terrestrial biodiversity in antarctica – recent advances and future challenges

13
Terrestrial biodiversity in Antarctica e Recent advances and future challenges Peter Convey British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 0ET, United Kingdom Received 7 December 2009; revised 22 February 2010; accepted 16 March 2010 Available online 25 March 2010 Abstract Although its major components have been known almost since the earliest exploring expeditions, even today the terrestrial biota of Antarctica is surprisingly poorly described in detail. It is clear that most currently ice-free ground in Antarctica would have been covered and scoured by glacial advances at the Last Glacial Maximum or previous maxima. Exceptions to this generalisation include parts of the Victoria Land Dry Valleys and some inland nunataks and mountain ranges at altitude, which host their own largely unique biota. However, as new baseline survey data have become available, in combination with the application of tech- niques of molecular biological analysis, new evidence has been obtained indicating that long-term persistence and regional isolation is a feature of the Antarctic terrestrial biota whose generality has not previously been appreciated. As well as creating a new paradigm in which to consider the evolution and adaptation of Antarctic terrestrial biota, this opens important new cross- disciplinary linkages in the field of understanding the geological and glaciological history of the continent itself. Superimposed on this emerging historical template of Antarctic biogeography, this biota now faces the twin challenges of responding to the complex processes of climate change facing some parts of the continent, and the direct impacts associated with human occupation and travel to and between the spatially very limited areas of terrestrial habitat. Ó 2010 Elsevier B.V. and NIPR. All rights reserved. Keywords: Biogeography; Biological invasions; Environmental change; Environmental manipulation; Phylogeography 1. Introduction Antarctica is the only one of the Earth’s continents not to have a long-term history of human contact and inhabitation. The continent itself was first landed on approximately two centuries ago (the northern Antarctic Peninsula), and East Antarctica only a little over a century ago, while the sub-Antarctic islands were mostly discovered and their marine-based living resources rapidly over-exploited over the last two to three centuries. Antarctica is an ice-bound continent, only 0.34% of its area is currently ever free of snow or ice in the form of terrestrial ecosystems including exposed nunataks, cliffs and seasonally snow and ice-free areas. Even then, most of these appear biologically barren, with the majority in areal terms being at higher altitude inland locations (see Table 1 in Convey et al., 2009a). Visibly developed terrestrial ecosystems are best represented in coastal areas, particularly along the Antarctic Peninsula, and the ‘oases’ of the East Antarctic coastline. Most ice- free areas are small and ‘island-like’ (Bergstrom and Chown, 1999), isolated by hostile ice and sea across E-mail address: [email protected]. 1873-9652/$ - see front matter Ó 2010 Elsevier B.V. and NIPR. All rights reserved. doi:10.1016/j.polar.2010.03.003 Available online at www.sciencedirect.com Polar Science 4 (2010) 135e147 http://ees.elsevier.com/polar/

Upload: peter-convey

Post on 29-Nov-2016

215 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

Available online at www.sciencedirect.com

Polar Science 4 (2010) 135e147http://ees.elsevier.com/polar/

Terrestrial biodiversity in Antarctica e Recent advances andfuture challenges

Peter Convey

British Antarctic Survey, High Cross, Madingley Road, Cambridge CB3 0ET, United Kingdom

Received 7 December 2009; revised 22 February 2010; accepted 16 March 2010

Available online 25 March 2010

Abstract

Although its major components have been known almost since the earliest exploring expeditions, even today the terrestrial biotaof Antarctica is surprisingly poorly described in detail. It is clear that most currently ice-free ground in Antarctica would have beencovered and scoured by glacial advances at the Last Glacial Maximum or previous maxima. Exceptions to this generalisationinclude parts of the Victoria Land Dry Valleys and some inland nunataks and mountain ranges at altitude, which host their ownlargely unique biota. However, as new baseline survey data have become available, in combination with the application of tech-niques of molecular biological analysis, new evidence has been obtained indicating that long-term persistence and regionalisolation is a feature of the Antarctic terrestrial biota whose generality has not previously been appreciated. As well as creatinga new paradigm in which to consider the evolution and adaptation of Antarctic terrestrial biota, this opens important new cross-disciplinary linkages in the field of understanding the geological and glaciological history of the continent itself. Superimposedon this emerging historical template of Antarctic biogeography, this biota now faces the twin challenges of responding to thecomplex processes of climate change facing some parts of the continent, and the direct impacts associated with human occupationand travel to and between the spatially very limited areas of terrestrial habitat.� 2010 Elsevier B.V. and NIPR. All rights reserved.

Keywords: Biogeography; Biological invasions; Environmental change; Environmental manipulation; Phylogeography

1. Introduction

Antarctica is the only one of the Earth’s continents notto have a long-term history of human contact andinhabitation. The continent itself was first landed onapproximately two centuries ago (the northern AntarcticPeninsula), and East Antarctica only a little overa century ago, while the sub-Antarctic islands weremostly discovered and their marine-based livingresources rapidly over-exploited over the last two to

E-mail address: [email protected].

1873-9652/$ - see front matter � 2010 Elsevier B.V. and NIPR. All rights

doi:10.1016/j.polar.2010.03.003

three centuries. Antarctica is an ice-bound continent,only 0.34% of its area is currently ever free of snow or icein the form of terrestrial ecosystems including exposednunataks, cliffs and seasonally snow and ice-free areas.Even then, most of these appear biologically barren, withthe majority in areal terms being at higher altitude inlandlocations (see Table 1 in Convey et al., 2009a). Visiblydeveloped terrestrial ecosystems are best represented incoastal areas, particularly along the Antarctic Peninsula,and the ‘oases’ of the East Antarctic coastline. Most ice-free areas are small and ‘island-like’ (Bergstrom andChown, 1999), isolated by hostile ice and sea across

reserved.

Page 2: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

Fig. 1. The three widely recognised terrestrial biogeographical zones

in Antarctica.

136 P. Convey / Polar Science 4 (2010) 135e147

a range of physical scales. The McMurdo Dry Valleys ofsouthern Victoria Land provide the major exception tothis generalisation, forming an area of w40,000 km2.Large parts of the continent can be described as frigiddesert (Sømme, 1995).

Biological diversity means ‘‘the variability amongliving organisms from all sources including, inter alia,terrestrial, marine and other aquatic ecosystems and theecological complexes of which they are part; thisincludes diversity within species, between species, andof ecosystems’’ (Anon., 2009). However it is considered,Antarctic terrestrial biodiversity is depauperate. Atspecies level (species ‘richness’) diversity is low, athigher taxonomic levels many groups are missing alto-gether, and in functional terms, many functions orservices are poorly or not represented (Convey, 2007a).The fauna consists entirely of invertebrates, and thenwith only two higher insect species present (bothDiptera) (Block, 1984), while plant communities arepredominantly cryptogamic (lower plants - mosses,liverworts, lichens) (Ochyra et al., 2008; Øvstedal andSmith, 2001). Only two higher plants are present onthe Antarctic continent, both restricted to coastal regionsof the Antarctic Peninsula, although a larger diversity ofboth higher plants and insects are found on the sub-Antarctic islands (Convey, 2007b). For most groups ofbiota, survey data and up-to-date taxonomic treatmentsare lacking and large gaps remain in knowledge of theirbiology and biogeography (e.g. Adams et al., 2006;Chown and Convey, 2007; Peat et al., 2007). This isparticularly true for the microbiota, which some recentstudies are indicating may be considerably more diversethan previously thought (e.g. Cowan et al., 2002; Pearceet al., 2009). Some of the simplest faunal communitieson the planet are found on the Antarctic continent(Convey and McInnes, 2005; Freckman and Virginia,1997). Trophic complexity of these ecosystems is alsogenerally simple, although few rigorous autecologicalstudies are available to justify this assumption in reality(Hogg et al., 2006). Until recently, few microbial datahave been available, although this is now changingrapidly with the application of cutting edge molecularbiological and ecophysiological techniques (e.g.Boenigk et al., 2006; Cowan et al., 2002; De Weveret al., 2009; Lawley et al., 2004; Taton et al., 2006;Vyverman et al., 2010; Yergeau et al., 2007).However, the spatial coverage of these studies at conti-nental scale remains very limited.

