MASARYK UNIVERSITY Faculty of science Jeffrey Clark Nekola INTERDISCIPLINARY INVESTIGATIONS INTO THE NATURE OF ECOLOGY AND EVOLUTION Postcards from along the road during my first quarter century of research Habilitation thesis Brno 2018 i JEFFREY CLARK NEKOLA. Interdisciplinary investigations into nature of ecology and evolution. ABSTRACT Why are there so many species of life on Earth? Why are they found where they are? How do they live together? How have they evolved? Why do they look the way they do? What do they tell us of the past and what do they suggest about the future? And how can they and humanity coexist in this finite world? These questions - and others - underlie my research. My primary focus is to understand the mechanisms which have given rise to our diverse world and the things we need to do to protect it. This thesis represents a current state of play across my interdisciplinary foray into the nature of ecology, accumulated during the first 25 years of research. This work begins with an overview of why I have been active in so many ecological fields, prefaced on the ancient Indian parable of the Blind Men and the Elephant. Following this, I have broken my working into seven major categories, ranging from Taxonomy, Faunistics, and Floristics through Population and Community Ecology, Spatio-temporal Ecology, Biogeography and Macroecology, Ecological Theory and Modeling, Conservation Biology, and Human Macroecology and Sustainability. The document ends with a concluding set of synthetic ideas that have emerged from this intergrated multi-scale series of analyses. ii ACKNOWLEDGEMENTS There have been innumerable people who has assisted my research efforts over the last quarter century – too many to individually name here. But among them a few rise to the surface. First, my PhD advisor Peter White has encouraged me from the start of my career to think broadly and holistically about ecological questions. I am grateful for the encouragement that Robert Cameron, Douglas Larson, Clifford Kraft, James Brown, and Brian McGill have provided over the years. I would likely not have stayed in the field without their efforts. I also greatly appreciate those over the last decade who have financially helped me keep the lights on in my lab and home, in particular Charles Drost, Ken Hotopp and Ben Hutchens. Lastly, I deeply thank Michal Horsák, Milan Chytrý and everyone else in the Department of Botany and Zoology at Masaryk University for offering me a career lifeline and opportunity to continue my research within the field. iii TABLE OF CONTENTS Page ABSTRACT ................................................................................................................................... i ACKNOWLEDGEMENTS ........................................................................................................ ii INTRODUCTION.........................................................................................................................2 SECTION I: Taxonomy, Floristics, and Faunistics Overview ............................................................................................................................ 3 Nekola, J.C. & B.F. Coles. 2016. Supraspecific taxonomy in the Vertiginidae (Gastropoda, Stylommatophora). Journal of Molluscan Studies. 82:208-212............ 6 Nekola, J.C., B.F. Coles & M. Horsák. 2015. Species assignment in Pupilla (Gastropoda: Pulmonata: Pupillidae): integration of DNA-sequence data and conchology. Journal of Molluscan Studies. 81:196-216............................................ 11 Nekola, J.C. & G. Rosenberg. 2013. Vertigo marciae (Gastropoda, Vertiginidae), a new land snail from Jamaica. The Nautilus. 127:107-114...................................... 32 Nekola, J.C. 1998. Butterfly (Lepidoptera: Lycaeniidae, Nymphalidae, and Satyridae) faunas of three peatland habitat types in the Lake Superior drainage basin of Wisconsin. Great Lakes Entomologist. 31:27-37......................................... 40 SECTION II: Population and Community Ecology Overview .......................................................................................................................... 52 Nekola, J.C. 2010. Acidophilic terrestrial gastropod communities of North America. Journal of Molluscan Studies. 76:144-156................................................. 54 Nekola, J.C. 2004. Vascular plant composition gradients within and between Iowa fens. Journal of Vegetation Science. 15:771-780.............................................. 67 Nekola, J.C. 2003. Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Diversity and Distribution. 9:55-71.......................................................................................................................... 77 Larson, D.W., U. Mattes-Sears, J.A. Gerrath, J.M. Gerrath, J.C. Nekola, G.L. Walker, S. Porembski, A. Charlton & N.W.K. Larson. 1999. Ancient stunted forests on cliffs. Nature. 398:382-383......................................................................... 94 SECTION III: Spatio-temporal Ecology Overview .......................................................................................................................... 96 Pigati, J.S., J.A. Rech & J.C. Nekola. 2010. Radiocarbon dating of small terrestrial gastropod shells in North America. Quaternary Geochronology. 5:519-532.............. 99 Nekola, J.C. & C.E. Kraft. 2002. Spatial constraint of peatland butterfly occurrences within a heterogeneous landscape. Oecologia. 130:62-71.................... 113 Nekola, J.C. 1999. Paleorefugia and neorefugia: the influence of colonization history on community pattern and process. Ecology. 80: 2459-2473....................... 122 iv SECTION IV: Biogeography and Macroecology Overview ........................................................................................................................ 137 Nekola, J.C. 2014. North American terrestrial gastropods through either end of a spyglass. Journal of Molluscan Studies. 80:238-248................................... 139 Nekola, J.C. & P.S. White. 1999. Distance decay of similarity in biogeography and ecology. Journal of Biogeography. 26:867-878................................................. 150 SECTION V: Ecological Theory and Modeling Overview ........................................................................................................................ 162 Nekola, J.C. & B. McGill. 2014. Scale dependency in the functional form of the distance decay relationship. Ecography. 37:309-320.......................................... 164 Nekola, J.C. & J.H. Brown. 2007. The wealth of species: ecological communities, complex systems, and the legacy of Frank Preston. Ecology Letters. 10:188-196..................................................................................................... 176 Bossenbroek, J.M., C.E. Kraft & J.C. Nekola. 2001. Prediction of long-distance dispersal using gravity models: Zebra Mussel invasion of inland lakes. Ecological Applications. 11:1178-1788...................................................................... 185 SECTION VI: Conservation Biology Overview ........................................................................................................................ 196 Nekola, J.C. 2012. The impact of utility corridors on terrestrial gastropod biodiversity. Biodiversity and Conservation. 21:781–795........................................ 197 McMillan, M., J.C. Nekola & D.W. Larson. 2003. Impact of recreational rock climbing on land snail communities of the Niagara Escarpment, southern Ontario, Canada. Conservation Biology. 17:616-621............................................................... 212 Nekola, J.C. 2002. Effects of fire management on the richness and abundance of central North American grassland land snail faunas. Animal Biodiversity and Conservation. 25:53-66........................................................................................ 218 SECTION VII: Human Macroecology and Sustainability Overview ........................................................................................................................ 232 Nekola, J.C., C.D. Allen, J.H. Brown, J.R. Burger, A.D. Davidson, T.S. Fristoe, M.J. Hamilton, S.T. Hammond, A. Kodric-Brown, N. Mercado-Silva & J.G. Okie. 2013. The Malthusian-Darwinian dynamic and the trajectory of civilization. Trends in Ecology and Evolution. 28:127-130........................................................... 234 Burger, J.R., J.H. Brown, C.D. Allen,, W.R. Burnside, A.D. Davidson, T.S. Fristoe, M.J. Hamilton, N. Mercado-Silva, J.C. Nekola, J.G. Okie & W. Zuo. 2012. The macroecology of sustainability. PLoS Biology. 10:e1001345..................................... 238 Brown, J.H., W.R. Burnside, A.D. Davidson, J.P. DeLong, W.C. Dunn, M.J. Hamilton, J.C. Nekola, J.G. Okie, N. Mercado-Silva, W.H. Woodruff & W. Zuo. 2011. Energetic Limits to Economic Growth. Bioscience. 61:19-26....................... 245 CONCLUSIONS ...................................................................................................................... 253 A group of blind men heard that a strange animal, called an elephant, had been brought to the town, but none of them were aware of its shape and form. Out of curiosity, they said: "We must inspect and know it by touch, of which we are capable". So, they sought it out, and when they found it they groped about it. In the case of the first person, whose hand landed on the trunk, said "This being is like a thick snake". For another one whose hand reached its ear, it seemed like a kind of fan. As for another person, whose hand was upon its leg, said, the elephant is a pillar like a tree-trunk. The blind man who placed his hand upon its side said, "an elephant is a wall". Another who felt its tail described it as a rope. The last felt its tusk, and stated an elephant is hard and smooth and like a spear. – Buddhist Udana 6.4 Six blind elephants were discussing what men were like. After arguing they decided to find one and determine it by direct experience. The first felt the man and declared, 'Men are flat.' After the other blind elephants felt the man, they agreed. – Mark Hartman We have to remember that what we observe is not nature in itself, but nature exposed to our method of questioning. – Werner Heisenberg 1 INTRODUCTION Why are there so many species of life on Earth? Why are they found where they are? How do they live together? How have they evolved? Why do they look the way they do? What do they tell us of the past and what do they suggest about the future? And how can they and humanity coexist in this finite world? These questions - and others - underlie my research. My primary focus is to understand the mechanisms which have given rise to our diverse world and the things we need to do to protect it. As intimated by the ancient Indian parable of the Blind Men and the Elephant, the ability to accurately interpret large-scale pattern – and identify potential underlying mechanism – is greatly enhanced by viewing systems through multiple lenses, preferably at different observational scales. This is especially important for ecological systems which range from sub-cellular to global scales and include multiple potential focal entities ranging from molecules to biomes. They are arranged into nested hierarchies of components, ranging from organic molecules and cells through populations, species, guilds, and trophic levels. They require the continual input of energy, material, and information to maintain their highly organized, far-from-equilibrium thermodynamic states. Even the simplest ecological system contains thousands to billions of individuals of tens to thousands of different species, ranging from unicellular prokaryotes, protists, and fungi to multicellular plants and animals. These individuals and species interact with each other and their extrinsic abiotic environment in inherently non-linear ways. Such relationships are neither entirely deterministic nor stochastic, include positive and negative feedbacks, and often possess long-lasting contingent effects across both space and time. Discovering general processes supporting and maintaining diversification thus by necessity requires an interdisciplinary focus. While this can certainly be accomplished via collaboration – and I have participated in many such projects – I have also found it useful to be interdisciplinary within my own brain. In this way I maintain hands-on, real-word experience across the full range of biodiversity topics, and can use this knowledge to identify common threads and understand potential analytical pitfalls. Often this leads to novel insights. While the specialization of research programs can lead to fundamental discoveries, it is important to remember that for almost a century biological and ecological research was conducted by investigators who held expertise across a wide breadth of the natural sciences. 19th Century field naturalists like Charles Darwin and Alfred Russell Wallace made lasting and often radical impacts on the field, not the least of which was the development of evolutionary theory. Given the complexity of ecological systems, some advances may only be possible via the inductive, holistic, and non-experimental techniques pioneered by these scientists. In combination with the modern tools of contemporary ecology, the field naturalist model can thus serve as an avenue for important future advances in ecological theory. Inspired by the astonishing breadth of knowledge accumulated by investigators such as Iowa’s Bohumil Shimek, who became a national authority in botany, malacology, ecology, paleoecology, and glacial geology (as well as being a close personal friend of Tomáš Masaryk and champion for Czechoslovakian independence), I have organized my research program upon the field naturalist model to address fundamental issues in the diversification and organization of ecological communities. As a result, I have vertically integrated my research activities to range 2 6 of the approximately 75 North American Vertigo species from organism taxonomy through population, community, and spatiotemporal ecology to biogeography, macroecology, and ecological theory and modeling. To ensure that my outlook does not become myopic, I have also horizontally integrated this program across three divergent taxa groups: vascular plants (sessile autotrophs), lepidoptera (mobile heterotrophs), and terrestrial gastropods (functionally sessile heterotrophs). In addition, I have maintained an active field research program which spans thousands of study sites on three continents. I am particularly interested in documenting findings that challenge current paradigms, as it is through such observations that scientific revolutions are based. The essential component of such an eclectic research program is to do what the blind men in the above parable did not: integrate their findings to allow documentation of large-scale, crosssystem reality. Accomplishing this task requires not only the ability to be expert within multiple frames of reference, but also the ability to take a cross-system philosophic perspective to identify potential correspondences and synergies. And like the elephants in the second parable, it is essential that the limitations of given observational approaches and deductive / inductive strategies be understood. Thus, as intimated by Heisenberg, by observing nature and biodiversity through multiple methods of questioning, it may be possible to glimpse the true nature of nature. Over the course of the last quarter century I have published 75 peer-reviewed works. As of October 27, 2017 these have generated 5266 citations with an H-index=26, a G-index=72, and an Age-Weighted Citation Rate=443.2. My research activities can be roughly grouped into seven categories: (1) Taxonomy, Floristics and Faunistics; (2) Population and Community Ecology; (3) Spatio-Temporal Ecology; (4) Biogeography and Macroecology; (5) Ecological Theory and Modeling; (6) Conservation Biology; and, (7) Human Macroecology. Each is represented in the following document by its own section. Each section leads off with a summary statement of my research activates in that field, followed by a list of publications, and then by representative examples of published articles. Section I: Taxonomy, Floristics, and Faunistics By being independent evolutionary units, species are one of the basic units underlying ecological pattern and process. An accurate understanding of taxonomy and the valid taxa known within a given system is thus a vital foundation for most ecological and biogeographic investigations. For this reason, I maintain an active interest in taxonomy, floristics, and faunistics. 3 Much of my recent work centers on terrestrial gastropods, where I have named new species, revised taxonomic concepts, and enumerated the taxa and distribution (both geographic and ecological) of both regional and local faunas based on nDNA and mtDNA sequence, shell features, ecological preferences, and biogeography. I have just completed a global phylogenetic analysis of the micro-mollusk genus Vertigo based on analysis of nDNA (ITS1 and ITS2), and mtDNA (CytB and 16S) amplicons from 91 putative taxa and 398 individuals. Using a consensus across both sets of DNA sequence, shell features, ecological preferences, and biogeography, six subgenera and 87 species or subspecies are empirically validated. This work demonstrated: (1) a suite of diagnostic shell features usually exists to demarcate each specieslevel taxon, though these are often not the features historically used; (2) ecological preferences evolve more slowly as compared to shell features; (3) subgeneric transcontinental ranges are ubiquitous; (4) 1/3 of all species possess continental or trans-continental ranges, with very few having range extents <1000 km.; (5) all subgenera and fully 2/3 of the global Vertigo taxa reside in North America, more than 2½ times the number found in central and eastern Asia, the second most diverse region. This is similar to several other molluscan groups for which North America serves as the global biodiversity hotspot. Earlier work also documents that diversification in the genus appears synchronous with past episodes of rapid global environment change. Determination of general trends and potential underlying mechanism requires that these analyses be replicated in other distantly related groups. The Pristilomatidae are an obvious target being not only in a different superfamily but also small (2-8 mm in diameter) and diverse, representing the fifth largest family within the North American fauna (64 recognized species). Paravitrea (42 species) is the sixth most speciose genus. Unlike Vertigo many Pristilomatidae have very limited ranges (often only a few counties in extent) with 100% of Pilsbryna species, 90% of Paravitrea species and of 73% Pristiloma species being listed as Globally Threatened or higher. Specieslevel diagnoses have historically relied on the number, shape, and placement of lamellae across shell ontogeny. This is troubling given the significant incidence of ecophenotypic and biogeographic plasticity in these features in other genera. As a result it is unclear how many valid species-level groups actually exist or what their true biodiversity, biogeographic and evolutionary patterns represent. Representative Publications [number of citations as of October 27, 2017] 1. Peer-reviewed Books: Schlicht, D.W., J.C. Downey & J.C. Nekola. 2007. The Butterflies of Iowa. University of Iowa Press. [26] 2. Peer-reviewed Articles: Nekola, J.C., S. Chiba, B.F. Coles, C.A. Drost, T. vonProschwitz & M. Horsák. In press. A phylogenetic overview of the genus Vertigo O. F. Müller, 1773 (Gastropoda: Pulmonata: Pupillidae: Vertigininae). Malacologia. Nekola, J.C. & B.F. Coles. 2016. Supraspecific taxonomy in the Vertiginidae (Gastropoda, Stylommatophora). Journal of Molluscan Studies. 82:208-212. [4] 4 Nekola, J.C., B.F. Coles & M. Horsák. 2015. Species assignment in Pupilla (Gastropoda: Pulmonata: Pupillidae): integration of DNA-sequence data and conchology. Journal of Molluscan Studies. 81:196-216. [12] Nekola, J.C. & G. Rosenberg. 2013. Vertigo marciae (Gastropoda, Vertiginidae), a new land snail from Jamaica. The Nautilus. 127:107-114. [2] Nekola, J.C., A. Jones, G. Martinez, S. Martinez, K. Mondragon, T. Lebeck, J. Slapcinsky & S. Chiba. 2012. Vertigo shimochii Kuroda & Amano 1960 synonymized with Gastrocopta servilis (Gould, 1843) based on conchological and DNA sequence data. Zootaxa. 3161:48-52. [3] Nekola, J.C. & B.F. Coles. 2010. Pupillid land snails of eastern North America. American Malacological Bulletin. 28:29-57. [38] Nekola, J.C., B.F. Coles & U. Bergthorsson. 2009. Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Molecular Phylogenetics and Evolution. 53:1010-1024. [38] Coles, B.F. & J.C. Nekola. 2007. Vertigo malleata, a new extreme calcifuge land snail from the Atlantic and Gulf coastal plains of the U.S.A. (Gastropoda, Vertiginidae). The Nautilus. 121:17-28. [10] Nekola, J.C. & M. Barthel. 2002. Morphometric analysis of the genus Carychium in the Great Lakes region of North America. Journal of Conchology. 37(5):515-531. [18] Nekola, J.C. & B.F. Coles. 2001. Systematics and ecology of Gastrocopta (Gastrocopta) rogersensis (Gastropoda: Pupillidae), a new species from the midwest of the United States of America. The Nautilus. 115:105-114. [16] Nekola, J.C. 1998. Butterfly (Lepidoptera: Lycaeniidae, Nymphalidae, and Satyridae) faunas of three peatland habitat types in the Lake Superior drainage basin of Wisconsin. Great Lakes Entomologist. 31:27-37. [10] Nekola, J.C. 1994. Environment and vascular flora of northeastern Iowa fen communities. Rhodora. 96:121-169. [23] 5 RESEARCH NOTE Supraspecific taxonomy in the Vertiginidae (Gastropoda: Stylommatophora) Jeffrey C. Nekola1 and Brian F. Coles2 1 Department of Biology, University of New Mexico, Albuquerque, NM 87131, USA; and 2 Mollusca Section, Department of Biodiversity, Amgueddfa Cymru - National Museum Wales, Cathays Park, Cardiff CF10 3NP, UK Correspondence: J. C. Nekola; e-mail: jnekola@unm.edu The genus Vertigo Mu¨ ller, 1774 consists of c. 100 species of terrestrial microsnails c. 1.5–3 mm in length with a rounded aperture and 0–6 (sometimes more) apertural lamellae at maturity. As currently defined, the genus is largely Holarctic in distribution with only a few Neotropical species being known (Pilsbry, 1948; Nekola & Rosenberg, 2013). Consensus does not exist concerning supraspecific taxonomy of the genus. Pilsbry (1919, 1927, 1948) placed Vertigo, Columella Westerlund, 1878 and Truncatellina Lowe, 1852 in the family Pupillidae, subfamily Vertigininae. Based largely on anatomy, he differentiated these from the subfamily Nesopupinae, into which he placed 11 mostly tropical genera: Bothriopupa Pilsbry, 1898, Campolaemus Pilsbry, 1892, Costigo Boettger, 1891, Cylindrovertilla Boettger, 1880, Lyropupa Pilsbry, 1900, Nesopupa Pilsbry, 1900, Pronesopupa Iredale, 1913, Ptychalaea Boettger, 1889, Pupisoma Stoliczka, 1873, Staurodon Lowe, 1854 and Sterkia Pilsbry, 1898. Bouchet & Rocroi (2005) assigned Vertigo to the family Vertiginidae, which in their scheme comprised three subfamilies: Vertigininae, Nesopupinae (both largely as designated by Pilsbry, 1927, but with Pupisoma being moved to the Valloniidae) and Gastrocoptinae (including Gastrocopta Wollaston, 1878 and 10 related genera). The Gastrocoptinae have subsequently been assigned to the Chondrinidae (Pokryszko et al., 2009). Pilsbry (1919, 1948) recognized four subgenera within Vertigo: Vertigo s.s. and the monotypic Angustula Sterki, 1888, Vertilla Moquin-Tandon, 1855 and Vertillaria Pilsbry, 1919. He further divided Vertigo s.s. into several informal sections and groups. This treatment was followed until Turgeon et al. (1998) and Sysoev & Schileyko (2009), respectively, considered the group Nearctula Sterki, 1892 and the subgenus Vertilla to be of generic rank. Similar intrageneric division has been prominent within Nesopupa; e.g. Gittenberger & van Bruggen (2013) recognized nine genera originally proposed as sections by Pilsbry & Cooke (Pilsbry, 1919): Afripupa, Cocopupa, Helenopupa, Indopupa, Infranesopupa, Insulipupa, Nesodagys, Nesopupa and Nesopupilla. To address these issues empirically, we have assembled and phylogenetically analysed DNA sequence data from the nuclear 28S ribosomal RNA (28S) and mitochondrial 16S ribosomal RNA (16S) genes. The species selected for study included the full conchological, biogeographic and ecological range of Vertigo, plus representatives of Columella, Gastrocopta, Nearctula and Truncatellina, and a variety of putative nesopupillids and other genera within the Pupillidae (sensu Pilsbry, 1927). We also included a representative sampling across the infraorder Orthurethra as designated by Wade, Mordan & Clarke (2001). Among the analysed taxa, the following represent the type species of their respective genera: Acanthinula aculeata (Mu¨ ller, 1774), Chondrina avenacea (Bruguie` re, 1792), Cochlicopa lubrica (Mu¨ ller, 1774), Columella edentula (Draparnaud, 1805), Helix pomatia Linne´ , 1758, Lauria cylindracea (da Costa, 1778), Leiostyla anglica (Wood, 1828), Nearctula californica (Rowell, 1862), Planogyra asteriscus (Morse, 1857), Pupilla muscorum (Linne´ , 1758) Pyramidula rupestris (Draparnaud, 1801), Solatopupa similis (Bruguie` re, 1792), Sterkia calamitosa (Pilsbry, 1889), Strobilops labyrinthica (Say, 1817), Vallonia costata (Mu¨ ller, 1774), Vertigo pusilla (Mu¨ ller, 1774), Vertilla angustior (Jeffreys, 1830) and Zoogenetes harpa (Say, 1824). Carychium tridentatum, Helix pomatia and Cornu aspersum were used as comparative outgroups. Archived GenBank sequences were used for 24 specimens, including data first reported by Armbruster et al. (2005) [AY546471], Dinapoli, Zinssmeister & Klussmann-Kolb (2010) [GU331954], Gaitan-Espitia, Nespolo & Opazo (2013) [JQ417194], Ketmaier et al. (2010) [jGU046389], Nekola & Rosenberg (2013) [KF214500, KF214496], Nekola et al. (2012) [JN941017, JN941032, JN941041, JN941044], Nekola, Coles & Bergthorsson (2009) [GQ921543], Wade, Mordan & Naggs (2006) [AY841284, AY841285, AY841286, AY841333], Wade et al. (2001) [AY014019, AY014020, AY014022, AY014023, AY014024, AY014025, AY014027, AY014028, AY014028, AY014030, AY014032, AY014033, AY014040, AY014148] and Weigand et al. (2013) [KC206171]. Sequences for the remaining 35 specimens (Table 1) were newly obtained. DNA extraction, purification, PCR amplification and sequencing were performed using previously published methods and primers (see Wade & Mordan, 2000; Nekola & Rosenberg, 2013). The amplified 28S region (Fig. 1) encompasses ITS-2, a region that cannot be aligned because of its intergeneric hypervariability. As a result, all sequence more than 284 bp upstream of the LSU2 primer of Wade & Mordan (2000) was excluded from analysis. The resultant analysed 28S amplicon ranged in length from 809 bp (Acanthinula aculeta) to 827 bp (Columella edentula). The entire 16S amplicon was used for analysis and ranged in length from 403 bp (Cornu aspersum) to 594 bp (Truncatellina cylindrica). Sequences were aligned using ClustalX, with adjustment by eye. Mega v. 5.0 was used to conduct neighbour-joining (NJ), maximum parsimony (MP) and maximum likelihood (ML) analyses separately for the 28S and 16S sequences. NJ analysis was based on maximum composite distance including transitions # The Author 2015. Published by Oxford University Press on behalf of The Malacological Society of London, all rights reserved Journal of The Malacological Society of London Molluscan Studies Journal of Molluscan Studies (2016) 82: 208–212. doi:10.1093/mollus/eyv034 Advance Access publication date: 20 July 2015 atUniversityofNewMexicoonFebruary15,2016http://mollus.oxfordjournals.org/Downloadedfrom 6 Table 1. Collection information for specimens analysed in this study. Taxon Country State or equivalent County or equivalent Site Lat./Long. Collection/ accession no. 16S GenBank acc. no. 28S GenBank acc. no. Afripupa bisulcata Ghana Fetish Grove 5.3356 N, 0.0801 W deWinter KT008314 KT008353 Bothriopupa tenuidens Costa Rica Puntarenas Saladera Lodge 8.6988 N, 83.3302 W Tattersfield KT008355 Columella edentula France Normandy Calvados La Vale´e des Vaux 48.9446 N, 0.4678 W BC00001 KT008310 KT008360 Leiostyla anglica UK Cumbria Meathop Cliff Seep 54.2066 N, 2.8729 W JCN KT008308 KT008363 Nearctula californica USA California Monterey Moss Landing Beach 36.8095 N, 121.7880 W JCN13934 KT008315 KT008349 Nearctula rowelli USA California Placer Auburn 38.9072 N, 121.0520 W JCN17232 KT008327 KT008347 Nesopupa newcombi USA Hawaii Honolulu Waianae Forest Reserve 21.5009 N, 158.1680 W UF469195 KT008307 KT008356 Planogyra asteriscus USA Maine Piscataquis Atkinson Mills 45.1421 N, 69.0312 W JCN10494 KT008305 KT008364 Pronesopupa boettgeri USA Hawaii Oahu Kalihi Valley 21.3585 N, 157.8580 W Hayes KT008306 KT008354 Ptychalaea tamagonori Japan Tokyo Ogasawara Chichijima 27.0948 N, 142.2166 E Chiba KT008323 KT008348 Pupilla muscorum Czech Rep. Moravia Brno 49.2509 N, 16.5738 E Horsak KT008313 KT008361 Pupisoma dioscoricola USA Florida Alachua Batram-Carr Woods 29.6436 N, 82.3456 W Slapcinsky KT008304 KT008357 Sterkia calamitosa USA California San Diego Sumner Canyon 32.8734 N, 117.2483 W JCN18169 KT008324 KT008350 Sterkia clementia USA California San Diego San Clemente Island 32.9953 N, 118.5516 W JCN19109 KT008325 KT008351 Sterkia hemphilli USA California San Diego Border Fields State Park 32.5428 N, 117.1061 W JCN19780 KT008331 KT008352 Strobilops labyrinthica USA Michigan Mackinac Brevort Lake 45.9829 N, 84.8618 W JCN17151 KT008309 KT008362 Truncatellina callicratis France Normandy Calvados Pont, N. of Falaise 48.9774 N, 0.0902 W BC00002 KT008302 KT008358 Truncatellina cylindrica Ukraine Crimea Rozovyi 44.6920 N, 34.3190 E Cameron KT008303 KT008359 Vallonia costata USA New Mexico Bernalillo Albuquerque 35.0727 N, 106.6160 W JCN KT008312 Vertigo alabamensis USA Florida Leon Wolf Trap Bay 30.3680 N, 84.5700 W JCN12323 KF214500 KT008339 Vertigo antivertigo UK Wales Anglesey Waun Eraud Fen 53.3008 N, 4.2411 W JCN KT008316 KT008338 Vertigo clappi USA Tennessee Carter Lynn Mountain 36.3545 N, 82.1535 W UF299578 KT008321 KT008337 Vertigo cristata USA Wisconsin Oneida