample intervals between 25 and 2 years [20].
Pennate–centric ratio—in sediments with low diatom concentrations, Minzoni et al. [15] use the prevalence of pennate diatoms in the assemblage to infer the presence of sea ice. The association between elevated contributions of pennate diatoms and heavier sea ice cover is based on analyses of the core top assemblages that are dominated by F. curta [83]. Applied to cores recovered from Ferrero Bay in the Amundsen Sea Embayment (Sites 10 and 11) where sedimentation rates are exceptionally low for the AP, the records of the pennate–centric ratio cover the whole of the last 2 ka at both sites with sample intervals of 500 to 1000 years (Figure 8 and Table 4).
Diatom concentrations—Sjunneskog and Taylor [42] and Michalchuk et al. [52] use diatom concentrations as a paleoproductivity proxy for the Palmer Deep and Firth of Tay marine cores (Sites 37 and 45) based on the dominant contribution of the sea ice melt bloom to the total diatom production which is reflected in the prevalence of Chaetoceros resting spores (60 to 90% of the total diatom content) in AP sediments [45,84].
Dark–light laminae—in laminated sediments from Edisto Inlet in the Ross Sea (Site 64) Tesi et al., [66] show that dark and light layers are characterized by distinctive diatom assemblages, biomarkers, and isotopic values that reflect the two dominant seasonal sea ice settings of the site: (1) sea ice break-up and thaw (dark) and (2) open surface waters (light). Tesi, et al.’s [66] record spans the last 2 ka with sample intervals of <20 years (Figure 8 and Table 4).

2.1.8. Diatom-Specific Highly Branched Isoprenoids (HBIs)

The use of HBIs as a source-specific biomarker for sea ice was first developed in the Arctic where IP25 (a mono-unsaturated C25 HBI, HBI-I) is produced by three or four sea ice diatoms [85,86]. Whilst IP25 has not been identified in Antarctic sediments, a di-unsaturated C25 HBI (diene, HBI-II) found to co-vary with IP25 in the Arctic [87,88,89,90] has been identified in sediments and sea ice from a variety of Antarctic locations [19,91,92,93,94,95,96,97]. In 2016, Belt et al. [10] established that the sympagic diatom Berkeleya adeliensis is a principal source of this diene (HBI-II) and proposed the term IPSO25 (ice proxy for the Southern Ocean with 25 carbon bonds) by analogy to the Arctic IP25. Analysis of its distribution in near-coastal Antarctic sediments led to the conclusion that IPSO25 may be a better proxy of the type of sea ice in which it is produced (platelet ice) rather than other parameters such as sea ice extent or seasonality [10,98].
A third tri-unsaturated C25 HBI (triene, HBI-III) was initially proposed to derive from open water taxa [94]. As such, the ratio between the diene and triene (HBI-II/HBI-III) was thought to reflect the relative contributions of sea ice and open ocean primary production [19,91,94]. Subsequent studies have suggested that the production of triene (HBI-III) may instead be concentrated in the marginal ice zone [50,93]. For more information on the use of HBI-II and HBI-III as sea ice proxies, readers are encouraged to consult the recent review paper by Belt [98].
Within this review, there are 14 HBI records for 12 sites across the Antarctic Peninsula, Prdyz Bay, the Adélie Land shelf, the coast of George V Land, and the western Ross Sea (Sites 16, 29, 33, 35, 37, 40, 43, 64, 65, 66, 67, and 68) (Figure 9 and Table 4). Of these 14 HBI records, 4 span from 2.0 ka BP to 1.0, 0.6, 0.3 and0.1 ka BP, 2 span to 0.0 ka BP, whilst the other 8 records begin at 0.42, 0.25, 0.22, 0.19, 0.17, 0.13, 0.08, and 0.03 ka BP, only covering the most recent centuries and decades. The lowest sample resolution for the 6 records covering older sediments range between 70 and <20 years, whilst the largest sample interval for the younger core archives ranges between 9 and 0.4 years.

2.2. Distribution of Diatom-Based Proxy Records

Whilst diatomaceous sediments are widely available in the deep Southern Ocean and the Antarctic continental shelf, there are large disparities in the distribution of records (Figure 1 and Figure 10). The highest concentrations of sites occur in the northern Antarctica Peninsula and the western Ross Sea. Meanwhile, there are currently no records on the continental shelf between the Larsen Shelf on the eastern Antarctic Peninsula and Enderby Land on the East Antarctic Margin (approx. 60° W to 60° E); there are only two records from the West Antarctic Margin between the Ross Ice Shelf and Alexander Island (approx. 170° E to 70° W); and there are only three coastal records from the Wilkes Land Margin (approx. 80° E to 135° E). This uneven distribution is typical of most Antarctic marine proxy records and largely reflects the availability (and collection) of marine sediment cores from the Antarctic continental shelf and Southern Ocean.

3. Advantages and Limitations

Diatom-based sea ice proxies from marine sediments provide the most direct link to the Antarctic sea ice environment. The fossilized frustules and biomarkers of diatoms living in or attached to sea ice offer the best evidence of sea ice in marine sediments. Whilst few of these ice-bound diatoms are preserved and most of them are specific to highly niche conditions or particular types of sea ice, there are several diatom species associated with sea ice and water in close proximity with sea ice, which are commonly preserved in Antarctic marine sediments. These taxa hold the greatest potential for a standard/generic sea ice proxy. The main pro’s and con’s of each proxy category are summarised in Table 8.

3.1. Relative Abundance

Except for HBIs, the relative abundance of species and species groups within the diatom assemblage provides baseline data in 77 of the proxy records used to reconstruct Antarctic sea ice over the past 2 ka.
The interpretation of sea ice proxies based on relative abundance is complicated by the differences in composition of the overall assemblage. Whilst the contribution of sea ice taxa to the total diatom assemblage in sediments from the deep Southern Ocean exhibits a strong latitudinal gradient correlated with sea ice cover [71], the relative abundance of sea ice taxa in Antarctic continental shelf sediment is more diverse. Where sea ice taxa comprise minor taxa (<10%), changes in the relative abundance of dominant/abundant taxa can erase, reverse, or enhance the relative abundance trends of the proxy [99]. In the Antarctic Peninsula, where Chaetoceros spp. frequently comprise >70% of the diatom assemblage [24,27,40,41,43,45,47], it is common practice to mitigate the dominance of Chaetoceros spp. by carrying out additional valve counts excluding Chaetoceros spp. in order to resolve the abundance patterns of minor taxa. The assemblage variability throughout the continental shelf makes it untenable to assign relative abundance thresholds to specific sea ice conditions (duration, concentration, etc.) and limits inter-site comparison to assess similarities in the timing and pattern of sea ice trends.
Ratios between the relative abundance of two diatom species or groups of species provide a way of normalizing data to mitigate the relative abundance differences between sites. A ratio between sea ice taxa and open ocean taxa reflects the prevalence of sea ice over ice-free conditions [45].

3.2. Statistical Approaches

At present, transfer functions (IKM, MAT, and GAM) provide the only quantitative approach to diatom-based sea ice reconstructions. Based on a reference data set of surface sediment samples from sites of varying sea ice conditions, transfer functions provide statistical estimates of past sea ice duration or concentration. The main advantage of transfer functions is that they yield quantitative results that are most easily applied in the set up and validation of climate models. The main limitations for the transfer functions are that the sea ice conditions and diatom assemblages of the continental shelf are typically too varied to be used in transfer functions. Moreover, in the case of IKM and MAT, the assumptions and types of assemblage data applied may not be suitable [100,101].
Other statistical methods applied in diatom-based sea ice reconstructions of the last 2 ka include principal component analyses (PCAs) and cluster analyses. These approaches are used to identify the individual and groups of taxa that are most strongly associated with sea ice and to determine similarities between samples based on their assemblage composition. Results gained from studies applying PCA and cluster analyses benefit from eliminating some of the subjectivity from the process of reconstructing sea ice and use statistical criteria that can be replicated at sites with different diatom assemblage data. The drawbacks with PCA and cluster analyses are that reconstructions are only qualitative and may reflect the multifaceted environmental gradients linked with sea ice that may complicate the proxy signal.

4. Discussion

4.1. Distribution

The current distribution of marine core sites with sea ice reconstructions for the past 2 ka shows that records are mostly located in the northern AP, the western Ross Sea, and the Prydz Bay area, highlighting the large expanse of the Antarctic continental shelf with no sea ice records of the past 2 ka (Figure 1). The concentration of records in these regions reflects a common bias towards the location of research bases and/or regular supply routes. Regions with fewer Antarctic bases and away from the primary logistic routes have few (Amundsen Sea, the Adelie coast, and the Wilkes Land Coast) or no sea ice records from marine core sites (Dronning Maud Land, Marie Byrd Land, and the southern Bellingshausen and Weddell Seas). Mapping the distribution, basal age, and resolution of proxy records also highlights areas where regional syntheses of centennial-scale sea ice history may be possible and reveals spatial trends in the choice of sea ice proxy that may provide a basis for standardization.
No single proxy is used at all sites. All eight proxy types are applied to sediment cores in the AP and F. curta % is used at all sites in the western Ross Sea (Figure 10). ‘Groups including F. curta’ is the most commonly applied proxy type and also the most widely distributed, with records throughout the AP, East Antarctic margin, and deep Southern Ocean (Figure 6). Although each of the 17 distinct groups within this category are only applied at a maximum of four sites (Table 5), the overall distribution attests to where ‘F. curta’ is considered useful in reconstructing sea ice. The area where F. curta is applied in sea ice reconstructions can be extended by including all proxy records that utilize F. curta. This combined distribution illustrates that F. curta is used at 57 of the 68 (in 77 of the 112 records) sites across Antarctica (Figure 11). Of the 11 sites where F. curta is omitted, 6 are from nearshore fjord and bay settings along the East Antarctic margin (Sites 19, 20, 29, 30, 31, and 64) where the sea ice proxies are tailored to fast ice [23,31,49], 2 are from the Amundsen Sea coast (10 and 11) where diatom concentrations are too low to produce reliable assemblage data [15] and the remaining 4 sites (42, 45, 51, 64) are from studies that promote other features of the sediment cores and/or focus on non-diatom proxies [49,52,57,66].
The paucity of the continental shelf records based on the more established F. curta + cylindrus% proxy is likely due to two principal factors that complicate the relative abundance patterns of this pairing. Firstly, the diversity of habitats on the continental shelf supports greater variation in the composition, species richness, and evenness of diatom assemblages such that relative abundances are also highly variable. Secondly, the production and distribution of F. cylindrus is influenced not only by sea ice but also by cold stratified waters [102,103,104]. Close to the continent, glacial discharge could produce cold stratified waters and exceed the sea ice influence on the occurrence of F. cylindrus, making it unreliable as a sea ice proxy in this setting.
Whilst HBI records have only been applied in recent years, the scope for more efficient analyses of samples and the potential to be included in standard geochemical processing may rapidly increase the application and distribution of this proxy over the next decade. HBIs may also be absent from the sedimentary record and the interpretation of such an absence is often complex. Possible causes include perennial sea ice cover, ice-free conditions, or ice conditions not suitable for diatoms (e.g., the sea ice may be too thick to allow a sufficient amount of light to pass through) [98]. Furthermore, after HBIs have been produced, they must avoid degradation before and after deposition. For example, the incorporation of sulfur into IPSO25 may impact some records [95,98] The challenges associated with interpreting such changes have restricted our ability to produce new HBI proxy records.

4.2. Age, Duration and Resolution of Records

Many of the Southern Ocean and Antarctic continental shelf records considered here extend beyond the 2 ka interval, with several covering the full Holocene or longer time intervals [13,14,24,25,26,36,39,41,57,63,105]. Since these sediment cores mostly target millennial to sub-millennial resolutions over a longer timeframe, some may be suitable for higher frequency sampling and dating to improve the resolution and age control of the youngest sediments.
Marine sedimentation rates at many of the sites presented here are considered too low to resolve sub-decadal records. There are rare sites of very high accumulation (~1 cm a−1) on the Antarctic continental shelf that can yield records of sub-decadal to annual resolution (e.g., Site 37/38: IODP 178-1098 and Site 33: IODP 318-1357). Furthermore, where varves are preserved at the highest accumulation sites (2–10 cm a−1), it is possible to resolve seasonal records, usually only for discrete horizons due to the number of samples required and the time-intensive nature of analyses [32,106].
The limits of resolution do not only depend on sedimentation rates but also on signal attenuation from bioturbation. Several studies have explored how bioturbation affects the amplitude, timing, and duration of climate signals in proxy records [107,108]. Whilst the techniques and specific results vary, the studies are broadly consistent in finding substantial attenuation (30 to 70%) in the amplitude of millennial scale climate signals when sedimentation rates are less than ~15 cm ka−1. Similar experiments with centennial and sub-centennial climate variations suggest that a signal is only preserved where the mixed layer depth is shallow (≤5 cm) and sedimentation rates are higher than ~15 cm ka−1 [108].
Of the 68 sites included in this study, only 7 have sedimentation rates less than 15 cm ka−1 (Table 4) with coarse minimum sample resolutions between 350 and 1000 years (Figure 12 and Table 4). The 13 proxy records with the highest resolution come from sites with sedimentation rates ranging from 1.4 m ka−1 to 18.13 m ka−1, where the depth and strength of climate signals are still modified within the sediment column but should preserve sub-millennial, centennial, and possibly sub-centennial variability.