These simple terrestrial ecosystems are thought to bevulnerable to environmental change, in particular in thecontext of changes to distributions and colonisation andinvasion by species not currently native to Antarctica

(Bergstrom and Chown, 1999; Convey, 2006; Freckmanand Virginia, 1997; Kennedy, 1995a, b; Walther et al.,2002). There is potential for new ecological niches tobe occupied and trophic functions to become estab-lished, inevitably changing structural and functionalaspects of the ecosystems. The native biota are not wellequipped to respond to such changes, being constrainedby their typically ‘‘adversity-selected’’ life historystrategies and lacking well-developed competitiveabilities (Convey, 1996a; Convey et al., 2006; Frenotet al., 2005).

2. Biogeography and history

Terrestrial biologists have generally recognisedthree biogeographic zones within Antarctica (Convey,2007a; Smith, 1984; Longton, 1988). These are mostfrequently described as the sub-, maritime and conti-nental (or frigid) Antarctic (Fig. 1), although otherterminology has been used, and generically separatethe terrestrial habitats of the main part of the continent,the Antarctic Peninsula, and the surrounding highlatitude Southern Ocean islands. The ecosystems andclimatic parameters of these three zones are distinc-tively different (Convey, 1996b; Walton, 1984).

Recently, the application of both classical biogeo-graphical analyses and modern molecular phylogenetic

Page 3: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

137P. Convey / Polar Science 4 (2010) 135e147

and phylogeographic techniques has been used to arguethat this visualisation of Antarctic biogeography is toosimplistic, and that elements of the contemporaryAntarctic terrestrial biota have a long-term history insitu in Antarctica over multi-million year timescales(Convey and Stevens, 2007; Convey et al., 2008, 2009b;Pugh and Convey, 2008). Examples are now availablepertinent to different timescales, although virtually allare inconsistent with a view widely held until recentlythat the majority of the Antarctic terrestrial biota mustbe recent colonists due to repeated extinction eventsthrough successive glacial maxima culminating in theLast Glacial Maximum (LGM) only w20,000 years ago(Fig. 2). These include, for instance, (1) the long-term(>c. 50 Myr) persistence of ‘sister’ chironomid midgespecies endemic to different and tectonically distinctelements of the Antarctic Peninsula and Scotia Arc(Allegrucci et al., 2006), on a timescale consistent withthe geological separation of Antarctica and SouthAmerica (cf. Livermore et al., 2007); (2) ancientdivergences between endemic springtails from the

Fig. 2. Schematic illustration of recent advances in evidence for long term

biota (taken from Convey et al., 2008, with the permission of Blackwell).

Transantarctic Mountains of East Antarctica, suggestingradiation on at least a Miocene timescale (21e11 Myr)(Stevens et al., 2006); (3) recent phylogeographicstudies demonstrating within-species divergence eventsthroughout the Pleistocene in springtails of VictoriaLand and the Antarctic Peninsula (McGaughran et al.,2010; Stevens et al., 2007); and (4) the recognition ofspecies-level endemism patterns, for instance inAntarctic nematode worms which, despite the general-isation of being highly capable of long-distancedispersal, present a fauna almost or entirely endemicto Antarctica (Adams et al., 2006; Andrassy, 1998;Maslen and Convey, 2006). Chown and Convey(2007) recognised a striking biogeographical boundarybetween the Antarctic Peninsula and the remainder ofWest and East Antarctica (named the Gressitt Line),with a complete lack of overlap at species-level in thefaunas of the two regions across the dominant terrestrialfaunal groups (particularly Acari, Collembola, Nem-atoda) strongly suggesting dissimilar but ancientorigins. Regionalisation is also now becoming apparent

persistence in Antarctica of most groups of contemporary terrestrial

Page 4: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

138 P. Convey / Polar Science 4 (2010) 135e147

at much smaller physical scales within Antarctica (Pughand Convey, 2008), providing a new and pressingchallenge for the conservation and management of thecontinent’s terrestrial biota.

It has been thought that microbial groups inAntarctica may not be as isolated from contact withlower latitudes as is the case of macroscopic organisms,through their apparently high potential for aerial trans-port (Finlay, 2002). Therefore, before the advent ofmolecular approaches, and based on morphologicaltaxomomy, it was often assumed or proposed that themajority of species present in Antarctica were cosmo-politan (e.g. see discussion in Pearce et al., 2009).Unfortunately, because many microbial species (andalso many microcscopic metazoans) are difficult toidentify using classical approaches it has until recentlybeen impossible to explore adequately whether‘uniqueness’ equates to ‘endemicity’, although rapidadvances are now being made through the application ofmolecular biological approaches (Boenigk et al., 2006;De Wever et al., 2009; Sands et al., 2008; Sjoling andCowan, 2003; Tindall, 2004; Taton et al., 2006;Vincent, 2000). Despite the general importance ofaerobiological dispersal for microbiota (Pearce et al.,2009) and confirmation of its occurrence at least atintra-regional scale within Antarctica (Hughes et al.,2004), the most recent studies of molecular diversityare starting to support a degree and timescale of isola-tion in Antarctic microbial taxa that are comparablewith those now proposed for the majority of macro-scopic groups (Vyverman et al., 2010).

3. Biological response to climate change inAntarctica

The combination of the magnitude of changes beingexperienced in parts of Antarctica and the generallysimple terrestrial ecosystems present is expected to leadto easily identifiable consequences (e.g. Bergstrom andChown, 1999; Convey, 2001, 2003, 2006; Kennedy,1995a; Walton et al., 1997; Wynn-Williams, 1994,1996). As a broad generalisation, environmentalamelioration (i.e. warmer temperatures and increasedwater availability) is predicted to lead to (i) increasedrates of local and long distance colonisation, and (ii)local-scale population expansion, leading to (iii)increased terrestrial diversity, biomass and trophiccomplexity, (iv) more complex ecosystem structure, and(v) a switch from the current dominance of physicalenvironmental variables to biotic factors (e.g. compe-tition, predation) driving ecosystem processes.However, in reality, the picture is far more complex as

these two environmental variables may interact toincrease abiotic stress levels (e.g. warming resulting inincreased desiccation, increased cloud cover leading tolower temperatures, reduced cloud cover leading tomore frequent freeze-thaw events, etc.). Changes inother stressors, such as increasing radiation linkedeither with changes in insolation/cloud cover or theformation of the ozone hole may also lead to negativeconsequences for biota and foodwebs, through requiringresource allocation to mitigation strategies.

Two lines of evidence have been applied topredictions of climate change responses, those beingobservational ecological studies, and a range of labo-ratory and field environmental manipulations. It shouldalso be noted that manipulation approaches are oftenprimarily used to examine ecophysiological orbiochemical responses to changes in environmentalstresses, rather than community level and biodiversityresponses. While manipulations are inevitably subjectto a range of methodological limitations and caveats(Kennedy, 1995b), they remain the only practicablemeans of achieving even partially realistic long-termstudies at remote and unhospitable locations, whilemore recent studies have made considerable advancesin overcoming the previous methodological limitations(Bokhorst et al., 2007a,b; Convey et al., 2002; Dayet al., 1999). Several reviews of the findings of thesestudies in the Antarctic have been published (Convey,2001, 2003; Kennedy, 1996).

However, and despite the considerable attentiondrawn by rapid trends of environmental change overrecent decades in parts of Antarctica, particularly theAntarctic Peninsula and some sub-Antarctic islands,there are surprisingly few robust scientific studies ofbiological responses to these changes in non-manipulated terrestrial ecosystems. The most widelyquoted example is that of rapid population expansionand local-scale colonisation by the two native floweringplants (Deschampsia antarctica and Colobanthus qui-tensis) in the maritime Antarctic (Fowbert and Smith,1994; Parnikoza et al., 2009; Smith, 1994). Thesestudies, in essence based on serendipitous samplingopportunities at a single location (the ArgentineIslands), currently remain the only repeat long-termmonitoring studies published of any terrestrial vegeta-tion or location in Antarctica. The population expan-sions observed are interpreted as warming and increasedwater availability in terrestrial habitats encouraging thegrowth and spreading of established plants, increasedfrequency of successful seed set, and increased germi-nation and establishment of seedlings. Anecdotally,similar changes are seen in the local distribution and

Page 5: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

139P. Convey / Polar Science 4 (2010) 135e147

development of typical cryptogamic vegetation of thisregion. Further, as these vegetation changes creates newhabitat, there are concurrent changes in the localdistribution and abundance of the invertebrate faunathat then colonises them. However, robust baselinesurvey data and monitoring studies capable of doc-umenting these changes remain critically lacking(Convey, 2006), and their establishment must now forman urgent priority.