Sugar Camp Bog 45.8499 N, 89.2958 W JCN12213 GQ921543 KT008344 Vertigo gouldii USA West Virginia Lincoln Big Ugly WMA 38.0859 N, 82.0009 W Doursen KT008330 KT008333 Vertigo cf. okinerabuensis Japan Iwate Ichinoseki Sarusawa 38.9869 N, 141.2550 E Chiba JN941044 KT008341 Vertigo malleata USA North Carolina Pender Holly Shelter Game Land 34.5492 N, 77.7817 W UF449308 KT008318 KT008345 Vertigo meramecensis USA Kentucky Bullitt Cave Hollow 37.9379 N, 85.6334 W JCN KT008329 KT008335 Vertigo milium USA Iowa Buchanan Rowley North fen 42.3764 N, 91.8507 W JCN5747 KT008328 KT008346 Vertigo moulinsiana Netherlands South Holland Kaag en Braassem Woubrugge, Widje Aa 52.1710 N, 4.6119 E JCN KT008326 KT008334 Vertigo pseudosubstriata Russia Altai Republic Seminski Pass S 50.9855 N, 85.6817 E JCN KT008317 KT008343 Vertigo pusilla Czech Rep. Moravia Podyji National Park 48.8586 N, 15.8960 E Horsak KF214496 KT008342 Vertigo pygmaea Czech Rep. Moravia Kotrle Fen 49.3779 N, 18.0236 E JCN KT008322 KT008340 Vertigo rugosula USA Arkansas Jefferson Lock & Dam #5 34.4040 N, 92.1020 W UF409059 KT008319 KT008332 Vertilla angustior Czech Rep. Moravia Pozdechov Fen 49.2339 N, 17.9864 E JCN KT008320 KT008336 Zoogenetes harpa Canada Quebec Manicougan Pointe-des-Montes 49.3256 N, 67.3700 W JCN13609 KT008311 Abbreviations: BC, Coles collection in National Museum of Wales, of which the accession numbers are to be preceded by ‘NMW.Z.2015.009’; JCN, Nekola collection in University of New Mexico; UF, material in Florida Museum of Natural History, Gainesville. RESEARCHNOTE 209 atUniversityofNewMexicoonFebruary15,2016 http://mollus.oxfordjournals.org/ Downloadedfrom 7 Figure 1. Map of the rRNA gene cluster with primer locations for the analysed segment of 28S in relation to the 5.8S and ITS-2 regions. Figure 2. Maximum-likelihood phylogenetic tree reconstruction based on nuclear 28S (left) and mitochondrial 16S (right) data. Nodes with strong to moderate support (!70) across all four phylogenetic reconstruction methods have been labelled to the left of that node by four support values: upper left (normal font) is for NJ; upper right (bold italic font) is for MP; lower left (bold font) is for Bayesian; lower right (italic font) is for ML. RESEARCH NOTE 210 atUniversityofNewMexicoonFebruary15,2016http://mollus.oxfordjournals.org/Downloadedfrom 8 and transversions with pairwise gap deletion. MP analysis used the close-neighbour interchange search option with the random addition of 10 replicate trees. ML analysis used all sites and was based on the Tamura-Nei substitution model, a five-category gamma distribution for substitution rates and the nearest neighbour interchange ML heuristic method. In all cases support values were estimated from 1,000 bootstrap replicates. Additionally, Bayesian trees were generated using MrBayes v. 3.1 (Huelsenbeck & Ronquist, 2001), with a generalised time reversible substitution model assuming gamma-shaped rate variation over 1,000,000 generations and a sampling frequency of once each 1,000 generations. Analysis of these data (Fig. 2) demonstrates poorer resolution using the mitochondrial 16S as compared with the nuclear 28S data, presumably due to higher base-pair saturation rates. Therefore, we used the 28S data to deduce phylogenetic relationships and the 16S data for quasi-independent corroboration. The 28S data show a highly-supported clade composed of two highly-supported sister clades, together including most of the putative Vertigininae (Clade V) and Nesopupinae (Clade N) included in this study. This ‘vertiginid’ clade in general corresponds with the Vertiginidae as outlined by Bouchet & Rocroi (2005) and Pokryszko et al. (2009), with the marked exceptions that: (1) Gastrocopta is excluded as it is neither closely related to Vertigo nor to the chondrinids, but rather represents a distinct branch within the Orthurethra, and (2) both Columella and Truncatellina are excluded as both are members of a moderately supported clade that includes Chondrina. Wade et al. (2001, 2006) have shown that the chondrinids are actually sister to the Orthurethra. The analysis of 16S data in general corroborates these findings, albeit without resolving the sister status of clades V and N. The 28S and 16S trees both demonstrate that Clade N consists only of tropical species historically assigned to the Nesopupinae. These are known to differ anatomically from the Vertigininae by possessing a penial appendix and a forked retractor muscle (Pilsbry, 1919). Clade V includes not only all analysed Vertigo, but also the putative genera Nearctula and Vertilla and members of some genera historically assigned to the Nesopupinae (Afripupa, Ptychalaea and Sterkia). Pilsbry (1919) lacked anatomical data for any species within these latter three genera and placed them outside of Vertigo solely on biogeographical (tropical/subtropical range) and conchological grounds (pustulate shell sculpture, and/or strength and location of angular lamella). Previous analyses have shown that these shell features are not reliable for use in supraspecific taxonomy of pupillids (Coles & Nekola, 2007; Nekola et al., 2009; Nekola, Coles & Horsa´ k, 2015). It should also be noted that for both the 28S and 16S trees, monophyly of Vertigo is only preserved when the taxonomic concept of the genus is expanded to include all members of Clade V. Because support levels for this clade are similar to other apparent genus-level clades (including Cochlicopa, Gastrocopta, Lauria and Truncatellina), our data suggest that it would be most parsimonious to assign all species within Clade V to the genus Vertigo. Our more limited sampling and representation of type species in the narrowly-constrained Nesopupinae (Clade N) makes it impossible for us to say how many generic-level entities might be supported in that group. However it is interesting to note that the genetic divergence between Bothriopupa and Nesopupa is well within the range found within Vertigo. It thus remains an open question how many genera within the Nesopupinae will ultimately be supported by DNA sequence data. These data do not support subgeneric classification within Vertigo, given that no nodes within Clade V possess support values .70 in either 28S or 16S analyses. However, the weakly supported monophyly of Nearctula/Sterkia suggests that useful subgeneric groupings may in fact exist. Consideration of these issues must await analysis of other amplicons that are more rapidly evolving (therefore with more taxonomically useful sites) than 28S, but more slowly evolving (thus with lower base-pair saturation rates) than 16S. If retained, the family Vertiginidae represents a highly supported clade possessing only two branches: Vertigo on one hand and a subset of nesopupids on the other. But how useful is recognition of so small a family? We agree with Puillandre et al. (2015) that such decisions are ultimately matters of taxonomic opinion that cannot be subjected to empirical criteria. In this we freely admit to erring on the side of conservatism, as we do not see the utility of erecting a large number of high-level taxonomic groups that constitute only single branches or simple two-branch entities when a more inclusive division containing more taxa is possible by simply stepping back one level in the tree. Even though outside our primary focus, these analyses also hint that other traditional supraspecific concepts within the Orthurethra may be in need of revision. In particular, the Valloniidae may be polyphyletic with both Pupisoma and Acanthinula/Planogyra each representing their own unique branches within the Orthurethra. The 16S data also suggest that Zoogenetes may be more closely related to Pupisoma. Pilsbry (1927: vii) appears to have been prescient when he stated “The division of the Orthurethra into families seems to be largely a matter of expediency.” These empirical data validate this statement and strongly suggest that formal reconsideration of supraspecific concepts across the entire infraorder are warranted, based upon DNA sequence data. ACKNOWLEDGEMENTS Specimens used in this analysis were generously provided by Robert Cameron, Satoshi Chiba, Dan Doursen, Ton de Winter, Ken Hayes, Michal Horsa´ k, John Slapcinsky, John Stanisic and Peter Tattersfield. Generous personal donations by Michal Horsa´ k covered laboratory expenses and this work would simply have been impossible without his assistance. REFERENCES ARMBRUSTER, G.F.J., BOEHME, M., BERNHARD, D. & SCHLEGEL, M. 2005. The H3/H4 histone gene cluster of land snails (Gastropoda: Stylommatophora): TS/TV ratio, GC3 drive and signals in Stylommatophoran phylogeny. Journal of Molluscan Studies, 71: 339–348. BOUCHET, P. & ROCROI, J.-P. 2005. Classification and nomenclature of gastropod families. Malacologia, 47: 1–397. COLES, B.F. & NEKOLA, J.C. 2007. Vertigo malleata, a new extreme calcifuge land snail from the Atlantic and Gulf coastal plains of the U.S.A. (Gastropoda, Vertiginidae). Nautilus, 121: 17–28. DINAPOLI, A., ZINSSMEISTER, C. & KLUSSMANN-KOLB, A. 2010. New insights into the phylogeny of the Pyramidellidae (Mollusca, Gastropoda). Journal of Molluscan Studies, 7: 1–7. GAITAN-ESPITIA, J.D., NESPOLO, R.F. & OPAZO, J.C. 2013. The complete mitochondrial genome of the land snail Cornu aspersum (Helicidae: Mollusca): intra-specific divergence of protein-coding genes and phylogenetic considerations within Euthyneura. PLoS One, 8: E67299. GITTENBERGER, E. & VAN BRUGGEN, A.C. 2013. Land snails of the islet of Misali, off Pemba Island, Zanzibar, Tanzania. Zoologische Mededelingen, 87: 235–273. HUELSENBECK, J.P. & RONQUIST, F. 2001. MRBAYES: Bayesian inference of phylogeny. Bioinformatics, 17: 754–755. KETMAIER, V., MANGANELLI, G., TIEDEMANN, R. & GIUSTI, F. 2010. Prei-Tyrrhenian phylogeography in the land snail Solatopupa guidoni (Pulmonata). Malacologia, 52: 81–96. NEKOLA, J.C., COLES, B.F. & BERGTHORSSON, U. 2009. Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Molecular Phylogenetics and Evolution, 53: 1010–1024. RESEARCH NOTE 211 atUniversityofNewMexicoonFebruary15,2016http://mollus.oxfordjournals.org/Downloadedfrom 9 NEKOLA, J.C., COLES, B.F. & HORSA´ K, M. 2015. Species assignment in Pupilla (Gastropoda: Pulmonata: Pupillidae): integration of DNAsequence data and conchology. Journal of Molluscan Studies, 81: 196–216. NEKOLA, J.C., JONES, A., MARTINEZ, G., MARTINEZ, S., MONDRAGON, K., LEBECK, T., SLAPCINSKY, J. & CHIBA, S. 2012. Vertigo shimochii Kuroda & Amano 1960 synonymized with Gastrocopta servilis (Gould, 1843) based on conchological and DNA sequence data. Zootaxa, 3161: 48–52. NEKOLA, J.C. & ROSENBERG, G. 2013. Vertigo marciae (Gastropoda, Vertiginidae), a new land snail from Jamaica. Nautilus, 127: 107–114. PILSBRY, H.A. 1919. Manual of conchology, second series. Vol. 25: Pupillidae (Vertigininae, Pupillinae). Academy of Natural Sciences, Philadelphia. PILSBRY, H.A. 1927. Manual of conchology, second series. Vol. 28: Geographic distribution of Pupillidae. Academy of Natural Sciences, Philadelphia. PILSBRY, H.A. 1948. Land Mollusca of North America (north of Mexico). Academy of Natural Sciences of Philadelphia Monograph, 3: vol. 2(2). POKRYSZKO, B.M., AUFFENBERG, K., HLAVA´ Cˇ , J. & NAGGS, F. 2009. Pupilloidea of Pakistan (Gastropoda: Pulmonata): Truncatellininae, Vertigininae, Gastrocoptinae, Pupillinae (in part). Annales Zoologici (Warszawa), 59: 423–458. PUILLANDRE, N., DUDA, T.F., MEYER, C., OLIVERA, B.M. & BOUCHET, P. 2015. One, four or 100 genera? A new classification of the cone snails. Journal of Molluscan Studies, 81: 1–23. SYSOEV, A. & SCHILEYKO, A. 2009. Land snails and slugs of Russia and adjacent Countries. Pensoft, Sofia. TURGEON, D.D., QUINN, J.F., JR., BOGAN, A.E., COAN, E.V., HOCHBERG, F.G., LYONS, W.G., MIKKELSEN, P., NEVES, R.J., ROPER, C.F.E., ROSENBERG, G., ROTH, B., SCHELTEMA, A., THOMPSON, F.G., VECCHIONE, M. & WILLIAMS, J.D. 1998. Common and scientific names of aquatic invertebrates from the United States and Canada: mollusks. Edn 2. American Fisheries Society Special Publication 26, Bethesda, MD. WADE, C.M. & MORDAN, P.B. 2000. Evolution within the gastropod molluscs; using the ribosomal RNA gene-clustre as an indicator of phylogenetic relationships. Journal of Molluscan Studies, 66: 565–570. WADE, C.M., MORDAN, P.B. & CLARKE, B. 2001. A phylogeny of the land snails (Gastropoda: Pulmonata). Proceedings of the Royal Society B, 268: 413–422. WADE, C.M., MORDAN, P.B. & NAGGS, F. 2006. Evolutionary relationships among the Pulmonate land snails and slugs (Pulmonata, Stylommatophora). Biological Journal of the Linnean Society, 87: 593–610. WEIGAND, A.M., JOCHUM, A., SLAPNIK, R., SCHNITZLER, J., ZARZA, E. & KLUSSMANN-KOLB, A. 2013. Evolution of microgastropods (Ellobioidea, Carychiidae): integrating taxonomic, phylogenetic and evolutionary hypotheses. BMC Evolutionary Biology, 13: 18. RESEARCH NOTE 212 atUniversityofNewMexicoonFebruary15,2016http://mollus.oxfordjournals.org/Downloadedfrom 10 Species assignment in Pupilla (Gastropoda: Pulmonata: Pupillidae): integration of DNA-sequence data and conchology Jeffrey C. Nekola1, Brian F. Coles2 and Michal Horsa´ k3 1 Biology Department, University of New Mexico, Albuquerque, NM 87131, USA; 2 Mollusca Section, Department of Biodiversity, Amgueddfa Cymru – National Museum Wales, Cathays Park, Cardiff CF10 3NP, UK; and 3 Department of Botany and Zoology, Masaryk University, Kotla´rˇska´ 2, Brno CZ-611 37, Czech Republic Correspondence: J.C. Nekola; e-mail: jnekola@unm.edu (Received 23 October 2013; accepted 2 September 2014) ABSTRACT Using the Pupilla faunas of Europe, North America, the Altai region of central Asia and eastern Asia, we consider whether the existing taxonomy based primarily on shell apertural characteristics correlates with relationships established on the basis of mitochondrial and nuclear DNA-sequence data. We obtained DNA sequence from nuclear ITS1 and ITS2 and mitochondrial COI and CytB from 80 specimens across 22 putative Pupilla taxa. The sequence data were analysed using maximum likelihood, maximum parsimony, Bayesian and neighbour-joining phylogenetic tree reconstruction, as well as base-pair substitution and insertion-deletion analysis. Revised species-level concepts were generated by identifying reciprocally monophyletic clades that exhibited unique conchological features. These analyses document that, although many previously described taxa have biological merit, the highly plastic nature of shell apertural features makes them unreliable indicators of species identity in several independent lineages. However, shell surface sculpture and architecture appear to provide more reliable diagnoses. Because of the traditional reliance of species-level taxonomy in Pupilla on plastic apertural features, too many species-level entities have been described in Europe and the Altai. Also, because taxonomically useful shell sculpture features have tended to be ignored, too few species have been described in eastern Asia and North America. As a result, confusion exists about species ranges, ecological tolerances and interpretation of Quaternary fossils within the genus. Based on these analyses three new species are described: P. alaskensis, P. hudsonianum and P. hokkaidoensis. INTRODUCTION The use of protein markers and DNA sequence data has shown that conchological traits can be poor indicators of relatedness, for example in eastern North American Polygyridae (Emberton, 1995), Thailand Gastrocoptinae (Tongkerd et al., 2004) and the clausiliid subfamily Alopiinae from Greece (Uit de Weerd et al., 2004). Similarly, our analysis of DNA sequence data in the genus Vertigo (Pupilloidea) indicates that while shell features do generally provide accurate assignment of genetically validated species-level entities, they are too labile to resolve evolutionary relatedness (Nekola et al., 2009). We now extend our molecular studies to include the genus Pupilla, a close relative of Vertigo. Both genera are in the Orthurethra and have long been considered to be in the same family, the Pupillidae (Pilsbry, 1948; Hubricht, 1985). Like Vertigo, Pupilla is distributed across the entire Holarctic and members of both genera commonly co-occur within the same sites. Pupilla species possess a minute shell (,5 mm in height) of conserved cylindrical-ovate form, although shell apertures show considerable variability, ranging from simple to callused and/or lamellate. Consequently, apertural features have been much used as diagnostic species-specific characters (Pilsbry, 1921, 1948). However, preliminary DNA sequence analysis of putative North American P. muscorum suggested that this emphasis on shell apertural characters for species assignment was unsatisfactory (Nekola et al., 2009). We have thus undertaken the present study to examine whether the classically used apertural traits used to identify Pupilla species are able accurately to assess taxonomy supported by mitochondrial and nuclear DNA sequence data across the Holarctic range of the genus. Because genitalic structure has been found to be of only limited utility in making species-level taxonomic distinctions in this genus (Pokryszko et al., 2009), DNA sequence data offer the only practical non-shell-based method for cross-validation of taxonomic concepts in Pupilla. METHODS Specimen selection and identification Specimens used for analysis were primarily obtained from collections made in 2000–2012 (Nekola, 2005; Horsa´ k et al., 2010, # The Author 2014. Published by Oxford University Press on behalf of The Malacological Society of London, all rights reserved Journal of The Malacological Society of London Molluscan Studies Journal of Molluscan Studies (2015) 81: 196–216. doi:10.1093/mollus/eyu083 Advance Access publication date: 21 December 2014 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 11 2012; Horsa´ k, Chytry´ & Axmanova´ , 2013; Nekola & Coles, 2010). These include most of the currently recognized Pupilla taxa from western and central Europe (east to the Ural Mountains), central Asia (Altai Republic), Japan (Hokkaido) and North America (Canada and the USA including Alaska), as established by original descriptions, authoritative accounts of regional molluscan faunas and monographs (Pilsbry, 1921, 1948; Kerney & Cameron, 1979; Schileyko, 1984; Meng & Hoffman, 2008; Pokryszko et al., 2009; von Proschwitz et al., 2009). We were unable to secure tissue samples from only two species within these target regions: P. seminskii Meng & Hoffman, 2009 (Altai Republic) and P. sterkiana Pilsbry, 1889 (North America). For each analysed taxon, an attempt was made to select multiple individuals from across their known geographic and ecological range. Six specimens also represent topotype or near-topotype material: AP2 (P. altaica), AP13 (P. alluvionica), ET7 (P. muscorum xerobia), P1 (P. hebes kaibabensis), P10 (P. syngenes) and P12 (P. sonorana). Archival museum material up to 65 years old up to was used to supplement the specimen set for P. triplicata (specimens H19-21; Table 1). Each specimen was taxonomically assigned using currently recognized diagnostic conchological features (Table 2) as reported by Pilsbry (1921, 1948), Kerney & Cameron (1979), Schileyko (1984), Meng & Hoffman (2008), Pokryszko et al. (2009) and von Proschwitz et al. (2009). In these works, apertural lamellar architecture has been given particular weight, with little variation being reported in their number, shape or placement within a given taxonomic concept. Apertural crest size, callus development and colour are also frequently used as diagnostic features. Shell sculpture, suture depth and shell apex shape have been used less frequently to distinguish some entities. Based on these diagnoses, shells from all analysed individuals and their respective populations were examined for nine conchological traits (see below). DNA extraction, PCR amplification and sequence analysis Live specimens of Pupilla were preserved in absolute ethanol, allowed to desiccate at ambient temperature and humidity, or in some cases were used before death. DNA was extracted using the Omega BioTek Mollusk DNA Extraction Kit. Because of the inability of water to displace air within these tightly coiled shells, shell destruction was required to allow access of proteinase to mummified tissue. Thus (with few exceptions) specimens were taken from lots containing multiple examples of each respective taxon, with the actual specimens used for DNA preparation being imaged at 15 magnification prior to shell destruction using methods described by Nekola, Coles & Bergthorsson (2009). The internal transcribed spacers (plus flanking sequence) of the nuclear ribosomal RNA complex (ITS1 and ITS2), and mitochondrial cytochrome oxidase subunit I (COI) and cytochrome b (CytB) were amplified using published methods with modifications as listed in Table 3. PCR products were sequenced in both forward and reverse directions using Perkin Elmer ABI Big Dye termination and standard protocols. COI and CytB sequences were also obtained from the GenBank database for data analysed by von Proschwitz et al. (2009) that could be unambiguously assigned to a single individual (P. muscorum, BadenWu¨ rtemberg, Germany; P. pratensis, Lagmansro, O¨ stergo¨ tland, Sweden; P. pratensis, Mecklenburg-Vorpommern, Germany) and for two outgroups (Vertigo pusilla and Gastrocopta cristata) that were previously analysed by Nekola & Rosenberg (2013). Phylogenetic analyses Sequences (excluding primer regions) were aligned using ClustalX with adjustment by eye for ITS1 and ITS2. COI and CytB were concatenated, and ITS1 and ITS2 sequences were analysed as a single construct by omitting 81 invariant base pairs from the intervening 5.8S region. Mega v. 5.0 was used to conduct neighbour-joining (NJ), maximum parsimony (MP) and maximum likelihood (ML) analyses separately for the concatenated nuclear and mitochondrial DNA sequences. NJ analysis was based on maximum composite distance including transitions and transversions with pairwise gap deletion. MP analysis used the close-neighbour interchange search option with the random addition of 10 replicate trees. ML analysis used all sites and was based on the Tamura-Nei substitution model, a five-category gamma distribution for substitution rates, and the nearest neighbour interchange ML heuristic method. In all cases support values were estimated from 1000 bootstrap replicates. Additionally, Bayesian trees were generated using MrBayes v. 3.1 (Huelsenbeck & Ronquist, 2001), using a GTR substitution model assuming gamma-shaped rate variation over 1,000,000 generations with a sampling frequency of once each 1000 generations. Because none of these methods makes full informatic use of insertions and deletions, we also constructed a matrix of all variable bases in both the ITS1 and ITS2 regions, including not only base-pair substitutions, but also insertions and deletions. Post-hoc species delimitation and conchology Identification of potential species-level (and higher) clades based on DNA sequence data was accomplished by examining the nDNA and mtDNA trees for highly supported, reciprocallymonophyletic clades. This approach was of limited value for the nDNA data, because of low node support due to limited variation of 90 informative sites across 1500 bp. To help resolve relationships using these data, we examined the matrix of variable sites by eye for base-pair substitutions, insertions and deletions held in common among groups of sequences. Apparent incongruencies in specimen placement between the nuclear vs mitochondrial sequences were identified as potential cases of interspecific mitochondrial introgression or incomplete lineage sorting. We have not used any of the various methods for species demarcation using single-locus analyses of base-pair variation (e.g. generalized mixed Yule-coalescent functions). Although we have previously used these methods (Nekola et al., 2009), they universally require generation of ultrametric trees, which assume constant evolutionary rates across all clades. As a result, these methods do not function well when base-pair substitution rates are clade-specific. Because assumption of rate homogeneity appeared unjustified within the current Pupilla dataset, we have instead opted for reciprocal-monophyly as our decision-rule to identify potential species-level clades based on genetic data. We then attempted to verify the biological validity of these potential genetically-supported species concepts by reanalysing shell features from the imaged shells as well as additional shells within each analysed population. The range of expressed shell variation within each reciprocally-monophyletic species-level clade was documented for nine conchological traits: height (mm), width (mm), shell form, apex shape, shell sculpture, suture depth, aperture shape, apertural crest and callus strength, and apertural lamellae number and configuration. Potential species-level clades were considered taxonomically validated when some subset of the above shell features was found to be unique to and thus diagnostic of that entity. Based on this revised taxonomy, we then updated biogeographic and ecological information for each species based on our extensive community ecology datasets (e.g. Horsa´ k et al., 2010; Nekola, 2014) in combination with other published accounts. TAXONOMY OF PUPILLA 197 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 12 Table1.InformationforspecimensofPupillausedinDNAsequenceanalysis. Taxon/locationLatitude/longitudeSpecimencodeCollectionandlotno.Namesupportedby DNAsequencedata (whendifferent) GenBankaccessionno. CO1CytBITS1ITS2 PupillaalluvionicaMeng&Hoffman,2008 Russia Belyashi,Altai49.26918N,87.98388EAP13JNKM518545KM518468KM518390KM518312 Pupillaalpicola(Charpentier,1837) Russia Belyashi,Altai49.52068N,88.01808EAP12JNKM518546KM518469KM518391KM518313 Slovakia Raksˇa48.87928N,18.89018EH6MHKM518547KM518470KM518392KM518314 Vazˇec49.06858N,19.99948EH11MHKM518548KM518471KM518393KM518315 Raksˇa48.87928N,18.89018EH12MHKM518549KM518472KM518394KM518316 Liptovska´Teplicˇka48.96328N,20.10448EH13MHKM518550KM518473KM518395KM518317 PupillaaltaicaMeng&Hoffman,2008 Russia Ustˇ-Muny,Altai51.72978N,85.73828EAP1JNP.turcmenicaKM518551KM518474KM518396KM518318 Kurai,Altai50.30808N,87.64858EAP2JNP.turcmenicaKM518552KM518475KM518397KM518319 Kurai,Altai50.23348N,87.78948EAP17MHP.turcmenicaKM518553KM518476KM518398KM518320 Pupillabigranata(Rossma¨ssler,1839) France Pont,Calvados48.97748N,0.09038WAP25BC2014.013.00081P.muscorumKM518554KM518477KM518399KM518321 PupillablandiMorse,1865 Canada Irvine,Alberta49.95958N,110.25118WAP34BC2014.013.00079KM518555KM518478KM518400KM518322 MooseJaw,Saskatchewan50.04348N,105.62468WAP35BC2014.013.00074KM518556KM518479KM518401KM518323 USA LoganCanyon,Utah41.74268N,111.76038WP7JN18327P.hebesKM518557KM518480KM518402KM518324 LasVegas,NewMexico35.64158N,105.18758WP15JN18417KM518558KM518403KM518325 RubyMountains,Nevada40.77518N,115.33648WP16JN18280P.hebesKM518559KM518481KM518404KM518326 WilsonArch,Utah38.27458N,109.37248WP18JN18237P.hebespithodesKM518560KM518482KM518405KM518327 PupillablandicharlestonensisPilsbry,1948 USA EastTinticRange,Utah39.96608N,112.04848WP2JN18257P.hebesKM518561KM518483KM518406KM518328 BullionCanyon,Utah38.41718N,112.31268WAP28JN17211P.hebespithodesKM518562KM518484KM518407KM518329 PupillablandipithodesPilsbry&Ferriss,1917 USA BullionCanyon,Utah38.41718N,112.31268WAP27JN17210KM518563KM518485KM518408KM518330 TusasRidge,NewMexico36.65198N,106.03818WAP38JN13013KM518564KM518486KM518409KM518331 Bland,NewMexico35.75048N,106.45698WP9JN14816KM518565KM518487KM518410KM518332 Pupillahebes(Ancey,1881) Japan J. C. NEKOLA ET AL. 198 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 13 Toyokoro,Nakagawa,Hokkaido42.60508N,143.55648EVH29JNP.hokkaidoensisKM518566KM518488KM518411KM518333 USA KnikIsland,Anchorage,Alaska61.50848N,149.03438WAP29JN15390P.alaskensisKM518567KM518489KM518412KM518334 UteCreekCanyon,Colorado38.58488N,105.96868WAP37JN12898P.blandiKM518568KM518490KM518413KM518335 HappyValley,Alaska69.33558N,148.73028WNS48JN15142P.alaskensisGQ921663KM518491KM518414KM518336 JarbidgeMountains,Nevada41.68678N,115.50618WP5JN18292P.hebespithodesKM518569KM518492KM518415KM518337 LoopeEast,California38.65918N,119.72228WP14JN17254KM518570KM518493KM518416KM518338 PupillahebeskaibabensisPilsbry&Ferriss,1911 U.S.A. KaibabPlateau,Arizona36.82998N,112.25428WP1JN18400P.hebesKM518571KM518494KM518417KM518339 PupillahebesnefasPilsbry&Ferriss,1910 USA SantaCatalinaMts.,Arizona32.41968N,110.73118WP6JN14052P.hebespithodesKM518572KM518495KM518418KM518340 Pupillacf.khunjerabicaAuffenberg&Pokryszko,2009 Russia Chagan-Uzun,Altai50.08698N,88.39418EAP11MHKM518573KM518496KM518419KM518341 Pupillacf.limataSchileyko,1989 Russia Bestyakh,Yakutia61.36248N,128.84338EAP20MHKM518574KM518497KM518420KM518342 Kapitonovka,Yakutia62.32928N,129.92828EAP39MHKM518575KM518498KM518421KM518343 PupillaloessicaLozˇek,1954 Russia Belyashi,Altai49.41868N,87.59288EAP5JNKM518576KM518499KM518422KM518344 Kosh-Agach,Altai49.99298N,88.54968EAP6JNKM518577KM518500KM518423KM518345 Belyashi,Altai49.28048N,87.49558EAP7JNKM518578KM518501KM518424KM518346 Kosh-Agach,Altai49.66098N,88.22788EAP8JNKM518579KM518502KM518425KM518347 Belyashi,Altai49.26918N,87.98388EAP9JNKM518580KM518503KM518426KM518348 Belyashi,Altai49.52068N,88.01808EAP10JNKM518581KM518504KM518427KM518349 Ulagan,Altai50.47678N,87.63018EAP19MHKM518582KM518505KM518428KM518350 Pupillamuscorum(Linnaeus,1758) Europe CzechRepublic Brno,Moravia49.25098N,16.57388EmtG-PupMHKM518583KM518506KM518429KM518351 USA(naturalized) CedarRapids,Iowa41.98668N,91.74008W22JN14592GQ921664 Syracuse,NewYork43.00748N,76.11058WAP26JN13955KM518584KM518507KM518430KM518352 Canada Cochrane,Alberta51.26428N,114.73268WAP36BC2014.013.00058P.hudsonianumKM518585KM518508KM518431KM518353 Churchill,Manitoba58.70868N,94.12308WP8JN11098P.hudsonianumKM518586KM518509KM518432KM518354 LaGrandePointe,Quebec50.20598N,63.39688WP13BC2014.013.00001P.hudsonianumKM518587KM518510KM518433KM518355 USA BullionCanyon,Utah38.41718N,112.33038WP17JN17219P.hebesKM518588KM518511KM518434KM518356 Continued TAXONOMY OF PUPILLA 199 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 14 Table1.Continued Taxon/locationLatitude/longitudeSpecimencodeCollectionandlotno.Namesupportedby DNAsequencedata (whendifferent) GenBankaccessionno. CO1CytBITS1ITS2 