4.3. Scope for Validation and Standardisation

Whilst the majority of diatom proxies covered in this review are based on ecological associations between the distribution of species in surface sediments and the mean oceanographic conditions above [10,70,71,81,82,94,100,109], validation (and/or calibration) against sea ice observations is problematic due to the relatively short 40-year instrumental record available for Antarctic sea ice, a timeframe that is not normally resolvable in marine sediment cores. Even at sites with exceptionally high sedimentation rates (>1 cm a−1), signal attenuation from bioturbation and dating uncertainties may still preclude accurate alignment with satellite records. Where calibration is attempted across regions with decreasing (increasing) sea ice trends over recent decades [4], surface proxies may still predispose down-core records to underestimate (overestimate) past sea ice [97].
In addition to the limitations on validation, the diversity of existing records, in terms of the proxies, resolution, and age uncertainties, makes it particularly difficult to compare data and reconstructions. It is also possible, and perhaps likely, that a single ‘standard’ proxy cannot encapsulate the heterogenous nature of the sea ice environment. The current diversity of approaches may therefore be seen in a positive light, providing more detailed and meaningful descriptions of a greater number of sea ice-related variables. Additionally, these records are tailored to represent sea ice conditions in different locations. Without evaluating how well proxies work in different regions, it is difficult to assess the feasibility of a standard circum-Antarctic sea ice proxy. Also, given the divergent trends in sea ice seen in different sectors of Antarctica over recent decades [4], initial comparisons would probably be best applied to regions with similar sea ice history.
A potential first step towards validation and standardization would be to compare different proxies within a single core and a single proxy between different cores located within areas of similar sea ice history (Figure 10). Whilst the use of Fragilariopsis curta in the majority of sea ice proxies and regions (Figure 11) makes it a prime candidate for evaluation as a standard Antarctic sea ice proxy, the analytical efficiency afforded by the relatively new HBIs also warrants further development to hone its ecological association with sea ice.

5. Conclusions

This review provides a thorough evaluation of the diatom-based proxies used to reconstruct Antarctic sea ice of the past 2 ka. The study collates 112 records from 68 core sites, with proxies based on more than 30 different combinations of diatom taxa as well as HBI biomarkers. Presenting the origins and tenets of the proxies indicates how their varied composition reflects the ecological and taxonomic diversity of the sea ice environment. The wide variety of diatom-based sea ice proxies may limit the opportunities and complicate data assimilation and proxy standardisation but this diversity also suggests that one ‘standard’ diatom-based proxy is probably not appropriate for the varied ecological conditions of Antarctic sea ice. The broad range of proxies currently available can provide richer, localized perspective, and therefore potential for a more detailed insight into the different ways that sea ice conditions have changed in the past.
The detailed inventory of proxies and cores presented within this review provides a valuable resource for future research since it summarizes essential reference data on 112 existing sea ice records into a single, readily accessible publication. This resource will permit easy identification of: where records with certain time frames and/or resolutions occur; where specific proxies are applied; where comparisons between proxies and records are feasible; and where suitable records for regional syntheses occur. Mapping the distribution of records also emphasizes the large areas of continental shelf with sparse or no records of past sea ice. This review of existing records will hopefully encourage more informed decisions on the choice of proxy and temporal framework, ultimately leading to greater compatibility and consistency in the diatom-based reconstructions of Antarctic sea ice.

Author Contributions

Conceptualisation: C.S.A.; Outline and approach: C.S.A.; Investigation & resources: C.S.A. and Z.C.W.; Writing: C.S.A. and Z.C.W.; Reviewing and editing: C.S.A. and Z.C.W.; Corrections and proof reading: C.S.A. and Z.C.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by British Antarctic Survey, with core funding from the Natural Environment Research Council and a PhD studentship from the National Environments Research Council Great Western 4+ Doctoral Training Partnership.

Data Availability Statement

Not applicable.

Acknowledgments

This work forms part of the British Antarctic Survey programme ‘Polar Science for Planet Earth’ and is a contribution to the PAGES 2k Network (through the CLIVASH2k project). Past Global Changes (PAGES) is supported by the US National Science Foundation and the Swiss Academy of Sciences. We thank Liz Thomas for her support in producing this review and we are grateful to the two anonymous reviewers for their constructive comments and suggestions which helped to improve the quality of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Information on diatom taxa included in Antarctic sea ice proxies:
1. Fragilariopsis curta is closely associated with sea ice throughout the Southern Ocean. F. curta is a common component of the diatom assemblage within the sea ice environment (often as a dominant species) in sea ice, in ice edge blooms, and in waters close to the sea ice [103,110,111,112,113,114,115,116,117,118,119]. This sea ice association is supported by plankton and sediment trap samples that show the distribution of F. curta to be broadly confined by the northern limit of sea ice [120,121]. In surface sediments, F. curta abundances increase southwards from the maximum limit of seasonal sea ice [81,82]. High abundances (>20%) are found in areas where summer sea surface temperatures (SSSTs) are below 0.5 °C [82] and where seasonal sea ice persists for 9–11 months/year [81]. The highest recorded sedimentary abundances (>60%) occur in the Ross Sea, Prydz Bay, and the Weddell Sea [12,81].
2. Fragilariopsis cylindrus is found in a range of sea ice settings and in waters closely influenced by sea ice, where they often occur as the dominant species [103,110,111,112,113,114,116,118,121,122]. Whilst widely accepted as a sea ice-related diatom, F. cylindrus has also been associated with melt waters and stratified conditions [31,49,104,111], especially in the Ross Sea [11,123]. In surface sediments, F. cylindrus abundances of >10% are found where SSSTs are below −0.5 °C and sea ice is present for 3–10 months/year [81,82,124].
3. Thalassiosira antarctica is not usually found in sea ice but is recognized as a sea ice associated diatom, living in waters adjacent to sea ice, in crack pools, and in waters with unconsolidated sea ice [110,112,119,123,125]. The highest abundances in the Ross Sea are found near to the coast and may indicate a link with ice shelf and tidewater glaciers [123]. The occurrence of T. antarctica in sediment traps and seasonally laminated diatomaceous sediments supports the assertion that T. antarctica is primarily a member of the late summer and/or autumn phytoplankton community [12,32,106,126]. The distribution of T. antarctica in surface sediments is broadly consistent with its occurrence in surface waters. The highest abundances (>25%) are typically observed in coastal regions that experience >6 months sea ice per year and have summer SSTs between −1 and 0 °C [13,81,82,123].
4. Fragilariopsis sublinearis follows a similar distribution to F. obliquecostata in surface waters. The highest abundances (5–10%) of F. sublinearis are located almost exclusively in Prydz Bay, the Ross Sea, the Weddell Sea, and along the Wilkes Land margin, close to and within the maximum February sea ice extent [71,112,120,127]. In sediments, relative abundances of F. sublinearis above 2% occur in locations where sea ice is present for >7.5 months/year [70,81,82].
5. Fragilariopsis kerguelensis is an endemic Southern Ocean, and pelagic marine diatom is common in the iron-limited waters of the Antarctic Circumpolar Current with peak productivity along the southern boundary of the Polar Front [128,129,130,131]. Experiments and observations show that the optimal growth of F. kerguelensis occurs in 4–5 °C waters—the range of summer sea surface temperatures in the Polar Front Zone [131,132]. In surface sediments, F. kerguelensis dominates diatom assemblages in the Antarctic and Polar Front Zone throughout most of the Southern Ocean [129,130,131] and is negatively correlated with sea ice concentrations [133].
6. Fragilariopsis obliquecostata is also associated with sea ice but is less common and typically confined to more southerly locations than either F. curta or F. cylindrus. F. obliquecostata has been recorded in pack ice, land fast ice, and within waters of the marginal ice zone of the Weddell and Amundsen Seas [110,112,114]. In surface sediments, F. obliquecostata is less abundant than F. cylindrus with elevated abundances of >5% generally found south of the Antarctic Divergence where SSSTs are below 0 °C [70,82]. Elevated abundances of F. obliquecostata in the sediments broadly correspond to the area of maximum perennial sea ice and/or locations that experience >7 m/y sea ice cover [81] with the highest abundances of >10% recorded off the Victoria Land coast in the Ross Sea, in the Weddell Sea, and in the Prydz Bay region [82,123].
7. Fragilariopsis rhombica is associated with regions close to the Antarctic coast and ice shelves [44,134]. Although it has been recorded in sea ice samples, F. rhombica is absent from descriptions of common sea ice taxa and is more abundant in the waters immediately north of the sea ice edge [110,116,135]. F. rhombica’s distribution in surface sediments is linked with the Antarctic Zone with maximum relative abundances of >10% found in areas where summer sea surface temperatures range from −1.5 to 1 °C and sea ice cover persists for 7–9 months of the year [81,82,124].
8. Fragilariopsis ritscheri is an endemic Antarctic species that is usually found at lower abundances than other Fragilariopsis species [82,120]. In sediments, F. ritscheri is recorded at highest abundances of 1.5–3% where surface water temperatures range between −2.0 and 1 °C, broadly consistent with the area south of the average winter sea ice extent [81,82,136].
9. Fragilariopsis vanheurckii is generally considered a sea ice diatom [135,137,138], and is also commonly found in waters close to the sea ice edge and in the stratified melt waters following sea ice melt [111,114]. Analyses of the diatom assemblage in sediment trap material from the northwest Weddell Sea showed that F. vanheurckii comprised the highest abundances of >15% when the sea ice edge was at or in close proximity to the mooring site [71]. In sediment assemblages, the highest abundances of >8% F. vanheurckii have been recorded in the Amundsen Sea [114] and northeast Antarctic Peninsula [139].
10. Actinocyclus actinochilus is linked with cool waters. In surface sediments, the distribution of A. actinochilus relates to surface water temperatures between −2 and 2 °C, broadly equivalent to the region between the minimum and maximum ice cover [81,82]. The highest abundances of A. actinochilus are reported from sediments in the Amundsen Sea (up to 14.4%, [114]) and Enderby Basin (9.38–13.33%, [131]). In other regions, A. actinochilus is more commonly present at abundances between 1 and 3% [20,81,82,123,140].
11. Porosira glacialis is a bipolar diatom species associated with cold shelf waters adjacent to sea ice [81,82]. P. glacialis is purported to grow in the open ocean beyond the perennial sea ice edge [81,82], and, while P. glacialis has been observed in waters with high concentrations of slush and wave-exposed shore ice, it has not been found in sea ice [110,125,141]. In sediments, P. glacialis is located shoreward of the maximum winter sea ice extent and reaches maximum abundances (>2%) beneath regions with February SSTs ranging from 0 to 0.5 °C [81] and at least 7.5 months per year sea ice cover [20,81]. While P. glacialis usually comprises only a minor component of the diatom assemblage, there are several records of diatomaceous laminae in which P. glacialis is visually one of the most noticeable taxa in late summer/autumn laminae from the Dumont d’Urville Trough, Mertz Ninnis Trough [32,142], Iceberg Alley [28], and the MacRobertson Shelf [106].
12. Entomoneis kjellmanii is a common member of the spring bottom ice community (particularly fast ice) prior to ice break out [74,143,144,145,146]. It is rarely observed in other sea ice habitats or in the surrounding waters, presenting a particularly robust affiliation with early spring, coastal sea ice conditions.
13. Navicula glaceii is a cryophilic, neritic diatom [141,147]. N. glaceii is particularly associated with the ‘slushy’ semi-frozen tideline habitat in fast ice and may be abundant in adjacent waters where it is seeded from melting and/or fragmentation of sea ice in the nearshore area [148].
14 and 15. Nitzschia stellata and Berkeleya adeliensis are categorized as ‘tube dwelling” sympagic diatoms, frequently reported in fast ice samples from coastal regions of Antarctica, particularly East Antarctica [117,149,150,151,152]. Both N. stellata and B. adeliensis can be dominant species within the sea ice, but are only rarely found in the adjacent waters or underlying sediments [115,149,153,154].
16. Thalassiosira australis is found in and around sea ice and can form dense mats on the underside of fast ice in coastal areas of East Antarctica [144,155,156,157]. T. australis resting spores are also found in sediments under fast ice and in sediments from seasonally open water sites closely ‘downstream’ of fast ice [158].
17. Pleurosigma directum is a planktic pennate species found at low (rare) abundances in many regions and described as “probably cosmopolitan” [159].
18. Pinnularia quadratarea is a pennate benthic species, found within and on congelation and fast ice. P. quadraterea is more heavily silicified and dissolution-resistant than many other sympagic diatoms, and, as such, is often present within the underlying sediments [44,153,160].
19. Fragilaria striulata is a marine species of this predominantly freshwater genus, widely distributed throughout the North Atlantic and Arctic Oceans. Mostly recorded as a benthic and epiphytic littoral species, it has also been reported as a neritic planktonic species [159,161]. In the Southern Hemisphere, F. striulata has been found in waters and surface sediments along the Chilean coast and at nearshore sites in Antarctica [161,162,163].
20. Synedropsis is an araphid bi-polar genus associated with sea ice. In the Antarctic, it is found predominantly within the bottom ice community or as an epiphyte on other diatoms [164,165].