Prompted by the difficulty in downscaling measuresof macroclimatic variables and trends to the micro-habitat scale relevant to most Antarctic terrestrial biota,Block and Convey (2001) and Convey et al. (2003)analysed a long term study of the water relations ofthe arthropod fauna on Signy Island (maritimeAntarctic), proposing these tiny organisms as a proxymeans of measuring water availability and changestherein. Their data suggested systematic changes inpatterns of water availability in these habitats that wereconsistent with local climate trends e in particularevidence for water being available earlier in the springand later in the autumn, and a period of increased waterstress in mid-summer.

Warming and changes in precipitation patterns haveincreased biological production in lakes, largely drivenby decreases in the duration and extent of lake ice coverand increased mixing of the water column (Quayleet al., 2002, 2003; Vincent et al., 2008). Lakes mayalso be sensitive to systematic changes in wind direc-tions leading to exposure to air masses from differentsources e for instance, some continental lakes haverecently become more saline due to drier conditionsleading to greater evaporation (Hodgson et al., 2006).

The limited observational data are supported bya number of field manipulation studies that have beencompleted at sub, maritime (the majority) and conti-nental Antarctic sites. Although earlier studies sufferedfrom serious methodological limitations (Kennedy,1995a), in simple terms, they demonstrated that thesoil microbial flora, bryophytes and invertebrate faunarespond rapidly to what were interpreted as improvedenvironmental conditions (Convey and Wynn-Williams,2002; Kennedy, 1994; Smith, 1990, 1993, 2001; Wynn-Williams, 1996) with greatly increased populations(Fig. 3). The adoption of more refined and provenmanipuation methodologies, with better replication andin some cases multivariate approaches to the manipu-lation of temperature, water and radiation regimes(e.g. Bokhorst et al., 2007a,b, 2008; Convey et al., 2002;Day et al., 1999, 2001; Sinclair, 2002) has shiftedthe emphasis from the simple description of rapidresponses to a higher level of integrated and improved

understanding. Responses to these manipulations havebeen quantified in terms of plant biochemistry,morphology, life history and ecology, and and atdifferent trophic levels, including the decompositioncycle, across the food web. It is now appreciated thatbiological responses to environmental change at indi-vidual organism or species level are often subtle, butthat they may integrate to give far greater impacts forthe community or ecosystem (Convey, 2003, 2006; Day,2001; Searles et al., 2001). There is also a clear need tocontinue such field experiments for periods long enoughto permit the often large and artefactual responses to theinitial perturbation to stabilise, almost inevitablya process far longer than the typical research fundingcycle, again highlighting the critical need for commit-ment to long term studies.

3.1. Direct human impacts on biodiversity

In global terms, the numbers of visitors who land orspend time on Antarctica would appear low relative toother continents. However, only 0.34% of the con-tinent’s area is ice-free (British Antarctic Survey,2004), and only a small proportion of that area isfound in the coastal regions where terrestrial ecosys-tems are best developed (Table 1 in Convey et al.,2009a), charismatic megafauna congregate, andresearch stations are preferentially constructed throughease of logistic access and construction, and proximityto research locations. These factors combine anddrastically magnify the potential for human impactupon the very ecosystems and biological communitiesthat are the target of research and public interest.

The contemporary intensity of human activity on theAntarctic continent and surrounding sub-Antarcticislands is in most cases greater than it has beenthroughout history since their discovery and initialexploration, only one to three centuries ago (Frenotet al., 2005; Tin et al., 2009), although the industrialexploitation of marine resources from certain sub-Antarctic islands, particularly South Georgia, provideexceptions to this generalisation (Convey andLebouvier, 2009). The research and associated logisticactivities of the 40þ national operators representingsignatory nations of the Antarctic Treaty Systemaccount for w5000 persons visiting the continent eachyear. Numerically, these are divided fairly evenlybetween operations in the northern Antarctic Peninsularegion (including the South Shetland Islands) where themajority of national research stations are established,and Victoria Land where, despite the fact that only threestations are present, one of these (McMurdo) has

Page 6: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

Fig. 3. (a) Field manipulation experiments have been completed at a range of locations encompassing the cold temperate Falkland Islands, sub-

Antarctic South Georgia, and maritime Antarctic locations between Signy Island (South Orkney Islands) and Alexander Island. (b) Lush bryo-

phyte growth within a simple cloche deployed for four years on Signy Island (South Orkney Islands) (Photograph: R.I.L. Smith). (c) ITEX-type

open top chambers (OTCs) deployed on Signy Island (South Orkney Islands) (Photograph: S. Bokhorst). (d) Two screen types deployed at Mars

Oasis, Alexander Island (Photograph: P. Convey). (e) Ventilated cloches deployed over swards of the grass Deschampsia antarctica on Leonie

Island, Ryder Bay, Adelaide Island (Photograph: British Antarctic Survey).

140 P. Convey / Polar Science 4 (2010) 135e147

a typical summer population of over 1000 staff. Otherresearch stations are dispersed widely along the EastAntarctic coastline and, increasingly, in the continentalinterior.

Tourist numbers have been increasing rapidly sincethe 1980s although have plateaud recently, most likelya temporary response to global economic recession.Currently, over 30,000 individuals each year visit andland in Antarctica as tourists, the large majority onorganised cruises, supported by a further 10e15,000ship’s crew and service personnel. The large majorityof these visit the northern Antarctic Peninsula andislands of the Scotia arc, typically landing at a smallnumber of well-known locations (Lynch et al., 2010).

Lynch et al.’s study highlights the concentrated natureof these landings, with 55% of landings in this areataking place at only 8 locations, the majority of thesereceiving approaching 10,000 individual visitors inrecent years, and two (Port Lockroy, Half MoonIsland) receiving up to 16,000. However, while thereare clearly more tourists than national operatorpersonnel expressed on either an annual or a specificlocation basis, the latter typically spend considerablylonger periods on the continent.

Few objective scientific reports have attempted toquantify human disturbance to Antarctic terrestrial andfreshwater ecosystems (Poland et al., 2003; Tejedoet al., 2009). The presence of stations, vehicles and

Page 7: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

Table 1

Overview of major trophic functions introduced to or significantly

augmented in sub-Antarctic ecosystems through the introduction of

non-indigenous terrestrial invertebrates (drawn from Frenot et al.,

2005, and additional unpublished data of P. Convey and R.S. Key).

Invertebrate group Trophic function

introduced or augmented

Earthworms (Annelida) (7 spp.) New detritivore pathways

Slugs (Mollusca) (3 spp.) Herbivory

Woodlice (Crustacea) (2 spp.) New detritivore pathways

Mites (Acari) (12 spp.) Range of function, though

mostly detritivores

Springtails (Collembola) (11 sp.) Detritivores

Aphids (Homoptera) (9 sp.) Herbivory, potential synergy

with native and non-indigenous

microbial disease-causing agents

and transfer

Beetles (Coleoptera) (3 sp.) New and aggressive predators

Flies (Diptera) (15 sp.) Detritivores, may release decay

bottlenecks

Plant pollinators, potential

synergy with insect-pollinated

non-indigenous plants

141P. Convey / Polar Science 4 (2010) 135e147

their operations clearly leads to local pollution throughdispersal of chemical pollutants, dust, and directdamage to vegetation, soil surfaces and freshwatersystems (Bargagli, 2005; Kaup and Burgess, 2002; Tinet al., 2009). Soil and freshwater ecosystems mayexperience eutrophication through human activities(Ohtani et al., 2000). Specific studies of human-formedtracks document considerable impact on plant speciesand soils on various sub-Antarctic islands (Gremmenet al., 2003; Scott and Kirkpatrick, 1994), and like-wise on soil properties and invertebrates in the mari-time Antarctic South Shetland Islands (Beyer andBolter, 2002; Tejedo et al., 2009), and soils in thecontinental Antarctic (Campbell et al., 1998). Acommon feature of these studies are comments to theeffect that recovery from these types of disturbance tovegetation and soils may take decades, at least.

3.2. Anthropogenic transfer

Antarctica and its terrestrial ecosystems are isolatedon a range of intra- and inter-continental scales, withinthe continent by the island-like nature of areas ofice-free ground, and inter-continentally by the atmo-spheric and oceanic circulations surrounding the conti-nent, combined with the vast distances and inhospitableconditions that must be survived while in transit and onarrival. Nevertheless, natural colonisation events havetaken place and continue to do so (Barnes et al., 2006;Clarke et al., 2005), utilising a range of dispersalmechanisms and routes (e.g. Hughes et al., 2006).Human-assisted transfers of biota overcome several ofthe barriers facing natural colonists, in particular beingpossible much more rapidly (hours to days) than thenatural processes, and in avoiding the extreme envi-ronmental stresses (e.g. low temperature, osmotic stress,desiccation, radiation) inherent in transfer at altitude inthe atmosphere, or on the ocean surface. It remainsdifficult to estimate the relative importance of naturaland human-assisted colonisation routes into theAntarctic, although at two remote Southern Oceanislands (Gough Island, Marion Island) it has been esti-mated that the latter has outweighed the former by atleast two orders of magnitude since their discovery(Gaston et al., 2003; Gremmen and Smith, 2004).