LakeBemidji,Minnesota47.53288N,94.82478W23JN9054P.hudsonianumGQ921662KM518357 LakeBemidji,Minnesota47.53288N,94.82478WAP33JN9054P.hudsonianumKM518589KM518512KM518435KM518358 PupillamuscorumxerobiaPilsbry,1914 USA BannonRanch,NewMexico36.91668N,103.78008WET7JN16491P.blandiKM518590KM518513KM518436KM518359 LakeBemidji,Minnesota47.53288N,94.82478WP4JN9054P.blandiKM518591KM518514KM518437KM518360 Pupillapratensis(Clessin,1871) Russia Aktash,Altai50.44728N,87.60788EH3MHP.alpicolaKM518592KM518515KM518438KM518361 Ulagan,Atlai50.47678N,87.62188EH4MHP.alpicolaKM518593KM518516KM518439KM518362 Europe CzechRepublic Pozdeˇchov,Moravia49.23398N,17.98648EH1MHP.alpicolaKM518594KM518517KM518440KM518363 Vysoke´My´to,Bohemia49.96118N,16.18928EH7MHP.alpicolaKM518518KM518441KM518364 Slovakia Za´vod48.53318N,16.99638EH5MHP.alpicolaKM518595KM518519KM518442KM518365 Pupillasonorana(Sterki,1899) USA SacramentoMountains,NewMexico32.71418N,105.75418WP12BC2005.011.03117KM518596KM518520KM518443KM518366 Pupillasterrii(Forster,1840) Albania PeriferiDibre41.81728N,20.50038EAP16MHKM518597KM518521KM518444KM518367 CzechRepublic Pavlov,Moravia48.87738N,16.66358EH8MHKM518598KM518522KM518445KM518368 Praha-Hlubocˇepy,Bohemia50.04198N,14.37618EH14MHKM518599KM518523KM518446KM518369 Klentnice,Moravia48.84678N,16.64058EAP22MHKM518600KM518524KM518447KM518370 Russia VerkhneBikberda,Bashkortostan52.32078N,56.80328EAP15MHKM518601KM518525KM518448KM518371 Slovakia Valaska´Dubova´49.15118N,19.64418EAP21MHKM518602KM518526KM518449KM518372 Pupillasyngenes(Pilsbry,1890) USA Mogollon,NewMexico33.39448N,108.80568WAP30BC2005.011.02961KM518603KM518527KM518450KM518373 KaibabPlateau,Arizona36.69188N,112.29898WP10BCKM518604KM518528KM518451KM518374 PupillasyngenesdextroversaPilsbry&Vanatta,1900 USA KaibabPlateau,Arizona36.69188N,112.29898WP11BCP.syngenesKM518605KM518529KM518452KM518375 Pupillatriplicata(Studer,1820) Russia J. C. NEKOLA ET AL. 200 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 15 OzeroKureevo,Altai52.48118N,85.76058EAP32MHKM518606KM518530KM518453KM518376 CzechRepublic Tocˇnı´k,Bohemia49.89058N,13.88668EH2MHKM518607KM518531KM518454KM518377 Pavlov,Moravia48.87738N,16.66358EH9MHKM518608KM518532KM518455KM518378 Milesˇov,Bohemia50.53448N,14.94378EH10MHKM518609KM518533KM518456KM518379 Kamy´k,Bohemia50.56438N,14.08878EH15MHKM518610KM518534KM518457KM518380 Chroustov,Moravia49.17188N,16.05028EH16MHKM518611KM518535KM518458KM518381 Hracholusky,Bohemia49.99748N,13.79048EH17MHKM518612KM518536KM518459KM518382 Louny,Bohemia(1948)H19VLKM518613KM518537KM518460KM518383 Kamy´k,Bohemia(1970)H20VLKM518614KM518538KM518461KM518384 Pupillatriplicata(Studer,1820) CzechRepublic Srdov,Bohemia(1950)H21VLKM518615KM518539KM518462 France Cahors,Dordogne44.47728N,1.43038EAP31BC2014.013.00082KM518616KM518540KM518463KM518385 Russia Nugush,Bashkortostan53.00668N,56.53568EAP23MHKM518617KM518541KM518464KM518386 Pupillaturcmenica(O.Boettger,1889) Russia Kosh-Agach,Altai50.07298N,88.72018EAP3MHKM518618KM518542KM518465KM518387 Belyashi,Altai49.29558N,87.73448EAP4MHKM518619KM518543KM518466KM518388 Kurai,Atlai50.23348N,87.78948EAP18MHKM518620KM518544KM518467KM518389 AllspecimensfromtheNekolacollection(JN)arecurrentlymaintainedattheUniversityofNewMexicobeforeultimatedepositionattheAcademyofNaturalSciencesatDrexelUniversity(ANSP).Allspecimensfromthe Colescollection(BC)arehousedattheNationalMuseumofWales;allaccessionnumbersforthismaterialareprecededby‘NMW.Z’.AllspecimensfromtheHorsa´kcollection(MH)arehousedattheDepartmentofBotany andZoologyatMasarykUniversity,Brno.MaterialfromtheVojenLozˇekcollection(VL)ishousedatCharlesUniversity,Prague. TAXONOMY OF PUPILLA 201 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 16 Table2.HistoricaltaxonomicandtraditionalconchologicalconceptsforanalysedPupilla. TaxonHeight (mm) Width (mm) ShellformApexshapeShellsculptureSuturedepthAperturalcrestAperturalcallusAperturallamellae alluvionica3.3–4.32.2–2.4WidecylindricalTapered/ domed Smooth/weakstriaeClearly evident Hardlythickened, clearlywhite 0–1(rareweakparietal) alpicola2.8–3.31.8–1.9WidecylindricalDomedObviousDeepWeak-absent0–1(rareweakparietal) altaica2.5–3.21.6–1.8CylindricalovoidDomedFineribswithdermal edges PronouncedThick,white2–3(angularpad;palatalrare) bigranata3.0–3.61.6CylindricalovoidTaperedAlmostsmooth;fine linesonly ShallowPresentThick,white2–3(rarecolumellar) blandi2.5–3.31.5–1.6CylindricalDomedDelicatestriationsDeepPresentYellow/tan3 blandi charlestonensis 3.01.4CylindricalDomedSlightlystriateDeepWell developed Brown3(palatal,parietalstrong,long) blandipithodes3.1–3.71.7–1.8WidecylindricalDomedDelicatestriaeDeepLow-strongWeak;brown3 hebes3.1–4.01.6–1.9SubcylindricDomedMinutestriaeDeepNone-weakNone0–1(rareparietal) hebeskaibabensis2.7–2.81.5SubcylindricDomedMinutestriaeDeepNone-weakNone0–1(rareparietal) hebesnefas3.2–4.21.7–1.9Subcylindric,sinistralDomedMinutestriaeDeepNone-weakNone1(parietal) cf.khunjerabica3.2–4.01.7–1.9CylindricalovoidTapered/ domed Almostsmooth,striae faint,irregular Moderately deep WeakThin-lacking0 cf.limata2.8–3.11.6–1.7CylindricalovoidTapered/ domed DelicatestriaeModerately deep Absent-weakNone0 loessica3.0–3.41.7– 1.75 CylindricalovoidTapered/ domed Finely,irregularlyribbedDeepAbsent-weakNone0 muscorum3.2–4.01.7CylindricalovoidTaperedAlmostsmooth;fine linesonly ShallowStrongThick,white0–2 muscorumxerobia2.3–2.81.4–1.5CylindricalovoidDomedAlmostsmooth;fine linesonly ShallowStrongThick,white1(parietal) pratensis3.5–4.51.9–2.1CylindricalovoidDomedFinestriationDeepWeakWeak,white0–2(weakifpresent);no depressionoverpalatal sonorana2.5–3.31.3–1.4CylindricalDomedFinestriaeDeepStrongThick,white3(alllong) sterrii2.8–3.51.6CylindricalTapered/ domed CoarsestriaeVerydeepWeak- moderate Moderatelythick, white 2(bothpeg-shaped) syngenes3.0–4.21.7–1.8Biconic,widestatc.4thwhorl;8 narrowwhorls;sinistral Tapered/ domed DelicatestriaeShallowStrong3 syngenes dextroversa 3.0–4.51.6–1.8Biconic,widestatc.4thwhorl;8 narrowwhorls;dextral Tapered/ domed DelicatestriaeShallowStrong3 triplicata2.2–2.8 (4) 1.4CylindricalDomedFine,closestriationDeepModerateDistinct,white3(bladeshaped) turcmenica3.0–3.21.1–1.4Cylindrical;shellthinnerTapered/ domed CoarsestriaeVerydeepWeakWeak,white0–2(weakparietaland/or palatal) J. C. NEKOLA ET AL. 202 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 17 RESULTS DNA sequence data A total of 80 specimens from 22 putative Pupilla taxa underwent DNA extraction (1–12 individuals/taxon, see Table 1). DNA sequences were obtained for 79 specimens for COI, 77 specimens for CytB and 77 specimens for ITS1 and ITS2. All COI and CytB amplicons consisted of 655 and 377 bp, respectively, and could be unambiguously aligned. The COI amplicon contained 240 and CytB 137 variable sites. The ITS1 amplicon length was 615–635 bp and the ITS2 amplicon length was 868–874 bp. All Pupilla ITS sequences could be unambiguously aligned, but those of the outgroup (Vertigo pusilla and Gastrocopta cristata) could not. The 5.8S region between the ITS amplicons (based on 23 Pupilla specimens for which it was determined) consisted of 81 invariant bases so that the entire contiguous sequence between the 50 -end of the ITS1 amplicon and the 30 -end of the ITS2 amplicon is 1569–1593 bp. The total informative sites in ITS consisted of 56 bp substitutions and 36 bp comprising 12 insertions/deletions. Phylogenetic reconstructions and supported taxonomic entities Phylogenetic tree reconstruction and base-pair variation maps based on concatenated COI þ CytB mtDNA and ITS1 þ ITS2 nDNA sequences support the presence of 17 putative reciprocallymonophyletic species or subspecies-level taxa (Figs 1 and 2, Table 4): P. alluvionica, P. alpicola, P. blandi, P. cf. khunjerabica, P. cf. limata, P. hebes, P. hebes pithodes, P. loessica, P. muscorum, Pupilla n. sp. (Alaska), Pupilla n. sp. (Hokkaido), Pupilla n. sp. (Hudsonian), P. sonorana, P. sterrii, P. syngenes, P. triplicata and P. turcmenica. The three new species identified by these analyses are formally described below and will be referred to hereafter as P. alaskensis, P. hokkaidoensis and P. hudsonianum, respectively. Pupilla alluvionica is a xeric rock outcrop and steppe species that differs from all other analysed Pupilla by possessing adult shells .2.1 mm in diameter with a smooth or very weakly striate shell, a crest and a white callus. The single individual analysed for DNA sequence possessed an ITS1 þ ITS2 sequence with four bases different from all other Pupilla (111C and 577A in ITS1; 333G and 490C in ITS2). However, its COI þ CytB sequence was part of the same highly supported clade defining P. turcmenica, which consistently co-occurs with P. alluvionica in the Altai. Pupilla alpicola is a wetland species whose shells are up to 2.1 mm wide with a shallow suture and a body whorl often slightly narrower than the penultimate whorl. This species is defined by 92C and usually 340C in ITS2. Two subpopulations are noted, one with AC at 171–172 in ITS1 and the other maintaining the consensus GT at these positions. While the former subpopulation is more prevalent in Europe and the latter in central Asia, individuals characteristic of either occur throughout its known range. Pupilla pratensis has been traditionally differentiated from P. alpicola by lacking a depression or flattening on the palatal wall of the aperture and having a rather pronounced shell apex (von Poschwitz et al., 2009). However, P. pratensis shares the same unique ITS2 bases as P. alpicola, with individuals referable to P. pratensis occurring in both of the unique ITS1 subpopulations. Additionally, P. pratensis mtDNA occurs throughout the same highly supported clade that contains all analysed P. alpicola. If P. pratensis is considered a shell form of P. alpicola, then P. alpicola is monophyletic for both nDNA and mtDNA. According to Pilsbry (1948), P. blandi is characterized by a cylindrical shell with a prominent crest, a yellow to tan callus and three apertural lamellae. Using these criteria, individuals conchologically assignable to P. blandi demonstrate polyphyly both in ITS1 þ ITS2 and COI þ CytB, occurring within three well-supported species-level clades. However, if these traditional conchological characteristics are abandoned in favour of surface sculpture, with P. blandi being differentiated by its irregular, very weak striae, shiny shell surface and shallow suture, this species becomes a strongly supported monophyletic entity for COI þ CytB that uniquely possesses a GAC insertion at 181– 183 in ITS2. Because this entity varies in apertural lamella number from 0 to 3, and has a crest and callus ranging from weak to strong, it includes a number of shell forms that were previously assigned to other taxa including P. hebes and P. muscorum xerobia. Pilsbry (1948) characterized P. hebes as possessing a minutely striate, subcylindric shell with no apertural callus, an absent to rarely weak apertural crest and absent to rarely weak parietal lamella. Shells displaying these traits demonstrate polyphyly among three different species-level clades. However, as with P. blandi, monophyletic grouping is apparent when a different suite of shell features is used to diagnose P. hebes, including a cylindrical-ovoid shell tapered for the upper 1/3-1/4 of the shell height, a normal to deep suture and possession of numerous sharp thread-like striae. Shells possessing these features all uniquely possess 508T in ITS1, while sharing 495A in ITS1 with P. blandi and P. hebes pithodes. Using the shell characters of Pilsbry (1948) for identification causes some individuals within this group to be incorrectly assigned to P. blandi, P. blandi charlestonensis and P. muscorum. It should be noted that COI þ CytB suggests that P. hebes exists as two discrete subpopulations, one ranging throughout the Great Basin from California to north-central Utah (samples P5, P7, P14 and P16) and the other being restricted to the canyonlands region of the Colorado Plateau (samples P1, P2 and P17). This latter subpopulation would equate to P. hebes kaibabensis of Pilsbry (1948). However, as the shells of this clade completely overlap with typical material as well as possessing identical ITS1 þ ITS2 sequence, it seemed best not to formally recognize this subpopulation at this time. Pilsbry (1948) characterized P. blandi pithodes as being wider than P. blandi, with a weak to absent crest and callus. He hypothesized that it was intermediate between P. blandi and P. hebes. ITS1 þ ITS2 indicate that in fact this entity is more Table 3. Primers used for genetic analysis. Region Direction Anneal Sequence Source COI f 458C 5′ -ATTCAACGAATCATAAAGATATTGG-3′ Author Design r 5′ -TATACTTCAGGATGACCAAAAAACCA-3′ Author Design CytB f 478C 5′ -TGAGGTGCAACAGTNATTAC-3′ Author Design r 5′ -GCAAATAAAAAGTATCACTCTGG-3′ Author Design ITS1 f 528C 5′ -TAACAAGGTTTCCGTATGTGAA-3′ Armbruster & Bernhard (2000) r 5′ -TCACATTAATTCTCGCAGCTAG-3′ Author Design ITS2 f 528C 5′ -CTAGCTGCGAGAATTAATGTGA-3′ Wade & Mordan (2000) r 5′ -GGTTTCACGTACTCTTGAAC-3′ Author Design TAXONOMY OF PUPILLA 203 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 18 Figure 1. Maximum-likelihood phylogenetic tree reconstruction for Pupilla based on concatenated ITS1 þ ITS2 data. Nodes with strong to moderate support across all four phylogenetic reconstruction methods have been labelled to the left of that node by four support values: upper left (normal font) is for NJ; upper right (bold italic font) is for MP; lower left (bold font) is for Bayesian; lower right (italic font) is for ML. Branch tip labels represent initial identifications based on traditional conchological features, whereas labels to the right of brackets represent valid names supported by both nDNA and mtDNA sequence analysis. J. C. NEKOLA ET AL. 204 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 19 Figure 2. Maximum-likelihood phylogenetic tree reconstruction for Pupilla based on concatenated COI þ CytB. Labelling conventions as in Figure 1. Specimens that have a significantly different topological location as compared with the nDNA tree are highlighted in gray. TAXONOMY OF PUPILLA 205 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 20 Table4.MatrixofvariablebasesinITS1andITS2. J. C. NEKOLA ET AL. 206 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 21 Forbothamplicons,numbersrefertothebasepairsdownstreamoftheterminationoftheforwardprimer.Basepairsinvariantacrossallsamplesareomitted,whilebasepairsthatdivergefromthegenusconsensusare highlighted.Dashindicatesbase-pairdeletion.Samplesaresortedbypost-hocsupportedspeciesnames(inbolditalic),whileinitialnamesusedpriortoDNAanalysis,andbasedontraditionalconchologicalfeatures,are listedintheSampleinformation.Potentialgroupingsofrelatedspeciesarenotedinthefarthest-rightcolumn,labelledSG.Notethattheinsertions:ITS168–71and72–75;andITS2181–183and504–506,are microsatelliterepeats. TAXONOMY OF PUPILLA 207 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 22 closely related to P. hebes, lacking the GAC insertion at 181–183 in ITS2. However, it is distinguished from this species by possessing 508G in ITS1. The COI þ CytB of P. hebes pithodes exists as a highly supported clade. Because we observed a gradation from barrel-shaped P. pithodes to more cylindrical P. hebes shells, we have chosen to consider P. pithodes as a subpopulation that occurs to the east of the main range of P. hebes, ranging from eastern Arizona and central Utah into Colorado, New Mexico and Texas. Use of traditional shell characteristics (Pilsbry, 1948) results in some P. hebes pithodes being incorrectly assigned to P. blandi charlestonensis. ITS1 þ ITS1 and COI þ CytB both indicate that the sinistral P. hebes nefas of southeastern Arizona should be regarded as a shell form of P. hebes pithodes. DNA analysis also demonstrates that P. alaskensis and P. hokkaidoensis, which have previously been confused with P. hebes, actually represent undescribed species related to the central Asian P. loessica, as all share 207A in ITS2. Both of these new species uniquely share 177G in ITS2, with P. hokkaidoensis differing from P. alaskensis based on an 11-bp deletion at 38–48 in ITS1, 111C in ITS1 and 79C in ITS2. Pupilla alaskensis shells differ from P. hebes in their widely spaced, coarse, somewhat anastomosing striae. Pupilla hokkaidoensis shells differ from P. hebes in their rotund cylindrical shape with shallow suture and anastomosing coarse striae (see taxonomic descriptions below). Pupilla cf. khunjerabica is represented in the sample by a single population from a riparian forest corridor in the Altai. This specimen possesses a cylindrical-ovoid shell tapered for the upper half of the shell height, very weak irregular thread-like striae, a moderately deep suture, a weak to absent callus or crest, and no apertural lamellae. Although its COI þ CytB sequence fell within the highly supported clade defining P. alaskensis, it uniquely possessed 137C in ITS2, and was the only taxon outside the southwestern North America P. syngenes/sonorana group to possess 156G in ITS2. Pupilla cf. limata was sampled from riparian forest in the Yakutia region of eastern Siberia and is characterized by a cylindrical-ovoid shell with numerous, sharp, somewhat anastomosing striae, a weak crest and no callus or apertural lamellae. While this taxon is highly supported as a monophyletic entity in COI þ CytB, one of the two analysed individuals possessed ITS1 þ ITS2 identical with P. hebes pithodes from western North America. The other individual uniquely possessed 444C in ITS1. Pupilla loessica is a steppe-tundra species that in modern times is restricted to central Asia, although Pleistocene fossils extend west into central Europe (e.g. Horsa´ k et al., 2010). It is distinguished by its numerous strong rounded anastomosing striae and lack of callus and apertural lamellae. It uniquely possessed 266C in ITS1 and shares 207A in ITS2 with P. alaskensis and P. hokkaidoensis. Pupilla loessica exists as a strongly supported clade in COI þ CytB. While one individual (AP6) possessed COI þ CytB characteristic of P. alaskensis, this individual had ITS1 þ ITS2 characteristic of P. loessica. Pupilla muscorum has traditionally been characterized by its cylindrical shell with smooth sculpture, shallow suture, strong crest, thick white callus and from zero to two parietal lamellae (Pilsbry, 1948; von Proschwitz et al., 2009). This species exists as a well defined clade in ITS1 þ ITS2 by uniquely possessing 375C and 609A in ITS1 and 336C in ITS2. It also exists as a strongly supported clade in COI þ CytB. Pupilla bigranata, which is distinguished from P. muscorum by its three strong apertural lamellae, was found to have COI þ CytB sequence identical with Moravian P. muscorum and ITS1 þ ITS2 sequence identical with New York P. muscorum. Individuals from the High Plains and Rocky Mountains of the western USA, which have been previously identified as P. muscorum using the above conchological characteristics, were shown by DNA sequence analysis actually to represent P. blandi or P. hebes. In North America P. muscorum was thought to exist in two disjunct populations, one ranging from the northeastern Atlantic seaboard west through the Great Lakes to Iowa, and the other ranging across the northern taiga (Oughton, 1948; Hubricht, 1985). This latter entity (P. hudsonianum) exists as a well defined monophyletic group uniquely possessing 56A, 541G and 583C in ITS2. It also exists as a strongly supported clade in COI þ CytB, although one individual (P13) possessed sequence characteristic of P. alaskensis. It is easily differentiated from P. muscorum by the strong taper over the top half of its shell height and its sculpture of dense thread-like striae (see taxonomic description below). Pupilla syngenes possesses a distinctive shell that is widest in the top half and has three apertural lamellae including a long, curved blade-like parietal lamella. This species of wooded, xeric rock outcrops in the southwestern USA uniquely possessed 83A in ITS2 and also exists as a strongly supported clade in COI þ CytB. The dextral individual of this typically sinistral species, termed P. syngenes dextroversa, had ITS1 þ ITS2 and COI þ CytB sequences identical to a typical individual within the same population. Pupilla sonorana was compared by Pilsbry (1948) with P. triplicata and distinguished by its small size, columnar shell, strong crest and three apertural lamellae including a curved blade-like parietal lamella. It appears closely related to P. syngenes by sharing 453T in ITS1 and 156G and 233A in ITS2. However, it differed from that species in possessing 218A and 538C and a GGCA insertion from 68 to 71 in ITS1 and 83G and 356A in ITS2. As only one individual was analysed, no species-level clade can be assigned in COI þ CytB. However, it differed from P. syngenes by 54 bp at these loci. Pupilla triplicata of rock outcrops from western Europe to central Asia has shells that differ from P. sonorana only by their weak, rounded (rather than sharp) striae. This species uniquely possesses 295A in ITS2. Most individuals also exhibit 617C in ITS1. COI þ CytB of this species form a highly supported clade. While two distinct subpopulations are suggested by 740A vs 740G in ITS2, there is no correspondence in COI þ CytB or in any noted conchological features. As such, this grouping appears to have no taxonomic merit. Pupilla sterrii of dry calcareous grasslands from central Europe to the Urals is characterized by its very deep suture and sharp, coarse, anastomosing striae. It is defined by uniquely possessing 509T in ITS1. The mtDNA of this species is highly variable, with individuals variously possessing COI þ CytB characteristic of P. alpicola, P. muscorum, P. triplicata or P. turcmenica. No analysed individual from Europe was found to possess mtDNA sequence with the expected topological position as sister to P. turcmenica. However the Urals specimen did and it may represent the only individual with both nDNA and mtDNA characteristic of P. sterrii. Pupilla turcmenica is a species of xeric calcareous grasslands that ranges across Asia Minor and central Asia. It is conchologically distinguished from the similar P. sterrii by its less deep sinus and more widely spaced striae. It uniquely displays 207G in ITS1, with its COI þ CytB forming a highly supported clade. Pupilla altaica has been recently differentiated from this species by its larger crest, more massive white callus and the presence of an angular pad on the parietal wall of the aperture (Meng & Hoffman, 2008). However, this entity did not possess any unique ITS1 þ ITS2 distinctions from P. turcmenica, being distributed throughout the same highly supported COI þ CytB clade encompassing that species. As such it appears to simply represent the high-calcificiation endpoint within the normal conchological range of P. turcmenica. These supported entities could be further associated into nine groups using ITS1 þ ITS2 data (used to order Table 4). Group J. C. NEKOLA ET AL. 208 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 23 1 consists of P. blandi, P. hebes, P. hebes pithodes and P. cf. limata and is characterized by insertion 495A in ITS1. Group 2 is represented only by P. hudsonianum and is characterized by the 56A insertion, 541G and 583C in ITS2. Group 3 consists of P. syngenes, P. sonorana and P. cf. khunjerabica and is characterized by 156G in ITS2. Group 4 consists of P. loessica, P. alaskensis and P. hokkaidoenis and is characterized by 207A in ITS2. Group 5 includes only P. alluvionica and is characterized by 111C and 577A in ITS1 and 333G and 490C in ITS2. Group 6 includes only P. alpicola and is characterized by 92C and typically 340C in ITS2. Group 7 consists only of P. muscorum and is characterized by 375C and 609A in ITS1 and 336C in ITS2. Group 8 is represented only by P. triplicata and is characterized by 295A in ITS2. Group 9 consists of P. sterrii and P. turcmenica and is characterized by 717T in ITS2. Because no variable bases are shared between groups, however, possible relationships between them cannot be inferred. The greater amount of variation within COI þ CytB allows resolution of deeper relationships. The nine interspecific groups suggested by ITS1 þ ITS2 are generally validated with high support by the mtDNA tree topology. The major exception is Group 1, whose members are spread across two major mtDNA clades: P. blandi, P. hebes hebes and P. hebes pithodes belong to one strongly supported clade, while P. cf. limata appears more related to P. loessica and P. alpicola in the mtDNA tree. Topological incongruence between mitochondrial and nuclear DNA phylogenies Comparison of the mtDNA tree with nDNA tree and base-pair variation matrix reveals topological incongruence in eleven specimens, or almost 15% of the total. These are largely limited to two groups: P. alaskensis and the P. sterrii/turcmenica clade. The highly supported mtDNA species-level clade containing both P. alaskensis specimens also harbours individuals with nDNA characteristic of P. loessica, P. cf. khunjerabica or P. hudsonianum. Individuals harbouring mtDNA characteristic of P. turcmenica may possess nDNA characteristic of either P. alluvionica or P. sterrii. In P. sterrii, specimens possessing nDNA characteristic of that species may harbour mtDNA characteristic of P. alpicola, P. triplicata, P. muscorum or P. turcmenica. While the current analysis is not capable of resolving the cause of these incongruencies, it does seem likely that mitochondrial introgression is responsible in the case of P. sterrii as European individuals variously possess mtDNA from all other known European species. Conchological variation: traditional vs other traits Comparison of conchological features among geneticallyidentified individuals demonstrates that the size of the apertural crest, degree of callus deposition, callus colour and number, shape and placement of apertural lamellae are of little taxonomic value (Table 5). For instance, in Europe P. alpicola, P. muscorum and P. triplicata all show variation ranging from zero to multiple apertural lamellae, and from absent/weak to strong crest and callus (Fig. 3). This same pattern is repeated in central Asia with P. turcmenica (Fig. 4) and in North America with P. blandi and P. hebes (Fig. 5). The reliance of traditional taxonomy on these traits is thus responsible for oversplitting in some regions (e.g. P. bigranata and P. pratensis in Europe and P. altaica in central Asia; Figs 3, 4) and for the abysmal initial sorting of western North American material (Fig. 5). However, other conchological traits do accurately reflect genetic relationships and are capable of accurately sorting individuals into species-level groups (Table 5; Figs 3–5). The most important of these are shell sculpture, including not only shape, strength, density and complexity of shell striae, but also the lustre of the underlying shell surface. Suture depth and apex architecture were also found to be valuable for species identification, as was shell width, which appeared to be relatively independent of shell height. SYSTEMATIC DESCRIPTIONS Pupillidae Pupilla Leach, in Fleming, 1828 Pupilla alaskensis Nekola & Coles, n. sp. (Fig. 6A–H, K) Types: Holotype (Fig. 6A–D, K): ANSP 458632, Happy Valley, North Slope Borough, Alaska, USA (698200 700 N, 1488430 4800 W). Paratypes: 10 shells, ANSP 458633, collected with holotype; 100 shells, NMW.Z.2014.013.00013, collected with holotype; 5 shells ANSP 458634, Sukakpak Mountain, Yukon-Koyukuk Census Area, Alaska, USA (678350 5500 N, 1498470 400 W); 60 shells NMW.Z.2014.013.00011, same loc. as preceding; 5 shells ANSP 458635, Livengood East, Yukon-Koyukuk Census Area, Alaska, USA (658270 5500 N, 1488200 4000 W); 3 shells ANSP 458636, Knik I., Matanuska-Susitna Borough, Alaska, USA (618300 3000 N, 149820 300 W). Zoobank registration: urn:lsid:zoobank.org:act:D98DD52C-BFAE- 4752-A9D4-18F116CE8957 Other material examined: NMW.Z.2005.011.01468, NMW.Z.2014. 