References

  1. Fetterer, F.; Knowles, K.; Meier, W.N.; Savoie, M.; Windnagel, A.K. Sea Ice Index, Version 3 (Southern Hemisphere); National Snow and Ice Data Center: Boulder, CO, USA, 2017. [Google Scholar]
  2. Stammerjohn, S.; Massom, R.; Rind, D.; Martinson, D. Regions of rapid sea ice change: An inter-hemispheric seasonal comparison. Geophys. Res. Lett. 2012, 39, L06501. [Google Scholar] [CrossRef] [Green Version]
  3. Holland, P.R. The seasonality of Antarctic sea ice trends. Geophys. Res. Lett. 2014, 41, 4230–4237. [Google Scholar] [CrossRef] [Green Version]
  4. Parkinson, C.L. A 40-y record reveals gradual Antarctic sea ice increases followed by decreases at rates far exceeding the rates seen in the Arctic. Proc. Natl. Acad. Sci. USA 2019, 116, 14414–14423. [Google Scholar] [CrossRef] [Green Version]
  5. Parkinson, C.L.; Cavalieri, D.J. Antarctic sea ice variability and trends, 1979–2010. Cryosphere 2012, 6, 871–880. [Google Scholar] [CrossRef] [Green Version]
  6. Rosenblum, E.; Eisenman, I. Sea Ice Trends in Climate Models Only Accurate in Runs with Biased Global Warming. J. Clim. 2017, 30, 6265–6278. [Google Scholar] [CrossRef]
  7. Fox-Kemper, B.; Hewitt, H.T.; Xiao, C.; Aðalgeirsdóttir, G.; Drijfhout, S.S.; Edwards, T.L.; Golledge, N.R.; Hemer, M.; Kopp, R.E.; Krinner, G.; et al. Ocean, Cryosphere and Sea Level Change. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Masson Delmotte, V., Zhai, P., Pirani, A., Connors, S.L., Péan, C., Berger, S., Caud, N., Chen, Y., Goldfarb, L., Gomis, M.I., et al., Eds.; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
  8. Thomas, E.R.; Allen, C.S.; Etourneau, J.; King, A.C.F.; Severi, M.; Winton, V.H.L.; Müller, J.; Crosta, X.; Peck, V.L. Antarctic Sea Ice Proxies from Marine and Ice Core Archives Suitable for Reconstructing Sea Ice over the past 2000 Years. Geosciences 2019, 9, 509. [Google Scholar] [CrossRef] [Green Version]
  9. Zonneveld, K.A.F.; Versteegh, G.J.M.; Kasten, S.; Eglinton, T.I.; Emeis, K.C.; Huguet, C.; Koch, B.P.; de Lange, G.J.; de Leeuw, J.W.; Middelburg, J.J.; et al. Selective preservation of organic matter in marine environments; processes and impact on the sedimentary record. Biogeosciences 2010, 7, 483–511. [Google Scholar] [CrossRef] [Green Version]
  10. Belt, S.T.; Smik, L.; Brown, T.A.; Kim, J.H.; Rowland, S.J.; Allen, C.S.; Gal, J.K.; Shin, K.H.; Lee, J.I.; Taylor, K.W.R. Source identification and distribution reveals the potential of the geochemical Antarctic sea ice proxy IPSO25. Nat. Commun. 2016, 7, 12655. [Google Scholar] [CrossRef] [Green Version]
  11. Leventer, A.; Dunbar, R.B.; DeMaster, D.J. Diatom Evidence for Late Holocene Climatic Events in Granite Harbor, Antarctica. Paleoceanography 1993, 8, 373–386. [Google Scholar] [CrossRef]
  12. Leventer, A.; Dunbar, R.B. Recent Diatom Record of McMurdo Sound, Antarctica: Implications for History of Sea Ice Extent. Paleoceanography 1988, 3, 259–274. [Google Scholar] [CrossRef]
  13. Cunningham, W.L.; Leventer, A.; Andrews, J.T.; Jennings, A.E.; Licht, K. Late Pleistocene-Holocene marine condtions in the Ross Sea, Antarctica: Evidence from the diatom record. Holocene 1999, 9, 129–139. [Google Scholar] [CrossRef]
  14. Mezgec, K.; Stenni, B.; Crosta, X.; Masson-Delmotte, V.; Baroni, C.; Braida, M.; Ciardini, V.; Colizza, E.; Melis, R.; Salvatore, M.C.; et al. Holocene sea ice variability driven by wind and polynya efficiency in the Ross Sea. Nat. Commun. 2017, 8, 1334. [Google Scholar] [CrossRef] [Green Version]
  15. Minzoni, R.T.; Majewski, W.; Anderson, J.B.; Yokoyama, Y.; Fernandez, R.; Jakobsson, M. Oceanographic influences on the stability of the Cosgrove Ice Shelf, Antarctica. Holocene 2017, 27, 1645–1658. [Google Scholar] [CrossRef]
  16. Hemer, M.A.; Harris, P.T. Sediment core from beneath the Amery Ice Shelf, East Antarctica, suggests mid-Holocene ice-shelf retreat. Geology 2003, 31, 127–130. [Google Scholar] [CrossRef]
  17. Berg, S.; Wagner, B.; Cremer, H.; Leng, M.; Melles, M. Late Quaternary environmental and climate history of Rauer Group, East Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 297, 201–213. [Google Scholar] [CrossRef]
  18. Crosta, X.; Crespin, J.; Swingedouw, D.; Marti, O.; Masson-Delmotte, V.; Etourneau, J.; Goosse, H.; Braconnot, P.; Yam, R.; Brailovski, I.; et al. Ocean as the main driver of Antarctic ice sheet retreat during the Holocene. Glob. Planet. Change 2018, 166, 62–74. [Google Scholar] [CrossRef]
  19. Denis, D.; Crosta, X.; Barbara, L.; Massé, G.; Renssen, H.; Ther, O.; Giraudeau, J. Sea ice and wind variability during the Holocene in East Antarctica: Insight on middle–high latitude coupling. Quat. Sci. Rev. 2010, 29, 3709–3719. [Google Scholar] [CrossRef]
  20. Pike, J.; Crosta, X.; Maddison, E.J.; Stickley, C.E.; Denis, D.; Barbara, L.; Renssen, H. Observations on the relationship between the Antarctic coastal diatoms Thalassiosira antarctica Comber and Porosira glacialis (Grunow) Jørgensen and sea ice concentrations during the late Quaternary. Mar. Micropaleontol. 2009, 73, 14–25. [Google Scholar] [CrossRef] [Green Version]
  21. Taylor, F.; McMinn, A. Late Quaternary Diatom Assemblages from Prydz Bay, Eastern Antarctica. Quat. Res. 2002, 57, 151–161. [Google Scholar] [CrossRef]
  22. McMinn, A. Late Holocene increase in sea ice extent in fjords of the Vestfold Hills, eastern Antarctica. Antarct. Sci. 2000, 12, 80–88. [Google Scholar] [CrossRef]
  23. McMinn, A.; Heijnis, H.; Harle, K.; McOrist, G. Late-Holocene climatic change recorded in sediment cores from Ellis Fjord, eastern Antarctica. Holocene 2001, 11, 291–300. [Google Scholar] [CrossRef]
  24. Allen, C.S.; Oakes-Fretwell, L.M.; Anderson, J.B.; Hodgson, D.A. A record of Holocene glacial and oceanographic variability in Neny Fjord, Antarctic Peninsula. Holocene 2010, 20, 551–564. [Google Scholar] [CrossRef]
  25. Peck, V.L.; Allen, C.S.; Kender, S.; McClymont, E.L.; Hodgson, D.A. Oceanographic variability on the West Antarctic Peninsula during the Holocene and the influence of upper circumpolar deep water. Quat. Sci. Rev. 2015, 119, 54–65. [Google Scholar] [CrossRef] [Green Version]
  26. Taylor, F.; McMinn, A. Evidence from diatoms for Holocene climate fluctuation along the East Antarctic margin. Holocene 2001, 11, 455–466. [Google Scholar] [CrossRef]
  27. Taylor, F.; Whitehead, J.; Domack, E. Holocene paleoclimate change in the Antarctic Peninsula: Evidence from the diatom, sedimentary and geochemical record. Mar. Micropaleontol. 2001, 41, 25–43. [Google Scholar] [CrossRef]
  28. Alley, K.; Patacca, K.; Pike, J.; Dunbar, R.; Leventer, A. Iceberg Alley, East Antarctic Margin: Continuously laminated diatomaceous sediments from the late Holocene. Mar. Micropaleontol. 2018, 140, 56–68. [Google Scholar] [CrossRef]
  29. Rathburn, A.E.; Pichon, J.J.; Ayress, M.A.; DeDeckker, P. Microfossil and stable-isotope evidence for changes in Late Holocene palaeoproductivity and palaeoceanographic conditions in the Prydz Bay region of Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1997, 131, 485–510. [Google Scholar] [CrossRef]
  30. Campagne, P.; Crosta, X.; Houssais, M.N.; Swingedouw, D.; Schmidt, S.; Martin, A.; Devred, E.; Capo, S.; Marieu, V.; Closset, I.; et al. Glacial ice and atmospheric forcing on the Mertz Glacier Polynya over the past 250 years. Nat. Commun. 2015, 6, 6642. [Google Scholar] [CrossRef] [Green Version]
  31. Cremer, H.; Gore, D.; Melles, M.; Roberts, D. Palaeoclimatic significance of late Quaternary diatom assemblages from southern Windmill Islands, East Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2003, 195, 261–280. [Google Scholar] [CrossRef]
  32. Maddison, E.J.; Pike, J.; Dunbar, R. Seasonally laminated diatom-rich sediments from Dumont d’Urville Trough, East Antarctic Margin: Late-Holocene Neoglacial sea-ice conditions. Holocene 2012, 22, 857–875. [Google Scholar] [CrossRef]
  33. Campagne, P.; Crosta, X.; Schmidt, S.; Noëlle Houssais, M.; Ther, O.; Massé, G. Sedimentary response to sea ice and atmospheric variability over the instrumental period off Adélie Land, East Antarctica. Biogeosciences 2016, 13, 4205–4218. [Google Scholar] [CrossRef] [Green Version]
  34. Crosta, X.; Etourneau, J.; Orme, L.C.; Dalaiden, Q.; Campagne, P.; Swingedouw, D.; Goosse, H.; Massé, G.; Miettinen, A.; McKay, R.M.; et al. Multi-decadal trends in Antarctic sea-ice extent driven by ENSO–SAM over the last 2000 years. Nat. Geosci. 2021, 14, 156–160. [Google Scholar] [CrossRef]
  35. Kulbe, T.; Melles, M.; Verkulich, S.R.; Pushina, Z.V. East Antarctic Climate and Environmental Variability over the Last 9400 Years Inferred from Marine Sediments of the Bunger Oasis. Arct. Antarct. Alp. Res. 2001, 33, 223–230. [Google Scholar] [CrossRef]
  36. Crosta, X.; Crespin, J.; Billy, I.; Ther, O. Major factors controlling Holocene delta C-13(org) changes in a seasonal sea-ice environment, Adelie Land, East Antarctica. Glob. Biogeochem. Cycles 2005, 19, 1–9. [Google Scholar] [CrossRef] [Green Version]
  37. Crosta, X.; Debret, M.; Denis, D.; Courty, M.A.; Ther, O. Holocene long- and short-term climate changes off Adélie Land, East Antarctica. Geochem. Geophys. Geosystems 2007, 8, Q11009. [Google Scholar] [CrossRef]
  38. Crosta, X.; Denis, D.; Ther, O. Sea ice seasonality during the Holocene, Adelie Land, East Antarctica. Mar. Micropaleontol. 2008, 66, 222–232. [Google Scholar] [CrossRef]
  39. Denis, D.; Crosta, X.; Schmidt, S.; Carson, D.S.; Ganeshram, R.S.; Renssen, H.; Bout-Roumazeilles, V.; Zaragosi, S.; Martin, B.; Cremer, M.; et al. Holocene glacier and deep water dynamics, Adélie Land region, East Antarctica. Quat. Sci. Rev. 2009, 28, 1291–1303. [Google Scholar] [CrossRef]
  40. Kim, S.; Yoo, K.-C.; Lee, J.I.; Khim, B.-K.; Bak, Y.-S.; Lee, M.K.; Lee, J.; Domack, E.W.; Christ, A.J.; Yoon, H.I. Holocene paleoceanography of Bigo Bay, west Antarctic Peninsula: Connections between surface water productivity and nutrient utilization and its implication for surface-deep water mass exchange. Quat. Sci. Rev. 2018, 192, 59–70. [Google Scholar] [CrossRef]
  41. Etourneau, J.; Collins, L.G.; Willmott, V.; Kim, J.H.; Barbara, L.; Leventer, A.; Schouten, S.; Sinninghe Damsté, J.S.; Bianchini, A.; Klein, V.; et al. Holocene climate variations in the western Antarctic Peninsula: Evidence for sea ice extent predominantly controlled by changes in insolation and ENSO variability. Clim. Past 2013, 9, 1431–1446. [Google Scholar] [CrossRef] [Green Version]
  42. Sjunneskog, C.; Taylor, F. Postglacial marine diatom record of the Palmer Deep, Antarctic Peninsula (ODP Leg 178, Site 1098) 1. Total diatom abundance. Paleoceanography 2002, 17, PAL-4. [Google Scholar] [CrossRef]
  43. Taylor, F.; Sjunneskog, C. Postglacial marine diatom record of the Palmer Deep, Antarctic Peninsula (ODP Leg 178, Site 1098) 2. Diatom assemblages. Paleoceanography 2002, 17, PAL-2. [Google Scholar] [CrossRef]