Across most sub-Antarctic islands, the last twocenturies of human contact have seen major impacts onterrestrial ecosystems from the deliberate and accidentalintroduction of many plants and animals (Convey,2007b; Convey et al., 2006; Convey and Lebouvier,2009; Frenot et al., 2005, 2008), providing an urgentwarning of the potential consequences should invasive

species become established on the Antarctic continent(Convey, 2008). Although very few establishmentevents have yet been recorded in this latter region, it isalready known that accidental transfers of biota dooccur, particularly associated with cargo, vehicles, foodand personal clothing (Sjoling and Cowan, 2000; Frenotet al., 2005; Hughes et al., 2010; Whinam et al., 2004).

During the early phase of human contact with thecontinent and surrounding islands there were manycases of deliberate import of vertebrates for transport,food and recreational purposes. On many sub-Antarcticislands these, along with the accidental introductions ofrodents, invertebrates and plants, have led to largeimpacts on native species and ecosystems (Bergstromet al., 2009; Bonner, 1984; Chapuis et al., 1994;Convey et al., 2006; Convey and Lebouvier, 2009;Frenot et al., 2005, 2008; Leader-Williams, 1988).Even after the establishment of the Antarctic TreatySystem, there were several botanical transplant experi-ments during the 1960s and 1970s, involving thetransfer and survival of a range of temperate, Arctic andsub-Antarctic species under maritime Antarctic condi-tions (Smith, 1996). While such experiments are nolonger permitted, and the original plant material wascompletely removed after the experiments, it was real-ised in the 1980s that an unintended consequence at onelocation had been the introduction and establishment ofsub-Antarctic terrestrial invertebrates (and, presumably,a still unknown number of microbial taxa) (Block et al.,1984). These species remain to the current day (Hughes

Page 8: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

142 P. Convey / Polar Science 4 (2010) 135e147

and Worland, 2010), with no realistic possibility ofremedial action to remove them. Overall, the presenceof all non-indigenous biota known to date can be mostplausibly be linked with ‘national’ operations and theirpreceding industrial exploitation industries with,despite expressed concerns, none to date linked with therapid expansion of the tourism industry withinAntarctica over the last few decades (Convey andLebouvier, 2009; Frenot et al., 2005; Tin et al., 2009).

There are currently far fewer non-indigenous speciesknown to be established on the Antarctic continent (fiveconfirmed, at least three others unconfirmed) than on thesub-Antarctic islands (w200) (Convey, 2008; Frenotet al., 2005, 2008; R.I.L. Smith and M. Richardson,unpubl. data). While several vertebrates, invertebratesand plants on various sub-Antarctic islands clearly have‘invasive’status (sensu Frenot et al., 2005), and others areidentified as having high potential to switch from theircurrent ‘persistent’ status under current regional envi-ronmental change scenarios, there is at present noevidence of any of the persistent non-indigenous speciesestablished on the continent becoming invasive.A common feature of many of the non-indigenousspecies already known to be established in the sub-Antarctic is that they belong to ecological functionalgroups, or introduce trophic functions, that are poorly ornot represented in the native communities. To an extentthis is obvious in the case of non-indigenous vertebrates,as no terrestrial vertebrate herbivores or predators arepresent in any Antarctic region, a fact that underlies theimportance of parts of the region as population centres formany of the charismatic vertebrates of the SouthernOcean (Woehler et al., 2001), and the evolution of uniquemegaherb-dominated plant communities (Mitchell et al.,1999). However, a large number of terrestrial inverte-brates has also been introduced to ecosystems within theAntarctic, and their impacts are less widely known, yetstill potentially fundamental to ecosystem functioning(Table 1). Additionally, anecdotal and a few publishedreports are available of yet further non-indigenous biotaexisting synanthropically for shorter or longer periods(i.e. directly in association with human activity, such aswithin station buildings, associated with foodstuffs, etc.)(Greenslade, 2006; Hughes et al., 2005).

In reality, knowledge of the presence, distributionand impacts of non-indigenous species in the Antarcticis currently far from perfect, and the available data onnumbers of such species are likely to be a considerableunderestimate, other than for the vertebrates. At themajority of locations baseline survey and monitoringdata are simply unavailable for most invertebrate andlower plant groups while, even for locations and groups

where data are available, there are no ongoing pro-grammes monitoring distribution and abundancechanges or impacts. The presence of non-indigenousmicrobiota is particularly poorly known (Convey,2008; Frenot et al., 2005), compounded by the chal-lenge of classifying any new microbial record as beinganthropogenically rather than naturally introduced, oralready native to the region under consideration.Circumstantial evidence is provided by the identifica-tion of microbial taxa present at human impacted sites(e.g. close to research stations) but not at neighbouringpristine areas (e.g. Azmi and Seppelt, 1998; Uptonet al., 1997). In recent years, the use of molecularbiological methodologies has started to improve thepotential of identifying non-indigenous microbes (e.g.Baker et al., 2003).

3.3. Vigilance

Both tourists and national operator personnel sharea common feature in the pattern of their contact withthe continent, this being to visit more than one location,and sometimes several, in succession over a shortperiod of the summer season. This often, for practicalreasons, involves progression from less extreme tomore extreme locations, in essence providing a series oflogistic and geographical ‘stepping stones’ for potentialcolonists. Thus, as well as the clear risk identifiedabove of inadvertent transfer of non-indigenous speciesinto the region, there is also the potential for transfer ofbiota between different locations within Antarctica,with many of these biota highly likely to be appropri-ately preadapted to permit establishment at the newlocation(s) (e.g. Chown and Convey, 2007; Convey,2008). In terms of impacts on regional biodiversity,given the recent recognition of clear and long standingevolutionary isolation and bioregionalisation withinAntarctica (including the sub-Antarctic islands), thislatter risk is at least as serious as the more widelyrecognised former risk, presenting a clear conservationmanagement challenge for the region (Chown et al.,2008; Grant et al., in press).

Until the recent International Polar Year programme‘Aliens in Antarctica’ (www.aliensinantarctica.aq)there had been no concerted effort to quantify theobjective risks associated with the various differentcommunities, logistic routes and means of movingpeople and cargo to and from the continent. At the sametime, the issue of ‘alien species’ has now been placed asa standing agenda item of the Committee for Environ-mental Protection of the Antarctic Treaty System. Thus,there has been a recent upsurge of research, national

Page 9: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

143P. Convey / Polar Science 4 (2010) 135e147

operator and public interest in the subject. Previouslyanecdotal knowledge of ‘accidental’ transfer events isnow being replaced by examples of careful documen-tation (e.g. Hughes et al., 2010; Lee and Chown,2009a,b; Whinam et al., 2004). Hughes et al.’s (2010)study highlights the very real risks associated withtransfer of cargo and equipment between successivelocations within Antarctica, in this case of mechanicalplant from sub-Antarctic South Georgia to maritimeAntarctic Adelaide Island. A clear failure of applicationof existing biosecurity procedures led to over 130 kg ofsoil, including live plants (phanerogams and crypto-gams), arthropods, nematodes, viable seeds andmicrobes being transferred, with this oversight notbeing noticed until a week after arrival. Several of thespecies identified during this study are either alreadynative to locations on the Antarctic Peninsula (i.e.potentially leading to dilution of regional geneticdiversty; Chown and Convey, 2007), or were previouslythe subject of 1960s transplant experiments confirmingtheir ability to survive in the more extreme maritimeAntarctic environment (even before recent climateamelioration trends) (Smith, 1996), while at least twospecies of invertebrate are already successfully estab-lished non-indigenous species elsewhere in the mari-time Antarctic.