013.00002 – 00023 c. 1000 shells from Alaska, USA; 30 lots (3739 individuals) from Alaska, USA in Nekola collection. Etymology: Specific name alaskensis refers to region in which species is known to occur. Diagnosis: Shell small, cylindrical-ovoid, similar to P. hebes but differing by its deeper suture and shell sculpture of widelyspaced, anastomosing radial striae. GenBank: GQ921663, KM518334, KM518336, KM518412, KM518414, KM518489, KM518491, KM518567. Description: Shell 2.6–3.3 mm tall  1.6–1.8 mm wide, opaque to translucent, yellowish-brown to cinnamon-brown; 6–6.5 whorls; apical whorls rounded-conical, remainder ovate-cylindrical to cylindrical; suture typically deep though sometimes of only normal depth with whorls consequently appearing swollen; shell surface silky in general appearance, the post-neanic whorls bearing sharp, irregular, often widely spaced, anastomosing radial striae occasionally developed into fine lamellae superimposed on a minutely and irregularly papillate surface (Fig. 6K); aperture 1/4 of shell height, ranging from slightly wider than tall (Fig. 6A, E, G) through circular (Fig. 6F) to slightly taller than wide (Fig. 6H), in profile ascending onto body whorl (Fig. 6B); umbilicus closed by preceding whorls (Fig. 6C); peristome interrupted by body whorl, apertural lip flared (Fig. 6B, D), shell slightly contracted behind (Fig. 6D); crest absent or weakly developed but not thickened or callused (Fig. 6A–H); apertural lamellae generally absent, though a vestigial, plate-shaped columellar is occasionally present (Fig. 6E, G). Geographical distribution: Currently known from just south of Arctic Ocean coastal plain in far northern Alaska to Pacific Coast near Anchorage, Alaska. It seems likely that this species will be found in adjacent areas of the Yukon and northwestern British Columbia, Canada. The published records for P. hebes TAXONOMY OF PUPILLA 209 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 24 Table5.PupillashellfeaturesforthosetaxonomicunitsvalidatedbyDNAsequencedata. TaxonHeight (mm) Width (mm) ShellformApexshapeShellsculptureSuturedepthApertureshapeAperturalcrestand callus Aperturallamellae alaskensis2.6–3.31.6–1.8Cylindricalovoidto cylindrical TaperedWidelyspaced,sharp, somewhatanastomosing coarsestriae;dullshell Normalto deep Slightlywiderthantallto slightlytallerthan wide Absenttoweakcrest; nocallus 0–1(vestigal,plate-like columellaronly) alluvionica3.3–4.32.0–2.4CylindricalovoidTapered/domedLow,rounded,irregular,wide spacedstriae;shellshiny Shallow/ normal TallerthanwideWeaktoverystrong crestandwhite callus 0–1(occasionalweak parietal) alpicola2.8–4.21.6–2.1Cylindricalovoid;body whorloftennarrower thanpenultimate TaperedFine,rounded,close, irregularstriae;shellshiny/ satiny ShallowRoundtoslightlywider thantall;smallin proportiontoshell size Weaktostrongcrest; absentto moderatewhite callus 0–2(occasionalweak parietalandpalatal) mostlypalatal depression blandi2.3–3.21.2–1.6CylindricalDomedIrregular,veryweakstriae; shellshiny ShallowRoundtoslightlywider thantall Weaktostrongcrest, absenttostrong, browntowhite callus 0–3 hebes2.6–3.51.4–1.7CylindricalovoidTaperedfor upper1/3–1/4of shellheight Sharp,numerousthread striae;shellsilky Normal/deepRoundtotallerthan wide Weaktostrongcrest; absenttostrong, browntowhite callus 0–3 hebespithodes2.9–3.31.6–1.8Cylindricalovoid;body whorloftennarrower thanpenultimate Taperedfor upper1/3–1/4of shellheight Sharp,numerousthread striae;shellsilky Normal/deepRoundtowiderthantallWeaktoverystrong crest;absentto strongbrownto whitecallus 1–3 hokkaidoensis3.0–3.11.7–1.8OvoidcylindricalTaperedAnastomosingcoarsestriae; shellsilkytodull ShallowRoundtowiderthantallWeakabsentcrest; nocallus 0 hudsonianum3.3–3.61.7–1.8CylindricalovoidStronglytapered forupper1/2 Densethreadstriae;shell silkytodull Normalto deep Tallerthanwide(rarely) towiderthantall Weaktostrong;white callus 0–2(weakparietaland vestigalcolumellar occasionallypresent) cf.khunjerabica2.9–3.51.7–1.8CylindricalovoidTaperedfor upper1/2of shellheight Veryweak,irregularthread striae;dullshell Moderately deep TallerthanwideThin-lacking0 cf.limata2.7–3.11.6–1.7CylindricalovoidTaperedSharp,numerous,somewhat anastamosingstriae;dull shell ShallowTallerthanwideWeakcrest;nocallus0 loessica2.6–3.61.6–2.0CylindricalovoidTaperedStrongbutrounded numerous,anastamosing striae;dullshell NormalRoundWeaktomoderate crest;nocallus 0 muscorum2.7–4.01.6–1.8CylindricalovoidTaperedLow,rounded,somewhat irregularstriae;shellshiny tosilky ShallowRoundStrongtoverystrong crestandwhite callus 1–2(columellarabsent) sonorana2.5–3.31.3–1.4Ovoidcylindrical,widest inupper1/2 DomedWeak,irregularsharpthread striae;shellsilky NormalRoundStrongcrest;strong toverystrongwhite callus 3(palatalrangingfrompeg tolongblade) J. C. NEKOLA ET AL. 210 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 25 from Anchorage, Alaska in Forsyth (2004) refer to material of the present authors and represent P. alaskensis. Habitat: This species has been found in upland and lowland tundra, taiga, fens, herb-rich meadows, coastal grasslands and riparian forest. Remarks: Shell reminiscent of P. loessica, but differs from that species by its widely-spaced, sharp and only slightly anastomosing striae, deeper suture and weaker (or absent) crest. It differs from P. hokkaidoensis in its deep suture, darker shell colour and more regular striae (Table 5). Pupilla hudsonianum Nekola & Coles, n. sp. (Fig. 7A–H, K) Types: holotype (Fig. 7A–D, K): ANSP 458637, Lake Bemidji State Park, Beltrami County, Minnesota, USA (478310 5800 N, 948490 2800 W). Paratypes: 10 shells, ANSP 458638, collected with holotype; 50 shells, NMW.Z.2005.011.00835, collected with holotype; 2 shells, ANSP 458639, highway 40 at Rabbit Hill Road, east of Benchlands (Calgary), Bighorn #8 Municipal District, Alberta, Canada (518150 5100 N, 1148430 5700 W); 100 shells, NMW.Z.2014.013.00058, same loc. as preceding; 5 shells, ANSP 458640, Goose Creek Road, Churchill, Manitoba, Canada (588420 3000 N, 94870 2200 W); 1 shell, ANSP 458641, La Grande Pointe, Duplessis District, Quebec, Canada (508120 2100 N, 638230 4800 W); 30 shells, NMW.Z.2014.013.00001, same loc. as preceding. Zoobank registration: urn:lsid:zoobank.org:act:B9E21337-42C7- 4BFD-87B9-EFE157D2A5A2. Other material examined: NMW.Z.2014.01300054-00060, 00066- 00071, 00080; c. 500 shells. Ten lots from Nekola collection (including one of Pleistocene fossil material; 651 individuals). Etymology: The specific name hudsonianum refers to Hudson Bay and to the Hudsonian life zone, which has been used to refer to the North American taiga, and which defines much of this species’ range. Diagnosis: Shell ovoid-cylindrical, similar to P. hebes, but differentiated by its more ovate shell shape with a surface sculpture consisting of densely packed radial thread-like striae, giving shell a silky lustre. GenBank: GQ921662, KM518353, KM518354, KM518355, KM518357, KM518358, KM518431, KM518432, KM518433, KM518435, KM518508, KM518509, KM518510, KM518512, KM518585, KM518586, KM518587, KM518589. Description: shell 3.3–3.6 mm tall  1.7–1.8 mm wide; opaque to translucent yellowish-brown; c. 6.5–7 whorls; apical whorls rounded-conical in outline, remainder cylindrical; suture moderately deep; shell surface silky in general appearance, postneanic whorls bearing irregular, dense, closely-spaced, weakly anastomosing radial thread-like striae superimposed on a minutely scaly surface, with minute papillae present between the scales (Fig. 7K); aperture c. 1 4 of shell height, approximately circular (Fig. 7F) to wider than tall (Fig. 7A, E, G), rarely taller than wide (Fig. 7H), in profile ascending onto body whorl (Fig. 7B); umbilicus closed by preceding whorls (Fig. 7C); peristome interrupted by body whorl; apertural lip expanded, shell slightly contracted behind; lip thickened by a weakly to strongly developed pale callus of shallow depth corresponding to a weakly to strongly developed crest (Fig. 7A, B, D, E); apertural sterrii2.6–3.51.5–1.8CylindricalovoidTapered/domedSharp,anastomosingcoarse striae;shellsilkytodull VerydeepRoundWeaktostrongcrest; absenttostrong whitecallus 1–2(pegshapedpalatal) syngenes3.0–4.51.6–1.8Biconic,widestinupper 1/3ofshell;8+whorls Tapered/domedWeakthreadstriae;shelldullShallowTallerthanwideVerystrongcrest; weaktomoderate browncallus 3 triplicata2.2–3.1 (rarely to4) 1.3–1.6Cylindricaltocylindrical ovoid TaperedLow,rounded,irregularstriae; shellshinytosilky Normalto deep RoundModeratetostrong crest;absentto strongwhitecallus 0–3(columellaroften absentorweak) turcmenica2.7–3.21.5–1.6CylindricalovoidTapered/domedRegularremote,somewhat anastomisingcoarsestriae; shelldull DeepRoundWeaktostrongcrest; absenttovery strongwhitecallus 0–3(columellarabsent; angularpadsometimes present) TAXONOMY OF PUPILLA 211 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 26 Figure 3. Pupilla species of primarily European distribution. Names are those supported by DNA sequence analysis. A–D. P. muscorum. A. Cedar Rapids, Iowa, USA (22). B. Brno, Moravia, Czech Republic (mtG-Pup) C. Syracuse, New York, USA (AP26). D. Pont, Calvados, France (AP25). E–H. P. alpicola. E. Raksˇ a, Slovakia (H6). F. Belyashi, Altai, Russia (AP12). G. Za´ vod, Slovakia (H5). H. Pozdeˇ chov, Moravia, Czech Republic (H1). I–M. P. sterrii. I. Verkhne Bikberda, Bashkortostan, Russia (AP15). J. Periferi Dibre, Albania (AP16). K. Klentnice, Moravia, Czech Republic (AP22). L. Pavlov, Moravia, Czech Republic (H8). M. Valaska´ Dubova´ , Slovakia (AP21). N–Q. P. triplicata. N. Hracholusky, Bohemia, Czech Republic (H17). O. Ozero Kureevo, Altai, Russia (AP32). P. Pavlov, Moravia, Czech Republic (H9). Q. Cahors, Dordogne, France (AP31). Figure 4. Pupilla species primarily of Asian/Beringian distribution. Names are those supported by DNA sequence analysis. A, B. P. loessica. A. Belyashi, Altai, Russia (AP7). B. Belyashi, Altai, Russia (AP9). C. P. alaskensis, Knik Island, Anchorage, Alaska, USA (AP29). D. P. hokkaidoensis, Toyokoro, Nakagawa, Hokkaido, Japan (VH29). E–H. P. turcmenica. E. Ust’-Muny, Altai, Russia (AP1). F. Kurai, Altai, Russia (AP17). G. Kurai, Altai, Russia (AP18). H. Kosh-Agach, Altai, Russia (AP3). I. P. alluvionica, Belyashi, Altai, Russia (AP13). J. P. cf. khunjerabica, Chagan-Uzun, Altai, Russia (AP11). K. P. cf. limata, Kapitonovka, Yakutia, Russia (AP39). J. C. NEKOLA ET AL. 212 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 27 lamellae generally absent (Fig. 7G, H), but a weak parietal (Fig. 7E) and vestigial plate-like columellar lamella (Fig. 7A, F) occasionally present. Geographical distribution: Currently documented from DNA sequence data from the foothills of the Rockies in western Alberta, Canada east through the northern (Churchill, Manitoba, Canada) and southern (Lake Bemidji, Minnesota, USA) taiga limits in central North America to the north shore of the Gulf of St Lawrence in Quebec. Shell lots at the Academy of Natural Sciences at Drexel University (ANSP 106909, 141759, 141770, 141776, 141783, 150006, 150026), Carnegie Museum (CM 86989, 87010, 62.20823), Museum of Comparative Zoology (MCZ 048304, 201542), National Museum of Canada (NMC 2892, 69132), Royal Ontario Museum (ROM 21464) and University of Michigan Museum of Zoology (UMMZ 55951, 109819, 109829, 168485, 180110, 180112) indicate that P. hudsonianum occurs across the southern shore of Hudson Bay in Ontario and along the Gulf of St Lawrence shore from the Gaspe´ and Anticosti Island in Quebec to the west shore of Newfoundland All reports of P. ‘muscorum’ from Pleistocene sediments in central North America (Hubricht, 1985) represent P. hudsonianum (Fig. 6J). Habitat: This species occurs has been found in mesic taiga, calcareous fens, dry sandy lakeshores and tundra-like turfs on shoreline limestone pavements. Remarks: Pupilla hudsonianum is most readily distinguished from P. muscorum, with which it has been previously confused, by its deeper suture, less massive and more yellow apertural callus, and sharp, fine striae which give the shell a matte luster (Table 5). Pupilla hokkaidoensis Nekola, Coles & S. Chiba, n. sp. (Fig. 8A–K) Types: holotype (Fig. 8A–D, K): ANSP 458642, Toyokoro, Nakagawa District, Hokkaido Prefecture, Japan (428360 1800 N, 1438330 2300 E). Paratypes: 10 shells, ANSP 458643, collected with holotype; 11 shells, NMW.Z.2014.013.00061, collected with holotype; 5 shells, ANSP 458644, Kushiro Marsh, Kushiro District, Hokkaido Prefecture, Japan (43820 200 N, 1448230 2400 E); 12 shells, NMW.Z.2014.013.00062, same loc. as preceding; 5 shells, ANSP 458645, Betsukai, Notsuke District, Hokkaido Prefecture, Japan (438200 5000 N, 1458190 600 E); 5 shells, ANSP 458646, Hama-koshimizu, Shari District, Hokkaido Prefecture, Japan (438560 100 N, 1448260 3800 E); 25 shells, NMW.Z.2014. 013.00064, same loc. as preceding. Zoobank registration: urn: lsid:zoobank.org:act:891CB0E8-E4AA- 43BB-9BDB-8F579461E48A. Other material examined: 6 shells NMW.Z.2005.011.03876, 03878; 4 lots from Nekola collection (176 individuals). Etymology: The specific name hokkaidoensis refers to the island of Hokkaido, where all known populations reside. Diagnosis: Shell small, ovoid-cylindrical, similar to P. hebes but differentiated by a more ovate shell with shallower sutures and shell surface sculpture of coarse, anastomosing, radial striae. GenBank: KM518566, KM518488, KM518411, KM518333. Description: Shell 3.0–3.1 mm tall  1.7–1.8 mm wide; opaque to translucent, yellow-brown; c. 6 whorls; apical whorls conical Figure 5. Pupilla species of North American distribution. Names are those supported by DNA sequence analysis. A–D. P. blandi. A. Ute Creek Canyon, Colorado, USA (AP37). B. Irvine, Alberta, Canada (AP34). C. Moose Jaw, Saskatchewan, Canada (AP35). D. Bannon Ranch, New Mexico, USA (image for ET7 was lost, so a similar shell from same population is figured). E–H. P. hebes pithodes. E. Tusas Ridge, New Mexico, USA (AP38). F. Bullion Canyon, Utah, USA (AP27). G. Bullion Canyon, Utah, USA (AP28). H. Bear Wallow, Arizona, USA (P6). I–M. P. hebes hebes. I. Loope East, California, USA (P14). J. Bullion Canyon, Utah, USA (P17). K. Ruby Mountains, Nevada, USA (P16). L. Kaibab Plateau, Arizona, USA (P1). M. East Tintic Range, Utah, USA (P2). N. P. sonorana, Sacramento Mountains, New Mexico, USA (P12). O. P. hudsonianum, Lake Bemidji, Minnesota, USA (AP33). P, Q. P. syngenes. P. Mogollon, New Mexico, USA (AP30). Q. Kaibab Plateau, Arizona, USA (P11). TAXONOMY OF PUPILLA 213 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 28 in outline, remainder ovoid-cylindrical giving shell slight barrel shape; suture shallow; shell surface silky in general appearance, post-neanic whorls bearing irregular, anastomosing radial striae most strongly developed on mid whorls, superimposed on a minutely and irregularly papillate surface (Fig. 8K); aperture c. 1 4 of shell height, ranging in shape from approximately circular (Fig. 8A, E, H) to taller than wide (Fig. 8F, G), in profile ascending onto body whorl (Fig. 8B); umbilicus closed by preceding whorls (Fig. 8C); peristome interrupted by body whorl; apertural lip flared (Fig. 8B–D), shell slightly contracted behind; crest absent or weakly developed (Fig. 8D), callus absent; apertural lamellae absent. Geographical distribution: Currently known only from the eastern coast of Hokkaido, Japan. Habitat: This species was found in beach grasslands, wetland margins and old fields. Remarks: Pupilla hokkaidoensis differs from P. loessica in its more ovate shell, more yellow shell colour, shallower suture and coarser, more widely spaced and irregular striae. It differs from P. alaskensis in its more yellow shell colour, shallower suture and more irregular striae. It differs from P. cf. limata from Yakutia, Siberia, in its larger size, lighter shell colour and presence of anastomosting striae. DISCUSSION These analyses show that in three widely separated geographic regions the understanding of species-level taxonomy within the genus Pupilla has been hampered by the traditional reliance on a suite of highly plastic shell apertural features that are of little taxonomic value. As a result, too many species have been described in Europe and central Asia, and too few species in North America and eastern Asia, with confusion existing about actual species ranges and ecological tolerances. However, DNA sequence analysis also confirms that most previously described taxa have biological merit, with alternative conchological traits such as shell sculpture and architecture being able accurately to distinguish these entities. Because traditional taxonomic concepts within Pupilla have been based on unstable shell features, larger patterns regarding biodiversity, biogeography and ecology must also be reconsidered. While we cannot deal here with these issues for the entire genus, the current analysis does allow for reconsideration within each of our three study regions. Reassessment of Pupilla biodiversity In Europe oversplitting has been predominant, with both P. bigranata and P. pratensis having been differentiated from P. Figure 6. A–H. Pupilla alaskensis, n. sp. A–D, K. Holotype, ANSP 458632; Happy Valley, North Slope Borough, Alaska, USA. E. Paratype, ANSP 458633; Happy Valley, North Slope Borough, Alaska, USA. F. Paratype, ANSP 458635; Livengood East, Yukon-Koyukuk Census Area, Alaska, USA. G. Paratype, ANSP 458634; Sukakpak Mountain, Yukon-Koyukuk Census Area, Alaska, USA. H. Paratype, ANSP 458636; Knik I., Matanuska-Susitna Borough, Alaska, USA. I, L. P. hebes, JCN 17254; Loope East, Alpine Co., California, USA. J. P. loessica, Belyashi, Altai Republic, Russia; 498160 800 N, 878590 200 E. Figure 7. A–H, J, K. Pupilla hudsonianum. A–D, K. Holotype, ANSP 458637; Lake Bemidji State Park, Beltrami County, Minnesota, USA. E. Paratype, ANSP 458638; Lake Bemidji State Park, Beltrami County, Minnesota, USA. F. Paratype, ANSP 458641; La Grande Pointe, Duplessis District, Quebec, Canada. G. Paratype, ANSP 458639; East of Benchlands, Bighorn #8 Municipal District, Alberta, Canada. H. Paratype, ANSP 458640; Goose Creek Road, Churchill, Manitoba, Canada. J. Pleistocene loess fossil, Wenig Road, Cedar Rapids, Linn Co., Iowa, USA; 42800 800 N, 918400 4000 W; JCN 3650. I, L. P. muscorum; Syracuse University South Campus, Syracuse, Onondaga Co., New York, USA; 43800 2700 N, 76860 3800 W; JCN 13955. J. C. NEKOLA ET AL. 214 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 29 muscorum and P. alpicola, respectively, based upon unstable shell apertural characters. It is fortunate that degree of apertural calcification has never been used to split P. triplicata, as its development of lamellae can range from three very strong (French Pyrenees) to absent (basalt talus in northern Bohemia). In central Asia, oversplitting has also been an issue. While some recently described taxa are strongly demarcated (e.g. P. alluvionica and P. cf. khujerabica), others (P. altaica and P. pratensis) appear to represent high-calcification endpoints in apertural development within previously described species (P. turcmenica and P. alpicola, respectively). Perhaps it is not surprising that such high-calcification shell forms tend to originate from drier, lower elevation sites, which would have higher calcium availability due to lower leaching and higher potential evapotranspiration rates (Lapenis et al., 2008). In North America, overlumping and ignorance of taxonomically valid shell traits has led to considerable confusion. First, P. muscorum is not a native North American species, with all examined putative native populations representing either P. hebes (southwestern USA) or P. hudsonianum (central/eastern taiga and tundra) and with P. muscorum xerobia being a junior synonym of P. blandi. Pupilla blandi should be distinguished from other North American species not by apertural features, but rather by its weak to obsolete striation and shining shell lustre. Pupilla hebes should be distinguished by its strong thread-like striae and narrow, columnar shell. Pupilla hebes pithodes is most closely related to P. hebes, but differs in its wider and more barrel-shaped shell. Additionally, P. alaskensis has been variously regarded as P. muscorum or P. hebes in spite of its coarser striation, ovate shell shape and deeper suture than either of these species. Reassessment of Pupilla biogeography In Europe, P. alpicola cannot be considered a central European endemic with a disjunct set of populations in the Altai (Horsa´ k et al., 2010). Rather, it extends continuously northwest into southern Scandinavia and Ireland (as the former P. pratensis) and east into central Asia. Although demarcation between the central Asian and European populations is evident in nDNA, the central Asian subpopulation extends at least as far west as Bohemia. The amount of mixing of these two populations during full glacial stages is thus unclear. Pupilla triplicata occurs as far east as the Altai in central Asia. Pupilla sterrii is the western sibling of P. turcmenica, with populations extending from central Europe east to the Urals. Pupilla muscorum is not a Holarctic species, but is a European endemic with confirmed Pleistocene fossil occurrences in loess deposits of central Europe. In the Altai, species status for two putative central Asian endemics (P. alluvionica and P. cf. khunjerabica) was established. However, another (P. altaica) was found to be simply a shell form within P. turcmenica, which ranges from western China and Tibet to the Iran–Turkmenistan border (Pilsbry, 1921). Pupilla loessica was shown to be a member of a Beringian group that also includes P. hokkaidoensis and P. alaskensis. In North America, the lack of true P. muscorum as a Pleistocene fossil suggests that it is an exotic species ranging from the western Great Lakes east to Virginia and north into the Canadian maritime provinces. The identical COI haplotype of the Brno and Cedar Rapids P. muscorum specimens suggest that both populations were sourced from the same pool. This is not surprising given that extensive immigration from the Czech Republic to eastern Iowa happened during the mid-1800s. Pupilla hudsonianum, which has been previously regarded as P. muscorum, extends west from the north shore of the St Lawrence River in Quebec to the southern border of Hudson Bay, northwestern Minnesota and the foothills of the Rockies in Alberta. It also represents the putative Pleistocene fossil ‘P. muscorum’ reported by Hubricht (1985) across the central Midwestern USA and Plains. Pupilla blandi is limited to the Plains (NE New Mexico to NW Minnesota to southern Saskatchewan and Alberta) and rarely penetrates west into the Rockies as far as the continental divide. Pupilla hebes is characteristic of the Great Basin from Arizona and California to Utah and Idaho, with a well-demarcated subpopulation from the Colorado Plateau being demonstrated by mtDNA. While this subpopulation would equate to P. hebes kaibabensis, its shells and nDNA do not differ in any meaningful way from typical P. hebes, and we have not chosen to recognize it here. Pupilla hebes pithodes is found to the south and east of typical P. hebes, ranging from eastern Arizona and SE Utah to the eastern foothills of the Rockies in Colorado and New Mexico. Pupilla alaskensis, formerly confused with P. hebes, is actually a sibling of the western Beringian P. loessica. Reassessment of Pupilla ecology The existence of so much apertural variation within species across Pupilla begs for an explanation. In particular, how much of these differences are due to genetic variation and how much to ecophenotypic response? Little empirical data exist to address this question. However, we typically noted limited variation in apertural features within populations. Populations expressing a poorly developed apertural callus and lamellae tended to be found in sites with low calcium availability, such as P. blandi Figure 8. A–H. Pupilla hokkaidoensis. A–D, K. Holotype, ANSP 458642; Toyokoro, Nakagawa District, Hokkaido Prefecture, Japan. E. Paratype, ANSP 458643; Toyokoro, Nakagawa District, Hokkaido Prefecture, Japan. F. Paratype, ANSP 458644; Kushiro Marsh, Kushiro District, Hokkaido Prefecture, Japan. G. Paratype, ANSP 458646; Hama-koshimizu, Shari District, Hokkaido Prefecture, Japan. H. Paratype, ANSP 458645; Betsukai, Notsuke District, Hokkaido Prefecture, Japan. I. P. hebes, Charleston, Elko Co., Nevada, USA; 418410 1200 N, 1158300 2200 W; JCN 18292. J. P. cf. limata, Kapitonovka, Yakutia Republic, Russia; 628190 4500 N, 1298550 4200 E. TAXONOMY OF PUPILLA 215 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 30 (AP37) on acid metamorphic rock in the Colorado Rockies, and P. triplicata on basalt talus slopes in Bohemia (e.g. H10). In contrast, the most heavily calcified P. turcmenica in the Altai tend to be restricted to xeric, low-elevation steppe, often on calcium-rich metamorphic rock or limestone. Individual age and the season when maturity is reached may also play important factors. While these observations suggest that ecophenotypic or developmental response is responsible for much of the observed variation, different shell forms have nevertheless been observed in co-occurring individuals that share identical mtDNA and nDNA haplotypes (e.g. AP27, AP28), suggesting that multiple factors may be operating. ACKNOWLEDGEMENTS We wish to thank Albuquerque Institute of Math & Science students Alex Jones, Gabe Martinez, Sarah Martinez and Khayman Mondragon for assisting with initial PCR and sequencing of the North American specimens. Japanese field work would not have been possible without the logistical support of Satoshi Chiba. Lucie Jurˇ icˇ kova´ (Charles University) kindly provided historical Pupilla collections by Vojen Lozˇ ek from Bohemia, Toma´ sˇ Cˇ ejka (Slovak Academy of Sciences) provided material of Slovakian P. pratensis and Zolta´ n Fehe´ r (Hungarian Natural History Museum) provided P. sterrii from Albania. Major funding for field work was provided by the USA National Science Foundation (EAR-0614963), the Czech Science Foundation (P504-11-0454) and the Japanese Society for the Promotion of Science (Kakenhi-24340128). We initially embarked on this project to discern whether P. triplicata with vestigal or absent lamellae from the basalt screes of northern Bohemia warranted erection as a new species. Had our initial hypothesis been validated, we intended to name this taxon after Dr Lozˇ ek. While this work ultimately documented an entirely different story, we would still like to thank Dr Lozˇ ek for his lifetime of work on Eurasian land snails, including the initial description of P. loessica. REFERENCES ARMBRUSTER, G.F.J. & BERNHARD, D. 2000. Taxonomic significance of ribosomal ITS-1 sequence markers in self-fertilizing land snails of Cochlicopa (Stylommatophora, Cochlicopidae). Zoosystematics & Evolution, 76: 11–18. EMBERTON, K.C. 1995. When shells do not tell—145-million years of evolution in North American polygyrid land snails, with a revision and conservation priorities. Malacologia, 37: 69–109. FORSYTH, R.G. 2004. Land snails of British Columbia. Royal BC Museum, Victoria, Canada. HORSA´ K, M., CHYTRY´ , M. & AXMANOVA´ , I. 2013. Exceptionally poor land snail fauna of central Yakutia (NE Russia): climatic and habitat determinants of species richness. Polar Biology, 36: 185–191. HORSA´ K, M., CHYTRY´ , M., POKRYSZKO, B.M., DANIHELKA, J., ERMAKOV, N., HA´ JEK, M., HA´ JKOVA´ , P., KINTROVA, K., KOCˇ I´ , M., KUBESˇ OVA´ , S., LUSTYK, P., OTY´ PKOVA´ , Z., PELA´ NKOVA´ , B. & VALACHOVICˇ , M. 2010. Habitats of relict terrestrial snails in southern Siberia: lessons for the reconstruction of palaeoenvironments of full-glacial Europe. Journal of Biogeography, 37: 1450–1462. HORSA´ K, M., HA´ JEK, M., SPITALE, D., HA´ JKOVA´ , P., DI´ TEˇ , D. & NEKOLA, J. 2012. The age of island-like habitats impacts habitat specialist species richness. Ecology, 93: 1106–1114. HUBRICHT, L. 1985. The distributions of the native land mollusks of the eastern United States. Fieldiana, New Series, 24: 1–191. HUELSENBECK, J.P. & RONQUIST, F. 2001. MRBAYES: Bayesian inference of phylogeny. Bioinformatics, 17: 754–755. KERNEY, M.P. & CAMERON, R.A.D. 1979. Field guide to the land snails of the British Isles and northwestern Europe. Collins, London. LAPENIS, A.G., LAWRENCE, G.B., BAILEY, S.W., APARIN, B.F. & SHIKLOMANOV, A.I. 2008. Climatically driven loss of calcium in steppe soil as a sink for atmospheric carbon. Global Biogeochemical Cycles, 22: GB2010, doi:10.1029/2007GB003077. MENG, S. & HOFFMAN, M.H. 2008. Pupilla altaica n. sp. and Pupilla alluvionica n. sp., two new species of the Pupillidae (Gastropoda: Pulmonata: Pupilloidea) from the Russian Altay. Mollusca, 26: 229–234. MENG, S. & HOFFMAN, M.H. 2009. Pupilla seminskii n. sp. (Gastropoda: Pulmonata: Pupilloidea) from the Russian Altay. Mollusca, 27: 83–86. NEKOLA, J.C. 2005. Latitudinal richness, evenness, and shell size gradients in eastern North American land snail communities. Records of the Western Australian Museum, Supplement, 68: 39–51. NEKOLA, J.C. 2014. North American terrestrial gastropods through either end of a spyglass. Journal of Molluscan Studies, 80: 238–248. NEKOLA, J.C. & COLES, B.F. 2010. Pupillid land snails of eastern North America. American Malacological Bulletin, 28: 29–57. NEKOLA, J.C., COLES, B.F. & BERGTHORSSON, U. 2009. Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Molecular Phylogenetics and Evolution, 53: 1010–1024. NEKOLA, J.C. & ROSENBERG, G. 2013. Vertigo marciae (Gastropoda, Vertiginidae), a new land snail from Jamaica. Nautilus, 127: 107–114. OUGHTON, J. 1948. A zoogeographical study of the land snails of Ontario. University of Toronto Studies. Biological Series, 57: 1–126. PILSBRY, H.A. 1921. Manual of conchology, second series. Vol. 26: Pupillidae (Vertigininae, Pupillinae). Academy of Natural Sciences, Philadelphia. PILSBRY, H.A. 1948. Land Mollusca of North America (north of Mexico). Academy of Natural Sciences of Philadelphia Monograph 3, vol. 2. POKRYSZKO, B.M., AUFFENBERG, K., HLAVA´ Cˇ , J.Cˇ . & NAGGS, F. 2009. Pupilloidea of Pakistan (Gastropoda: Pulmonata): Truncatellininae, Vertigininae, Gastrocoptinae, Pupillinae (in part). Annales Zoologici (Warszawa), 59: 423–458. SCHILEYKO, A.A. 1984. Nazemnye molljuski podotryda Pupillina fauny SSSR (Gastropoda, Pulmonata, Geophila). Fauna SSSR, Molljuski, 3: 1–399 (In Russian). TONGKERD, P., LEE, T., PANHA, S., BURCH, J.B. & O´ FOIGHIL, D. 2004. Molecular phylogeny of certain Thai Gastrocoptinae micro land snails (Stylommatophora: Pupillidae) inferred from mitochondrial and nuclear ribosomal DNA sequences. Journal of Molluscan Studies, 70: 139–147. UIT DE WEERD, D.R., PIEL, W.H. & GITTENBERGER, E. 2004. Widespread polyphyly among Alopiinae snail genera: when phylogeny mirrors biogeography more closely than morphology. Mollecular Phylogenetics and Evolution, 33: 533–548. VON PROSCHWITZ, T., SCHANDER, C., JUEG, U. & THORKILDSEN, S. 2009. Morphology, ecology, and DNA-barcoding distinguish Pupilla pratensis (Clessen, 1871) from Pupilla muscorum (Linnaeus, 1758) (Pulmonata: Pupillidae). Journal of Molluscan Studies, 75: 315–322. WADE, C.M. & MORDAN, P.B. 2000. Evolution within the gastropod molluscs; using the ribosomal RNA gene-clustre as an indicator of phylogenetic relationships. Journal of Molluscan Studies, 66: 565–570. J. C. NEKOLA ET AL. 216 atUniversityofNewMexicoonMay4,2015http://mollus.oxfordjournals.org/Downloadedfrom 31 Vertigo marciae (Gastropoda: Vertiginidae), a new land snail from Jamaica Jeffrey C. Nekola Department of Biology University of New Mexico Albuquerque, NM 87131 USA jnekola@unm.edu Gary Rosenberg Department of Malacology Academy of Natural Sciences of Drexel University 1900 Benjamin Franklin Parkway Philadelphia, PA 19103 USA rosenberg@ansp.org ABSTRACT Vertigo marciae, a new species of gastropod mollusk (Pupilloidea: Vertiginidae), is described from Jamaica. This species is known in the Recent fauna only from John Crow Peak in the Blue Mountains, but also occurs as a Pleistocene fossil at Red Hills Road Cave. Vertigo marciae has been confused with Vertigo gouldii, but differs by its smaller shell size, lack of distinct shell striation, lack of an angular lamella, and presence of a flared aperture base. DNA sequence analyses document that V. marciae possesses unique mtDNA and nDNA sequences and is most closely allied with Vertigo alabamensis, Vertigo hebardi, and Vertigo oscariana. This group of species comprises a highly supported clade whose members are limited either to the Caribbean or the southeastern USA. Additional Keywords: Vertigo, biogeography, southeastern USA, Caribbean, DNA sequence analysis INTRODUCTION In their synopsis of the Jamaican land snail fauna, Rosenberg and Muratov (2006) reported three species from the genus Vertigo, all of which possess ranges extending into the eastern half of North America (Nekola and Coles, 2010): V. gouldii, V. milium, and V. ovata. However, images of putative Jamaican ‘V. gouldii’ on the Discover Life website (http://pick4 .pick.uga.edu/mp/20q?search=Vertigoþgouldii&guide¼1) illustrate a shell quite unlike V. gouldii from eastern North America (Nekola and Coles, 2010) in its lightyellow shell color, lack of sharp shell striation, and absence of a basal lamella. Observation of the three known Jamaican lots for this entity in the Academy of Natural Sciences of Philadelphia (ANSP) collections confirmed these differences. Although they looked most like V. hebardi Vanatta, 1912 of the Florida Keys, these lots clearly differed from that species by their taller shell, reduced striation, and absence of an angular lamella. Thus, rather than representing a population of V. gouldii isolated by 2000 km from its nearest neighbors (Nekola and Coles, 2010), these specimens appear to represent an undescribed new species. Based on shell descriptions, reports of Vertigo gouldii from the Pleistocene Red Hills Road Cave deposits of Jamaica (Paul and Donovan, 2005) also appeared to represent this new species. This conclusion was subsequently confirmed via observation of digital images. We therefore investigated this putative new species via analyses of both shell morphology and DNA sequence data, and are now in the unusual situation of describing and reporting a new species simultaneously from both living and fossil material, while also being able to report on its phylogenetic relationships. MATERIALS AND METHODS Field Collection: The three known Recent lots of the putative new Jamaican species (ANSP 402244, 403039, and 403040) were collected on John Crow Peak in the Blue Mountains of eastern Jamaica on May 22, 1999, during the Jamaican Biotic Survey. Collecting methods were detailed in Rosenberg and Muratov (2006), and included drying of soil litter samples over a Berlese Funnel. This encouraged snails to enter aestivation, allowing them to mummify upon death. For this reason, it was possible to successfully extract, amplify, and sequence selected amplicons from both their mitochondrial and nuclear genomes (Nekola et al., 2009). Shell Measurements: Measurements were determined in 0.1 mm increments for adult shells with aperture facing up, using a dissecting microscope with a calibrated ocular micrometer. Height and width were measured as the dimensions of a bounding box with long axis parallel to the shell axis and sides tangent to the tip of the protoconch, the base of the lip, the right-most margin of the aperture and the left-most margin of the body whorl. Imaging: Shells were imaged at 20 magnification using a digital camera attached to a stereomicroscope. THE NAUTILUS 127(3):107–114, 2013 Page 107 32 Approximately 14 separate 1388Â1040 pixel images were made of each specimen with the image focal lengths positioned at 75 mm increments from the front to back of the shell. CombineZ5 freeware (http://www .hadleyweb.pwp.blueyonder.co.uk/CZ5/combinez5.htm) was used to assemble a final image from the focused parts of each separate image. Assembled images were imported into Adobe Photoshop, where brightness and contrast were optimized and the background made uniformly black. These images were then compiled into a single plate. DNA Sequence Analysis: Mitochondrial cytochrome b (CytB), 16S ribosomal RNA (16S) and the internal transcribed spacers 1 (ITS–1) and 2 (ITS-2) of the nuclear ribosomal RNA gene were investigated to test the distinctiveness of the putative new Jamacian species and to resolve its nearest evolutionary neighbors. The mitochondrial cytochrome oxidase subunit 1 (CO1) was not analyzed as we were unable to amplify this gene in the Jamaican species. Sixteen specimens were chosen for analysis (Table 1). This set includes not only two specimens of the Jamaican species but also Vertigo alabamensis, V. gouldii, V. hebardi, and V. oscariana. We attempted to maximally spread these individuals across the known geographic range of each species (see Nekola and Coles, 2010). Because V. conecuhensis is likely a simple shell form of V. alabamensis (Nekola and Coles, 2010), only a single individual of this taxon was included. Topotype or near-topotype material was selected for V. conecuhensis (Pond Creek, Covington Co., Alabama, about 50 km from the type locality in Evergreen, Alabama) and V. hebardi (Long Key, Florida which is the type locality). Previously analyzed 16S sequence data from three Vertigo gouldii specimens analyzed by Nekola et al. (2009) were retrieved from GeneBank (Table 1). Vertigo pusilla, the type species of the genus, was also included, as were Gastrocopta cristata and Pupilla muscorum for outgroup comparisons. Genomic DNA was extracted from live, ethanolpreserved, or mummified material using the OmegaBioTek Mollusk DNA Extraction Kit. Because shell destruction was required, all shells were imaged prior to extraction using the methods detailed above. PCR amplification and sequencing of CytB, 16S, ITS-1, and ITS-2 were accomplished using standard methods (Nekola et al., 2009). Resultant traces were examined, primer ends removed, and aligned by eye. Because of their lower evolutionary rates, the nDNA ITS-1 and ITS-2 regions were concatenated to provide a roughly similar number of nucleotide differences to either of the mtDNA amplicons. Mega 5.0 (Tamura et al., 2011) was used to calculate the average number of nucleotide differences in the CytB, 16S, and concatenated ITS-1 þ ITS-2 regions between all putative species-level taxa. Substitutions included both transitions and transversions with pairwise gap deletion. Table 1. Specimen information for material used in DNA sequence analysis. Taxon Location Latitude/Longitude GENBANK Accession number CytB 16S ITS-1 ITS-2 Vertigo marciae new species John Crow Peak, St. Andrew Parish, Jamaica 18.1132 N., 76.6685 W. KF214513 KF214503 John Crow Peak, St. Andrew Parish, Jamaica 18.1132 N., 76.6685 W. KF214514 KF214502 KF214489 KF214477 Vertigo alabamensis Clapp, 1915 Johnson Mill Bay, Bladen Co., North Carolina, USA 34.7125 N., 78.5261 W. KF214515 KF214501 KF214490 KF214478 Wolf Trap Bay, Leon Co., Florida, USA 30.3680 N., 84.5700 W. KF214516 KF214500 KF214491 KF214479 Vertigo conecuhensis Clapp, 1915 Pond Creek, Covington Co., Alabama 31.1036 N., 86.5343 W. KF214517 KF214499 KF214492 KF214480 Vertigo gouldii (A. Binney, 1843) Brush Creek Canyon, Fayette Co., Iowa, USA 42.7796 N., 91.6890 W. KF214508 GQ921506 KF214484 KF214472 Panther Creek, Searcy Co., Arkansas, USA 36.0858 N., 92.5649 W. KF214509 GQ921507 KF214485 KF214473 11-Point River, Oregon Co., Missouri, USA 36.7931 N., 91.3334 W. KF214510 KF214506 KF214486 KF214474 Vertigo hebardi Vanatta, 1912 Long Key, Monroe County, Florida, USA 24.8146 N., 80.8211 W. KF214511 KF214505 KF214487 KF214475 Elliott Key, Miami-Dade County, Florida, USA 25.4553 N., 80.1925 W. KF214512 KF214504 KF214488 KF214476 Vertigo oscariana (Sterki, 1890) Blanchard Springs, Stone Co., Arkansas, USA 35.9582 N., 92.1778 W. KF214518 KF214498 KF214493 KF214481 Wadboo Creek, Berkeley Co., South Carolina, USA 33.1971 N., 79.9461 W. KF214519 KF214497 KF214494 KF214482 Vertigo pusilla Mu¨ller, 1774 Podyji National Park, Moravia, Czech Republic 48.8586 N., 15.8960 E. KF214520 KF214496 KF214495 KF214483 Gastrocopta cristata (Pilsbry and Vanatta, 1900) Albuquerque, Bernalillo Co., New Mexico, USA 35.0727 N., 106.6160 W. KF214522 JN941032 Pupilla muscorum (Linnaeus, 1758) Masaryk University, Brno, Moravia, Czech Republic 49.2509 N., 16.5738 E. KF214521 KF214507 Page 108 THE NAUTILUS, Vol. 127, No. 3 33 Mega 5.0 was used to conduct nearest–neighbor joining (NNJ), maximum parsimony (MP), and maximum likelihood (ML) trees for the three focal regions. NNJ was based on Maximum Composite Distance including transitions and transversions with pairwise gap deletion. MP used the close neighbor interchange search option with the random addition of 10 replicate trees. ML used all sites and was based on the Tamura–Nei substitution model, a five–category Gamma Distribution for substitution rates, and the Nearest Neighbor Interchange ML heuristic method. In all cases support values were estimated from 1000 bootstrap replicates. Additionally, Bayesian trees were generated using MrBayes 3.1 (Huelsenbeck and Ronquist, 2001) using a GTR substitution model assuming gamma-shaped rate variation over 1,000,000 generations with a sampling frequency of once each 1000 generations. Nomenclature: Taxonomic names and concepts follow Nekola and Coles (2010). Apertural lamellae and fold nomenclature follows that of Pilsbry (1948: 869, fig. 469), i.e., parietal “teeth” are referred to as “folds” and all other “teeth” are termed “lamellae”, whatever their form. Also, we follow Pilsbry (1948) by referring to the raised riblets on the surface of Vertigo shells as “striae.” SYSTEMATICS Class Gastropoda Subclass Pulmonata Order Stylommatophora Family Vertiginidae Genus Vertigo Mu¨ller, 1773 Vertigo marciae new species (Figures 1–6; Tables 1, 2) GenBank Accessions KF214477, KF214489, KF214502, KF214503, KF214513, KF214514 Diagnosis: Minute shell reminiscent of Vertigo hebardi but taller, less striate, and lacking angular lamella; shell surface smooth and glossy with indistinct wrinkles; aperture flared toward bottom; four lamellae/folds present, including parietal, columellar, and two palatals. Description: Shell 1.4–1.6 mm tall  0.8–1.0 mm wide (holotype 1.5  0.8 mm), columnar-ovoid to ovoid, approximately 4–4.5 whorls, with moderately shallow suture and domed apex. Translucent, pale yellow-brown color. Body whorl approximately 60% of total height (Figures 1–3, 5, 6). Protoconch and neanic whorls smooth (Figure 4), with subsequent whorls having irregular, infrequent, and weak wrinkles. Immediately behind aperture sculpture takes the form of irregular low striae (Figure 2). Aperture flared on bottom, making it taller than wide and approximately 1/3 of shell height. Lip unthickened and slightly reflexed, sinulus moderate-weak, sometimes expressed as simple flattening of palatal wall. Basally aperture abruptly inflates to form rounded swelling but not crest (Figure 4). Umbilicus closed (Figure 3). Aperture with four lamellae/folds: parietal lamella strong, slightly sinuous (Figures 1, 2, 5, 6); columellar lamella downward-sloping, peg-shaped; two palatal folds with lower being approximately twice as long as upper and extending approximately 0.2 whorls into body whorl, lower slightly more immersed than upper, both highest at mid length (Figures 1, 2, 5, 6). Apertural end of lower palatal fold coincides with abrupt inflation of basal aperture. Externally shell only slightly impressed over palatal folds (Figure 4). Type Material: Holotype (Figures 1–4), ANSP 450580, from type locality, May 22, 1999; paratypes (Figures 5–6), ANSP 402244, 8 shells and Institute of Jamaica, 2 shells, from type locality; ANSP 403039, Jamaica, St. Andrew Parish, John Crow Peak, Blue Mountains, elfin forest with bamboo near summit, 18 050 4500 N, 76 400 08.400 W, altitude 1755 m, May 22, 1999; sta. JBS4a, 16 shells; ANSP 403040, Jamaica, St. Andrew Parish, Vinegar Hill Trail near head of Clyde River, 18 050 N, 76 390 1800 W, altitude 1520 m, May 22, 1999, sta. JBS 5, 3 shells. Type Locality: Jamaica, St. Andrew Parish, John Crow Peak, Blue Mountains; litter sample collected at base of limestone cap, 18 50 5000 N, 76 400 500 W, altitude 1550 m, May 22, 1999, sta. JBS4c. Other Material Examined: Digital images of two specimens from Red Hills Road Cave, Jamaica (Paul and Donovan, 2005), Paul collection. Etymology: The specific name marciae refers to Dr. Marcia Mundle, then of the Jamaica Conservation and Development Trust. Dr. Mundle arranged for the vehicle used and park ranger guide that accompanied the expedition to the type location, and was present when the species was first collected. RESULTS AND DISCUSSION Variation: Vertigo marciae is rather constant in general appearance in terms of its shape, color, sculpture, and development of the apertural lamellae. While some variation in size (especially height) was noted, this was minor—only 16% difference between the largest and smallest shells was observed. Comparison with Other Vertigo Species: Vertigo marciae differs from all other Vertigo species by its small (<1.7 mm in height) yellow shell with indistinct striae/wrinkles, flared aperture base, and lack of angular and basal lamellae. It is closest in appearance to V. hebardi of the Florida Keys (Figure 7), with which it shares a small yellow shell, the lack of a basal lamella, and preference for accumulations of tropical forest leaf litter. However, it differs from this species in its taller shell with less distinct striae and absence of an angular lamella. It is also reminiscent of Vertigo marki Gulick, J.C. Nekola and G. Rosenberg, 2013 Page 109 34 Figures 1–12. Stereomicroscope images of Vertigo marciae and related taxa. 1–4. Vertigo marciae, holotype, Jamaica, St. Andrew Parish, John Crow Peak, Blue Mountains, 18 50 4500 N, 76 400 800 W to 18 50 5000 N, 76 400 500 W, ANSP 450580. 1. Apertural view. 2. Profile. 3. Umbilical view showing parietal and upper palatal lamellae. 4. Apical view showing protochonch and the basal apertural dilation. 5. Vertigo marciae, ANSP 402244, Paratype, second specimen from the type locality, exhibiting a shorter shell. 6. Vertigo marciae, ANSP 402244, Paratype, third specimen from the type locality, exhibiting a slightly more worn shell. 7. Vertigo hebardi, Long Key, Monroe County, Florida, 24 480 5200 N, 80 490 1400 W, JCN 17479. 8. Vertigo gouldii (small southern form), Tellico Gorge, Monroe Co., Tennessee, 35 190 4900 N, 84 100 5900 W, BFC 1332. 9. Vertigo alabamensis, Lanier Quarry, Pender Co., North Carolina, 34 370 4900 N, 77 400 2700 W, JCN 10781. 10. Vertigo “conecuhensis,” Pond Creek seep, Covington Co., Alabama, 31 60 1200 N, 86 320 300 W, JCN 12364. 11. Vertigo oscariana, Wadboo Creek, Berkeley Co., South Carolina, 33 110 5000 N, 79 560 4600 W, JCN 10908. 12. Vertigo gouldii (normal form), Deer Creek, Fillmore Co., Minnesota, 43 430 5600 N, 92 200 3900 W, JCN 14646. Page 110 THE NAUTILUS, Vol. 127, No. 3 35 1904 from Bermuda, with which it shares a smooth, yellowish shell that lack both basal and angular lamellae (Pilsbry, 1919). However, V. marciae differs from this species by its smaller shell height and lack of a callus on the palatal wall of the aperture. Vertigo oscariana from the southeastern USA (Figure 11) also has a small, smooth yellow shell without a basal lamella, but this species also lacks an upper palatal fold, has a callus on the outer apertural margin, and a shell that is widest in the middle, tapering both to base and apex. Vertigo alabamensis of the southeastern North American coastal plain (Figures 9, 10) also has a yellowish shell lacking striae, but this species is larger (generally >¼1.8 mm tall), possessing both angular and basal lamellae, a strong crest on the apertural margin, and a strong sinulus on the palatal wall of the aperture. Rosenberg and Muratov (2006) identified Vertigo marciae as Vertigo gouldii (Figures 8, 12) because of its general shell shape and placement of the parietal lamella and palatal folds, and also because V. gouldii was reported from Jamaica by Pilsbry and Cooke (1919: 99). However, V. gouldii has a larger shell of brown color, has a duller shell luster from the presence of abundant strong but irregular striae, and always possesses a basal lamella. Specimens documenting the Pilsbry and Cooke record have not been traced, and are apparently based on personal communication between Victor Sterki and Pilsbry. Bequaert and Miller (1973: 95) rejected Antillean records of V. gouldii. While Vertigo milium—which also occurs in Jamacian tropical forest leaf litter—possesses a shell <1.7 mm tall; this species is easily distinguished from V. marciae in its dark red-brown color, presence of both angular and basal lamellae, and its long, curved lower palatal fold which deeply enters the shell. Geographic Distribution and Ecology: Vertigo marciae is currently known in the Recent fauna only from the crest of John Crow Peak and its immediate vicinity in the Blue Mountains of eastern Jamaica at elevations of 1520–1755 m. John Crow Peak is capped by an isolated limestone outlier known to have distinctive plant communities, including endemic species (Grossman et al., 1993). (Note that John Crow Peak is not the same location as the John Crow Mountains, the easternmost range of Jamaica.) Vertigo marciae is found mainly on or adjacent to limestone boulders and outcrops in tropical as well as scrub forest with bamboo, but has also been found at one site (JBS 5) that lacks exposed limestone. As a Pleistocene fossil, it is known only from Red Hills Road Cave, which is about 21 km west of Rosenberg and Muratov’s sites on John Crow Peak and 1000 m lower in elevation (520 m, Paul and Donovan, 2005). Vertigo marciae joins the two Radiodiscus species and the Punctum reported by Paul and Donovan (2005) on the basis of personal communication from Rosenberg as species known in Jamaica only from John Crow Peak and from the Red Hills Road Cave. Paul and Donovan interpreted the faunal changes from the Red Hills Road Cave fauna to the recent fauna around the cave as suggesting drying of the climate since the Pleistocene. As Vertigo marciae was found nowhere else in Jamaica among the hundreds of sites sampled by Rosenberg and Muratov, its current occurrence might represent a relict distribution. Although it must currently be considered a Jamaican endemic, the general lack of local endemism in the genus Vertigo (Nekola, 2009) suggests the possibility that it may occur elsewhere in the adjacent Caribbean, especially on carbonate substrates in mid- to high-elevation montane forest. Table 2. Average number of base-pair differences between all pairwise combinations of Vertigo marciae and related species. A. CytB amplicon (355 base pairs) V. oscariana 53.1 V. alabamensis 65.6 54.3 V. hebardi 62.3 52.0 8.3 V. marciae 63.8 46.5 38.7 36.0 V. gouldii V. oscariana V. alabamensis V. hebardi B. 16S amplicon (443–447 base pairs) V. oscariana 40.9 V. alabamensis 37.3 44.5 V. hebardi 35.8 44.5 4.0 V. marciae 42.5 42.0 18.0 18.0 V. gouldii V. oscariana V. alabamensis V. hebardi C. Concatenated ITS-1 and ITS-2 amplicons (1274–1284 base pairs) V. oscariana 34.8 V. alabamensis 32.3 20.5 V. hebardi 34.3 22.5 5.0 V. marciae 37.3 23.5 9.0 12.0 V. gouldii V. oscariana V. alabamensis V. hebardi J.C. Nekola and G. Rosenberg, 2013 Page 111 36 Genetic Distinctness and Phylogenetic Relationships: DNA sequence analysis clearly demonstrates that V. marciae is distinct at the species level. It possesses an average of 36 and 39 base pair differences with V. hebardi and V. alabamensis, respectively, in the 355 bp CytB amplicon (Table 2). This equates to a 10–11% variation across the entire amplicon. In addition, it possessed an average of 46.5 base pair differences in CytB (13%) with V. oscariana and 63.8 differences (18%) with V. gouldii. In the more slowly evolving 443 to 447 bp 16S region V. marciae differed by 18 bases (4.0%) from V. hebardi and V. alabamensis, 42 bases (9.4%) from V. oscariana, and 42.5 bases (9.6%) from V. gouldii. In the concatenated 1274–1284 bp ITS-1 þ ITS-2 nDNA amplicon, V. marciae possessed an average of 9 basepair (0.7%) differences with V. alabamensis, 12 (0.9%) with V. hebardi, 23.5 (1.8%) with V. oscariana, and 37.3 (2.9%) with V. gouldii. These levels of difference were considerably larger than those exhibited between the conchologically distinct V. alabamensis and V. hebardi (Figures 7, 9, 10), which amounted to an average difference of 8.3 bases (2.3%) in CytB, 4.0 (0.9%) in 16S, and 5.0 (0.4%) in concatenated ITS-1 and ITS-2. The topologies across phylogenetic reconstructions were largely compatible, with the same highly supported nodes being identified in all cases (Figures 13–15). These demonstrate that V. marciae is clearly a member of the genus Vertigo, being within the same highly supported clade in CytB and 16S that includes V. pusilla, the type species of the genus. While it was not possible to root the concatenated ITS-1 þ ITS-2 tree because of profound differences with both Gastrocopta cristata and Pupilla muscorum sequences, V. marciae was easily aligned with all Vertigo sequences. The phylogenetic reconstructions also demonstrate that V. marciae is a member of a highly supported clade that includes both V. alabamensis and V. hebardi. The existence of a single ancestor to all of these species implies long distance dispersal from the southeastern United States to Jamaica, perhaps with migrating birds as a vector. This is not an unreasonable scenario, given that much longer feats of long distance dispersal via migrating birds have been documented in the eastern Atlantic (Gittenberger et al., 2006). Figure 13. The phylogenetic relationships of Vertigo marciae as reconstructed by maximum-likelihood analysis from the CytB amplicon. Nodes with strong to moderate support across all four phylogenetic reconstruction methods have been labeled to the left of that node by four support values: The upper left (normal font) is for nearest neighbor joining. The upper right (bold italic font) is for maximum parsimony. The lower left (bold font) is for Bayesian. The lower right (italic font) is for maximum likelihood. Page 112 THE NAUTILUS, Vol. 127, No. 3 37 Figure 14. The phylogenetic relationships of Vertigo marciae as reconstructed by maximum-likelihood analysis from the 16S amplicon. Nodes with strong to moderate support across all four phylogenetic reconstruction methods have been labeled to the left of that node by four support values: The upper left (normal font) is for nearest neighbor joining. The upper right (bold italic font) is for maximum parsimony. The lower left (bold font) is for Bayesian. The lower right (italic font) is for maximum likelihood. Figure 15. The phylogenetic relationships of Vertigo marciae as reconstructed by maximum-likelihood analysis from the concatenation of the ITS-1 and ITS-2 amplicons. Nodes with strong to moderate support across all four phylogenetic reconstruction methods have been labeled to the left of that node by four support values: The upper left (normal font) is for nearest neighbor joining. The upper right (bold italic font) is for maximum parsimony. The lower left (bold font) is for Bayesian. The lower right (italic font) is for maximum likelihood. J.C. Nekola and G. Rosenberg, 2013 Page 113 38 While resolution of deeper nodes within the genus Vertigo was not possible in the CytB and 16S mtDNA amplicons (Figures 13, 14), presumably due to basepair saturation, the concatenated ITS-1þ ITS-2 data (Figure 15) demonstrated a very highly supported node connecting the V. marciae / V. hebardi / V. alabamensis clade to V. oscariana. This radiation includes some of the most distinct members of the genus (Pilsbry, 1948). Additional field work across the Caribbean—especially in Cuba, Hispaniola, Puerto Rico, and Bermuda—will be required to determine the actual number of species contained within this group. ACKNOWLEDGMENTS John Slapcinsky of the Florida Museum of Natural History provided ethanol-preserved tissue samples of Vertigo hebardi, while Michal Horsak of Masaryk University in Brno, the Czech Republic, provided live samples for Vertigo pusilla and Pupilla muscorum. CytB and 16S sequences for V. pusilla, P. muscorum, and Gastrocopta cristata were retrieved from full mitochondrial genomic sequences provided by Jason Marquardt and Ulfar Bergthorsson of the University of New Mexico. Christopher Paul, University of Bristol, provided digital images of the Red Hills Road Cave fossils. Field work which established the persistence of Vertigo hebardi in the Florida Keys was supported by The Bailey-Matthews Shell Museum through an R.T. Abbott Visiting Curatorship to JCN. Additional funding for PCR and sequencing analysis was provided by Michal Horsak. The field work which resulted in the Recent samples of the new species being collected was supported by NSF Grant DEB-9870233 to Gary Rosenberg. LITERATURE CITED Bequaert, J.C. and W.B. Miller. 1973. The Mollusks of the Arid Southwest. University of Arizona, Tucson, xviþ271 pp. Gittenberger, E., D.S.J. Groenenberg, B. Kokshoorn and R.C. Preece. 2006. Molecular trails from hitch-hiking snails. Nature 439: 409. Grossman, D.H., S. Iremonger, and D.M. Muchoney. 1993. Jamaica: map of natural communities and modified vegetation types. Jamaica: a rapid ecological assessment. Phase 1: An island-wide characterization of mapping of natural communities and modified vegetation types. The Nature Conservancy, Washington, D.C., USA. Huelsenbeck, J.P. and F. Ronquist. 2001. MRBAYES: Bayesian inference of phylogeny. Bioinformatics 17: 754–755. Nekola, J.C. and B.F. Coles. 2010. Pupillid land snails of eastern North America. American Malacological Bulletin 28: 29–57. Nekola, J.C. 2009. Big ranges from small packages: North American vertiginids more widespread than thought. The Tentacle 17: 26–27. Nekola, J.C., B.F. Coles, and U. Bergthorsson. 2009. Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Molecular Phylogenetics and Evolution 53: 1010–1024. Paul, C.R.C. and S.K. Donovan. 2005. Quaternary and Recent land snails (Mollusca: Gastropoda) from Red Hills Road Cave, Jamaica Bulletin of the Mizunami Fossil Museum 32: 109–144. Pilsbry, H.A. 1919. American Species of Vertigo. Manual of Conchology. Second Series (Pulmonata) 25: 74–150, pls. 6–13. Pilsbry, H.A. 1948. Land Mollusca of North America (North of Mexico). Academy of Natural Sciences of Philadelphia Monographs 3, vol. 2(2): i–xlvii, 521–1113. Rosenberg, G. and I.V. Muratov. 2006. Status report on the terrestrial mollusca of Jamaica. Proceedings of the Academy of Natural Sciences of Philadelphia 155: 117–161. Tamura, K., D. Peterson, N. Peterson, G. Stecher, M. Nei, and S. Kumar. 2011. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution 28: 2731–2739. Page 114 THE NAUTILUS, Vol. 127, No. 3 39 40 41 42 43 44 45 46 47 48 49 50 51 Section II: Population and Community Ecology A vital component of understanding ecological systems is determining the factors controlling species abundance, range and community composition at local scales. I have conducted such work on vascular plants, lepidoptera, and terrestrial gastropods. The focus for many of these studies is determination of the local environmental factors that most strongly regulate occurrence and composition patterns, especially for regionally rare taxa at or near their geographic range limits. Often these studies have employed multivariate statistical techniques such as multidimensional scaling ordination and model-based cluster analysis. I have also been involved in determining the age structure of cliff-face tree populations (including millennium-aged red and white cedars) along the Niagaran Escarpment in the Great Lakes region. These types of studies not only help focus future biological surveys, but also help determine optimum conservation strategies. They also allow for large-scale comparisons of the factors influencing range limits by allowing for comparison of distribution and abundance drivers across a species distribution. Using my community biodiversity database which currently represents 1823 sites, ~600 species (½ of the continental fauna) and ¾ million individuals ranging from Prudhoe Bay to San Diego east to Labrador and Key West (all personally sampled using uniform protocols) I have begun to investigate continental αdiversity. Two unexpected patterns emerge: First, site richness does not monotonically increase towards the tropics, with a local maximum generally occurring between 35°-45° N. Thus, central Appalachian and Ozark sites tend to support more species than those to their immediate south. Second, a strong linear and almost order of magnitude increase is noted from west to east between 30°-40° N, with a similar though weaker (roughly 60%) increase from 40°-50° N. Because regional species pools are roughly similar in size, it is likely that these differences are rooted in contemporaneous ecological process rather than environmental and/or evolutionary history. The upper limit of α-diversity in fact does appear to be strongly influenced by current climate, in particular Aridity Index (AI), peaking at values around 1, which occurs when annual precipitation = annual potential evaportranspiration. This result can be related back to land snail physiology, which demands both water to make mucus required for movement and Ca for shell construction. At low AI, land snails will be stressed by droughtinduced inactivity, while at high AI they will be stressed by low [Ca] caused by leaching via excess precipitation. Given that land snail physiology is uniform across the biosphere, similar relationships must exist elsewhere in the globe. Initial analysis of European and New Zealand Alpha-diversity (community species richness) vs. Aridity Index (yearly precipitation/yearly potential evapotranspiration) for North American land snails. The upper limit was determined using quantile regression at tau=0.99. For the domain AI<0.85 the best fit two parameter model is a modified logit function; for AI>0.85 it is a power-law function. 52 land snail communities suggests that this pattern does represent a general global expectation. Additional work is also underway to investigate this relationship in east and southeast Asia, Australia, Africa, and the Neotropics. This work is being supported in part by grants from the National Park Service and Texas Parks and Wildlife. Representative Publications [number of citations as of October 27, 2017] Růžička, V., M. Zacharda, L. Němcová, P. Šmilauer &. J.C. Nekola. 2012. Periglacial microclimate in low altitude scree slopes supports relict biodiversity. Journal of Natural History. 46:2145-2157. [23] McClain, C.R., J.C. Nekola, L. Kuhnz, P.N. Slattery & J.P. Barry. 2011. Local-Scale Faunal Turnover on the Deep Pacific Seafloor. Marine Ecology Progress Series. 422:193-200. [15] Nekola, J.C. 2010. Acidophilic terrestrial gastropod communities of North America. Journal of Molluscan Studies. 76:144-156. [29] Stanisic, J, R.A.D. Cameron, B.M. Pokryszko & J.C. Nekola. 