  44. Leventer, A. The Fate of Antarctic “Sea-ice diatoms” and their use as palaeoenvironmental indicators. In Antarctic Sea Ice Biological, Processes, Interactions and Variability; Union, A.G., Ed.; American Geophysical Union: Washington, DC, USA, 1998; pp. 121–137. [Google Scholar]
  45. Leventer, A.; Domack, E.W.; Ishman, S.E.; Brachfeld, S.; McClennen, C.E.; Manley, P. Productivity cycles of 200–300 years in the Antarctic Peninsula region: Understanding linkages among the sun, atmosphere, oceans, sea ice, and biota. Geol. Soc. Am. Bull. 1996, 108, 1626–1644. [Google Scholar] [CrossRef]
  46. Roberts, S.J.; Monien, P.; Foster, L.C.; Loftfield, J.; Hocking, E.P.; Schnetger, B.; Pearson, E.J.; Juggins, S.; Fretwell, P.; Ireland, L.; et al. Past penguin colony responses to explosive volcanism on the Antarctic Peninsula. Nat. Commun. 2017, 8, 14914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Barbara, L.; Crosta, X.; Schmidt, S.; Masse, G. Diatoms and biomarkers evidence for major changes in sea ice conditions prior the instrumental period in Antarctic Peninsula. Quat. Sci. Rev. 2013, 79, 99–110. [Google Scholar] [CrossRef]
  48. Yoon, H.I.; Park, B.-K.; Kim, Y.; Kang, C.Y. Glaciomarine sedimentation and its paleoclimatic implications on the Antarctic Peninsula shelf over the last 15,000 years. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2002, 185, 235–254. [Google Scholar] [CrossRef]
  49. Minzoni, R.T.; Anderson, J.B.; Fernandez, R.; Wellner, J.S. Marine record of Holocene climate, ocean, and cryosphere interactions: Herbert Sound, James Ross Island, Antarctica. Quat. Sci. Rev. 2015, 129, 239–259. [Google Scholar] [CrossRef]
  50. Barbara, L.; Crosta, X.; Leventer, A.; Schmidt, S.; Etourneau, J.; Domack, E.; Masse, G. Environmental responses of the Northeast Antarctic Peninsula to the Holocene climate variability. Paleoceanography 2016, 31, 131–147. [Google Scholar] [CrossRef] [Green Version]
  51. Heroy, D.C.; Sjunneskog, C.; Anderson, J.B. Holocene climate change in the Bransfield Basin, Antarctic Peninsula: Evidence from sediment and diatom analysis. Antarct. Sci. 2008, 20, 69–87. [Google Scholar] [CrossRef]
  52. Michalchuk, B.R.; Anderson, J.B.; Wellner, J.S.; Manley, P.L.; Majewski, W.; Bohaty, S. Holocene climate and glacial history of the northeastern Antarctic Peninsula: The marine sedimentary record from a long SHALDRIL core. Quat. Sci. Rev. 2009, 28, 3049–3065. [Google Scholar] [CrossRef]
  53. Barcena, M.A.; Fabres, B.; Isla, E.; Flores, J.A.; Sierro, F.J.; Canals, M.; Palanques, A. Holocene neoglacial events in the Bransfield Strait (Antarctica). Palaeocenographic and palaeoclimatic significance. Sci. Mar. 2006, 70, 607–619. [Google Scholar] [CrossRef] [Green Version]
  54. Barcena, M.A.; Isla, E.; Plaza, A.; Flores, J.A.; Sierro, F.J.; Masque, P.; Sanchez-Cabeza, J.A.; Palanques, A. Bioaccumulation record and paleoclimatic significance in the Western Bransfield Strait. The last 2000 years. Deep-Sea Res. Part Ii-Top. Stud. Oceanogr. 2002, 49, 935–950. [Google Scholar] [CrossRef]
  55. Kyrmanidou, A.; Vadman, K.J.; Ishman, S.E.; Leventer, A.; Brachfeld, S.; Domack, E.W.; Wellner, J.S. Late Holocene oceanographic and climatic variability recorded by the Perseverance Drift, northwestern Weddell Sea, based on benthic foraminifera and diatoms. Mar. Micropaleontol. 2018, 141, 10–22. [Google Scholar] [CrossRef]
  56. Barcena, M.A.; Gersonde, R.; Ledesma, S.; Fabres, J.; Calafat, A.M.; Canals, M.; Sierro, F.J.; Flores, J.A. Record of Holocene glacial oscillations in Bransfield Basin as revealed by siliceous microfossil assemblages. Antarct. Sci. 1998, 10, 269–285. [Google Scholar] [CrossRef]
  57. Milliken, K.T.; Anderson, J.B.; Wellner, J.S.; Bohaty, S.M.; Manley, P.L. High-resolution Holocene climate record from Maxwell Bay, South Shetland Islands, Antarctica. Geol. Soc. Am. Bull. 2009, 121, 1711–1725. [Google Scholar] [CrossRef]
  58. Yoon, H.I.; Yoo, K.C.; Park, B.K.; Kim, Y.; Khim, B.K.; Kang, C.Y. The origin of massive diamicton in Marian and Potter coves, King George Island, West Antarctica. Geosci. J. 2004, 8, 1–10. [Google Scholar] [CrossRef]
  59. Yoo, K.-C.; Yoon, H.I.; Kim, J.-K.; Khim, B.-K. Sedimentological, geochemical and palaeontological evidence for a neoglacial cold event during the late Holocene in the continental shelf of the northern South Shetland Islands, West Antarctica. Polar Res. 2009, 28, 177–192. [Google Scholar] [CrossRef]
  60. Yoon, H.I.; Yoo, K.C.; Bak, Y.S.; Lim, H.S.; Kim, Y.; Lee, J.I. Late Holocene cyclic glaciomarine sedimentation in a subpolar fjord of the South Shetland Islands, Antarctica, and its paleoceanographic significance: Sedimentological, geochemical, and paleontological evidence. Geol. Soc. Am. Bull. 2010, 122, 1298–1307. [Google Scholar] [CrossRef]
  61. Divine, D.V.; Koç, N.; Isaksson, E.; Nielsen, S.; Crosta, X.; Godtliebsen, F. Holocene Antarctic climate variability from ice and marine sediment cores: Insights on ocean-atmosphere interaction. Quat. Sci. Rev. 2010, 29, 303–312. [Google Scholar] [CrossRef]
  62. Ferry, A.J.; Crosta, X.; Quilty, P.G.; Fink, D.; Howard, W.; Armand, L.K. First records of winter sea ice concentration in the southwest Pacific sector of the Southern Ocean. Paleoceanography 2015, 30, 1525–1539. [Google Scholar] [CrossRef] [Green Version]
  63. Xiao, W.S.; Esper, O.; Gersonde, R. Last Glacial-Holocene climate variability in the Atlantic sector of the Southern Ocean. Quat. Sci. Rev. 2016, 135, 115–137. [Google Scholar] [CrossRef]
  64. Bianchi, C.; Gersonde, R. Climate evolution at the last deglaciation: The role of the Southern Ocean. Earth Planet. Sci. Lett. 2004, 228, 407–424. [Google Scholar] [CrossRef]
  65. Orme, L.C.; Crosta, X.; Miettinen, A.; Divine, D.V.; Husum, K.; Isaksson, E.; Wacker, L.; Mohan, R.; Ther, O.; Ikehara, M. Sea surface temperature in the Indian sector of the Southern Ocean over the Late Glacial and Holocene. Clim. Past 2020, 16, 1451–1467. [Google Scholar] [CrossRef]
  66. Tesi, T.; Belt, S.T.; Gariboldi, K.; Muschitiello, F.; Smik, L.; Finocchiaro, F.; Giglio, F.; Colizza, E.; Gazzurra, G.; Giordano, P.; et al. Resolving sea ice dynamics in the north-western Ross Sea during the last 2.6 ka: From seasonal to millennial timescales. Quat. Sci. Rev. 2020, 237, 106299. [Google Scholar] [CrossRef]
  67. Vorrath, M.E.; Müller, J.; Rebolledo, L.; Cárdenas, P.; Shi, X.; Esper, O.; Opel, T.; Geibert, W.; Muñoz, P.; Haas, C.; et al. Sea ice dynamics in the Bransfield Strait, Antarctic Peninsula, during the past 240 years: A multi-proxy intercomparison study. Clim. Past 2020, 16, 2459–2483. [Google Scholar] [CrossRef]
  68. Ashley, K.E.; Crosta, X.; Etourneau, J.; Campagne, P.; Gilchrist, H.; Ibraheem, U.; Greene, S.E.; Schmidt, S.; Eley, Y.; Massé, G.; et al. Exploring the use of compound-specific carbon isotopes as a palaeoproductivity proxy off the coast of Adélie Land, East Antarctica. Biogeosciences 2021, 18, 5555–5571. [Google Scholar] [CrossRef]
  69. Hogg, A.G.; Heaton, T.J.; Hua, Q.; Palmer, J.G.; Turney, C.S.M.; Southon, J.; Bayliss, A.; Blackwell, P.G.; Boswijk, G.; Bronk Ramsey, C.; et al. SHCal20 Southern Hemisphere Calibration, 0–55,000 Years cal BP. Radiocarbon 2020, 62, 759–778. [Google Scholar] [CrossRef]
  70. Esper, O.; Gersonde, R. New tools for the reconstruction of Pleistocene Antarctic sea ice. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 399, 260–283. [Google Scholar] [CrossRef]
  71. Gersonde, R.; Zielinski, U. The reconstruction of late Quaternary Antarctic sea-ice distribution-the use of diatoms as a proxy for sea-ice. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2000, 162, 263–286. [Google Scholar] [CrossRef]
  72. Crosta, X.; Shukla, S.K.; Ther, O.; Ikehara, M.; Yamane, M.; Yokoyama, Y. Last Abundant Appearance Datum of Hemidiscus karstenii driven by climate change. Mar. Micropaleontol. 2020, 157, 101861. [Google Scholar] [CrossRef]
  73. Crosta, X.; Sturm, A.; Armand, L.; Pichon, J.J. Late Quaternary sea ice history in the Indian sector of the Southern Ocean as recorded by diatom assemblages. Mar. Micropaleontol. 2004, 50, 209–223. [Google Scholar] [CrossRef]
  74. Lazzara, L.; Nardello, I.; Ermanni, C.; Mangoni, O.; Saggiomo, V. Light environment and seasonal dynamics of microalgae in the annual sea ice at Terra Nova Bay, Ross Sea, Antarctica. Antarct. Sci. 2007, 19, 83–92. [Google Scholar] [CrossRef] [Green Version]
  75. Perrin, R.A.; Lu, P.; Marchant, H.J. Seasonal variation in marine phytoplankton and ice algae at a shallow antarctic coastal site. Hydrobiologia 1987, 146, 33–46. [Google Scholar] [CrossRef]
  76. Ryan, K.G.; Hegseth, E.N.; Martin, A.; Davy, S.K.; O’Toole, R.; Ralph, P.J.; McMinn, A.; Thorn, C.J. Comparison of the microalgal community within fast ice at two sites along the Ross Sea coast, Antarctica. Antarct. Sci. 2006, 18, 583–594. [Google Scholar] [CrossRef] [Green Version]
  77. Kaczmarska, I.; Barbrick, N.E.; Ehrman, J.M.; Cant, G.P. Eucampia Index as an indicator of the Late Pleistocene oscillations of the winter sea-ice extent at the ODP Leg 119 Site 745B at the Kerguelen Plateau. Hydrobiologia 1993, 269/270, 103–112. [Google Scholar] [CrossRef]
  78. Fryxell, G.A.; Prasad, A.K.S.K. Eucampia antarctica var. recta (Mangin) stat. nov. (Biddulphiaceae, Bacillariophyceae): Life stages at the Weddell Sea ice edge. Phycologia 1990, 29, 27–38. [Google Scholar] [CrossRef]
  79. Allen, C.S. Proxy development: A new facet of morphological diversity in the marine diatom Eucampia antarctica (Castracane) Mangin. J. Micropalaeontology 2014, 33, 131–142. [Google Scholar] [CrossRef] [Green Version]
  80. Leventer, A.; Domack, E.; Barkoukis, A.; McAndrews, B.; Murray, J. Laminations from the Palmer Deep: A diatom-based interpretation. Paleoceanography 2002, 17, 1–15. [Google Scholar] [CrossRef]
  81. Armand, L.K.; Crosta, X.; Romero, O.; Pichon, J.-J. The biogeography of major diatom taxa in Southern Ocean sediments: 1. Sea ice related species. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 223, 93–126. [Google Scholar] [CrossRef]
  82. Zielinski, U.; Gersonde, R. Diatom distribution in Southern Ocean surface sediments (Atlantic sector): Implications for palaeoenvironmental reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol. 1997, 129, 213–250. [Google Scholar] [CrossRef]
  83. Świło, M.; Majewski, W.; Minzoni, R.T.; Anderson, J.B. Diatom assemblages from coastal settings of West Antarctica. Mar. Micropaleontol. 2016, 125, 95–109. [Google Scholar] [CrossRef]
  84. Crosta, X.; Pichon, J.; Labracherie, M. Distribution of Chaetoceros resting spores in modern peri-Antarctic sediments. Mar. Micropaleontol. 1997, 29, 283–299. [Google Scholar] [CrossRef]
  85. Belt, S.T.; Masse, G.; Rowland, S.J.; Poulin, M.; Michel, C.; LeBlanc, B. A novel chemical fossil of palaeo sea ice: IP25. Org. Geochem. 2007, 38, 16–27. [Google Scholar] [CrossRef]