4. Concluding discussion

Recent advances in knowledge of Antarctic terres-trial diversity have been substantial, and sufficient tounderpin paradigm shifts in the patterns recognised andtheir implications both within biology and beyond(Convey and Stevens, 2007; Convey et al., 2008,2009a). However, there remain considerable gaps inknowledge, both in taxonomic (i.e. within all highertaxonomic groups present, even those that are ‘betterknown’) and geographic (i.e. spatial distribution ofsurvey effort) contexts (Adams et al., 2006; Chown andConvey, 2007; Peat et al., 2007). For virtually allbiological groups, it remains typical that available dataare skewed towards the physical vicinity of a smallnumber of research locations, in particular followingthe historical patterns of activity of a small number ofindividual specialist biologists, and that often in theearly decades of Antarctic biological exploration(Adams et al., 2006; Chown and Convey, 2007). Thus,there is an urgent need to rectify the absence of effortdevoted to high quality baseline biodiversity survey,which must underpin any pure or applied researcheffort aimed at identifying patterns, even beforeattention is given to the potential consequences of

global or regional environmental change and otherhuman impacts on this biodiversity, or to the increas-ingly recognised need for conservation managementapproaches within the region. Recent advances in andapplication of modern techniques within molecularbiology, in concert with classical approaches, are nowleading to rapid and fundamental changes in our viewof Antarctic terrestrial diversity, even in previouslyintractable microbial groups (e.g. Bridge et al., 2008;Pugh, 1993, 2004; Pugh et al., 2002; Sjoling andCowan, 2003; Vyverman et al., 2010).

Antarctica is, rightly, at the forefront of researcheffort into the processes and consequences of globalenvironmental change (Convey et al., 2009a; Turneret al., 2009a), with the Antarctic Peninsula in partic-ular being one of the three most rapidly changingregions on the planet, and the most rapid in theSouthern Hemisphere. Paradoxically, the majority ofthe continent has been protected from the full impact ofthese global climatic change processes over the lastseveral decades by the formation of the Antarctic ozonehole (Turner et al., 2009b), itself a separate conse-quence of anthropogenic pollution of the atmosphere.This protection is forecast to wane over the nextcentury, as the ozone hole repairs, meaning that theentire continent is forecast to experience considerableclimatic changes over coming decades. The terrestrialand freshwater ecosystems of Antarctica are uniquelyplaced to provide understanding and early warningof the likely effects of environmental change forecosystem processes and services, of global relevanceand application. Clear changes in some native ecosys-tems are already apparent (Bergstrom et al., 2006),although in reality limited to few specific terrestrial(Convey et al., 2003; Parnikoza et al., 2009) or limnetic(Quayle et al., 2002, 2003) ecosystems, communities orspecies. Likewise, the importance of direct humanimpacts (habitat damage and modification, pollution,species introductions) as potentially compromisingecosystem structure and function has only recentlystarted to receive recognition and research attention(Frenot et al., 2005; Tejedo et al., 2009; Tin et al.,2009), despite one of the founding principles of theAntarctic Treaty being to ensure the preservation of theenvironment of Antarctica. In the context of research tounderpin knowledge both of biological responses toenvironmental change and other direct human impactsthere has been a long-term and critical failure atnational level in recognition and funding of appropriatemonitoring or manipulation/modelling programmesapplied to key communities and locations, whichrequires urgent rectification.

Page 10: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

144 P. Convey / Polar Science 4 (2010) 135e147

There have been considerable advances in knowl-edge and understanding of Antarctic terrestrial diver-sity in recent years (e.g. Adams et al., 2006;Greenslade, 1995; Ochyra et al., 2008; Øvstedal andSmith, 2001; Pugh, 1993, 2004; Pugh and Convey,2008; Pugh et al., 2002). In part this has been ach-ieved through the impetus from ‘umbrella’ researchprogrammes, as exemplified by the SCAR RiSCC(Regional Sensitivity to Climate Change in AntarcticTerrestrial and Limnetic Ecosystems, 2000e2005) andEBA (Evolution and Biodiversity in Antarctica, www.eba.aq; 2005e2013) programmes, which have beencentral in encouraging international collaboration, datasharing and analyses, and the establishment of robustterrestrial biodiversity databases. Even now, the fullpotential of these approaches is yet to be realised, andthe terrestrial research community would do well tofollow the example set in the marine realm by twofurther SCAR-endorsed research efforts e the SCAR-MarBIN marine biodiversity database (www.scarmarbin.be) and the IPY programme ‘Census ofAntarctic Marine Life (CaML; www.caml.aq).

Acknowledgements

I thank the organisers of the Xth SCAR BiologySymposium, Sapporo, for the opportunity and supportto present this paper at the meeting. This paper is anoutput of the BAS ‘Polar Science for Planet Earth’ corescience programme, and also contributes to the SCAREBA programme.

References

Adams, B., Bardgett, R.D., Ayres, E., Wall, D.H., Aislabie, J.,

Bamforth, S., Bargagli, R., Cary, C., Cavacini, P., Connell, L.,

Convey, P., Fell, J., Frati, F., Hogg, I., Newsham, N.,

O’Donnell, A., Russell, N., Seppelt, R., Stevens, M.I., 2006.

Diversity and distribution of Victoria Land biota. Soil Biol.

Biochem. 38, 3003e3018.

Allegrucci, G., Carchini, G., Todisco, V., Convey, P., Sbordoni, V.,

2006. A molecular phylogeny of Antarctic Chironomidae and its

implications for biogeographical history. Polar Biol. 29, 320e326.

Andrassy, I., 1998. Nematodes in the sixth continent. J. Nematode

Syst. Morphol. 1, 107e186.

Anon., 2009. Convention on biological diversity. http://www.cbd.int/.

Azmi, O.R., Seppelt, R.D., 1998. The broad-scale distribution of

microfungi in the Windmill Islands region, Antarctica. Polar

Biol. 19, 92e100.

Baker, G., Ah Tow, L., Cowan, D.A., 2003. Detection of non-

indigenous micro-organisms in ‘pristine’ environments. J.

Microbiol. Methods 53, 157e164.

Bargagli, R., 2005. Antarctic Ecosystems. Environmental Contami-

nation, Climate Change, and Human Impact. In: Ecological

Studies 175. Springer.

Barnes, D.K.A., Hodgson, D.A., Convey, P., Allen, C.S., Clarke, A.,

2006. Incursion and excursion of Antarctic biota: past, present

and future. Global Ecol. Biogeogr. 15, 121e142.

Bergstrom, D.M., Chown, S.L., 1999. Life at the front: history,

ecology and change on southern ocean islands. Trends Ecol.

Evol. 14, 472e476.

Bergstrom, D.M., Convey, P., Huiskes, A.H.L. (Eds.), 2006. Trends

in Antarctic Terrestrial and Limnetic Ecosystems: Antarctica as

a Global Indicator. Springer, Dordrecht.

Bergstrom, D.M., Lucieer, A., Kiefer, K., Wasley, J., Belbin, L.,

Pedersen, T.K., Chown, S.L., 2009. Indirect effects of invasive

species removal devastate World Heritage Island. J. Appl. Ecol.

46, 73e81.

Beyer, L., Bolter, M., 2002. Geoecology of Antarctic Ice Free

Coastal Landscapes. Springer, Berlin.

Block, W., 1984. Terrestrial microbiology, invertebrates and

ecosystems. In: Laws, R.M. (Ed.), Antarctic Ecology. Academic

Press, London, pp. 163e236.

Block, W., Burn, A.J., Richard, K.J., 1984. An insect introduction to

the maritime. Antarctic. Biol. J. Linn. Soc. 23, 33e39.

Block, W., Convey, P., 2001. Seasonal and long-term variation in

body water content of an Antarctic springtail - a response to

climate change? Polar Biol. 24, 764e770.

Boenigk, J., Pfandl, K., Garstecki, T., Harms, H., Novarino, G.,

Chatzinotas, A., 2006. Evidence for geographic isolation and

signs of endemism within a protistan morphospecies. Appl.

Environm. Microbiol. 72, 5159e5164.

Bokhorst, S., Huiskes, A., Convey, P., Aerts, R., 2007a. The effect of

environmental change on vascular plant and cryptogam communi-

ties along a latitudinal gradient from the Falkland Islands to the

Maritime Antarctic. BMC Ecol 7, 15. doi:10.1186/1472-6785-7-15.

Bokhorst, S., Huiskes, A., Convey, P., Aerts, R., 2007b. Climate

change effects on decomposition rates in ecosystems from the

Maritime Antarctic and Falkland Islands. Global Change Biol.

13, 2642e2653.

Bokhorst, S., Huiskes, A., Convey, P., van Bodegom, P.M., Aerts, R.,

2008. Climate change effects on soil arthropod communities from

the Falkland Islands and the Maritime Antarctic. Soil Biol.

Biochem. 40, 1547e1556.

Bonner, W.N., 1984. Introduced mammals. In: Laws, R.M. (Ed.),

Antarctic Ecology, vol. 1. Academic Press, London, pp. 237e278.