2007. Forest snail faunas from S.E. Queensland and N.E. New South Wales (Australia): Patterns of local and regional richness and differentiation. Malacologia. 49:445-462. [26] Nekola, J.C. 2004. Vascular plant composition gradients within and between Iowa fens. Journal of Vegetation Science. 15:771-780. [57] Nekola, J.C. 2003. Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Diversity and Distribution. 9:55-71. [52] Nekola, J.C. 2001. Distribution and ecology of Vertigo cristata (Sterki, 1919) in the western Great Lakes region. American Malacological Bulletin. 16:47-52. [8] Nekola, J.C. & P.A. Massart. 2001. Distribution and ecology of Vertigo nylanderi Sterki, 1909 in the western Great Lakes region. American Malacological Bulletin. 16:53-60. [5] Larson, D.W., U. Mattes-Sears, J.A. Gerrath, N.W.K. Larson, J.M. Gerrath, J.C. Nekola, G.L. Walker, S. Porembski & A. Charlton. 2000. Evidence for the widespread occurrence of ancient forests on cliffs. Journal of Biogeography. 27:319-331. [42] Nekola, J.C. 1999. Terrestrial gastropod richness of carbonate cliff and associated habitats in the Great Lakes region of North America. Malacologia. 41:231-252. [36] Nekola, J.C. & T.A. Smith. 1999. Terrestrial gastropod richness patterns in Wisconsin carbonate cliff communities. Malacologia. 41:253-269. [55] Larson, D.W., U. Mattes-Sears, J.A. Gerrath, J.M. Gerrath, J.C. Nekola, G.L. Walker, S. Porembski, A. Charlton & N.W.K. Larson. 1999. Ancient stunted forests on cliffs. Nature. 398:382-383. [40] 53 ACIDOPHILIC TERRESTRIAL GASTROPOD COMMUNITIES OF NORTH AMERICA JEFFREY C. NEKOLA Biology Department, University of New Mexico, Albuquerque, NM 87131, USA Correspondence: J.C. Nekola, e-mail: jnekola@unm.edu (Received 10 September 2009; accepted 19 October 2009) ABSTRACT Habitats with soil pH , 4 and Ca , 100 ppm, such as pocosins, Sphagnum bogs and heathlands, would appear inimical to land-snail biodiversity. Nevertheless, a survey of 1,356 sites, c. 1/2 million individuals and over 240 species (c. 1/5 of the continental fauna) across North America shows that c. 10% of species appear to favour such highly or moderately acidic sites, spread from subtropical forests of the Gulf of Mexico coast to the arctic tundra. Nonmetric multidimensional scaling ordination of faunas from 292 sites that support at least five co-occurring species documents that the principal axis of compositional variation in highly and moderately acidic habitats is significantly (P , 0.0005) correlated with latitude, while the second axis is significantly (P , 0.0005) correlated with moisture level. Composition was found to vary continuously along both axes, implying that discrete acidophilic communities are not present. While highly and moderately acidic sites were shown to have significantly (P , 0.000000005) lower richness and abundance compared with neutral/calcareous habitats, even the most acidic sites still typically supported 5–10 species. Abundance distributions in highly acidic habitats were found to be more uneven those of neutral/calcareous sites. The greater richness of the North American acidophilic land snail fauna compared with that in Europe has allowed communities to display replacement-driven compositional turnover. These results demonstrate that it is vitally important for biodiversity surveys of North American land snails not to ignore acidic habitats, because they harbour an important and surprisingly diverse fauna. INTRODUCTION It is well established that terrestrial gastropod communities in lime-rich habitats such as carbonate bedrock outcrops, fens, chalk grasslands and rich upland and lowland forests often support abundant and diverse land-snail communities (e.g. Boycott, 1934; Burch, 1962; Kerney & Cameron, 1979; Stanisic, 1997; Barker & Mayhill, 1999; Nekola, 1999; Schilthuizen, 2004; Horsa´ k, 2005). In fact, one of the principle global trends identified in land-snail ecology is the strong positive correlation between individual abundance, species richness and the pH of the organic litter in which land-snail communities reside (Valovirta, 1968; Wa¨ reborn, 1970; Walde´ n, 1981; Outeiro, Agu¨ era & Parejo, 1993; Barker & Mayhill, 1999; Nekola & Smith, 1999; Hermida, Ondina & Rodreiguez, 2000; Pokryszko & Cameron, 2005; Horsa´ k, 2006). This is not surprising given the high metabolic calcium demands of land snails not only for shell generation but also for egg production (Wa¨ reborn, 1970; Ga¨ rdenfors, 1992). Across the entire northwestern European fauna of about 300 species, only five (Columella aspera, Vertigo ronnebyensis, Vertigo lilljeborgi, Zonitoides excavatus and Zoogenetes harpa) have been noted to favour acidic habitats (Boycott, 1934; Kerney & Cameron, 1979). Highly acidic habitats, such as Sphagnum bogs, heathlands and pine forest have thus generally been regarded as unfavourable for high land-snail biodiversity (Karlin, 1961; Kerney & Cameron, 1979; Horsa´ k & Ha´ jek, 2003). No strongly acidophilic/calcifugic land-snail fauna had previously been reported from North America (Pilsbry, 1948; Hubricht, 1985) when Vertigo malleata was described and shown to be limited to ombotrophic peatlands of the Atlantic coastal plain (Coles & Nekola, 2007). These sites possess soils with organic litter pH ranging from 3.5 to 4.0 and Ca concentrations of only 25–100 ppm (Woodwell, 1958, Richardson, 1983). An additional 35 co-occurring species, representing about one-fifth of the entire North American Atlantic coastal plain fauna, were noted from these and adjacent upland pineflatwood sites, which have soil pH ranging from 3.5 to 4.5 and Ca ppm values of 20–500 (Binkley et al., 1992). Some of the species observed in these habitats, such as Triodopsis soelneri and Vertigo alabamensis, are thought to be among the continent’s rarest species (Hubricht, 1985; NatureServe, 2009). Does this apparent acidophilic land-snail fauna from the Atlantic seaboard represent an idiosyncratic outlier, or do similar faunas occur across North America? How many species should be considered acidophilic/calcifugic within the continental fauna? What major gradients in community composition can be identified, and with which environmental variables do these correlate? How does community richness and abundance vary between base-rich and base-poor habitats across the major land-snail habitat types? To provide a better understanding of the community ecology of acidophilic land-snail communities within North America these issues are addressed in the present paper. MATERIAL AND METHODS Study sites Land-snail faunas were analysed from 1,356 sites ranging from the Alaskan North Slope and southern California coast eastwards to central Quebec, and the Atlantic seaboard from Maine to Georgia (Fig. 1). Matrix vegetation varies from tundra and taiga in the north, to tallgrass prairie in the Midwest, to mixed deciduous-conifer forest in the southwest and to mixed boreal-hardwood, deciduous and evergreen subtropical forest in the east and south (Barbour & Billings, 1988). Using the principal land-snail community types identified by Nekola (2003), this sample includes 377 rock outcrops, Journal of Molluscan Studies (2010) 76: 144–156. Advance Access Publication: 17 January 2010 doi:10.1093/mollus/eyp053 # The Author 2010. Published by Oxford University Press on behalf of The Malacological Society of London, all rights reserved atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 54 357 upland forests, 286 lowland forests, 114 upland grasslands and 222 lowland grasslands. Field methods Documentation of terrestrial gastropod faunas from representative 100–1,000 m2 areas within sites was accomplished by hand collection of larger shells and litter sampling for smaller taxa. Particular focus was placed on the latter approach as it provides the most complete assessment of site faunas (Oggier, Zschokke & Baur, 1998; Cameron & Pokryszko, 2005). As suggested by Emberton, Pearce & Randalana (1996), collections were made at places of high micromollusc density such as loosely compacted leaf litter lying on top of highly compacted damp mineral soil or humus. Approximately 500 ml of sieved litter ranging from 0.425 to 9.5 mm in minimum particle size was retrieved from each site, either by field sieving or collection of c. 4 l of unsorted leaf litter. These unsorted bagged samples were slowly and completely dried with the c. 500 ml of sorted litter extracted from each sample in the laboratory by use of careful but vigorous water disaggregation. A half litre of sieved litter was set as the sample goal, as this volume typically captured 10 times the number of individuals as species, which is advocated by Cameron & Pokryszko (2005) for accurate documentation of land-snail community composition. The latitude–longitude position of each site was determined using either USGS 7.5 min topographic maps or a hand-held GPS. Each site was scored in one of the five principal land-snail habitat categories identified by Nekola (2003). Sites were also scored along a moisture gradient from 1 (dry vertical rock outcrops and xeric uplands) to 5 (completely saturated vegetation mats). Scores of 2 represented dry-mesic uplands, 3 represented mesic sites and 4 lowland sites with nearly saturated soils. The soil acidity of each site was assessed by observation of exposed bedrock, vascular plant and bryophyte cover. Bryophyte composition was particularly noted as this group serves as one of the best bio-indicators of soil-water solution acidity (Vitt & Chee, 1990). Sites were categorized as ‘highly acidic’ when exposed bedrock consisted of silica sand, shale, quartzite, schist, gneiss or felsic igneous rock, when vascular plant cover was dominated by pines and/or oaks, the shrub layer was limited to various heaths, Myrica and/or Ilex, and when bryophytes were limited to acidophilic genera such as Leucobryum, Polytrichum and Sphagnum (Crum & Anderson, 1981). Sites were categorized as ‘moderately acidic’ when exposed bedrock consisted of CaCO3-cemented sandstone or intermediate to intermediate-felsic igneous rock or when vascular plant and bryophyte cover included not only acidophilic taxa but also more base-demanding genera such as Climacium, Hypnum, Mnium, Rhodobryum, Rhytididelphus and Thuidium (Crum & Anderson, 1981). Sites were characterized as ‘neutral or calcareous’ when exposed bedrock consisted of carbonates or mafic/ultramafic igneous rock and/or when acidophilic vascular plants and bryophytes were absent. Laboratory procedures Retained leaf litter fractions were passed through a standard sieve series [ASTME 3/8" (9.5 mm), #10 (2.0 mm), #20 (0.85 mm) and #40 (0.425 mm) mesh screens], and then hand picked against a neutral-brown background. All shells and shell fragments were removed, and assigned to species (or subspecies) using the author’s reference collection. The total numbers of shells per species per site were recorded, as were the number of unassignable immature individuals and fragmentary shells. Nomenclature is based principally on Hubricht (1985), with updates by Turgeon et al. (1998), Nekola (2004) and Nekola & Coles (in press). Authors and dates for each species are given in Table 1. Statistical procedures Occurrences for all species encountered were analysed in terms of their presence/absence in highly acidic, moderately acidic, or neutral/calcareous soils. The species limited only to highly acidic and/or moderately acidic sites were identified, as well as those species limited to neutral/calcareous sites. Additionally, species found across the entire base-status spectrum were enumerated. The impact of base-status on population size was documented by first recording the observed population sizes for each species across all sites. These abundances were then grouped by their soil acidity category, with statistical difference in abundances between these groups being calculated using the nonparametric Kruskal–Wallis test. To provide appropriate statistical power to mitigate potential type-II errors, tests were limited to the 128 species with at least 10 total occurrences and 30 total individuals. A Bonferroni correction was used to modify the significance threshold of these tests to 0.05/128 or 0.00039. Gradients in composition of acidic habitat land-snail communities were identified via global nonmetric multidimensional scaling (NMDS) using DECODA (Minchin, 1990). NMDS was used as it makes no assumptions regarding the underlying nature of distributions of species along compositional gradients. As such, NMDS is the most robust form of ordination for detection of ecological patterns (Minchin, 1987). To ordinate sites, a matrix of dissimilarity coefficients was calculated between all pairwise combinations of highly or moderately acidic sites with five or more species, using the Czekanowski (Bray–Curtis) index (Faith, Minchin & Belbin, 1987) on species abundance data which had been doubly standardized, first to make all species maxima ¼ 1, and then to equalize the total number of individuals per site. All species (including the most rarely encountered) were considered. NMDS in one through four dimensions was then performed, with 200 Figure 1. Location of all 1,356 sampling sites. Black circles represent highly or moderately acidic habitats; open circles represent neutral/ calcareous stations. ACIDOPHILIC LAND-SNAIL COMMUNITIES 145 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 55 Table 1. Calcifuge/calcicole status of species, based on their presence or absence across the soil acidity spectrum. Occurring only in highly acidic sites: Daedalochila leporina (Gould, 1848) Glyphyalinia n. sp. Mesomphix perlaevis (Pilsbry, 1900) Neohelix divesta (Gould, 1851) Neohelix solemi Emberton, 1988 Triodopsis soelneri (J.B. Henderson, 1907) Vertigo oralis Sterki, 1898 Occurring only in moderately acidic sites: Lobosculum pustula (Fe´russac, 1822) Pristiloma arcticum (Lehnert, 1884) Occurring only in highly or moderately acidic sites: Glyphyalinia luticola Hubricht, 1966 Paravitrea petrophila (Bland, 1883) Triodopsis juxtidens (Pilsbry, 1894) Vertigo alabamensis Clapp, 1915 Vertigo malleata Coles & Nekola, 2007 Occurring in both acidic and neutral/calcareous habitats: Anguispira alternata (Say, 1817) Anguispira fergusoni (Bland, 1861) Carychium exiguum (Say, 1822) Carychium exile H.C. Lea, 1842 Carychium mexicanum Pilsbry, 1891 Catinella avara (Say, 1824) Cepaea hortensis (Mu¨ller, 1774) Cochlicopa lubrica (Mu¨ller, 1774) Cochlicopa lubricella (Porro, 1838) Cochlicopa morseana (Doherty, 1878) Columella columella alticola (Ingersoll, 1875) Columella simplex (Gould, 1841) Daedalochila dorfeuilliana (I. Lea, 1838) Deroceras leave (Mu¨ller, 1774) Discus catskillensis (Pilsbry, 1898) Discus cronkhitei (Newcomb, 1865) Discus patulus (Deshayes, 1830) Discus patulus edentulus (Pilsbry, 1948) Discus shimeki (Pilsbry, 1890) Euchemotrema fraternum (Say, 1824) Euchemotrema leai leai (A. Binney, 1840) Euconulus alderi (Gray, 1840) Euconulus chersinus (Say, 1821) Euconulus dentatus (Sterki, 1893) Euconulus fulvus (Mu¨ller, 1774) Euconulus fulvus alaskensis (Pilsbry, 1899) Euconulus polygyratus (Pilsbry, 1899) Euconulus trochulus (Reinhardt, 1883) Gastrocopta contracta (Say, 1822) Gastrocopta holzingeri (Sterki, 1889) Gastrocopta pentodon (Say, 1821) Gastrocopta riparia Pilsbry, 1948 Gastrocopta rupicola (Say, 1821) Gastrocopta tappaniana (C.B. Adams, 1842) Glyphyalinia indentata (Say, 1823) Glyphyalinia rhoadsi (Pilsbry, 1899) Glyphyalinia solida (H.B. Baker, 1930) Continued Table 1. Continued Glyphyalinia wheatleyi (Bland, 1883) Guppya sterkii (Dall, 1888) Haplotrema concavum (Say, 1821) Hawaiia minuscula (A. Binney, 1840) Helicodiscus inermis H.B. Baker, 1929 Helicodiscus notius Hubricht, 1962 Helicodiscus parallelus (Say, 1817) Helicodiscus shimeki Hubricht, 1962 Hendersonia occulta (Say, 1831) Inflectarius inflectus (Say, 1821) Mesodon elevatus (Say, 1821) Mesodon zaletus (A. Binney, 1837) Mesomphix capnodes (W.G. Binney, 1857) Mesomphix friabilis (W.G. Binney, 1857) Mesomphix globosus (MacMillan, 1940) Mesomphix subplanus (A. Binney, 1842) Neohelix albolabris (Say, 1816) Neohelix alleni (Wetherby in Sampson, 1883) Nesovitrea binneyana (Morse, 1864) Nesovitrea electrina (Gould, 1841) Oxyloma retusa (I. Lea, 1834) Paravitrea andrewsae (W.G. Binney, 1858) Paravitrea lamellidens (Pilsbry, 1898) Paravitrea multidentata (A. Binney, 1840) Paravitrea significans (Bland, 1866) Paravitrea simpsoni (Pilsbry, 1889) Patera binneyanus (Pilsbry, 1899) Patera perigraptus (Pilsbry, 1894) Planogyra asteriscus (Morse, 1857) Pomatiopsis lapidaria (Say, 1817) Punctum blandianum Pilsbry, 1900 Punctum californicum Pilsbry, 1898 Punctum minutissimum (I. Lea, 1841) Punctum n. sp. Punctum randolphi (Dall, 1895) Punctum vitreum H.B. Baker, 1930 Pupilla hebes (Ancey, 1881) Pupilla muscorum (Linne´, 1758) Pupisoma dioscoricola (C.B. Adams, 1845) Pupisoma macneilli (Clapp, 1918) Stenotrema barbatum (Clapp, 1904) Stenotrema labrosum (Bland, 1862) Stenotrema stenotrema (Pfeiffer, 1819) Stenotrema unciferum (Pilsbry, 1900) Striatura exigua (Stimpson, 1847) Striatura ferrea Morse, 1864 Striatura meridionalis (Pilsbry & Ferriss, 1906) Striatura milium (Morse, 1859) Strobilops aenea Pilsbry, 1926 Strobilops affinis Pilsbry, 1893 Strobilops labyrinthica (Say, 1817) Strobilops texasiana Pilsbry & Ferriss, 1906 Succinea ovalis Say, 1817 Succinea strigata Pfeiffer, 1855 Trichia hispida (Linne´, 1798) Triodopsis hopetonensis (Shuttelworth, 1852) Vallonia costata (Mu¨ller, 1774) Continued J. C. NEKOLA 146 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 56 Table 1. Continued Vallonia gracilicosta Reinhardt, 1883 Vallonia pulchella (Mu¨ller, 1774) Ventridens brittsi Pilsbry, 1892 Ventridens cerinoideus (Anthony, 1865) Ventridens intertextus (A. Binney, 1841) Ventridens pilsbryi Hubricht, 1964 Vertigo AK 1 Vertigo AK 2 Vertigo AK 3 Vertigo AK 5 Vertigo arthuri (von Martens, 1884) Vertigo bollesiana (Morse, 1865) Vertigo coloradensis (Cockerell, 1891) Vertigo columbiana Sterki, 1892 Vertigo cristata Sterki, 1919 Vertigo elatior Sterki, 1894 Vertigo aff. genesii (Gredler, 1865) Vertigo gouldii (A. Binney, 1843) Vertigo hannai Pilsbry, 1919 Vertigo hubrichti Pilsbry, 1934 Vertigo milium (Gould, 1840) Vertigo modesta hoppii (Mo¨ller, 1842) Vertigo modesta modesta (Say, 1824) Vertigo modesta ultima Pilsbry, 1948 Vertigo nylanderi Sterki, 1909 Vertigo oscariana (Sterki, 1890) Vertigo oughtoni Pilsbry, 1948 Vertigo ovata Say, 1822 Vertigo paradoxa Sterki, 1900 Vertigo perryi Sterki, 1905 Vertigo pygmaea (Draparnaud, 1801) Vertigo rugosula Sterki, 1890 Vertigo ventricosa (Morse, 1865) Vitrina alaskana Dall, 1905 Vitrina limpida Gould, 1850 Vitrinizonites latissimus (Lewis, 1875) Zonitoides arboreus (Say, 1816) Zonitoides nitidus (Mu¨ller, 1774) Zoogenetes harpa (Say, 1824) Occurring only in neutral/calcareous sites: Allogona profunda (Say, 1821) Anguispira kochi (Pfeiffer, 1845) Anguispira strongylodes (Pfeiffer, 1854) Angusipira jessica Kutchka, 1938 Appalachina sayana (Pilsbry, 1906) Carychium clappi Hubricht, 1959 Carychium nannodes Clapp, 1905 Catinella exile (Leonard, 1972) Catinella gelida (F.C. Baker, 1927) Catinella vermeta (Say, 1829) Catinella wandae (Webb, 1953) Cepaea nemoralis (Linne´, 1798) Daedalochila bisontes Coles & Walsh, 2006 Daedalochila lithica (Hubricht, 1961) Daedalochila peregrina (Rehder, 1932) Discus macclintockii (F.C. Baker, 1928) Continued Table 1. Continued Discus rotundatus (Mu¨ller, 1774) Euchemotrema hubrichti (Pilsbry, 1940) Euchemotrema leai aliciae (Pilsbry, 1893) Gastrocopta abbreviata (Sterki, 1909) Gastrocopta armifera (Say, 1821) Gastrocopta corticaria (Say, 1816) Gastrocopta cristata (Pilsbry & Vanatta, 1900) Gastrocopta pellucida (Pfeiffer, 1841) Gastrocopta procera (Gould, 1840) Gastrocopta rogersensis Nekola & Coles, 2001 Gastrocopta similis (Sterki, 1909) Gastrocopta sterkiana Pilsbry, 1912 Glyphyalinia caroliniensis (Cockerell, 1890) Glyphyalinia cumberlandiana (Clapp, 1919) Hawaiia n. sp. Helicina orbiculata (Say, 1818) Helicodiscus n. sp. Helicodiscus roundyi (Morrison, 1935) Helicodiscus singleyanus (Pilsbry, 1890) Inflectarius edentatus (Sampson, 1889) Inflectarius subpalliatus (Pilsbry, 1893) Lamellaxis gracilis (Hutton, 1834) Lobosculum pustuloides (Bland, 1851) Mesodon clausus clausus (Say, 1821) Mesodon normalis (Pilsbry, 1900) Mesodon thyroidus (Say, 1816) Mesomphix cupreus (Rafinesque, 1831) Mesomphix inornatus (Say, 1821) Mesomphix rugeli (W.G. Binney, 1879) Microphysula cookei (Pilsbry, 1922) Neohelix dentifera (A. Binney, 1837) Neohelix major (A. Binney, 1837) Oxychylus cellarius (Mu¨ller, 1774) Oxychylus draparnaudi (Beck, 1837) Oxyloma haydeni (W.G. Binney, 1858) Oxyloma peoriensis (Wolf, in Walker, 1892) Oxyloma verrilli Bland, 1865 Paravitrea AR 1 Paravitrea tridens Pilsbry, 1946 Patera clenchi (Rehder, 1932) Patera pennsylvanicus (Green, 1827) Pilsbryna quadrilamellata Slapcinsky & Coles, 2002 Polygyra cereolus (Muhlfeld, 1818) Polygyra semptemvolva Say, 1818 Pomatiopsis cincinnatiensis (I. Lea, 1840) Punctum conspectum (Bland, 1865) Punctum smithi Morrison, 1935 Pupoides albilabris (C.B. Adams, 1821) Rabdotus dealbatus (Say, 1821) Stenotrema altispira (Pilsbry, 1894) Stenotrema hirsutum (Say, 1817) Stenotrema pilsbryi (Ferriss, 1900) Succinea indiana Pilsbry, 1905 Succinea putris (Linne´, 1798) Trichia striolata (Pfeiffer, 1828) Triodopsis cragini Call, 1886 Triodopsis discoidea (Pilsbry, 1904) Continued ACIDOPHILIC LAND-SNAIL COMMUNITIES 147 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 57 maximum iterations, a stress ratio stopping value of 0.9999 and a small stress stopping value of 0.01. Output was scaled in halfchange units, so that an interpoint distance of 1.0 will correspond, on average, to a 50% turnover in species composition. Because a given NMDS run may locate a local (rather than the global) stress minimum, multiple NMDS runs were conducted on a given set of data from different initial random starting points to assess the stability of an individual solution (Minchin, 1987). In this analysis, DECODA used a total of 10 random starting configurations. Solutions in each of the four dimensions were compared using a Procrustes transformation to identify those that were statistically identical. The number of unique solutions, and number of runs which fell into each, was then calculated across each dimension. The modal solution out of 10 runs was identified, and was considered to be a global optimum when it was achieved in at least 50% of starts. Compositional variation across the diagram was documented by dividing it into regions whose boundaries were either 0.5 standard deviation above or below the ordination centroid along all major axes of variation. Within each of these regions, all taxa found within at least 25% of included sites were identified. Environmental correlations with this diagram were determined by calculating in DECODA the maximum correlation vector for site latitude and moisture level. The significance of each was estimated through Monte-Carlo simulations using 1,000 replications. Full two-way ANOVAs were used to test for differences in the central tendency of richness and log-transformed abundance across the three acidity levels and five major terrestrial gastropod habitat types. ANOVA was performed as richness and log-transformed abundance were essentially normally distributed across all three categories of habitat acidity. Even though log-transformation greatly alters the appearance of the species abundance distribution within sites, this technique is appropriate when comparisons of identically transformed data are made between sites (Nekola et al., 2008). The patterns among habitat types and acidity levels were visualized using box plots. Abundance distribution variation between highly acidic, moderately acidic and neutral/calcareous categories within each major habitat type was displayed using rank–frequency (dominance–diversity) plots. These were created by calculating the proportional abundance for each species within a given site. Abundances were then sorted from most to least abundant. The median proportional abundance for the most abundant species was then determined across all sites of a given base-status within that habitat type. This process was repeated for the second, third, fourth etc. to the least abundant species. These abundance values were plotted on a log-axis vs rank position. This process was repeated for all three base-status levels within a given habitat type. RESULTS Even though almost two-thirds of sites sampled in California and the Southwest occurred on highly felsic igneous outcrops, acidophilic vegetation was never noted, with vascular plants and bryophytes appearing essentially identical across the bedrock acidity spectrum. In addition, no differences were noted between the land-snail community compositions of these sites. This result is almost certainly due to the high rates of dryfall Ca inputs (Waring & Schlesinger, 1985) which have been estimated to range between 190 and 200 mg/m2 /year in southern California (Ellis, Verfaillie & Kummerow, 1983). Because of low leaching rates, at least 98% of soil Ca in the Southwest originates from atmospheric deposition (Capo & Chadwick, 1999), allowing ample base to be present in all soils, no matter what the parent material. For this reason, all 162 sites sampled in Arizona, California, Colorado and New Mexico have been eliminated from further analysis, because none actually represents acidophilic conditions, regardless of the underlying bedrock type. Calcifuge/calcicole status Of the 1,194 analysed sites outside of the Southwestern USA, land snails were observed from 1,188. Of these, 391 were observed to be highly or moderately acidic: 91 rock outcrops, 67 upland forests, 124 lowland forests, 22 upland grasslands and 87 lowland grasslands. In total, 241 species (c. 40% of the eastern and northern North American fauna and c. 20% of the entire North American fauna) and 486,153 identified individuals were encountered across these sites. Essentially all wideranging eastern and northern taxa were observed, with most of the unencountered species representing regional endemics, principally in the Polygyridae and Zonitidae (Hubricht, 1985). Seven species only occurred in highly acidic habitats, two only in moderately acidic habitats and a remaining five only in either moderately or highly acidic habitats (Table 1). While 93 species were restricted to neutral/calcareous habitats, the remaining 134 were found both in neutral/calcareous and moderately/highly acidic sites. Kruskal–Wallis analysis of abundances between highly acidic, moderately acidic and neutral/calcareous sites for the 128 species with at least 10 occurrences and 30 individuals (Table 2) documented 10 species with abundances that were significantly (P , 0.00039) greater in highly acidic sites, and an additional seven that significantly (P , 0.00039) favoured moderately acidic sites. One species tended (0.05 , P, 0.00039) to favour moderately acidic sites. A total of 34 species was shown to display no significant (P . 