  86. Brown, T.A.; Belt, S.T.; Tatarek, A.; Mundy, C.J. Source identification of the Arctic sea ice proxy IP25. Nat. Commun. 2014, 5, 4197. [Google Scholar] [CrossRef] [PubMed]
  87. Cabedo-Sanz, P.; Belt, S.T.; Knies, J.; Husum, K. Identification of contrasting seasonal sea ice conditions during the Younger Dryas. Quat. Sci. Rev. 2013, 79, 74–86. [Google Scholar] [CrossRef]
  88. Fahl, K.; Stein, R. Modern seasonal variability and deglacial/Holocene change of central Arctic Ocean sea-ice cover: New insights from biomarker proxy records. Earth Planet. Sci. Lett. 2012, 351, 123–133. [Google Scholar] [CrossRef] [Green Version]
  89. Xiao, X.; Fahl, K.; Stein, R. Biomarker distributions in surface sediments from the Kara and Laptev seas (Arctic Ocean): Indicators for organic-carbon sources and sea-ice coverage. Quat. Sci. Rev. 2013, 79, 40–52. [Google Scholar] [CrossRef] [Green Version]
  90. Massé, G.; Rowland, S.J.; Sicre, M.A.; Jacob, J.; Jansen, E.; Belt, S.T. Abrupt climate changes for Iceland during the last millennium: Evidence from high resolution sea ice reconstructions. Earth Planet. Sci. Lett. 2008, 269, 565–569. [Google Scholar] [CrossRef] [Green Version]
  91. Barbara, L.; Crosta, X.; Massé, G.; Ther, O. Deglacial environments in eastern Prydz Bay, East Antarctica. Quat. Sci. Rev. 2010, 29, 2731–2740. [Google Scholar] [CrossRef]
  92. Cárdenas, P.; Lange, C.B.; Vernet, M.; Esper, O.; Srain, B.; Vorrath, M.-E.; Ehrhardt, S.; Müller, J.; Kuhn, G.; Arz, H.W.; et al. Biogeochemical proxies and diatoms in surface sediments across the Drake Passage reflect oceanic domains and frontal systems in the region. Prog. Oceanogr. 2018, 174, 72–88. [Google Scholar] [CrossRef]
  93. Collins, L.G.; Allen, C.S.; Pike, J.; Hodgson, D.A.; Weckström, K.; Massé, G. Evaluating highly branched isoprenoid (HBI) biomarkers as a novel Antarctic sea-ice proxy in deep ocean glacial age sediments. Quat. Sci. Rev. 2013, 79, 87–98. [Google Scholar] [CrossRef]
  94. Massé, G.; Belt, S.T.; Crosta, X.; Schmidt, S.; Snape, I.; Thomas, D.N.; Rowland, S.J. Highly branched isoprenoids as proxies for variable sea ice conditions in the Southern Ocean. Antarct. Sci. 2011, 23, 487–498. [Google Scholar] [CrossRef] [Green Version]
  95. Sinninghe Damsté, J.S.; Rijpstra, W.I.C.; Coolen, M.J.L.; Schouten, S.; Volkman, J.K. Rapid sulfurisation of highly branched isoprenoid (HBI) alkenes in sulfidic Holocene sediments from Ellis Fjord, Antarctica. Org. Geochem. 2007, 38, 128–139. [Google Scholar] [CrossRef]
  96. Smik, L.; Belt, S.T.; Lieser, J.L.; Armand, L.K.; Leventer, A. Distributions of highly branched isoprenoid alkenes and other algal lipids in surface waters from East Antarctica: Further insights for biomarker-based paleo sea-ice reconstruction. Org. Geochem. 2016, 95, 71–80. [Google Scholar] [CrossRef] [Green Version]
  97. Vorrath, M.E.; Muller, J.; Esper, O.; Mollenhauer, G.; Haas, C.; Schefuss, E.; Fahl, K. Highly branched isoprenoids for Southern Ocean sea ice reconstructions: A pilot study from the Western Antarctic Peninsula. Biogeosciences 2019, 16, 2961–2981. [Google Scholar] [CrossRef] [Green Version]
  98. Belt, S.T. Source-specific biomarkers as proxies for Arctic and Antarctic sea ice. Org. Geochem. 2018, 125, 277–298. [Google Scholar] [CrossRef] [Green Version]
  99. Allen, C.S.; Pike, J.; Pudsey, C.J.; Leventer, A. Submillennial variations in ocean conditions during deglaciation based on diatom assemblages from the southwest Atlantic. Paleoceanography 2005, 20, 16. [Google Scholar] [CrossRef]
  100. Armand, L.; Ferry, A.J.; Leventer, A. Advances in palaeo sea ice estimation. In Sea Ice; Thomas, D.N., Ed.; John Wiley & Sons Ltd.: Chichester, UK, 2017; pp. 600–629. [Google Scholar]
  101. Ferry, A.J.; Prvan, T.; Jersky, B.; Crosta, X.; Armand, L.K. Statistical modeling of southern ocean marine diatom proxy and winter sea ice data: Model comparison and developments. Prog. Oceanogr. 2015, 131, 100–112. [Google Scholar] [CrossRef]
  102. Kang, S.H.; Fryxell, G.A. Fragilariopsis cylindrus (Grunow) Krieger-the most abundant diatom in water column assemblages of Antarctic marginal ice-edge zones. Polar Biol. 1992, 12, 609–627. [Google Scholar] [CrossRef]
  103. Pike, J.; Allen, C.S.; Leventer, A.; Stickley, C.E.; Pudsey, C.J. Comparison of contemporary and fossil diatom assemblages from the western Antarctic Peninsula shelf. Mar. Micropaleontol. 2008, 67, 274–287. [Google Scholar] [CrossRef]
  104. von Quillfeldt, C.H. The diatom Fragilariopsis cylindrus and its potential as an indicator species for cold water rather than for sea ice. Vie Milieu 2004, 54, 137–143. [Google Scholar]
  105. Ashley, K.E.; McKay, R.; Etourneau, J.; Jimenez-Espejo, F.J.; Condron, A.; Albot, A.; Crosta, X.; Riesselman, C.; Seki, O.; Massé, G.; et al. Mid-Holocene Antarctic sea-ice increase driven by marine ice sheet retreat. Clim. Past 2021, 17, 1–19. [Google Scholar] [CrossRef]
  106. Stickley, C.E.; Pike, J.; Leventer, A.; Dunbar, R.; Domack, E.W.; Brachfeld, S.; Manley, P.; McClennan, C. Deglacial ocean and climate seasonality in laminated diatom sediments, Mac.Robertson Shelf, Antarctica. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 227, 290–310. [Google Scholar] [CrossRef]
  107. Anderson, D.M. Attenuation of millennial-scale events by bioturbation in marine sediments. Paleoceanography 2001, 16, 352–357. [Google Scholar] [CrossRef]
  108. Liu, H.; Meyers, S.R.; Marcott, S.A. Unmixing deep-sea paleoclimate records: A study on bioturbation effects through convolution and deconvolution. Earth Planet. Sci. Lett. 2021, 564, 116883. [Google Scholar] [CrossRef]
  109. Crosta, X.; Pichon, J.J.; Burckle, L.H. Application of modern analog technique to marine Antarctic diatoms: Reconstruction of maximum sea-ice extent at the Last Glacial Maximum. Paleoceanography 1998, 13, 284–297. [Google Scholar] [CrossRef]
  110. Garrison, D.L. Antarctic Sea Ice Biota. Am. Zool. 1991, 31, 17–33. [Google Scholar] [CrossRef] [Green Version]
  111. Hegseth, E.N.; Von Quillfeldt, C.H. Low phytoplankton biomass and ice algal blooms in the Weddell Sea during the ice-filled summer of 1997. Antarct. Sci. 2002, 14, 231–243. [Google Scholar] [CrossRef]
  112. Kang, S.H.; Kang, J.S.; Lee, S.; Chung, K.H.; Kim, D.; Park, M.G. Antarctic phytoplankton assemblages in the marginal ice zone of the northwestern Weddell Sea. J. Plankton Res. 2001, 23, 333–352. [Google Scholar] [CrossRef] [Green Version]
  113. Kang, S.-H.; Fryxell, G.A. Phytoplankton in the Weddell Sea, Antarctica: Composition, abundance and distribution in water-column assemblages of the marginal ice-edge zone during austral autumn. Mar. Biol. 1993, 116, 335–348. [Google Scholar] [CrossRef]
  114. Kellogg, D.E.; Kellogg, T.B. Microfossil distributions in modern Amundsen Sea sediments. Mar. Micropaleontol. 1987, 12, 203–222. [Google Scholar] [CrossRef]
  115. Leventer, A.; Dunbar, R. Diatom flux in McMurdo Sound, Antarctica. Mar. Micropaleontol. 1987, 12, 49–64. [Google Scholar] [CrossRef]
  116. Ligowski, R.; Godlewski, M.; Lukowski, A. Sea ice diatoms and ice edge planktonic diatoms at the northern limit of the Weddell Sea pack ice. Proc. NIPR Symp. Polar Biol. 1992, 5, 9–20. [Google Scholar]
  117. Saggiomo, M.; Poulin, M.; Mangoni, O.; Lazzara, L.; De Stefano, M.; Sarno, D.; Zingone, A. Spring-time dynamics of diatom communities in landfast and underlying platelet ice in Terra Nova Bay, Ross Sea, Antarctica. J. Mar. Syst. 2017, 166, 26–36. [Google Scholar] [CrossRef]
  118. Schloss, I.; Estrada, M. Phytoplankton composition in the Weddell-Scotia Confluence area during austral spring in relation to hyrdrography. Polar Biol. 1994, 14, 77–90. [Google Scholar] [CrossRef]
  119. Smetacek, V.; Scharek, R.; Gordon, L.I.; Eicken, H.; Fahrbach, E.; Rohardt, G.; Moore, S. Early spring phytoplankton blooms in ice platelet layers of the southern Weddell Sea, Antarctica. Deep-Sea Res. Part A-Oceanogr. Res. Pap. 1992, 39, 153–168. [Google Scholar] [CrossRef]
  120. Cefarelli, A.O.; Ferrario, M.E.; Almandoz, G.O.; Atencío, A.G.; Akselman, R.; Vernet, M. Diversity of the diatom genus Fragilariopsis in the Argentine Sea and Antarctic waters: Morphology, distribution and abundance. Polar Biol. 2010, 33, 1463–1484. [Google Scholar] [CrossRef] [Green Version]
  121. Rigual-Hernández, A.S.; Pilskaln, C.H.; Cortina, A.; Abrantes, F.; Armand, L.K. Diatom species fluxes in the seasonally ice-covered Antarctic Zone: New data from offshore Prydz Bay and comparison with other regions from the eastern Antarctic and western Pacific sectors of the Southern Ocean. Deep Sea Res. Part II Top. Stud. Oceanogr. 2019, 161, 92–104. [Google Scholar] [CrossRef]
  122. Garrison, D.L.; Buck, C.E.; Fryxell, G. Algal assemblages in Antarctic pack ice and in ice-edge plankton. J. Phycol. 1987, 23, 564–572. [Google Scholar] [CrossRef]
  123. Cunningham, W.L.; Leventer, A. Diatom assemblages in surface sediments of the Ross Sea: Relationship to present oceanographic conditions. Antarct. Sci. 1998, 10, 134–146. [Google Scholar] [CrossRef]
  124. Leventer, A. Modern distribution of diatoms in sediments from the George-V-Coast, Antarctica. Mar. Micropaleontol. 1992, 19, 315–332. [Google Scholar] [CrossRef]
  125. Gleitz, M.; Grossmann, S.; Scharek, R.; Smetacek, V. Ecology of diatom and bacterial assemblages in water associated with melting summer sea ice in the Weddell Sea, Antarctica. Antarct. Sci. 1996, 8, 135–146. [Google Scholar] [CrossRef]
  126. Maddison, E.J.; Pike, J.; Leventer, A.; Domack, E.W. Deglacial seasonal and sub-seasonal diatom record from Palmer Deep, Antarctica. J. Quat. Sci. 2005, 20, 435–446. [Google Scholar] [CrossRef]
  127. Garrison, D.L.; Jeffries, M.O.; Gibson, A.; Coale, S.L.; Neenan, D.; Fritsen, C.; Okolodkov, Y.B.; Gowing, M.M. Development of sea ice microbial communities during autumn ice formation in the Ross Sea. Mar. Ecol. Prog. Ser. 2003, 259, 1–15. [Google Scholar] [CrossRef] [Green Version]
  128. Assmy, P.; Smetacek, V.; Montresor, M.; Klaas, C.; Henjes, J.; Strass, V.H.; Arrieta, J.M.; Bathmann, U.; Berg, G.M.; Breitbarth, E.; et al. Thick-shelled, grazer-protected diatoms decouple ocean carbon and silicon cycles in the iron-limited Antarctic Circumpolar Current. Proc. Natl. Acad. Sci. USA 2013, 110, 20633–20638. [Google Scholar] [CrossRef] [Green Version]
  129. Bathmann, U.V.; Scharek, R.; Klaas, C.; Dubischar, C.D.; Smetacek, V. Spring development of phytoplankton biomass and composition in major water masses of the Atlantic sector of the Southern Ocean. Deep-Sea Res. Part Ii-Top. Stud. Oceanogr. 1997, 44, 51–67. [Google Scholar] [CrossRef] [Green Version]
  130. Crosta, X.; Romero, O.; Armand, L.K.; Pichon, J.-J. The biogeography of major diatom taxa in Southern Ocean sediments: 2. Open ocean related species. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 223, 66–92. [Google Scholar] [CrossRef]