Bridge, P.D., Spooner, B.M., Roberts, P.J., 2008. Non-lichenized

fungi from the Antarctic region. Mycotaxon 106, 485e490.

British Antarctic Survey, 2004. Antarctica 1:10 000 000 Map. BAS

(misc) 11. British Antarctic Survey, Cambridge.

Campbell, I.B., Claridge, G.G.C., Balks, M.R., 1998. Short and long-

term impacts of human disturbance on snow-free surfaces in

Antarctica. Polar Record 34 (188), 15e24.

Chapuis, J.L., Bousses, P., Barnaud, G., 1994. Alien mammals,

impact and management in the French subantarctic islands. Biol.

Conserv. 67, 97e104.

Chown, S.L., Convey, P., 2007. Spatial and temporal variability

across life’s hierarchies in the terrestrial Antarctic. Phil. Trans.

Roy. Soc. Lond. B 362, 2307e2331.

Chown, S.L., Lee, J.E., Shaw, J.D., 2008. Conservation of Southern

Ocean Islands: invertebrates as exemplars. J. Insect. Conserv. 12,

277e291.

Clarke, A., Barnes, D.K.A., Hodgson, D.A., 2005. How isolated is

Antarctica? Trends Ecol. Evol. 20, 1e3.

Convey, P., 1996a. The influence of environmental characteristics on

the life history attributes of Antarctic terrestrial biota. Biol. Rev.

71, 191e225.

Page 11: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

145P. Convey / Polar Science 4 (2010) 135e147

Convey, P., 1996b. Overwintering strategies of terrestrial inverte-

brates from Antarctica e the significance of flexibility in

extremely seasonal environments. Eur. J. Entomol. 93, 489e505.

Convey, P., 2001. Terrestrial ecosystem response to climate changes

in the Antarctic. In: Walther, G.-R., Burga, C.A., Edwards, P.J.

(Eds.), ‘‘Fingerprints’’ of Climate Change e Adapted Behaviour

and Shifting Species Ranges. Kluwer, New York, pp. 17e42.

Convey, P., 2003. Maritime Antarctic climate change: signals from

terrestrial biology. In: Domack, E., Burnett, A., Leventer, A.,

Convey, P., Kirby, M., Bindschadler, R. (Eds.), Antarctic Penin-

sula Climate Variability: Historical and Palaeoenvironmental

Perspectives. Antarctic Research Series, vol. 79. American

Geophysical Union, Washington, D.C, pp. 145e158.

Convey, P., 2006. Antarctic climate change and its influences on

terrestrial ecosystems. In: Bergstrom, D.M., Convey, P.,

Huiskes, A.H.L. (Eds.), Trends in Antarctic Terrestrial and

Limnetic Ecosystems: Antarctica as a Global Indicator. Springer,

Dordrecht, pp. 253e272.

Convey, P., 2007a. Antarctic Ecosystems. In: Levin, S.A. (Ed.),

Encyclopedia of Biodiversity, second ed.). Elsevier, San Diego.

doi:10.1016/B0-12-226865-2/00014-6.

Convey, P., 2007b. Influences on and origins of terrestrial biodiversity

of the sub-Antarctic islands. Pap. Proc. Roy. Soc. Tasmania 141,

83e93.

Convey, P., 2008. Non-native species in Antarctic terrestrial and

freshwater environments: presence, sources, impacts and

predictions. In: Rogan-Finnemore, M. (Ed.), Non-native Species

in the Antarctic Proceedings. Gateway Antarctica, Christchurch,

pp. 97e130.

Convey, P., Bindschadler, R.A., di Prisco, G., Fahrbach, E., Gutt, J.,

Hodgson, D.A., Mayewski, P., Summerhayes, C.P., Turner, J.,

2009a. Antarctic climate change and the environment. Antarct.

Sci. 21, 541e563.

Convey, P., Block, W., Peat, H.J., 2003. Soil arthropods as indicators

of water stress in Antarctic terrestrial habitats? Global Change

Biol. 9, 1718e1730.

Convey, P., Frenot, F., Gremmen, N., Bergstrom, D., 2006. Biological

invasions. In: Bergstrom, D.M., Convey, P., Huiskes, A.H.L.

(Eds.), Trends in Antarctic Terrestrial and Limnetic Ecosystems:

Antarctica as a Global Indicator. Springer, Dordrecht, pp.

193e220.

Convey, P., Gibson, J., Hillenbrand, C.-D., Hodgson, D.A.,

Pugh, P.J.A., Smellie, J.L., Stevens, M.I., 2008. Antarctic

terrestrial life e challenging the history of the frozen continent?

Biol. Rev. 83, 103e117.

Convey, P., Lebouvier, M., 2009. Environmental change and human

impacts on terrestrial ecosystems of the sub-Antarctic islands

between their discovery and the mid-Twentieth Century. Pap.

Proc. Roy. Soc. Tasmania 143, 33e44.

Convey, P., McInnes, S.J., 2005. Exceptional, tardigrade dominated,

ecosystems from Ellsworth Land, Antarctica. Ecology 86, 519e527.

Convey, P., Pugh, P.J.A., Jackson, C., Murray, A.W., Ruhland, C.T.,

Xiong, F.S., Day, T.A., 2002. Response of Antarctic terrestrial

arthropods to long-term climate manipulations. Ecology 83,

3130e3140.

Convey, P., Stevens, M.I., 2007. Antarctic biodiversity. Science 317,

1877e1878.

Convey, P., Stevens, M.I., Hodgson, D.A., Smellie, J.L.,

Hillenbrand, C.-D., Barnes, D.K.A., Clarke, A., Pugh, P.J.A.,

Linse, K., Cary, S.C., 2009b. Exploring biological constraints on

the glacial history of Antarctica. Quatern. Sci. Rev. 28, 3035e3048.

Convey, P., Wynn-Williams, D.D., 2002. Antarctic soil nematode

response to artificial environmental manipulation. Eur. J. Soil

Biol. 38, 255e259.

Cowan, D.A., Russell, N.J., Mamais, A., Sheppard, D.M., 2002.

Antarctic dry valley mineral soils contain unexpectedly high

levels of microbial biomass. Extremophiles 6, 431e436.

Day, T.A., 2001. Multiple trophic levels in UV-B assessments ecompleting the ecosystem. New Phytol. 152, 183e186.

Day, T.A., Ruhland, C.T., Xiong, F., 2001. Influence of solar UV-B

radiation on Antarctic terrestrial plants: results from a 4-year

field study. J. Photochem. Photobiol. B: Biol. 62, 78e87.

Day, T.A., Ruhland, C.T., Grobe, C.W., Xiong, F., 1999. Growth and

reproduction of Antarctic vascular plants in response to warming

and UV radiation reductions in the field. Oecologia 119, 24e35.

De Wever, A., Leliaert, F., Verleyen, E., Vanormelingen, P., Van der

Gucht, K., Hodgson, D.A., Sabbe, K., Vyverman, W., 2009.

Hidden levels of phylodiversity in Antarctic green algae: further

evidence for the existence of glacial refugia. Proc. Roy. Soc.

Lond. B. doi:10.1098/rspb.2009.0994.

Finlay, B.J., 2002. Global dispersal of free-living microbial eukaryote

species. Science 296, 1061e1063.

Fowbert, J.A., Smith, R.I.L., 1994. Rapid population increase in

native vascular plants in the Argentine Islands, Antarctic Penin-

sula. Arct. Alpine Res. 26, 290e296.

Freckman, D.W., Virginia, R.A., 1997. Low-diversity Antarctic soil

nematode communities: distribution and response to disturbance.

Ecology 78, 363e369.

Frenot, Y., Chown, S.L., Whinam, J., Selkirk, P., Convey, P.,

Skotnicki, M., Bergstrom, D., 2005. Biological invasions in the

Antarctic: extent, impacts and implications. Biol. Rev. 80,

45e72.

Frenot, Y., Convey, P., Lebouvier, M., Chown, S.L., Whinam, J.,

Selkirk, P.M., Skotnicki, M., Bergstrom, D.M., 2008. Antarctic

biological invasions: sources, extents, impacts and implications.

In: Rogan-Finnemore, M. (Ed.), Non-native Species in the

Antarctic Proceedings. Gateway Antarctica, Christchurch, pp.

53e96.

Gaston, K.J., Jones, A.G., Hanel, C., Chown, S.L., 2003. Rates of

species introduction to a remote oceanic island. Proc. Roy. Soc.