0.05) changes in abundance across the acidity spectrum. Of these, 23 had only between 10 and 30 total occurrences, making it probable that some of these nonsignificant test scores represent type-II errors. This is particularly likely for Discus macclintockii, Hawaiia n. sp. and Oxyloma verrilli, which had their twelve or fewer occurrences limited to neutral/calcareous habitats. Likewise, it seems Table 1. Continued Triodopsis neglecta (Pilsbry, 1899) Triodopsis obsoleta (Pilsbry, 1894) Triodopsis tridentata (Say, 1816) Triodopsis vulgata (Pilsbry, 1948) Vallonia excentrica Sterki, 1893 Vallonia parvula Sterki, 1892 Vallonia perspectiva Sterki, 1892 Ventridens acerra (Lewis, 1870) Ventridens ligera (Say, 1821) Vertigo AK 4 Vertigo binneyana Sterki, 1890 Vertigo clappi Brooks & Hunt, 1936 Vertigo meramecensis Van Devender, 1979 Vertigo morsei Sterki, 1894 Vertigo parvula Sterki, 1890 Vertigo tridentata Wolf, 1870 Webbhelix multilineata (Say, 1821) Xolotrema denotata (Fe´russac, 1821) Xolotrema fosteri (F.C. Baker, 1932) Zonitoides limatulus (W.G. Binney, 1840) J. C. NEKOLA 148 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 58 Table 2. Base-status preferences based upon Kruskal–Wallis analysis of site abundance. Species P-value Species significantly favouring highly acidic sites: Euconulus chersinus 0.00000039 Glyphyalinia luticola 0 Glyphyalinia n. sp. 0 Glyphyalinia solida 0 Striatura meridionalis 0.00000009 Strobilops texasiana 0 Ventridens cerinoideus 0.00000002 Vertigo alabamensis 0 Vertigo malleata 0 Vertigo modesta hoppii 0.00004532 Species significantly favouring moderately acidic sites: Planogyra asteriscus 0.00000001 Striatura exigua 0.00000002 Striatura ferrea 0.00002965 Vertigo cristata 0 Vertigo perryi 0 Vertigo ventricosa 0 Zoogenetes harpa 0 Species tending to favour moderately acidic sites Vertigo nylanderi 0.02831 Species demonstrating no preference in base status of habitat (30 occurrences): Cochlicopa morseana 0.1217 Euconulus fulvus alaskensis 0.9992 Glyphyalinia rhoadsi 0.2121 Pupilla hebes 0.08355 Striatura milium 0.06981 Succinea strigata 0.4538 Vertigo AK 2 0.906 Vertigo coloradensis 0.1796 Vertigo hannai 0.5831 Vertigo modesta modesta 0.1135 Vertigo ovata 0.3306 Species demonstrating no preference in base status of habitat (,30 occurrences): Discus macclintockii 0.08603 Euconulus dentatus 0.404 Euconulus trochulus 0.7056 Glyphyalinia wheatleyi 0.0512 Hawaiia n. sp. 0.08603 Helicodiscus inermis 0.0714 Helicodiscus notius 0.1385 Inflectarius inflectus 0.2089 Neohelix albolabris 0.0578 Oxychylus cellarius 0.08603 Oxyloma verrilli 0.05241 Paravitrea significans 0.1658 Pomatiopsis lapidaria 0.0666 Punctum blandianum 0.1177 Pupilla muscorum 0.3372 Strobilops aenea 0.508 Ventridens brittsi 0.9471 Vertigo AK 1 0.3169 Continued Table 2. Continued Species P-value Vertigo AK 3 0.1944 Vertigo AK 5 0.9488 Vertigo genesii 0.819 Vertigo modesta ultima 0.2797 Vertigo oscariana 0.2958 Species significantly favouring neutral/calcareous sites: Allogona profunda 0.00001154 Anguispira alternata 0 Carychium exiguum 0 Carychium exile 0 Catinella avara 0.00000003 Cochlicopa lubrica 0 Cochlicopa lubricella 0 Columella simplex 0 Deroceras laeve 0 Discus catskillensis 0.00000001 Discus cronkhitei 0.00000013 Euchemotrema fraternum 0 Euchemotrema leai leai 0 Euconulus alderi 0.0001001 Euconulus fulvus 0.00000592 Euconulus polygyratus 0 Gastrocopta armifera 0 Gastrocopta contracta 0 Gastrocopta corticaria 0 Gastrocopta holzingeri 0 Gastrocopta pentodon 0 Gastrocopta similis 0.00000013 Vallonia perspectiva 0 Vallonia pulchella 0.00000001 Vertigo arthuri 0.00000276 Vertigo bollesiana 0.00000102 Vertigo elatior 0 Vertigo gouldi 0 Gastrocopta tappaniana 0.00000005 Glyphyalinia indentata 0 Guppya sterkii 0.0001567 Haplotrema concavum 0.0001865 Hawaiia minuscula 0 Helicodiscus parallelus 0 Helicodiscus shimeki 0.0001685 Helicodiscus singleyanus 0.0001555 Hendersonia occulta 0 Nesovitrea binneyana 0 Nesovitrea electrina 0 Oxyloma retusa 0.00000399 Paravitrea multidentata 0.00000049 Punctum vitreum 0 Pupoides albilabris 0.00000021 Stenotrema barbatum 0.0001403 Strobilops affinis 0.00008082 Strobilops labyrinthica 0 Succinea ovalis 0.00008805 Vallonia costata 0 Vallonia gracilicosta 0 Vallonia parvula 0.0000716 Continued ACIDOPHILIC LAND-SNAIL COMMUNITIES 149 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 59 probable that with additional observations Punctum blandianum will be found strongly to favour acidic sites, as its population sizes are more than five times greater on highly acidic as compared to neutral/calcareous sites. The remaining 76 taxa either significantly (P , 0.00039) favoured (56 species) or tended (0.05 , P ,0.00039) to favour (20 species) neutral/calcareous sites. Compositional gradients Visual observation of the optimal NMDS result (the twodimensional solution) using all 305 highly or moderately acidic sites with at least five observed species indicated a group of 13 outliers along Axis 1 that represented tundra or near-tundra sites in Alaska and northern Manitoba. To assist gradient interpretation, these sites were assigned to a unique compositional group, and then removed from analysis. NMDS of the remaining 292 sites in one dimension generated a minimum stress configuration of 0.2603, which was achieved from nine of 10 random starting configurations. Procrustes analysis showed that these solutions fell into only a single group. NMDS along two dimensions generated a minimum stress configuration of 0.1934, which was achieved from all 10 random starting configurations. Procrustes analysis showed that these solutions fell into three unique categories, of which one was achieved in six different starts. NMDS along three dimensions generated a minimum stress configuration of 0.1489, which was achieved from six of the 10 random starting configurations. Procrustes analysis showed that these minimum stress solutions fell into two unique groups, each achieved in three different starting configurations. NMDS along four dimensions generated a minimum stress configuration of 0.1220, which was achieved from nine of 10 random starting configurations. Procrustes analysis showed that each of these solutions was unique. Based on these results the modal configuration from the twodimensional NMDS was chosen as the most stable for depicting land-snail community composition trends (Fig. 2). Approximately two complete turnovers in species composition are indicated along the first axis, with a single complete turnover being noted along the second. Environmental biplot analysis (Fig. 3) demonstrates that the first axis strongly (P , 0.0005) correlates with site latitude, with a maximum observed correlation of r ¼ 0.8366. The second axis strongly (P , 0.0005) correlates with water level, with a maximum observed correlation of r ¼ 0.6753. Ordination region one is found in the upper left of the diagram, and principally represents low latitude wetlands such as pocosins and bay-head swamps. Twelve species were found to occur in at least 25% of these sites (Table 3). Ordination region two is found at the mid-left side of the diagram, and principally represents low latitude mesic sites such as pine forest, bay forest and heathlands. Nine species were found to occur in at least 25% of these sites. Ordination region three is found on the lower left side of the diagram, and principally represents low latitude xeric sites such as dry bedrock outcrops, talus slopes and heath balds. Nine species were found to occur in at least 25% of these sites. Ordination region four is in the upper-mid portion of the diagram, and principally represents mid-latitude wetlands such as sedge meadows, and swamps of tamarack, northern white cedar, willow, red maple and alder. Twenty-three species were found to occur in at least 25% of these sites. Ordination region five is found in the centre of the diagram, and principally represents mid-latitude mesic sites such as pine, hemlock and yellow birch forests as well as some mesic bedrock outcrops. Fourteen species were found to occur in at least 25% of these sites. Ordination region six is on the bottom margin of the diagram, and principally represents mid-latitude xeric sites such as exposed bedrock outcrops and dry upland forest. Fourteen species were found to occur in at least 25% of these sites. Ordination region seven is on the upper right of the diagram, and principally represents high-latitude wetlands such as cattail swamps, sedge meadows and Sphagnum bogs. Ten species were found to occur in at least 25% of these sites. Ordination region eight is on the right-centre margin of the diagram, and principally represents high-latitude mesic sites such as heathlands, spruce-fir forest and some mesic rock outcrops. Eleven species were found to occur in at least 25% of these sites. Ordination region nine is found on the lower right of the diagram, and principally represents high-latitude xeric sites such as exposed bedrock outcrops. Nine species were found in at least 25% of these sites. Lastly, ordination region 10 represents the outlier sites found along the middle of Axis 2 to the right of the ordination diagram. These sites represent tundra sites across the full extent of the moisture gradient, ranging from sedge meadows to exposed rock outcrops. Ten species were found in at least 25% of these sites. Richness and abundance patterns ANOVA of richness and abundance patterns between the three soil acidity classes within the five major land-snail habitat types demonstrate that soil acidity imposed a highly significant (P , 0.000000005) negative impact on both site richness and log-transformed abundance (Table 4). Median richness levels ranged from 4.5 to 10 species per site in highly acidic soils, increased to 7–15 in moderately acidic soils, and were 11.5–20 in neutral/calcareous sites (Fig. 4). Median total abundance Table 2. Continued Species P-value Vertigo hubrichti 0 Vertigo milium 0 Vertigo paradoxa 0.00000027 Vertigo pygmaea 0.0001343 Vertigo tridentata 0 Zonitoides arboreus 0 Species tending to favour neutral/calcareous sites: Catinella exile 0.0005619 Catinella ’gelida’ 0.000435 Catinella ’vermeta’ 0.03188 Columella columella alticola 0.04251 Gastrocopta procera 0.002008 Gastrocopta rogersensis 0.00552 Mesodon clausus clausus 0.0007254 Mesodon thyroidus 0.004289 Neohelix alleni 0.002879 Oxyloma haydeni 0.02484 Punctum minutissimum 0.00181 Punctum n. sp. 0.007416 Triodopsis tridentata 0.0071 Webbhelix multilineata 0.00552 Vertigo meramecensis 0.001558 Vertigo morsei 0.002587 Vertigo oughtoni 0.02095 Vitrina limpida 0.001014 Xolotrema fosteri 0.04088 Zonitoides nitidus 0.001456 The significance threshold has been modified to P  0.00039 based on a Bonferroni correction. J. C. NEKOLA 150 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 60 ranged from 50.5 to 119.5 individuals in highly acidic soils to 57–292.5 in moderately acidic soils and 211–440 in neutral/ calcareous sites (Fig. 4). Habitat type also had a highly significant (P , 0.000000005) impact on site richness, but only a marginally significant (P ¼ 0.04) impact on log-transformed total abundance. Bedrock outcrop sites had the highest observed richness with a median score ranging from 10 to 20 species, followed by upland forests (8–15), lowland forests (6– 15), lowland grasslands (5–12) and upland grasslands (4.5– 11.5). While the interaction between habitat type and soil acidity significantly impacted site richness (P ¼ 0.001), no effect on log-transformed abundance was noted (P ¼ 0.69). The significant interaction appears due to the fact that richness of highly and moderately acidic bedrock outcrop sites was identical, as were the richness of moderately acidic and neutral/calcareous upland forests (Fig. 4). Rank–abundance plots across all five habitat types demonstrate a clear tendency for highly acidic sites to have more uneven abundance distributions than for neutral/calcareous sites (Fig. 5), with the commonest species in highly acidic habitats tending to have a higher proportional median abundance and with scores rapidly decreasing to much smaller values for a given rank. In most cases, curves from moderately acidic sites were intermediate between highly acidic and neutral/calcareous sites. Exceptions to this were noted in lowland forest and upland grasslands, in which the curves for moderately acidic and neutral/calcareous sites were largely coincident. DISCUSSION The North American acidophilic land-snail fauna These analyses clearly document that Vertigo malleata is not an idiosyncratic anomaly within the North American land-snail fauna. Almost two-thirds of all analysed northern and eastern North American species were found to occur at least occasionally in highly to moderately acidic habitats. Of these, 16 (Daedalochila leporina, Euconulus chersinus, Glyphyalinia luticola, Glyphyalinia n. sp., Glyphyalinia solida, Mesomphix perlaevis, Neohelix divesta, Neohelix solemi, Striatura meridionalis, Strobilops texasiana, Triodopsis soelneri, Ventridens cerinoideus, Vertigo alabamensis, V. malleata, Vertigo modesta hoppii and Vertigo oralis) only occurred or had statistically higher abundance in highly acidic sites. An additional 10 (Lobosculum pustula, Planogyra asteriscus, Pristiloma arcticum, Striatura exigua, Striatura ferrea, Vertigo cristata, Vertigo nylanderi, Vertigo perryi, Vertigo ventricosa and Zoogenetes harpa) only occurred or had statistically higher abundance in moderately acidic sites. Paravitrea petrophila and Triodopsis juxtidens were also only seen in moderately and highly acidic sites. Thus, about 12% of the sampled fauna exhibits clear signs of acidophilic/calcifugic affinities. Almost all of these taxa are wide-ranging, with only one (Triodopsis solneri) being a local endemic. While other presumed acidophiles (e.g. Catinella oklahomarum, Inflectarius downieanus, Mesodon normalis, Neohelix dentifera, Praticolella bakeri, Praticolella jejuna, Praticolella lawae and Praticolella mobiliana) have been reported previously (Hubricht, 1985), they have either not yet been encountered or have too few observed occurrences to draw firm conclusions regarding their preferred base status. The actual number of acidophilic/calcifugic land-snail species in North America is thus certainly higher than is reported here. It is therefore vitally important for North American land-snail biodiversity surveys not to ignore acidic habitats, because they harbour an important and surprisingly diverse fauna. Acidophilic tendencies are not limited to a few closely related groups, with the taxa listed above representing at least eight families (Euconulidae, Gastrodontidae, Oxychilidae, Polygyridae, Pristolomatidae, Valloniidae and Vertiginidae of Bouchet et al., 2005). With additional sampling, two more (Punctidae and Succineidae) will also likely be added to this list. Most of this fauna shares little phylogenetic overlap with the known European acid-tolerant fauna: only two of the above genera (Vertigo and Zoogenetes) possess acidophilic taxa in both the Old and New Worlds (Kerney & Cameron, 1979). Instead, most of the North American acidophilic fauna represents endemic families or genera, such as the Polygyridae (representing one-fifth of identified acidophilic species) and the gastrodontid genus Striatura (all species demonstrated acidophilic or generalist tendencies). Biogeographically, acidophilic species were found to range across the entire continent from semitropical forests of the Gulf Coast to arctic tundra on the Alaska North Slope. These factors all strongly suggest that the evolution of acidophilic traits has occurred in situ multiple times within North America. A few clear analogues between the European and North American acidophilic faunas can be observed, however. For instance, Vertigo ronnebyensis not only shares a very similar shell with V. cristata, but also an apparent preference for acidic taiga soils (Kerney & Cameron, 1979). Additionally, Vertigo lilljeborgi not only shares roughly similar shell characteristics with V. perryi, but similar affinities with tussock-side sedge litter accumulations in acidic wet meadows. The Holarctic Zoogenetes harpa also appears to demonstrate acidophilic tendencies across its range. It remains unknown how close the evolutionary relationships between these taxa and populations may be, given that at least some supposed Holarctic land-snail species can be shown to be members of only distantly related clades based on their mitochondrial DNA sequences (Nekola, Coles & Bergthorsson, in press). Figure 2. Optimum two-dimensional NMDS solution, showing locations of the 10 regions used to determine compositional changes. Figure 3. Environmental biplot diagram showing the vectors of maximum correlation for latitude and moisture levels. ACIDOPHILIC LAND-SNAIL COMMUNITIES 151 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 61 Table 3. Ordered matrix of species frequency within each of the 10 identified NMDS ordination regions, with the reported number representing the per cent of sites within that region which supported populations of that given species. Species Region 1 Region 2 Region 3 Region 4 Region 5 Region 6 Region 7 Region 8 Region 9 Region 10 Vertigo malleata 64.3 41.4 Strobilops texasiana 57.1 55.2 Punctum minutissimum 57.1 82.8 35.3 86.5 88.4 65.4 Gastrocopta pentodon 50.0 62.1 70.6 27.9 Vertigo milium 35.7 Striatura milium 35.7 34.5 78.9 95.4 84.6 Striatura meridionalis 35.7 48.3 76.4 Euconulus chersinus 35.7 Gastrocopta tappaniana 35.7 80.8 36.0 Glyphyalinia solida 35.7 Hawaiia minuscula 35.7 Gastrocopta contracta 28.6 Zonitoides arboreus 34.5 58.8 44.2 53.5 88.5 51.7 74.2 Vertigo alabamensis 31.0 Glyphyalinia indentata 27.6 47.1 Columella simplex 29.4 80.8 72.1 36.5 28.0 51.7 Haplotrema concavum 29.4 Helicodiscus parallelus 29.4 27.9 Guppya sterkii 29.4 Nesovitrea electrina 92.3 32.6 80.0 58.6 Striatura exigua 78.9 65.1 51.9 27.6 29.0 Carychium exiguum 71.2 Strobilops labyrinthica 67.3 81.4 42.3 Euconulus fulvus 67.3 79.1 36.5 48.0 93.1 71.0 Euconulus alderi 65.4 32.0 Vertigo cristata 59.6 74.4 67.3 72.4 80.7 Vertigo nylanderi 51.9 Vertigo ventricosa 50.0 28.0 Nesovitrea binneyana 50.0 79.1 71.2 32.0 55.2 74.2 Planogyra asteriscus 46.2 32.6 Striatura ferrea 34.6 41.9 27.6 Discus catskillensis 32.7 65.1 76.9 74.2 Vertigo elatior 32.7 Punctum n. sp. 30.8 Succinea ovalis 30.8 Helicodiscus shimeki 30.8 39.5 36.5 Carychium exile 26.9 Vertigo bollesiana 34.9 Zoogenetes harpa 44.2 48.3 38.7 Vertigo paradoxa 38.5 38.7 Anguispira alternata 26.9 Discus cronkhitei 40.0 51.7 38.7 Vertigo perryi 32.0 Vertigo modesta modesta 28.0 69.2 Vitrina limpida 27.6 Euconulus fulvus alaskensis 76.9 Vertigo modesta hoppii 76.9 Vertigo AK 2 61.5 Vertigo hannai 46.2 Succinea strigata 46.2 Columella columella alticola 38.5 Pupilla hebes 38.5 Vertigo modesta ultima 38.5 Vertigo genesii 30.8 Only species with a frequency of at least 25% within a region are listed. J. C. NEKOLA 152 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 62 Acidic land-snail habitats Acidophilic land-snail faunas were observed in a number of habitat types across the continent. In the south, some of the most diverse assemblages occurred in ombotrophic pocosin peatlands, bayhead seepage swamps, Atlantic white cedar forests, pine-wiregrass savanna, pine barrens, heath balds and upland oak woods. To the north, other important reservoirs for acidophilic land-snail diversity included Sphagnum bogs, swamps of base-poor cattail, sedge, northern white cedar, red maple, balsam fir and tamarack, wooded igneous rock outcrops and base-poor upland forests. Acidophilic land snails were also abundantly found across North America in base-poor taiga and acidic tundra over base-poor rock such as sandstone, quartzite and gneiss. It is astonishing not only that land-snail communities exist on such sites, but also that many of the included species possess large populations (up to 2,000/m2 ; Coles & Nekola, 2007) and strongly calcified shells (e.g. S. texasiana, V. alabamensis). Because of the very small Ca pools in such habitats (Binkley et al., 1992), it is clear that these populations must be successfully capturing Ca that is fluxing rapidly through the environment. Ca sources in these sites are undoubtedly various organically derived Ca salts, including Ca citrate and Ca oxalate (Wa¨ reborn, 1970), which are leached from various tree and shrub species such as aspen, dogwood or cedar (Karlin, 1961; Nation, 2007). Because of the high densities of individuals observed in their litter accumulations, it appears that heaths such as Chamaedaphne, Gaylussacia, Lyonia, Kalmia, Vaccinium and Zenobia, and other shrubs such as Ilex, Gordonia, Magnolia and Myrica, must also be important Ca sources. No matter the source, all acidophilic land snails must have evolved highly efficient Ca-pumps that are able to move ions counter to very strong osmotic gradients. If these pumps require low Ca concentrations or pH for proper functioning, it might provide a metabolic explanation for the calcifugic behaviour of some species. Because of the very high soil acidity at many of these sites (e.g. pH , 4; Woodwell, 1958; Binkley et al., 1992), the shells of many acidophilic snails are being actively dissolved even on living individuals. For instance, it is difficult to find living V. malleata or V. alabamensis individuals that do not possess at least one shell ulceration. Incipient ulcers can be seen in the individuals figured by Coles & Nekola (2007) as opaque white patches on otherwise translucent shells. Shell erosion is so rapid in these habitats that no adult shells of the seasonally active V. alabamensis could be found during autumn sampling, even on sites that supported colonies of .200 individuals/m2 in spring (Nekola & Coles, in press). Acidophilic land-snail community composition gradients The dispersion of sites across the optimal two-dimensional NMDS ordination diagram is continuous along both axes, implying that there are no hard compositional breaks within the fauna. This can be readily seen in the continuous variation exhibited in pineland and ombotrophic bog faunas along the Atlantic seaboard, with sites in the New Jersey Pine Barrens being intermediate between eastern Maine raised bogs and Carolina pocosins (Coles & Nekola, 2007). Similar transitions can be found between Atlantic coastal plain and mixed-deciduous forest faunas in central Massachusetts and southern Maine, between mixed-deciduous and taiga faunas along the Cote Nord in eastern Quebec and the western shore of Lake Superior, and between taiga and tundra faunas in the Alaskan interior, at Churchill, Manitoba, and in central Quebec. Analysis of environmental covariates identifies essentially the same organizing factors as have been previously shown in calcareous sites in the Great Lakes region (Nekola, 2003): Table 4. Summary statistics for ANOVA on the impacts of soil acidity and habitat type on community richness and log-transformed individual abundance. Predictor variable df Sum-of- squares Mean squares F-value P A. Site richness Habitat type 4 10,181 2,545 86.459 ,0.000000005 Soil acidity 2 11,602 5,801 197.064 ,0.000000005 Habitat type * soil acidity 1 302 302 10.260 0.001396 B. Log-transformed total abundance Habitat type 4 17.97 4.5 2.5178 0.03981 Soil acidity 2 451.26 225.6 126.4315 ,0.000000005 Habitat type * soil acidity 0.28 0.3 0.1596 0.68959 Figure 4. Box-plot diagrams demonstrating the impact of habitat type and soil acidity on site richness (top) and log-transformed site abundance (bottom). ACIDOPHILIC LAND-SNAIL COMMUNITIES 153 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 63 latitude was most strongly correlated with the first axis, while water level was strongly related to the second. As a result, it appears that at a continental scale acidophilic faunas respond in similar ways to the same environmental variables as do calciphilic faunas. Although found to be very important in base-rich communities, the impact of soil surface architecture was not documented in the current analysis, because thick organic litter accumulations (e.g. ‘duff soils’) were present on essentially all sites. However, acidic soils did tend towards turf-style architecture with increasing water levels. Diversity patterns While faunas of acidic habitats were both more depauperate in species and individuals than those of neutral/calcareous sites, median richness fell by no more than half, while abundance fell by no more than an order of magnitude. Median richness of even the most acidic sites ranged up to 10 species. This contrasts with patterns observed in Europe, where moderately acidic sites may often harbour as many species as typical base-rich sites (Bishop, 1977, 1980; Tattersfield, 1990; Cameron & Greenwood, 1991), and the most acidic sites typically have maximum diversity levels of ,10 (Kerney & Cameron, 1979). As in Europe (Cameron & Pokryszko, 2005), faunas were found to become increasingly uneven in their abundance distributions with increasing soil acidity. Acidic site faunas thus tend to be dominated by a relatively few common species, with the remainder being rare. This pattern is common among a large number of high-stress communities (Whittaker, 1975). The steepness of the rank–abundance curve on acidic sites also indicates that adequate documentation of site faunas may require extra field collection effort as compared to sites with higher base status (Cameron & Pokryszko, 2005). Acidophilic land-snail communities in a global perspective This documentation of acidophilic land-snail faunas is not without precedent. Bishop (1977), Tattersfield (1990) and Cameron & Greenwood (1991) have reported on land-snail community patterns from acidic sites in southwestern Ireland, the English Pennines and Scottish Highlands, respectively. Valovirta (1968), Wa¨ reborn (1970), Walde´ n (1981), Ga¨ rdenfors (1992) and Ma¨ nd, Ehlvest & Kristaja (2001) have documented acidophilic land-snail faunas in Scandinavia and the Baltic Republics. Ago´ csy (1968) and Sˇ tamol (1991) have documented acidophilic land-snail faunas from eastern Europe, while Bishop (1980) also demonstrated the existence of such faunas in Italy. The existence of a low-elevation acidophilic fauna from the North Island of New Zealand is also suggested by the work of Barker & Mayhill (1999). The species found on these sites can make up a not-inconsequential fraction of their respective regional faunas: Boycott (1934) reported roughly half of the British land-snail fauna to be ‘indifferent to lime’, occurring freely in both acidic and calcareous sites, while Bishop (1977) reported that about 40% of all Irish land snails occur in acidic woodlands. What makes the North American fauna outstanding, however, is the presence of so many clearly acidophilic/calcifugic species. While these make up an estimated 10% of the sampled northern and eastern North American fauna, Boycott (1934) could only identify one such species from Britain. Even with updating by Kerney & Cameron (1979), it is clear that Figure 5. Rank–abundance (a.k.a. dominance–diversity) plots for median proportional abundance across all highly acidic (dotted line), moderately acidic (dashed line) and neutral/calcareous (solid line) sites within a given major habitat type. J. C. NEKOLA 154 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 64 perhaps 1% or less of the northwestern European land-snail fauna can be considered acidophilic. This has profound implications in terms of composition gradients. First, the lack of acid-preferring species allows for considerable stochastic compositional variation between adjacent European acidic sites (Cameron & Pokryszko, 2005). In North America, however, the presence of acidophilic species allows for much greater consistency in site composition: the standard deviations of Axis 1 and Axis 2 scores for Carolina pocosins were only 0.36 and 0.25, respectively (implying c. 82-87% similarity), 0.26 along both Axis 1 and 2 for New England conifer swamps (c. 87% similarity), and 0.20 and 0.19 (c. 90% similarity) for western Lake Superior igneous outcrop sites. Acidic-habitat faunas throughout Europe also tend to be characterized by a suite of ubiquitous species which occur across the entire base-status gradient. The major compositional change with increasing base-status is the continuous addition of increasingly calciphilic species (Valovirta, 1968; Walde´ n, 1981; Cameron & Greenwood, 1991; Horsa´ k & Ha´ jek, 2003) until a few taxa drop out at the sites of highest base-status (Horsa´ k, 2006). Among European studies, only Ago´ csy (1968) did not recognize this pattern, instead noting that about 60% of acidic-habitat snails were restricted to these sites, whereas there was only 20% similarity between the faunas of adjacent acidic and calcareous habitats. European land-snail community compositions can thus be generally characterized as being nested along a base-status axis. Although also suggested from coastal British Columbia forests by Cameron (1986), this pattern appears not to be typical for the rest of North America. Rather, acidophilic faunas are made not only of species ‘indifferent’ to lime, but also a group of species (c. 20% of site faunas) that actually favour highly or moderately acidic sites. As a result, replacement-driven turnover in species occurrence can be seen along the base-status axis across much of North America. An important question that remains unanswered, however, is why North America and Europe differ so much in their frequency of acidophilic land-snail species. Given the apparently frequent presence of acidophiles in land-snail faunas of West Africa (De Winter & Gittenberger, 1998) and Borneo (Schilthuizen & Rutjes, 2001), it seems possible that a productive line of inquiry may be directed at determining why Europe has an unusually depauperate fauna, as opposed to concentrating on why that of North America appears strangely enriched. ACKNOWLEDGEMENTS Major funding for this project was provided by the National Science Foundation (EAR-0614963), the Maine Department of Inland Fisheries and Wildlife, Wildlife Resource Assessment Section, the Massachusetts Natural Heritage and Endangered Species Program, and the Minnesota Nongame Wildlife Tax Checkoff and Minnesota State Park Nature Store Sales through the Minnesota Department of Natural Resources Natural Heritage and Nongame Research Program. Brian Coles provided innumerable insights regarding field sampling and microhabitat preferences for many of these species, as well as thoughts regarding potentially analogous species in Europe and North America. Invaluable suggestions on earlier drafts were provided by Linda Fey, Robert Cameron and two anonymous reviewers. REFERENCES AGO´ CSY, P. 1968. Data to quantitative conditions in the mollusk faunas of two different substrates in central Hungary. Acta Zoologica Academiae Scientiarum Hungaricae, 14: 1–6. BARBOUR, M.G. & BILLINGS, W.D. 1988. North American terrestrial vegetation. Cambridge University Press, New York. BARKER, G.M. & MAYHILL, P.C. 1999. Patterns of diversity and habitat relationships in terrestrial mollusc communities of the Pukeamaru Ecological District, northeastern New Zealand. Journal of Biogeography, 26: 215–238. BINKLEY, D., RICHTER, D., DAVID, M.B. & CALDWELL, B. 1992. Soil chemistry in a Loblolly/Longleaf Pine forest with interval burning. Ecological Applications, 2: 157–164. BISHOP, M.J. 1977. The Mollusca of acid woodland in west Cork and Kerry. Proceedings of the Royal Irish Academy, Section B, 77: 227–244. BISHOP, M.J. 1980. The Mollusca of acid woodland in the Italian province of Novara. Journal of Conchology, 30: 181–188. BOUCHET, P., ROCROI, J.