  131. Mohan, R.; Quarshi, A.A.; Meloth, T.; Sudhakar, M. Diatoms from the surface waters of the Southern Ocean during the austral summer of 2004. Curr. Sci. 2011, 100, 1323–1327. [Google Scholar]
  132. Fischer, G.; Gersonde, R.; Wefer, G. Organic carbon, biogenic silica and diatom fluxes in the marginal winter sea-ice zone and in the Polar Front Region: Interannual variations and differences in composition. Deep-Sea Res. Part Ii-Top. Stud. Oceanogr. 2002, 49, 1721–1745. [Google Scholar] [CrossRef]
  133. Burckle, L.H.; Cirilli, J. Origin of diatom ooze belt in the Southern Ocean; implications for late Quaterary paleoceanography. Micropaleontology 1987, 33, 82–86. [Google Scholar] [CrossRef]
  134. Mohan, R.; Shanvas, S.; Thamban, M.; Sudhakar, M. Spatial distribution of diatoms in surface sediments from the Indian sector of Southern Ocean. Curr. Sci. 2006, 91, 1495–1502. [Google Scholar]
  135. Medlin, L.; Priddle, J. (Eds.) Polar Marine Diatoms; British Antarctic Survey: Cambridge, UK, 1990. [Google Scholar]
  136. Esper, O.; Gersonde, R. Quaternary surface water temperature estimations: New diatom transfer functions for the Southern Ocean. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 414, 1–19. [Google Scholar] [CrossRef]
  137. Escutia, C.; Warnke, D.; Acton, G.D.; Barcena, A.; Burckle, L.; Canals, M.; Frazee, C.S. Sediment distribution and sedimentary processes across the Antarctic Wilkes Land margin during the Quaternary. Deep-Sea Res. Part Ii-Top. Stud. Oceanogr. 2003, 50, 1481–1508. [Google Scholar] [CrossRef]
  138. Garrison, D.L.; Buck, K.R. Sea-ice algal communities in the Weddell Sea: Species composition in ice and plankton assemblages. In Marine Biology of Polar Regions and Effects of Stress on Marine Organisms; Gray, J.S., Christiansen, M.E., Eds.; Wiley & Jons Ltd: Chichester, UK, 1985; pp. 103–122. [Google Scholar]
  139. Buffen, A.; Leventer, A.; Rubin, A.; Hutchins, T. Diatom assemblages in surface sediments of the northwestern Weddell Sea, Antarctic Peninsula. Mar. Micropaleontol. 2007, 62, 7–30. [Google Scholar] [CrossRef]
  140. Taylor, F.; McMinn, A.; Franklin, D. Distribution of diatoms in surface sediments of Prydz Bay, Antarctica. Mar. Micropaleontol. 1997, 32, 209–229. [Google Scholar] [CrossRef]
  141. Krebs, W.N.; Lipps, J.H.; Burckle, L.H. Ice diatom floras, Arthur Harbor, Antarctica. Polar Biol. 1987, 7, 163–171. [Google Scholar] [CrossRef]
  142. Maddison, E.J.; Pike, J.; Leventer, A.; Dunbar, R.; Brachfeld, S.; Domack, E.W.; Manley, P.; McClennen, C. Post-glacial seasonal diatom record of the Mertz Glacier Polynya, East Antarctica. Mar. Micropaleontol. 2006, 60, 66–88. [Google Scholar] [CrossRef]
  143. Arrigo, K.R. Sea Ice Ecosystems. Annu. Rev. Mar. Sci. 2014, 6, 439–467. [Google Scholar] [CrossRef]
  144. McMinn, A.; Hodgson, D. Summer phytoplankton succession in Ellis Fjord, eastern Antarctica. J. Plankton Res. 1993, 15, 925–938. [Google Scholar] [CrossRef]
  145. Rózanska, M.; Gosselin, M.; Poulin, M.; Wiktor, J.M.; Michel, C. Influence of environmental factors on the development of bottom ice protist communities during the winter–spring transition. Mar. Ecol. -Prog. Ser. 2009, 386, 43–59. [Google Scholar] [CrossRef] [Green Version]
  146. van Leeuwe, M.A.; Tedesco, L.; Arrigo, K.R.; Assmy, P.; Campbell, K.; Meiners, K.M.; Rintala, J.-M.; Selz, V.; Thomas, D.N.; Stefels, J. Microalgal community structure and primary production in Arctic and Antarctic sea ice: A synthesis. Elem. Sci. Anthr. 2018, 6, 1–25. [Google Scholar] [CrossRef] [Green Version]
  147. Ahn, I.Y.; Chung, H.; Kang, J.S.; Kang, S.H. Diatom composition and biomass variability in nearshore waters of Maxwell Bay, Antarctica, during the 1992/1993 austral summer. Polar Biol. 1997, 17, 123–130. [Google Scholar] [CrossRef]
  148. Whitaker, T.M.; Richardson, M.G. Morphology and Chemical composition of a natural population of an ice-associated Antarctic marine diatom Navicula glaciei. J. Phycol. 1980, 16, 250–257. [Google Scholar] [CrossRef]
  149. Scott, P.; McMinn, A.; Hosie, G. Physical parameters influencing diatom community structure in Eastern Antarctic sea-ice. Polar Biol. 1994, 14, 507–517. [Google Scholar] [CrossRef]
  150. Medlin, L.; Hasle, G. Some Nitzschia and related diatom species from fast ice samples in the Arctic and Antarctic. Polar Biol. 1990, 10, 451–479. [Google Scholar] [CrossRef]
  151. Fiala, M.; Kuosa, H.; Kopczynska, E.E.; Oriol, L.; Delille, D. Spatial and seasonal heterogeneity of sea ice microbial communities in the first-year ice of Terre Adelie area (Antarctica). Aquat. Microb. Ecol. 2006, 43, 95–106. [Google Scholar] [CrossRef] [Green Version]
  152. Medlin, L. The nomenclature and type locality of Berkeleya adeliensis (Bacillariophyceae): A correction. Plant Ecol. Evol. 2019, 152, 409–411. [Google Scholar] [CrossRef]
  153. Tanimura, Y.; Fukuchi, M.; Watanabe, K.; Moriwaki, K. Diatoms in Water Column and Sea-ice in Lutzhow-Holm Bay, Antarctica, and their Preservation in the Underlying Sediments. Bull. Natl. Sci. Museum. Tokyo Ser. C 1990, 18, 15–39. [Google Scholar]
  154. McMinn, A.; Bleakley, N.; Steinburner, K.; Roberts, D.; Trenerry, L. Effect of permanent sea ice cover and different nutrient regimes on the phytoplankton succession of fjords of the Vestfold Hills Oasis, eastern Antarctica. J. Plankton Res. 2000, 22, 287–303. [Google Scholar] [CrossRef] [Green Version]
  155. Fryxell, G.A. Thalssiosira australis Peragallo and T. lentiginosa (Janisch) G. Fryxell, comb. nov.: Two Antarctic diatoms (Bacillariophyceae). Phycologia 1977, 16, 95–104. [Google Scholar] [CrossRef]
  156. Ichinomiya, M.; Nakamachi, M.; Fukuchi, M.; Taniguchi, A. Population dynamics of an ice-associated diatom, Thalassiosira australis Peragallo, under fast ice near Syowa Station, East Antarctica, during austral summer. Polar Biol. 2008, 31, 1051–1058. [Google Scholar] [CrossRef]
  157. McMinn, A. Preliminary investigation of the contribution of fast-ice algae to the spring phytoplankton bloom in Ellis Fjord, eastern Antarctica. Polar Biol. 1996, 16, 301–307. [Google Scholar] [CrossRef]
  158. Ichinomiya, M.; Gomi, Y.; Nakamachi, M.; Honda, M.; Fukuchi, M.; Taniguchi, A. Temporal variations in the abundance and sinking flux of diatoms under fast ice in summer near Syowa Station, East Antarctica. Polar Sci. 2008, 2, 33–40. [Google Scholar] [CrossRef] [Green Version]
  159. Hasle, G.R.; Syvertsen, E.E. Chapter 2: Marine Diatoms. In Identifying Marine Phytoplankton; Tomas, C.R., Ed.; Academic Press: San Diego, CA, USA, 1997. [Google Scholar]
  160. Leventer, A.; Armand, L.; Harwood, D.; Jordan, R.W.; Ligowski, R. (Eds.) New Approaches and Progress in the Use of Polar Marine Diatoms in Reconstructing Sea-Ice Distribution. In Antarctica: A Keystone in a Changing World – Online Proceedings of the 10th ISAES X, Santa Barbara, CA, USA, 26 August–1 September 2007; USGS Open-File Report 2007-1047, Extended Abstract 005, 4p; USGS: Liston, VA, USA, 2007. [Google Scholar]
  161. Cremer, H.; Roberts, D.; McMinn, A.; Gore, D.; Melles, M. The Holocene Diatom Flora of Marine Bays in the Windmill Islands, East Antarctica. Bot. Mar. 2003, 46, 82–106. [Google Scholar] [CrossRef]
  162. Rivera, P.; Cruces, F. Fragilaria striatula lyngbye: Una diatomea marina muy poco conocida para chile. Gayana. Botánica 2002, 59, 35–41. [Google Scholar] [CrossRef]
  163. Kang, J.S.; Kang, S.H.; Lee, J.H.; Lee, S. Seasonal variation of microalgal assemblages at a s fixed station in King George Island, Antarctica, 1996. Mar. Ecol. Prog. Ser. 2002, 229, 19–32. [Google Scholar] [CrossRef]
  164. Hasle, G.R.; Medlin, L.K.; Syvertsen, E.E. Synedropsis gen. nov., a genus of araphid diatoms associated with sea ice. Phycologia 1994, 33, 248–270. [Google Scholar] [CrossRef]
  165. Cefarelli, A.O.; Ferrario, M.E.; Vernet, M. Diatoms (Bacillariophyceae) associated with free-drifting Antarctic icebergs: Taxonomy and distribution. Polar Biol. 2016, 39, 443–459. [Google Scholar] [CrossRef]
Figure 1. Map showing sites (red-filled circles) of published marine sediment core records with diatom-based sea ice proxies for all or part of the past 2 ka. Sites are numbered corresponding to the map references given in Table 1. AS—Amundsen Sea; BHS—Bellingshausen Sea; DUS—Dumont d’Urville Sea; PB—Prydz Bay. Winter and summer sea ice limits depict the 1981–2010 median September and February sea ice extents (Source: /DATASETS/NOAA/G02135/south/monthly/).
Figure 1. Map showing sites (red-filled circles) of published marine sediment core records with diatom-based sea ice proxies for all or part of the past 2 ka. Sites are numbered corresponding to the map references given in Table 1. AS—Amundsen Sea; BHS—Bellingshausen Sea; DUS—Dumont d’Urville Sea; PB—Prydz Bay. Winter and summer sea ice limits depict the 1981–2010 median September and February sea ice extents (Source: /DATASETS/NOAA/G02135/south/monthly/).
Geosciences 12 00282 g001
Figure 2. Map showing cores sites where F. c + cy% is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 2. Map showing cores sites where F. c + cy% is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g002
Figure 3. Map showing cores sites where F. curta % is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Figure 3. Map showing cores sites where F. curta % is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Geosciences 12 00282 g003
Figure 4. Map showing cores sites where F. c + cy/T. antarctica is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 4. Map showing cores sites where F. c + cy/T. antarctica is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g004
Figure 5. Map showing cores sites where F. curta/F. kerguelensis is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 5. Map showing cores sites where F. curta/F. kerguelensis is used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g005
Figure 6. Map showing cores sites where groups of taxa including F. curta are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Figure 6. Map showing cores sites where groups of taxa including F. curta are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Geosciences 12 00282 g006
Figure 7. Map showing cores sites where groups of taxa excluding F. curta are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 7. Map showing cores sites where groups of taxa excluding F. curta are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g007
Figure 8. Map showing cores sites where all other proxies based on diatom taxa are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 8. Map showing cores sites where all other proxies based on diatom taxa are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g008