Lond. B 270, 1091e1098.

Grant, S., Convey, P., Hughes, K.A., Phillips, R.A., Trathan, P.N.,

Conservation and management of Antarctic ecosystems. In:

Rogers, A.D., Murphy, E., Clarke, A., Johnston, N.M. (Eds.),

Antarctica: An Extreme Environment in a Changing World, in

press.

Greenslade, P., 1995. Collembola from the Scotia Arc and Antarctic

Peninsula including descriptions of two new species and notes on

biogeography. Polskie Pismo Entomologiczne 64, 305e319.

Greenslade, P., 2006. The Invertebrates of Macquarie Island.

Australian Antarctic Division, Kingston.

Gremmen, N., Smith, V., 2004. The Flora of Marion and Prince

Edward Islands. Data Analyse Ecologie, Diever.

Gremmen, N.J.M., Smith, V.R., van Tongeren, O.F.R., 2003. Impact

of trampling on the vegetation of subantarctic Marion Island.

Arct. Antarct. Alpine Res. 35, 442e446.

Hodgson, D.A., Roberts, D., McMinn, A., Verleyen, E., Terry, B.,

Corbett, C., Vyverman, W., 2006. Recent rapid salinity rise in

three East Antarctic lakes. J. Paleolimnol. 36, 385e406.

Hogg, I.D., Cary, S.C., Convey, P., Newsham, K., O’Donnell, T.,

Adams, B.J., Aislabie, J., Frati, F.F., Stevens, M.I., Wall, D.H.,

2006. Biotic interactions in Antarctic terrestrial ecosystems: are

they a factor? Soil Biol. Biochem 38, 3035e3040.

Page 12: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

146 P. Convey / Polar Science 4 (2010) 135e147

Hughes, K.A., Convey, P., Maslen, N.R., Smith, R.I.L., 2010. Acci-

dental transfer of non-native soil organisms into Antarctica on

construction vehicles. Biol. Invasions 12, 875e891.

Hughes, K.A., McCartney, H.A., Lachlan-Cope, T.A., Pearce, D.A.,

2004. A preliminary study of airborne microbial biodiversity over

peninsular Antarctica. Cell. Molec. Biol. 50, 537e542.

Hughes, K.A., Ott, S., Bolter, M., Convey, P., 2006. Colonisation

processes. In: Bergstrom, D.M., Convey, P., Huiskes, A.H.L.

(Eds.), Trends in Antarctic Terrestrial and Limnetic Ecosystems:

Antarctica as a Global Indicator. Springer, Dordrecht, pp. 35e54.

Hughes, K.A., Walsh, S., Convey, P., Richards, S., Bergstrom, D.,

2005. Alien fly populations established at two Antarctic research

stations. Polar Biol. 28, 568e570.

Hughes, K.A., Worland, M.R., 2010. Spatial distribution, habitat

preference and colonization status of two alien terrestrial inver-

tebrate species in Antarctica. Antarct. Sci.. doi:10.1017/

S0954102009990770.

Kaup, E., Burgess, J.S., 2002. Surface and subsurface flows of

nutrients in natural and human impacted catchments on Broknes,

Larsemann Hills, Antarctica. Antarct. Sci. 14, 343e352.

Kennedy, A.D., 1994. Simulated climate change: a field manipulation

study of polar microarthropod community response to global

warming. Ecography 17, 131e140.

Kennedy, A.D., 1995a. Antarctic terrestrial ecosystem response to

global environmental change. Annu. Rev. Ecol. Syst. 26, 683e704.

Kennedy, A.D., 1995b. Simulated climate change: are passive

greenhouses a valid microcosm for testing the biological effects

of environmental perturbations? Global Change Biol. 1, 29e42.

Kennedy, A.D., 1996. Antarctic fellfield response to climate change:

a tripartite synthesis of experimental data. Oecologia 107,

141e150.

Lawley, B., Ripley, S., Bridge, P., Convey, P., 2004. Molecular

analysis of geographic patterns of eukaryotic diversity in

Antarctic soils. Appl. Environm. Microbiol. 70, 5963e5972.

Leader-Williams, N., 1988. Reindeer on South Georgia: The Ecology

of An Introduced Population. Cambridge University Press,

Cambridge.

Lee, J.E., Chown, S.L., 2009a. Breaching the dispersal barrier to

invasion: quantification and management. Ecol. Appl. 19,

1944e1959.

Lee, J.E., Chown, S.L., 2009b. Quantifying the propagule load

associated with the construction of an Antarctic research station.

Antarct. Sci. 21, 471e475.

Livermore, R., Hillenbrand, C.-D., Meredith, M., Eagles, G., 2007.

Drake Passage and Cenozoic climate: an open and shut case?

Geochem. Geophys. Geosystems 8, Q01005. doi:10.1029/

2005GC001224.

Longton, R.E., 1988. The Biology of Polar Bryophytes and Lichens.

Cambridge University Press, Cambridge, 391 pp.

Lynch, H.J., Crosbie, K., Fagan, W.F., Naveen, R., 2010. Spatial

patterns of tour ship traffic in the Antarctic Peninsula Region.

Antarct. Sci. 22, 123e130.

Maslen, N.R., Convey, P., 2006. Nematode diversity and distribution

in the southern maritime Antarctic e clues to history? Soil Biol.

Biochem. 38, 3141e3151.

McGaughran, A., Toricelli, G., Carapelli, A., FratiStevens, M.I.,

Convey, P., Hogg, I.D., 2010. Contrasting phylogeographic patterns

for springtails reflect different evolutionary histories between the

Peninsula and continental Antarctica. J. Biogeog. 37, 103e119.

Mitchell, A.D., Meurk, C.D., Wagstaff, S.J., 1999. Evolution of

Stilbocarpa, a megaherb from New Zealand’s sub-antarctic

islands. NZ J. Bot. 37, 205e211.

Ochyra, R., Smith, R.I.L., Bednarek-Ochyra, H., 2008. The Illus-

trated Moss Flora of Antarctica. Cambridge University Press,

Cambridge.

Ohtani, S., Suvama, K., Kanda, H., 2000. Environmental monitoring by

means of soil algae and microorganisms in the vicinity of Syowa

Station. Nankyoku Shiryo (Antarctic Record) 44, 265e276.

Øvstedal, D.O., Smith, R.I.L., 2001. Lichens of Antarctica and South

Georgia. A Guide to Their Identification and Ecology. Cambridge

University Press, Cambridge.

Parnikoza, I., Convey, P., Dykyy, I., Trakhimets, V., Milinevsky, G.,

Tyschenko, O., Inozemtseva, D., Kozeretska, I., 2009. Current

status of the Antarctic herb tundra formation in the central

Argentine Islands. Global Change Biol. 15, 1685e1693.

Pearce, D.A., Bridge, P.D., Hughes, K.A., Sattler, B., Psenner, R.,

Russell, N.J., 2009. Microorganisms in the atmosphere over

Antarctica. FEMS Microbiol. Ecol. 69, 143e157.

Peat, H.J., Clarke, A., Convey, P., 2007. Diversity and biogeography

of the Antarctic flora. J. Biogeog. 34, 132e146.

Poland, J.S., Riddle, M.J., Zeeb, B.A., 2003. Contaminants in the

Arctic and the Antarctic: a comparison of sources, impacts, and

remediation options. Polar Record 39, 369e383.

Pugh, P.J.A., 1993. A synonymic catalogue of the Acari from

Antarctica, the sub-Antarctic Islands and the Southern Ocean. J.

Nat. Hist. 27, 323e421.

Pugh, P.J.A., 2004. Biogeography of spiders (Araneae: Arachnida) on

the islands of the Southern Ocean. J. Nat. Hist. 38, 1461e1487.

Pugh, P.J.A., Convey, P., 2008. Surviving out in the cold: Antarctic

endemic invertebrates and their refugia. J. Biogeog. 35,

2176e2186.

Pugh, P.J.A., Dartnall, H.J.G., McInnes, S.J., 2002. The non-marine

Crustacea of Antarctica and the islands of the Southern Ocean:

biodiversity and biogeography. J. Nat. Hist. 36, 1047e1103.

Quayle, W.C., Convey, P., Peck, L.S., Ellis-Evans, J.C., Butler, H.G.,

Peat, H.J., 2003. Ecological responses of maritime Antarctic

lakes to regional climate change. In: Domack, E., Burnett, A.,

Leventer, A., Convey, P., Kirby, M., Bindschadler, R. (Eds.),

Antarctic Peninsula Climate Variability: Historical and Palae-

oenvironmental Perspectives. Antarctic Research Series, vol. 79.