Figure 9. Map showing cores sites where highly branched isoprenoids (HBIs) are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Figure 9. Map showing cores sites where highly branched isoprenoids (HBIs) are used to reconstruct sea ice conditions. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP.
Geosciences 12 00282 g009
Figure 10. Maps of regions with greatest density of sea ice records from marine sediment cores, showing the diversity of proxy types and sample resolution. The symbol size reflects the resolution of the record and the color of the symbol indicates the proxy type (see Table 1 and Table 3).
Figure 10. Maps of regions with greatest density of sea ice records from marine sediment cores, showing the diversity of proxy types and sample resolution. The symbol size reflects the resolution of the record and the color of the symbol indicates the proxy type (see Table 1 and Table 3).
Geosciences 12 00282 g010
Figure 11. Map of sites that include F. curta in at least one proxy record. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Figure 11. Map of sites that include F. curta in at least one proxy record. The symbol size reflects the resolution of the record and the color of the symbol indicates the basal age of the record between 0 and 2 ka BP. The hollow circle symbols are used for discrete and/or discontinuous records from laminated sediments to illustrate the resolution only.
Geosciences 12 00282 g011
Figure 12. Plot showing the age, duration (horizontal lines, upper axis), and minimum resolution (open circles, bottom axis) for all downcore proxy records (site numbers, left axis) of sea ice included in this review.
Figure 12. Plot showing the age, duration (horizontal lines, upper axis), and minimum resolution (open circles, bottom axis) for all downcore proxy records (site numbers, left axis) of sea ice included in this review.
Geosciences 12 00282 g012
Table 1. Marine core sites with diatom-based sea ice proxy records for all or part of the past 2 ka.
Table 1. Marine core sites with diatom-based sea ice proxy records for all or part of the past 2 ka.
Map Ref.Core/Site IDLatLongLocation NameReferences
1WG35−77.989162.853Granite Harbor, Western Ross Sea[11]
2Multiple−77.668165.500McMurdo Sound; Western Ross Sea[12]
3WG17−77.000162.850Granite Harbor, Western Ross Sea[11]
4KC208.09−76.972162.876Granite Harbor, Western Ross Sea[11]
5KC31−75.700165.418Western Ross Sea[13]
6KC37−74.499167.744Western Ross Sea[13]
7KC39−74.474173.474Western Ross Sea[13]
8BAY05-43c−74.000166.050Wood Bay, Western Ross Sea[14]
9ANTA99-cJ5−73.817175.650Joides Basin, Western Ross Sea[14]
10KC17−73.420−102.827Ferrero Bay, Amundsen Sea Embayment[15]
11KC15−73.360−101.836Ferrero Bay, Amundsen Sea Embayment[15]
12BAY05-20c−72.300170.050Cape Hallet, Ross Sea[14]
13AM02−69.71372.640Amery Ice Shelf, Prydz Bay[16]
14CO1011−68.82777.760Flag Island Inlet, Prydz Bay[17]
15CO1010−68.81777.833Filla Island Inlet, Prydz Bay[17]
16JPC24−68.69476.709Svenner Channel, Prydz Bay[18,19,20]
17KROCK-15-GC29−68.66476.696Prydz Bay[21]
18Abel Bay−68.65078.400Long Fjord, Ingrid Christensen Coast[22]
19Watts Basin−68.60378.213Ellis Fjord, Ingrid Christensen Coast[23]
20Deep Basin−68.56078.199Ellis Fjord, Ingrid Christensen Coast[23]
21Platcha Bay−68.51578.478Long Fjord, Ingrid Christensen Coast[22]
22JPC43−68.257−66.962Neny Fjord, Marguerite Bay, AP[24]
23TPC522−67.856−68.205Marguerite Bay, AP[25]
24KROCK-125-GC2−67.47464.973Nielsen Bay, MacRobertson Land[26]
25GC1−67.180−66.797Lallemand Fjord, AP[27]
26JPC41−67.13162.990Iceberg Alley, MacRobertson Land[28]
27GC 5−67.05969.016MacRobertson Shelf, Prydz Bay[29]
28KROCK-128-GC1−66.98363.154Iceberg Alley, MacRobertson Land[26]
29CB2010−66.906142.436Commonwealth Bay, Adélie Land[30]
30PG1433−66.465110.572Browning Bay, Windmill Is., Wilkes Land[31]
31PG1430−66.453110.498Peterson Inlet, Windmill Is., Wilkes Land[31]
32MD03-2597−66.412140.421Dumont d’Urville Trough, Adélie Land[32]
33DTCI2010; 318-1357B−66.411140.445Dumont d’Urville Trough, Adélie Land[33,34]
34PG1173−66.267100.750Bunger Oasis, Wilkes Land[35]
35MD03-2601−66.052138.557Dumont d’Urville Trough, Adélie Land[20,36,37,38,39]
36WAP13-GC47−65.613−64.759Bigo Bay, AP[40]
37JPC10/178-1098−64.883−64.200Palmer Deep, AP[41,42,43]
38PD92-30−64.862−64.208Palmer Deep, AP[44,45]
39TC46/GC47−64.588−64.805Anvers Shelf, AP[46]
40MTC18A−64.772−62.829Andvord Drift, Gerlache Strait, AP[47]
41GC 02−64.000−64.000Anvers Shelf, AP[48]
42KC2B−63.971−57.759Herbert Sound, James Ross Island, AP[49]
43MTC38C, JPC38−63.717−57.411Vega Drift, Prince Gustav Channel, AP[47,50]
44PC61−63.389−60.319Western Basin, Bransfield Strait, AP[51]
45JPC02−63.343−55.887Firth Of Tay, Joinville Island, AP[52]
46A-3−63.168−59.302Orleans Trough, Bransfield Strait, AP[53,54]
47JPC36−63.089−55.411Perseverance Drift, Joinville Island, AP[55]
48GC 03−63.000−64.000Anvers Shelf, AP[48]
49A-6−62.912−59.97Western Basin, Bransfield Strait, AP[53,54]
50Gebra-1−62.589−58.542Central Basin, Bransfield Strait, AP[53,56]
511B−62.282−58.754Maxwell Bay, South Shetland Is.[57]
52MC-01−62.202−58.727Marian Cove, South Shetland Is.[58]
53WB2−62.200−60.700Outer Shelf, South Shetland Is.[59]
54CB2−62.191−58.833Collins Harbour, South Shetland Is.[60]
55Gebra-2−61.943−55.17Eastern Basin, Bransfield Strait, AP[53,56]
5613PC (TN057-13PC4)−53.2005.1000Atlantic-Indian Ridge, S Atlantic Ocean[61]
57E27-23−59.618155.238Emerald Basin, SE Indian Ocean[62]
58PS1652-2−53.6645.100Atlantic-Indian Ridge, S Atlantic Ocean[63]
59PS1768-8−52.5934.476Atlantic-Indian Ridge, S Atlantic Ocean[63]
60177-1094/PS2090-1−53.1795.132Atlantic-Indian Ridge, S Atlantic Ocean[63,64]
61PS2102-2−53.0734.986Atlantic-Indian Ridge, S Atlantic Ocean[63]
6217PC (TN057-17PC1)−50.0006.000Cape Basin, S Atlantic Ocean[61]
63COR1GC−54.26739.766Conrad Rise, SW Indian Ocean[65]
64HLF17-1−72.308−172.054Edisto Inlet, Ross Sea[66]
65PS97/056-1−64.757−60.442Gilbert Strait, Bransfield Strait, AP[67]
66PS97/068-2−63.168−59.302Orleands Trough, Bransfield Strait, AP[67]
67PS97/072-2−62.007−56.065Eastern Basin, Bransfield Strait, AP[67]
68DTGC2011−66.408140.441Dumont d’Urville Trough, Adélie Land[68]
AP—Antarctic Peninsula
Table 2. The most common diatom taxa in Antarctic sea ice proxies of the past 2 ka.
Table 2. The most common diatom taxa in Antarctic sea ice proxies of the past 2 ka.
Diatom Speciesn
Fragilariopsis curta77
Fragilariopsis cylindrus54
Thalassiosira antarctica21
Fragilariopsis sublinearis21
Fragilariopsis kerguelensis19
Fragilariopsis obliquecostata19
Fragilariopsis rhombica16
Fragilariopsis ritscheri14
Fragilariopsis vanheurckii14
Actinocyclus actinochilus12
Porosira glacialis12
Fragilariopsis separanda10
Table 3. Number of records for different proxy categories.
Table 3. Number of records for different proxy categories.
Categoriesn
1.F. curta + F. cylindrus (F. c + cy) %10
2.F. curta %23
3.F. c + cy/T. antarctica3
4.F. curta/F. kerguelensis4
5.Groups including F. curta36
6.Groups excluding F. curta7
7.Other15
8.HBIs14
TOTAL112
Table 4. Summary information on the ages, resolution and proxy types of all sites.
Table 4. Summary information on the ages, resolution and proxy types of all sites.
Map Ref.Core/Site IDApproximate Dates *Approx. SR (m/ka)Resolution (Lowest) Resolution (Highest) Proxy Categories References
1WG351.2 to ~0.6 ka BP1.6760322[11]
2Multiple0.5 to ~0 ka BP1100562[12]
3WG171.2 to ~0.6 ka BP1.1760322[11]
4KC208.091.3 to ~0 ka BP2.4665332[11]
5KC312.0 to ~0 14C ka BP0.194002222[13]
6KC372.0 to ~0 14C ka BP0.114002222[13]
7KC392.0 to ~0 14C ka BP0.334002222[13]
8BAY05-43c2.0 to ~0.0 ka BP2.1338152[14]
9ANTA99-cJ52.0 to ~1.0 ka BP0.562392[14]
10KC172.0 to ~0 ka BP0.1310005007[15]
11KC152.0 to ~0 ka BP0.1310005007[15]
12BAY05-20c2.0 to ~0.1 ka BP1.5848192[14]
13AM022.0 to ~0.0 14C ka BP0.054002222[16]
14CO10112.0 to ~0 ka BP0.61000, 1000500, 5005, 7[17]
15CO10102.0 to ~0 ka BP2400, 400222, 2225, 7[17]
16JPC242.0 to ~0.6 ka BP1.4370, 70, 70, 7036, 36, 36, 365, 5, 7, 8[18,19,20]
17KROCK-15-GC292.0 to ~0 14C ka BP0.134002225[21]
18Abel Bay2.0 to ~0 14C ka BP0.45200, 20051, 512, 6[22]
19Watts Basin2.0 to ~0.2 14C ka BP1.3945186[23]
20Deep Basin2.0 to ~0.8 14C ka BP2.4230126[23]
21Platcha Bay2.0 to ~0.0 14C ka BP0.4200, 200105, 1052, 6[22]
22JPC432.0 to ~0 ka BP2.35100515[24]
23TPC5222.0 to ~0.8 ka BP0.332401334[25]
24KROCK-125-GC22.0 to ~0 14C ka BP0.9100515[26]
25GC12.0 to ~0 ka BP0.5200, 200105, 1052, 5[27]
26JPC412.0 to ~0 ka BP^12<1^ 2[28]
27GC 51.3 to ~0 14C ka BP0.147001752[29]
28KROCK-128-GC12.0 to ~0.2 14C ka BP0.65100515[26]
29CB20100.25 to ~0 ka BP1.46, 63, 37, 8[30]
30PG14332.0 to ~0.3 ka BP2.1885446[31]
31PG14302.0 to ~0 ka BP2.65100516[31]
32MD03-25972.0 to ~0.7 ka BP^20.54<1^ 5[32]
33DTCI2010; 318-1357B0.04 to ~0 ka BP; 2,0 to 0.1 ka BP **18.13, 19.1419, 0.4, 19, 0.410, <0.4, 3, <0.41, 7, 8, 8[33,34]
34PG11732.0 to ~0 ka BP1.4550205[35]
35MD03-26012.0 to ~1.0 ka BP510, 25, 10, 25, 25, 50<10, 10, <10, 10, 10, 261, 4, 5, 5, 7, 8[20,36,37,38,39]
36WAP13-GC472.0 to ~0.1 ka BP0.841901004[40]
37JPC10/178-10982.0 to ~0.2 ka BP, 2.0 to 0 ka BP **1.16, 2.00173, 100, 100, 4391, 51, 51, 171, 5, 7, 8[41,42,43]
38PD92-302.0 to ~0.3 14C ka BP2.8285, 8544, 441, 3[44,45]
39TC46/GC472.0 to ~0.1 ka BP0.421001904[46]
40MTC18A0.13 to ~0 ka BP3.463, 31, 15, 8[47]
41GC 022.0 to ~0.9 14C ka BP0.16220, 220122, 1223, 5[48]
42KC2B2.0 to ~0 ka BP1.152001056[49]
43MTC38C, JPC380.08 to ~0 ka BP, 2.0 to ~0 ka BP **5.75, 2.52, 2, 50, 20<1, <1, 20, <205, 8, 5, 8[47,50]
44PC612.0 to ~0 ka BP0.52001052[51]
45JPC022.0 to ~0 ka BP112001057[52]
46A-31.7 to ~0 ka BP2.5985445[53,54]
47JPC360.8 to ~0 ka BP17.540, 4021, 215, 7[55]
48GC 032.0 to ~0.1 14C ka BP0.32188, 18899, 993, 5[48]
49A-61.8 to ~0.1 ka BP0.8180415[53,54]
50Gebra-12.0 to ~0.2 ka BP0.7290465[53,56]
511B2.0 to ~0.1 ka BP4.219504757[57]
52MC-011.7 to ~0 14C ka BP2.1817, 17<17, <172, 7[58]
53WB21.5 to ~0 14C ka BP1.838152[59]
54CB22.0 to ~0 14C ka BP2.220, 20<20, <202, 7[60]
55Gebra-21.6 to ~0.1 ka BP2.9390155[53,56]
5613PC (TN057-13PC4)2.0 to ~0 ka BP0.3520<205[61]
57E27-232.0 to ~1.5 ka BP0.55001005[62]
58PS1652-22.0 to ~0.6 ka BP0.64280, 280156, 1561, 5[63]
59PS1768-82.0 to ~0.8 ka BP0.1600, 600300, 3001, 5[63]
60177-1094/PS2090-12.0 to 1.0 ka BP, 2.0 to ~0.8 ka BP **0.25, 0.30240, 500, 240133, 250, 1331, 1, 5[63,64]
61PS2102-22.0 to ~0.2 ka BP0.32190, 190100, 1001, 5[63]
6217PC (TN057-17PC1)2.0 to ~0.0 ka BP0.2540165[61]
63COR1GC2.0 to 1.0 ka BP0.2552, 5226, 265, 5[65]
64HLF17-12.0 to 0.0 ka BP7.2520, 20<20, <207, 8[66]
65PS97/056-10.17 to ~0 ka BP217, 17, 99, 9, 42, 5, 8[67]
66PS97/068-20.22 to ~0 ka BP1.9511, 11, 66, 6, 22, 5, 8[67]
67PS97/072-20.19 to ~0 ka BP2.0510, 10, 56, 6, 22, 5, 8[67]
68DTG20110.42 to ~0 ka BP11.174, 4<4, <41, 8[68]
SR—sedimentation rate(s); AP—Antarctic Peninsula; 14C ka BP = corrected 14C age, ka BP = calendar age (based on calibrated C14 dates, Pb210 ages and/or extrapolated from sedimentation rates). * Where more than one age model published, the one in calendar years or the most recent is used; ** Different age ranges are listed in order of publication. ^ Discrete sample horizons from laminated sediments—date range is not representative and resolution is discontinuous. ∆ Resolution calculated from minimum & maximum number of samples (2 to 5, 6 to 10, 11 to 20, 21 to 40, 41 to 100, >100) divided by the age interval (≤2 ka). † Proxy categories are listed in Table 3.
Table 5. Composition of proxy groups including Fragilariopsis curta.
Table 5. Composition of proxy groups including Fragilariopsis curta.
CompositionSitesReferences
Fragilariopsis curta, F. cylindrus, Navicula glaceii, and F. rhombica.14, 15[17]
Fragilariopsis curta, F. cylindrus, F. sublinearis, F. obliquecostata, F. vanheurckii, and Porosira glacialis16, 35[18,19]
Cluster groups: Coastal-CRS—Fragilariopsis curta, F. cylindrus, F. rhombica, and Pseudonitzschia turgiduloides; Shelf-CRS—F. curta, F. cylindrus, F. rhombica, Pentalamina corona a, Porosira glacialis, and Thalassiosira antarctica17, 24, 28[21,26]
PCA: Fragilariopsis curta, F. cylindrus, F. obliquecostata, F. ritscheri, F. sublinearis, and F. vanheurckii22[24]
Cluster group 1: dominated by Thalassiosira antarctica (T1 and T2) and Fragilariopsis curta, with F. cylindrus, F. rhombica, Navicula spp., Pentalamina corona a, Pseudonitzschia turgiduloides, Rhizosolenia spp., and Synedra spp. also present25[27]
Spring sea ice represented by laminae types A1, A2, and A3: CRS and Fragilariopsis spp. (with F. curta, F. cylindrus, and F. rhombica dominant)32[32]
Actinocyclus actinochiIus, Berkeleya rutilans, Entomoneis kufferathiib, Eucampia antarctica, Fragilariopsis angulatac, F. curta, F. cylindrus, F. obliquecostata, F. ritscheri, F. sublinearis, Porosira glacialls, P. pseudodenticulata, and Distephanus speculumd34[35]
The Fragilariopsis curta group: F. curta, F. cylindrus, and F. vanheurckii; and the Fragilariopsis cryophilic group: F. obliquecostata, F. ritscheri, and F. sublinearis35[38]
Cluster groups: Cocconeis assemblage dominated by Fragilariopsis curta and Thalassiosira antarctica, with Cocconeis as a unique indicator, and Corethron assemblage dominated by F. curta and F. cylindrus, with Corethron spp. and Pseudonitzschia turgiduloides as indicators37[43]
Fragilariopsis curta, F. cylindrus, and F. vanheurckii40, 43[47]
Sea ice taxa (not itemised) assume at least: Fragilariopsis curta and F. cylindrus (based on the use of ‘F. c + cy/T. antarctica’ ratio in the same publication)41, 48[48]
Fragilariopsis curta, F. cylindrus, F. sublinearis, and F. vanheurckii43, 47[50,55]
The sea ice taxa group: Fragilariopsis curta, F. cylindrus, F. sublinearis, F. obliquecostata, and F. vanheurckii46, 49, 50, 55[53,54,56]
MAT (31–33 taxa): Actinocyclus actinochilus, Alveus marinus e, Azpeitia tabularis, the Chaetoceros resting spore group, Fragilariopsis curta, F. cylindrus, F. doliolus, F. kerguelensis, F. obliquecostata, F. rhombica, F. ritscheri, F. separanda, F. sublinearis, Hemidiscus cuneiformis, Porosira glacialis, P. pseudodenticulata, Rhizosolenia antennata f. semispina, R. styliformis, Roperia tesselata, Stellarima microtrias, Thalassionema nitzschioides, T. nitzschioides var. lanceolata, T. nitzschioides var. parva, the Thalassiosira antarctica group (warm and cold morphologies e), the T. eccentrica group, Shionodiscus gracilis f, T. lentiginosa, S. oestrupii f, T. oliveriana, T. tumida, Thalassiothrix spp., and Trichotoxon reinboldii [72,73]56, 62, 63[61,65]
GAM: Actinocyclus actinochilus, Fragilariopsis curta, F. cylindrus, and Thalassiosira lentiginosa57[62]
MAT (28 taxa): Actinocyclus actinochilus, Azpeitia tabularis, Fragilariopsis curta, F. cylindrus, F. doliolus, F. kerguelensis, F. obliquecostata, F. rhombica, F. ritscheri, F. separanda, F. sublinearis, Hemidiscus cuneiformis, Nitzschia bicapitata, Porosira pseudodenticulata, Pseudonitzschia turgiduloides, Rhizosolenia spp., R. antennata f. semispina, R. bergonii, Roperia tesselata, Thalassionema nitzschioides f. 1, T. nitzschioides var. parva, T. nitzschioides var. lanceolata + T. nitzschioides var. capitulata, Thalassiosira antarctica, Shionodiscus gracilis e, T. lentiginosa, S. oestrupii e, T. oliveriana, and Thalassiothrix antarctica [70]58, 59, 60, 61, 65, 66, 67[63,67]
The sea ice group: Fragilariopsis curta, F. cylindrus, F. obliquecostata, F. ritscheri, Porosira glacialis, and Thalassiosira tumida63[65]
a—Bolidophyceae Parmales; b—Synonym: Amphiprora kufferathii: c—Synonym: F. rhombica; d—Silicoflagellate; e—excluded from MAT-31 f—previously assigned to Thalassiosira genus.
Table 6. Composition of proxy groups excluding Fragilariopsis curta.
Table 6. Composition of proxy groups excluding Fragilariopsis curta.
CompositionSitesReferences
Fast ice index: Entomoneis kjellmannii, Nitzschia stellata, Berkeleya adelienses, Thalassiosira australis, Pleurosigma directum and Pinnularia quadreata18–21[22,23]
Fragilaria striulata, Navicula glaceii, and Synedropsis spp.30, 31[31]
Prolonged sea ice inferred from Navicula spp.42[49]
Table 7. Composition of all ‘other’ diatom proxies.
Table 7. Composition of all ‘other’ diatom proxies.
CompositionSitesReferences
Entomoneis kjellmanii %14 and 15[17]
Eucampia index51[57]
F. cylindrus %29, 33, 52, and 54[30,33,58,60]
F. group/T. antarctica (T2)47[55]
Porosira glacialis/T antarctica16 and 35[20]
Pennate–centric ratios10 and 11[15]
Diatom concentrations37 and 45[42,52]
Frequency of dark–light laminae64[66]
Table 8. Summary of main advantages and limitations of the eight proxy types.
Table 8. Summary of main advantages and limitations of the eight proxy types.
Proxy TypeAdvantagesLimitations
1. F. curta + F. cylindrusWell-established links between F. curta and F. cylindrus with sea ice melt waters, relative abundance in sediments over large areas of the deep Southern Ocean consistently linked with seasonal sea ice, based on large reference dataset of sediment traps and core tops throughout the deep South AtlanticProxy only ‘calibrated’ for Atlantic sector of the deep Southern Ocean
Both F. curta and F. cylindrus are widely preserved in marine sediments of the deep and continental shelf areas of the Southern OceanF. cylindrus not exclusive to sea ice meltwater but also found in glacial meltwaters and in stratified waters of the Ross Sea
Can be successfully applied at several sites in the deep South Atlantic and over a variety of late Quaternary timescalesF. nana often not differentiated from F. cylindrus
Well-defined morphology that minimizes risks of misidentificationExcludes broader assemblage information
Small changes in the relative abundances of dominant species can have a pronounced impact on the percentages of minor taxa
2. F. curtaWell-established association with seasonal sea ice throughout the deep Southern Ocean and the continental shelfAssociation with seasonal sea ice fails over short/annual timeframe
Widely preserved and common in sediments throughout the deep Southern Ocean and the continental shelfRelative abundances vary greatly from site to site
Well-defined morphology that minimizes risks of misidentificationExcludes broader assemblage information
3. F. c + cy/T. antarcticaRatio reduces influence of % changes in dominant speciesSeveral morphotypes of T. antarctica and ambiguity over exact ecological associations—complicate identification and interpretation of ratio
Well-established association between F. c + cy and seasonal sea ice (see above)Subjective boundaries between T. antarctica morphotypes compound identification difficulties and in particular compromise cross-site comparison
4. F. curta/F.kerguelensisRatio reduces influence of % changes in dominant speciesAbundance of F. kerguelensis (strongly silicified) can be increased in sediments affected by dissolution
F. kerguelensis is well established as an open ocean diatom, so the ratio is a robust indicator for the relative influence of open ocean versus sea ice conditions
5. Groups including F. curtaGroups incorporate information of broader assemblageDifficult to have a group that is appropriate for a wide range of continental shelf sites
Groups can be tailored to local ocean conditions and/or assemblage compositionTransfer functions primarily built on reference data from deep ocean sites where seasonal sea ice expands and retreats along the north–south axis are not necessarily appropriate for continental shelf sites where sea ice distribution is more complex
Many groups are statistically definedTransfer functions only reliable within area of reference sites
Transfer functions incorporate most species and produce quantitative results
6. Groups excluding F. curtaGroups mostly tailored to specific nearshore conditions or particular types of sea iceMost groups not suitable for application in offshore regions of the continental shelf or deep ocean regions
Sensitive to different types of sea iceSensitive to different types of sea ice rather than duration or distribution of sea ice cover
7. OtherMany based on routine information that can be applied alongside other species or assemblage proxies (e.g., Diatom concentration, pennate–centric, P. glaciailis/T. antarctica)Most are only indirectly linked to sea ice
Can be widely applied throughout the continental shelf and deep ocean sitesMay require additional analyses beyond the standard assemblage composition (e.g., morphometrics of the marine diatom Eucampia antarctica)
8. HBIsLess time-intensive analyses, creating generally higher resolution records than traditional diatom assemblage dataStill requires considerable effort and personnel time to produce records
Absence of diene (HBI-II) can result from both open ocean and permanent sea ice conditions
Diene (HBI-II) linked to the sea ice diatom Berkeleya adeliensisEnvironmental controls on the production of triene (HBI-III) are still ambiguous
Widely preserved in sediments and robust at timescales up to at least 100 kaGenerally requires some validation with diatom assemblage data to aid interpretation
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Allen, C.S.; Weich, Z.C. Variety and Distribution of Diatom-Based Sea Ice Proxies in Antarctic Marine Sediments of the Past 2000 Years. Geosciences 2022, 12, 282. https://doi.org/10.3390/geosciences12080282

AMA Style

Allen CS, Weich ZC. Variety and Distribution of Diatom-Based Sea Ice Proxies in Antarctic Marine Sediments of the Past 2000 Years. Geosciences. 2022; 12(8):282. https://doi.org/10.3390/geosciences12080282

Chicago/Turabian Style

Allen, Claire S., and Zelna C. Weich. 2022. "Variety and Distribution of Diatom-Based Sea Ice Proxies in Antarctic Marine Sediments of the Past 2000 Years" Geosciences 12, no. 8: 282. https://doi.org/10.3390/geosciences12080282

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