American Geophysical Union, Washington, D.C, pp. 159e170.

Quayle, W.C., Peck, L.S., Peat, H., Ellis-Evans, J.C., Harrigan, P.R.,

2002. Extreme responses to climate change in Antarctic lakes.

Science 295, 645.

Sands, C.J., McInnes, S.J., Marley, N., Goodall-Copestake, W.P.,

Convey, P., Linse, K., 2008. Phylum Tardigrada: an ‘‘individual’’

approach. Cladistics 24, 861e871.

Scott, J.J., Kirkpatrick, J.B., 1994. Effects of human trampling on the

sub-Antarctic vegetation of Macquarie Island. Polar Record 30

(174), 207e220.

Searles, P.S., Kropp, B.R., Flint, S.D., Caldwell, M.M., 2001. Influ-

ence of solar UV-B radiation on peatland microbial communities

of southern Argentina. New Phytol. 152, 213e221.

Sinclair, B.J., 2002. Effects of increased temperatures simulating

climate change on terrestrial invertebrates on Ross Island,

Antarctica. Pedobiologia 46, 150e160.

Sjoling, S., Cowan, D.A., 2000. Detecting human bacterial contam-

ination in Antarctic soils. Polar Biol. 23, 644e650.

Sjoling, S., Cowan, D.A., 2003. High 16S rDNA bacterial diversity in

glacial meltwater lake sediment, Bratina Island, Antarctica.

Extremophiles 7, 275e282.

Smith, R.I.L., 1984. Terrestrial biology of the Antarctic and sub-

Antarctic. In: Laws, R.M. (Ed.), Antarctic Ecology. Academic

Press, London, pp. 61e162.

Page 13: Terrestrial biodiversity in Antarctica – Recent advances and future challenges

147P. Convey / Polar Science 4 (2010) 135e147

Smith, R.I.L., 1990. Signy Island as a paradigm of biological and

environmental change in Antarctic terrestrial ecosystems. In:

Kerry, K.R., Hempel, G. (Eds.), Antarctic Ecosystems, Ecological

Change and Conservation. Springer-Verlag, Berlin, pp. 32e50.

Smith, R.I.L., 1993. The role of bryophyte propagule banks in

primary succession: case study of an Antarctic fellfield soil. In:

Miles, J., Walton, D.W.H. (Eds.), Primary Succession on Land.

Blackwell, Oxford, pp. 55e78.

Smith, R.I.L., 1994. Vascular plants as indicators of regional

warming in Antarctica. Oecologia 99, 322e328.

Smith, R.I.L., 1996. Introduced plants in Antarctica: potential

impacts and conservation issues. Biol. Conserv 76, 135e146.

Smith, R.I.L., 2001. Plant colonization response to climate change in

the Antarctic. Folia Fac. Sci. Nat. Univ. Masarykianae Brunensis,

Geographia 25, 19e33.

Sømme, L., 1995. Invertebrates in Hot and Cold Arid Environments.

Springer-Verlag, Berlin.

Stevens, M.I., Frati, F., McGaughran, A., Spinsanti, G., Hogg, I.D.,

2007. Phylogeographic structure suggests multiple glacial refugia

in northern Victoria Land for the endemic Antarctic springtail

Desoria klovstadi, (Collembola, Isotomidae). Zool. Scripta 36,

201e212.

Stevens, M.I., Greenslade, P., Hogg, I.D., Sunnucks, P., 2006.

Examining Southern Hemisphere springtails: could any have

survived glaciation of Antarctica? Molec. Biol. Evol. 23,

874e882.

Taton, A., Grubisic, S., Balthasart, P., Hodgson, D.A., Laybourn-

Parry, J., Wilmotte, A., 2006. Biogeographical distribution and

ecological ranges of benthic cyanobacteria in East Antarctic

lakes. FEMS Microbiol. Ecol. 57, 272e289.

Tejedo, P., Justel, A., Benayas, J., Rico, E., Convey, P., Quesada, A.,

2009. Soil trampling in an Antarctic Specially Protected Area:

tools to assess levels of human impact. Antarct. Sci. 21,

229e236.

Tin, T., Fleming, Z., Hughes, K.A., Ainley, D., Convey, P.,

Moreno, C., Pfeiffer, S., Scott, J., Snape, I., 2009. Impacts of

local human activities on the Antarctic environment: a review.

Antarct. Sci. 21, 3e33.

Tindall, B.J., 2004. Prokaryotic diversity in the Antarctic: the tip of

the iceberg. Microb. Ecol. 47, 271e283.

Turner, J., Bindschadler, R., Convey, P., di Prisco, G., Fahrbach, E.,

Gutt, J., Hodgson, D., Mayewski, P., Summerhayes, C. (Eds.),

2009a. Antarctic Climate Change and the Environment. Scientific

Committee on Antarctic Research. Cambridge.

Turner, J., Comiso, J.C., Marshall, G.J., Lachlan-Cope, T.A.,

Bracegirdle, T.J., Maksym, T., Meredith, M.P., Wang, Z., Orr, A.,

2009b. Non-annular atmospheric circulation change induced by

stratospheric ozone depletion and its role in the recent increase of

Antarctic sea ice extent. Geophys. Res. Lett.. doi:10.1029/

2009GL037524.

Upton, M., Pennington, T.H., Haston, W., 1997. Detecting

commensals in the area around an Antarctic research station.

Antarct. Sci. 9, 156e161.

Vincent, W.F., 2000. Evolutionary origins of Antarctic microbiota:

invasion, selection and endemism. Antarct. Sci. 12, 374e385.

Vincent, W.F., MacIntyre, S., Spigel, R.H., Laurion, E., 2008. The

physical limnology of high-latitude lakes. In: Vincent, W.F.,

Laybourn-Parry, J. (Eds.), Polar Lakes and Rivers e Limnology

of Arctic and Antarctic Aquatic Ecosystems. Oxford University

Press, Oxford, pp. 65e81.

Vyverman, W., Verleyen, E., Wilmotte, A., Hodgson, D.A.,

Willem, A., Peeters, K., Van de Vijver, B., De Wever, A.,

Leliaert, F., Sabbe, K., 2010. Evidence for widespread endemism

among Antarctic micro-organisms. Polar Sci. 4 (2), 103e113.

Walther, G.-R., Post, E., Convey, P., Parmesan, C., Menzel, M.,

Beebee, T.J.C., Fromentin, J.-M., Hoegh-Guldberg, O.,

Bairlein, F., 2002. Ecological responses to recent climate change.

Nature 416, 389e395.

Walton, D.W.H., 1984. The terrestrial environment. In: Laws, R.M.

(Ed.), Antarctic Ecology. Academic Press, London, pp. 1e60.

Walton, D.W.H., Vincent, W.F., Timperley, M.H., Hawes, I., Howard-

Williams, C., 1997. Synthesis: polar deserts as indicators of

change. In: Lyons, W.B., Howard-Williams, C., Hawes, I. (Eds.),

Ecosystem Processes in Antarctic Ice-Free Landscapes. Balkema,

Rotterdam, pp. 275e279.

Whinam, J., Chilcott, N., Bergstrom, D.M., 2004. Subantarctic

hitchhikers: expeditioners as vectors for the introduction of alien

organisms. Biol. Conserv. 121, 207e219.

Woehler, E.J., Cooper, J., Croxall, J.P., Fraser, W.R., Kooyman, G.L.,

Miller, G.D., Nel, D.C., Patterson, D.L., Peter, H.-U., Ribic, C.A.,

Salwicka, K., Trivelpiece, W.Z., Weimerskirch, H., 2001. A

statistical assessment of the status and trends of Antarctic and

subantarctic seabirds. Report on SCAR BBS Workshop on

Southern Ocean Seabird Populations, 1e43.

Wynn-Williams, D.D., 1994. Potential effects of ultraviolet radiation

on Antarctic primary terrestrial colonizers: cyanobacteria, algae,

and cryptogams. Antarct. Res. Ser. 62, 243e257.

Wynn-Williams, D.D., 1996. Response of pioneer soil microalgal

colonists to environmental change in Antarctica. Microb. Ecol.

31, 177e188.

Yergeau, E., Bokhorst, S., Huiskes, A.H.L., Boschker, H.T.S.,

Aerts, R., Kowalchuk, G.A., 2007. Size and structure of bacterial,

fungal and nematode communities along an Antarctic environ-

mental gradient. FEMS Microbiol. Ecol. 59, 436e451.