-P., FRY´ DA, J., HAUSDORF, B., PONDER, W., VALDE´ S, A´ . & WARE´ N, A. 2005. Classification and nomenclator of gastropod families. Malacologia, 47: 1–397. BOYCOTT, A.E. 1934. The habitats of land Mollusca in Britain. Journal of Ecology, 22: 1–38. BURCH, J.B. 1962. The Eastern land snails. W. C. Brown Co., Dubuque, Iowa. CAMERON, R.A.D. 1986. Environment and diversities of forest snail faunas from coastal British Columbia. Malacologia, 27: 341–355. CAMERON, R.A.D & GREENWOOD, J.J.D. 1991. Some montane and forest molluscan faunas from eastern Scotland: effects of altitude, disturbance and isolation. In: Proceedings of the Tenth International Molluscan Congress, Tu¨bingen, 1989 (C. Meier-Brook, ed.), pp. 437–442. Unitas Malacologica, Tu¨ bingen. CAMERON, R.A.D. & POKRYSZKO, B.M. 2005. Estimating the species richness and composition of land mollusc communities. Journal of Conchology, 38: 529–547. CAPO, R.C. & CHADWICK, O.A. 1999. Sources of strontium and calcium in desert soil and calcrete. Earth and Planetary Science Letters, 170: 61–72. COLES, B.F. & NEKOLA, J.C. 2007. Vertigo malleata, a new extreme calcifuge land snail from the Atlantic and Gulf coastal plains of the U.S.A. (Gastropoda, Vertiginidae). Nautilus, 121: 17–28. CRUM, H.A. & ANDERSON, L.E. 1981. Mosses of eastern North America. Columbia University Press, New York. DE WINTER, A.J. & GITTENBERGER, E. 1998. The terrestrial gastropod fauna of a square kilometer patch of rainforest in southwestern Cameroon: high species richness, low abundance, and seasonal fluctuations. Malacologia, 40: 231–250. ELLIS, B.A., VERFAILLIE, J.R. & KUMMEROW, J. 1983. Nutrient gain from wet and dry atmospheric deposition and rainfall acidity in southern California chaparral. Oecologica, 60: 118–121. EMBERTON, K.C., PEARCE, T.A. & RANDALANA, R. 1996. Quantitatively sampling land-snail species richness in Madagascan rainforests. Malacologia, 38: 203–212. FAITH, D.P., MINCHIN, P.R. & BELBIN, L. 1987. Compositional dissimilarity as a robust measure of ecological distance. Vegetatio, 69: 57–68. GA¨ RDENFORS, U. 1992. Effects of artificial liming on land snail populations. Journal of Applied Ecology, 29: 50–54. HERMIDA, J., ONDINA, M.P. & RODREIGUEZ, T. 2000. The relative importance of edaphic factors on the distribution of some terrestrial gastropod species: autecological and synecological approaches. Acta Zoologica Scientiarum Hungaricae, 46: 265–274. HORSA´ K, M. 2005. Molluscs. In: Ecology and palaeoecology of spring fens of the west Carpathians (A. Poulı´cˇ kova, M. Ha´ jek & K. Rybnı´cˇ ek, eds), pp. 197–208. Olomouc, Czech Republic. HORSA´ K, M. 2006. Mollusc community patterns and species response curves along a mineral richness gradient: a case study of fens. Journal of Biogeography, 33: 98–107. HORSA´ K, M. & HA´ JEK, M. 2003. Composition and species richness of molluscan communities in relation to vegetation and water chemistry in the western Carpathian spring fens: the poor-rich gradient. Journal of Molluscan Studies, 69: 349–357. HUBRICHT, L. 1985. The distributions of the native land mollusks of the eastern United States. Fieldiana, new series, 24: 1–191. ACIDOPHILIC LAND-SNAIL COMMUNITIES 155 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 65 KARLIN, E.J. 1961. Ecological relationships between vegetation and the distribution of land snails in Montana, Colorado and New Mexico. American Midland Naturalist, 65: 60–66. KERNEY, M.P. & CAMERON, R.A.D. 1979. Field guide to the land snails of the British Isles and northwestern Europe. Collins Press, London. MA¨ ND, R., EHLVEST, A. & KRISTAJA, P. 2001. Land snails in an afforested oil-shale mining area. Proceedings of the Estonian Academy of Science: Biology and Ecology, 50: 37–41. MINCHIN, P.R. 1987. An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio, 69: 89–108. MINCHIN, P.R. 1990. DECODA software. Australian National University, Canberra, Australia. NATION, T.H. 2007. The influence of flowering dogwood (Cornus florida) on land snail diversity in a southern mixed hardwood forest. American Midland Naturalist, 157: 137–148. NATURESERVE. 2009. NatureServe Explorer: an online encyclopedia of life. Version 7.0. NatureServe, Arlington, Virginia. http://www .natureserve.org/explorer. (accessed 5 September 2009). NEKOLA, J.C. 1999. Terrestrial gastropod richness of carbonate cliff and associated habitats in the Great Lakes region of North America. Malacologia, 41: 231–252. NEKOLA, J.C. 2003. Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Diversity and Distribution, 9: 55–71. NEKOLA, J.C. 2004. Terrestrial gastropod fauna of northeastern Wisconsin and the southern Upper Peninsula of Michigan. American Malacological Bulletin, 18: 21–44. NEKOLA, J.C. & COLES, B.F. In press. Pupillid land snails of eastern North America. American Malacological Bulletin. NEKOLA, J.C., COLES, B.F. & BERGTHORSSON, U. Evolutionary pattern and process in the Vertigo gouldii group of minute North American land snails. Molecular Phylogenetics and Evolution, 53: 1010–1024. NEKOLA, J.C., Sˇ IZLING, A.L., BOYER, A.G. & STORCH, D. 2008. Artifactions in the log-transformation of species abundance distributions. Folia Geobotanica, 43: 259–468. NEKOLA, J.C. & SMITH, T.A. 1999. Terrestrial gastropod richness patterns in Wisconsin carbonate cliff communities. Malacologia, 41: 253–269. OGGIER, P., ZSCHOKKE, S. & BAUR, B. 1998. A comparison of three methods for assessing the gastropod community in dry grasslands. Pedobiologia, 42: 348–357. OUTEIRO, A., AGU¨ ERA, D. & PAREJO, C. 1993. Use of ecological profiles and canonical correspondence analysis in a study of the relationship of terrestrial gastropods and environmental factors. Journal of Conchology, 34: 365–375. PILSBRY, H.A. 1948. Land Mollusca of North America (north of Mexico). Academy of Natural Sciences of Philadelphia Monograph vol. 1(1) and 2(2). POKRYSZKO, B.M. & CAMERON, R.A.D. 2005. Geographic variation in the composition and richness of forest land snail faunas in northern Europe. Records of the Western Australian Museum, Supplement, 68: 115–132. RICHARDSON, C.J. 1983. Pocosins: vanishing wastelands or valuable wetlands? Bioscience, 33: 626–633. SCHILTHUIZEN, M. 2004. Land snail conservation in Borneo: limestone outcrops act as arks. Journal of Conchology, Special Publication, 3: 149–154. SCHILTHUIZEN, M. & RUTJES, H.A. 2001. Land-snail diversity in a square kilometre of tropical forest in Sabah, Malaysian Borneo. Journal of Molluscan Studies, 67: 417–423. Sˇ TAMOL, V. 1991. Coenological study of snails (Mollusca: Gastropoda) in forest phytocoenoses of Medvednica mountain (NW Croatia, Yugoslavia). Vegetatio, 95: 33–54. STANISIC, J. 1997. An area of exceptional land snail diversity: the Macleay Valley, north-eastern New South Wales. Memoirs of the Museum of Victoria, 56: 441–448. TATTERSFIELD, P. 1990. Terrestrial mollusc faunas from some south Pennine woodlands. Journal of Conchology, 33: 355–374. TURGEON, D.D., QUINN, J.F., JR, BOGAN, A.E., COAN, E.V., HOCHBERG, F.G., LYONS, W.G., MIKKELSEN, P., NEVES, R.J., ROPER, C.F.E., ROSENBERG, G., ROTH, B., SCHELTEMA, A., THOMPSON, F.G., VECCHIONE, M. & WILLIAMS, J.D. 1998. Common and scientific names of aquatic invertebrates from the United States and Canada: mollusks. Edn. 2. American Fisheries Society Special Publication vol. 26. Bethesda, Maryland. VALOVIRTA, I. 1968. Land molluscs in relation to acidity on hyperite hills in central Finland. Annales Zoologici Fennici, 5: 245–253. VITT, D.H. & CHEE, W.L. 1990. The relationships of vegetation to surface water chemistry and peat chemistry in fens of Alberta, Canada. Vegetatio, 89: 87–106. WALDE´ N, H.W. 1981. Communities and diversity of land molluscs in Scandinavian woodlands. I. high diversity communities in taluses and boulder slopes in SW Sweden. Journal of Conchology, 30: 351–372. WA¨ REBORN, I. 1970. Environmental factors influencing the distribution of land molluscs of an oligotrophic area in southern Sweden. Oikos, 21: 285–291. WARING, R.H. & SCHLESINGER, W.H. 1985. Forest ecosystems: concepts and management. Academic Press, New York. WHITTAKER, R.H. 1975. Communities and ecosystems. MacMillan Publishing Co., New York. WOODWELL, G.M. 1958. Factors controlling growth of Pond Pine seedlings in organic soils of the Carolinas. Ecological Monographs, 28: 219–236. J. C. NEKOLA 156 atUniversityofNewMexicoonApril28,2010http://mollus.oxfordjournals.orgDownloadedfrom 66 - Vascular plant compositional gradients within and between Iowa fens - 771 Journal of Vegetation Science 15: 771-780, 2004 © IAVS; Opulus Press Uppsala. Abstract Question: What is the nature and relative importance of compositional gradients within- and between fens? Location: Iowa, USA. Methods: 506 0.5 m × 0.5 m quadrats were sampled from 31 fens across a 550 km extent. Presence/absence of all vascular plant taxa, plus the non-vascular genera Sphagnum and Chara, and values for 24 environmental variables were noted. Global Non-Metric Multidimensional Scaling and Monte Carlo tests were used to describe compositional variation and identify significant environmental co-variables. Model-based cluster analysis was used to identify the optimal number of groups supported by the data, while k-means clustering was used to assign each quadrat to a group. The number of occurrences (and frequency) of each species within each group was calculated. Two-dimensional 95% Gaussian confidence intervals, ANOVA, correlation coefficient homogeneity tests, log-linear modelling, and Fisher’s exact tests were used to document patterns of compositional change. Results: Two stable axes of variation were identified: the first being most closely correlated with soil pH, Mg, Ca, P, S, vegetation height, surface and –10 cm soil temperature, site area, perimeter, perimeter/area ratio, growing season, and air temperature, with the second being most correlated to soil moisture, N, disturbance level, % organic matter, hummock height, N-S coordinate, and precipitation. Individual sites harboured between 20-47% of total compositional variation, with 28% of Axis 1 and 55% of Axis 2 scores being contained within-sites. Five compositional regions were identified that differed in the proportion of calciphile and hydrophile species. Compositional groups differed significantly between geologic types. Conclusions: While the principal axis of variation (corresponding to the rich-poor fen gradient) is present largely between sites, the second axis (corresponding to water level) is largely repeated within sites. Documentation and protection of vegetation patterns and species diversity within Iowa fens will thus require consideration of multiple sites across the landscape. Keywords: Cluster analysis; Community ecology; Gradient analysis; Landscape pattern; Multidimensional scaling; North America; Peatland. Nomenclature: Nekola (1994). Vascular plant compositional gradients within and between Iowa fens Nekola, Jeffery C. Department of Natural and Applied Sciences; University of Wisconsin - Green Bay; Green Bay, Wisconsin 54311, USA; E-mail nekolaj@uwgb.edu Introduction To understand the full range of environmental factors influencing peatland vegetation, simultaneous documentation of both within- and between-site gradients must take place. Unfortunately, previous investigations have focused either on numerous (> 80) quadrats from individual sites (van der Valk 1975; Bernard et al. 1983; Stewart 1987; Singsaas 1989; Gerdol 1990; Glaser et al. 1990; Magnússon & Magnússon 1990; Mooney & O’Connell 1990; Wassen et al. 1990a; Stoynoff 1993; McCormac & Schneider 1994; Fojt & Harding 1995; Bowlesetal.1996;Choesin&Boerner2000;Gunnarsson et al. 2000; Choesin & Boerner 2002; Southall et al. 2003), or on few (< 10) quadrats sampled from multiple sites in the landscape (Charman 1993; Johnson & Leopold 1994; Motzkin 1994; Slack 1994; Bergamini et al. 2001; Vanderpuye et al. 2002; Stammel et al. 2003). The following study considers the relative importance of compositional gradients within- and between-site fens based on moderate numbers (10-30) of quadrats sampled from multiple sites (31) across Iowa. These data will be used to address the following questions: 1. What are the major compositional gradients? 2. What environmental factors underlie these trends? 3. How much compositional variation is captured within vs. between sites? Methods Study region Fens are peatland habitats with saturated (not inundated) soils whose source water has been enriched in nutrients by passage through the ground (Sjörs 1952). Fewer than 200 fens remain in the northeastern half of Iowa out of an original ca. 2400 pre-settlement sites (Nekola 1994). Extant examples are most common on the Iowan Erosional Surface and the northwestern margin of the Des Moines Lobe (Fig. 1). The Iowan Erosional Surface was formed through intense periglacial erosion during the Wisconsinan, while the Des Moines Lobe was formed ca. 14000-12000 yr ago during a final 67 772 Nekola, J.C. surge of Wisconsinan ice sheet following the onset of climatic warming (Prior 1991). These fens can be broadly classified into five geological types, depending upon aquifer type (Thompson et al. 1992): glacial till, bedrock, eolian sand, fluvial sand, and former lake basins (pothole or oxbow lakes). While statistical differences exist in peat chemistry between these groups (e.g. percent organic matter, pH, P, Ca, and Mg), no more than 35% of observed variance is accounted for by them (Nekola 1994). Study sites In total 31 fens from 23 counties across a 550 km extent (Fig. 1) were selected for analysis based on their natural integrity and geographical location. Investigation was limited to high-quality, undisturbed sites that maximized geographic spread across the landscape. While an attempt was made to maintain at least 40 km separation, some very high-quality sites were selected even though they were more closely positioned. Selected fens represent all five major geologic groups (22 till, 4 basin, 2 bedrock, 2 eolian, and 1 fluvial) and range in size from 0.1 - 28.8 ha (Table 1). Field methods A total of 506 0.5 m × 0.5 m quadrats were sampled in June and July, 1989. While this quadrat size is smaller than some previous North American investigations (i.e. Motzkin 1994), it is consistent with many other peatland vegetation studies (e.g. Bellamy 1967; Singaas 1989; Magnússon & Magnússon 1990; Stoynoff 1993; Bowles et al. 1996). This grain size is adequate for ordination analysis as it captures an average of 12 co-occurring species, or ca. 15% of an individual site flora. Sites were sampled at two intensity levels. One group (10 sites) was more intensively sampled at 30 quadrats per site. An effort was made to geographically stratify these across the landscape. At Mt. Auburn fen, only 26 quadrats were sampled due to weather conditions; 10 quadrats per site were recorded from the second group (21 sites). Only 7 and 5 quadrats were recorded from Rochester South and Rowley North, respectively, due to weather conditions. Each site was divided into five concentric zones of equal width positioned from the centre to edge. While zone widths differed between sites, their moisture levels were similar. Stratified-random sampling was used to position quadrats within each zone. When this sampling protocol missed major vegetation types or rare species, additional quadrats were non-randomly placed to capture them. Such subjective placement was limited to less than 10 quadrats (< 2% of total). All vascular plant species, plus Sphagnum and Chara spp., were noted from each quadrat. Surface and subsurface (–10 cm) soil temperature, vegetation height, and maximum hummock height were also recorded. Level of disturbance was noted using a subjective score ranging from 1 (least) to 3 (most). Site locations, perimeters and areas were calculated through digitization of USDA Soil Conservation Service county soil maps. Soil chemistry A 100-cm3 soil sample was taken from the centre of each quadrat. These were transported in water-tight plastic bags and not allowed to come in contact with sulfur-rich paper products. Samples were either oven dried within eight hours of collection, or frozen for later drying. Following the methodologies of Dahnke (1988), percent organic matter, pH, NO3-N, extractable P, exchangeable K, extractable SO4-S, exchangeable Ca, and exchangeable Mg were determined by Minnesota Valley Testing Laboratory of Nevada, Iowa. The percent water mass present in each wet sample was determined. However, because the water holding capacity of soils is positively correlated with organic matter content (Hudson 1994), many waterlogged but mineral-rich samples had lower water mass percentages than seemingly drier samples with higher organic matter levels. To help correct for this, soil moisture (SM) was calculated for each sample based on the following: SM = W * (1 – O) (1)Fig. 1. Landform regions of Iowa with location of sampled fen sites. Site letter codes correspond to those in Table 1. 68 - Vascular plant compositional gradients within and between Iowa fens - 773 Table 1. List and locations for sampled fen sites, including their size, geologic type, sampled quadrats, and percent compositional overlap with all other sites. Size Geologic Quadrats Site County Site name Location (ha) type sampled % overlap code Allamakee County Clear Creek 1 91° 25' 28" W., 43° 27' 30" N. 1.3 Fluvial 11 80.4 J Benton County Mt. Auburn 92° 5' 59" W., 42° 14' 50" N. 28.8 Till 26 22.3 R Black Hawk County Hammond Road 92° 20' 36" W., 42° 20' 48" N. 1.6 Till 10 23.8 P Bremer County Brayton-Horsley 92° 6' 27" W., 42° 48' 35" N. 4.7 Till 30 35.7 A Bushing 92° 27' 1" W., 42° 49' 48" N. 2.0 Till 12 43.5 I Buchanan County Cutshall Access 92° 4' 17" W, 42° 32' 47" N. 4.0 Basin 10 46.6 C Rowley 91° 51' 7" W., 42° 22' 27" N. 2.2 Till 30 22.9 D Rowley North 91° 51' 5" W., 42° 22' 36" N. 0.3 Till 5 70.1 Y Cedar County Rochester South 91° 8' 12" W., 41° 36' 59" N. 0.3 Eolian 7 23.4 X Cerro Gordo County Buffalo Slough 93° 11' 54" W., 43° 10' 36" N. 12.6 Basin 30 36.6 O Neuhring 93° 18' 39" W., 43° 13' 3" N. 1.0 Till 10 45.8 L Chickasaw County Chickasaw 1 92° 31' 3" W., 43° 0' 33" N. 18.4 Till 10 65.9 K Kleiss 92° 6' 20" W., 43° 0' 58" N. 1.6 Till 10 28.8 B Clay County Gillett 95° 0' 52" W., 43° 0' 53" N. 3.1 Till 10 48.6 % Clayton County Postville 91° 34' 0" W., 43° 2' 3" N. 37.5 Till 10 84.7 M Clinton County Toronto 2 90° 48' 9" W., 41° 53' 49" N. 1.3 Eolian 30 87.0 V Delaware County Hawker 91° 32' 38" W., 42° 30' 47" N. 7.3 Till 30 45.6 G Dickinson County Silver Lake 95° 21' 55" W., 43° 26' 16" N. 2.6 Till 12 84.4 # Lower Gar Lake 95° 6' 34" W., 43° 20' 48" N. 3.3 Till 10 61.7 $ Emmet County O’Brien 94° 50' 36" W., 43° 26' 16" N. 0.1 Till 10 31.5 * Fayette County Hunter Creek 91° 56' 50" W., 42° 39' 32" N. 0.9 Till 10 18.3 T Smithfield Township Hall 91° 47' 18" W., 42° 46' 57" N. 3.5 Till 30 21.2 Q Franklin County Maynes Creek 93° 1' 49" W., 42° 42' 24" N. 1.0 Bedrock 10 35.1 N Grundy County New Hartford 92° 38' 22" W., 42° 33' 16" N. 1.4 Till 10 34.7 S Howard County Staff Creek 92° 30' 34" W., 43° 26' 40" N. 2.4 Till 13 54.4 Z Linn County Matus 91° 29' 4" W., 42° 7' 57" N. 18.5 Till 30 55.9 E Western College 91° 37' 2" W., 41° 51' 56" N. 0.6 Till 10 31.9 F Mitchell County St. Ansgar 92° 55' 27" W, 43° 23' 21" N. 2.0 Bedrock 30 55.2 H Muscatine County Nichols 91° 16' 54" W., 41° 27' 24" N. 49.3 Basin 30 17.9 W Palo Alto County Dead Man’s Lake 93° 33' 59" W., 43° 14' 57" N. 1.2 Basin 10 11.3 @ Winneshiek County Jackson Juncton 92° 2' 52" W., 43° 7' 15" N. 1.4 Till 10 46.0 U where W = percent water mass of wet soil sample; and O = percent organic mass of dry soil sample. When soil organic content is low SM remains close in value to the observed percent water mass. However, as soil organic matter content increases, SM will fall accordingly. Meaningful comparisons between sites were made possible by severe drought conditions that allowed differences to reflect differential groundwater seepage rates as opposed to timing of rainfall events. Water chemistry was not measured for two principal reasons: First, as standing water was not present in all quadrats it could have been sampled only through the digging of pits that will cause significant alterations to water chemistry (Glaser et al. 1990). Second, soil chemistry is a much better correlate for vascular plant composition (Sjörs & Gunnarsson 2002) while water chemistry principally impacts bryophytes (Vitt & Chee 1990). Climate analysis Principal Components Analysis in conjunction with a varimax rotation was conducted on 30-yr mean values for 22 climate variables measured from 96 Iowa recording stations. Four major gradients, accounting for over 95% of the total climate variation, were identified: air temperature (correlated with Average Yearly, Summer, Fall, and Spring Temperatures, Heat Stress Days/Year, Heat Stress Degrees/Year, Heating Degree Days, Cooling Degree Days, Growing Degree Days), non-summer precipitation (correlated with Average Yearly, Winter, Spring and Fall Precipitation, Average Winter Temperature, Maximum 24-hr Precipitation), growing season (correlated with 32º, 30º, 28º, 26º, and 24º Growing Seasons, number of Freezing Days/Year), and average summer precipitation (Nekola 1994). PCA scores along each of these axes were determined for each recording station, with punctual kriging (Burgess & Webster 1980) being used to estimate site-specific values along each axis based upon the nearest 20 stations. Statistical procedures Ordination Species lists from each quadrat were subjected to global non-metric multidimensional scaling (NMDS) using the DECODA software program (Minchin 1990). 69 774 Nekola, J.C. NMDS was used as it is the most robust form of ordination for detection of ecological pattern (Minchin 1987). Using all species, a dissimilarity matrix was calculated using the Kulczynski index (Faith et al. 1987) for all pairwise combinations of sites. NMDS in one through four dimensions was then performed with 200 maximum iterations, a stress ratio stopping value of 0.999900, and a small stress stopping value of 0.010000. Output was scaled in half-change units. Because a given NMDS run may locate a local (rather than the global) stress minimum, multiple NMDS runs starting from different initial random points must be compared to determine solution stability (Minchin 1987). For this ordination, DECODA used a total of 20 random starting configurations. Solutions in each of the four dimensions were compared using Procrustes transformations to identify those that were statistically distinct. The number of unique solutions, and runs that fell into each was then calculated across each of the four dimensions. The modal solution out of twenty runs was identified and was considered stable when it was achieved in at least 25% of starts. Identification of compositional groups Clustering was performed on the selected ordination output rather than raw data as the former are less susceptible to sampling or other inadvertent errors (Equihua 1990). Model-based cluster analysis (Banfield & Raftery 1992) was used to identify the number of groups most supported by the data. A sum-of-squares model was employed as it generates spherical clusters that will be of maximal compositional similarity. The approximate weight of evidence for k clusters (AWEk) was calculated for k = 1 to n–1 clusters (where n = the total number of ordinated sites) via the S+ MCLUST algorithm (Anon. 1995); k-means iterative relocation (Hartigan 1975) was then used to assign each site to a cluster as it is also based on sum-of-squares criteria. Ordination interpretation The number of occurrences (and frequency) of each species within each k-means cluster was calculated. The 20 most frequent taxa, and rare taxa (see Nekola 1994) reaching modal frequency within each cluster, were then determined. Two-dimensional 95% Gaussian confidence intervals (Sokal & Rohlf 1981) were calculated for each site and used to estimate compositional overlap between sites. Quadrats from other sites were considered statistically similar in composition when they fell within a given site’s 95% confidence ellipse. Quantification of Axis 1 and Axis 2 stand score variation between sites was accomplished via ANOVA and represented using box plots. The significance of differences in the variation accounted for between sites was estimated using a correlation coefficient homogeneity test (Sokal & Rohlf 1981). The maximum correlation vectors for all environmental variables were calculated by DECODA and their significance estimated through Monte-Carlo tests. Variables were assigned into one of two groups: those that varied within sites (pH, NO3-N, P, K, SO4-S, Ca, Mg, % organic matter, vegetation height, hummock height, surface soil temperature, –10 cm soil temperature, disturbance level, soil moisture), and those that only varied between sites (N-S Coordinate, E-W Coordinate, Perimeter, Area, Perimeter/Area Ratio, NonSummer Precipitation, Summer Precipitation, Growing Season, Air Temperature). Significance of differences in frequency of each kmeans cluster between the five geologic fen types was estimated via log-linear modelling. As predicted values were sparse (< 5) in more than one-fifth of cells (Zar 1984), Fisher’s Exact test was used to identify those compositionalclusterssignificantlyfavouredwithineach geologic group. Because the total number of quadrats sampled from glacial till sites exceeded 200, log-linear modelling was used to estimate significance for this group. The significance threshold for these analyses were adjusted to p = 0.01 using a Bonferroni correction. Results Site ordination A total of 217 taxa (including Sphagnum and Chara) were identified (App. 1). NMDS demonstrated that in one dimension the most stable solution (minimum stress configuration = 0.3607) was achieved in 14 starts; in two dimensions the most stable solution (minimum stress configuration = 0.2467) was achieved in five starts; in three dimensions the most stable solution (minimum stress configuration = 0.1923) was achieved in two starts; and in four dimensions all starts provided different solutions (minimum stress configuration = 0.1594). The most stable two-dimensional solution was chosen for further analysis because it possessed a relatively low stress level while also being achieved in at least 25% of starts. Visual observation of this solution demonstrated essentially constant variation across both axes (Fig. 2). One major outlier (from Silver Lake Fen and the lone quadratlocatedinahypercalcarouswatertrack)ispresent approximately 0.75 axis units to the right of the general distribution. This microhabitat is quite rare in Iowa, being limited to this and two adjacent sites. 70 - Vascular plant compositional gradients within and between Iowa fens - 775 Compositional variation within and between sites Considerable compositional overlap was noted between sites, ranging from 11.3% of all other quadrats (Dead Man’s Lake) to 87% (Toronto 2), with a median of 43.5% (Table 1). These values are skewed slightly to the right, with the bulk ranging from 20 - 47%. ANOVA identified significant (p < 0.0005) partitioning of both Axis 1 and Axis 2 scores between sites. However, the amount of variance explained significantly (p < 0.000005) varies from 72% for Axis 1 to 45% for Axis 2 (Fig. 3). Identification and description of compositional clusters The maximum AWEk score (2426.9) was achieved at 73 compositional clusters. As this represents too many groups for generalization of compositional trends, AWEk scores from k = 1 to 15 were calculated, along with the percent increase in AWEk from k–1 to k clusters. After six clusters, the percent increase in AWEk fell below 10%, and decreased steadily to the 2.5% range by cluster 15. However, as the sixth k-means cluster contained only a single outlying quadrat (the Silver Lake fen water track), it was considered a statistical artifact with low inferential power. Each quadrat was thus assigned to one of five clusters. Cluster A (100 quadrats) is located on the lower left of the ordination diagram (Fig. 2) and is dominated by plants such as Campanula aparinoides, Impatiens capensis, Aster puniceus, Carex prairea, and Lycopus americanus that demand high moisture levels but can tolerate non-calcareous conditions (Table 2). In this cluster 14 rare Iowa fen species reach their modal frequencies (Table 3) including Aster junciformis, Berula erecta, Galium labradoricum, Menyanthes trifoliata, Salix pedicellaris and Solidago uliginosa. Cluster B (88 quadrats) is located in the upper left of the ordination diagram (Fig. 2) and is dominated by two ferns (Thelypteris palustris and Onoclea sensibilis) and other species such as Polygonatum sagittatum and Triadenum fraseri that can tolerate somewhat drier, non-calcareous conditions (Table 2). Four rare Iowa fen species/genera reach modal frequencies in this cluster (Table 3) including Drosera rotundifolia, Potentilla palustris, Rubus pubescens, Sphagnum spp. and Triadenum fraseri. Cluster C (85 quadrats) is located in the upper centre of the ordination diagram (Fig. 2) and is dominated by typical sedge meadow species such as Carex stricta, Aster puniceus, Thelypteris palustris, Pycnanthemum virginianum, and Solidago altissima that require rich, moderately wet soils (Table 2). Only three rare Iowa fen plants (Cypripedium parviflorum, Ophioglossum vulgatum, Viola pallens) reach modal frequencies in this cluster (Table 3). Cluster D (169 quadrats) is located to the lower right of Cluster C (Fig. 2). This cluster demarcates the most commonly encountered vegetation of Iowa fen mats, andisdominatedbyAsterpuniceus,Lycopusamericanus, Muhlenbergia glomerata, Viola nephrophylla and Lysimachia quadriflora (Table 2). Carex species are present but do not dominate, with the flora representing a wide mix of relatively low-statured forbs and graminoids. Nine rare Iowa fen plants reach modal frequencies in this cluster (Table 3) including such characteristicMidwesternUSAfentaxaasBetulapumila, Carex prairea, Cypripedium candidum, Salix candida, Solidago riddellii and Valeriana edulis. ClusterE(64quadrats)islocatedinthecentre-rightof the ordination diagram (Fig. 2) and is dominated by species requiring both relatively high moisture levels and calcareous conditions such as Muhlenbergia glomerata, Parnassia glauca, Eleocharis elliptica, Rhynchospora capillacea, and Lobelia kalmii (Table 2). In cluster E 24 rare Iowa fen plants reach their modal frequencies in this cluster (Table 3) including Carex sterilis, Eleocharis pauciflora, Juncus alpinus, Platanthera hyperborea, Rhynchospora capillacea, Scleria triglomerata, Spiranthes romanzoffiana, Triglochin maritimum, T. palustre and Utricularia minor. Fig. 2. NMDS ordination of 506 Iowa fen quadrats. Letters correspond to the compositional group that each quadrat was assigned via k-means clustering. 71 776 Nekola, J.C. Table 3. Rare taxa (as listed in Nekola 1994) reaching modal frequencies in each compositional cluster. Cluster A Angelica atropurpurea Aster junciformis Berula erecta Carex rostrata Cirsium muticum Eleocharis smallii Equisetum fluviatile Galium labradoricum Menyanthes trifoliata Mimulus glabratus Rumex orbiculatus Salix pedicellaris Solidago uliginosa Cluster B Drosera rotundifolia Potentilla palustris Rubus pubescens Sphagnum spp. Triadenum fraseri Cluster C Cypripedium parviflorum Ophioglossum vulgatum Viola pallens Cluster D Betula pumila Bromus ciliatus Carex prairea Cypripedium candidum Salix candida Salix × rubella Salix × clarkei Solidago riddellii Valeriana edulis Cluster E Carex granularis Carex sartwellii Carex tetanica Carex sterilis Chara spp. Eleocharis elliptica Eleocharis pauciflora Eriophorum angustifolium Gentiana crinita Gentiana procera Gerardia paupercula Juncus alpinus Juncus balticus Liparis loeselii Lobelia kalmii Muhlenbergia glomerata Parnassia glauca Platanthera hyperborea Rhynchospora capillacea Scleria verticillata Spiranthes romanzoffiana Triglochin maritimum Triglochin palustre Utricularia minor Table 4. Contingency table analysis showing the number of quadrats within each of the five compositional clusters represented in each of the geologic fen types. Log-linear test statistic represents significance across entire table. Cluster Geologic fen type Till Bedrock Fluvial Basin Eolian A 41 10 4 37 8 B 40 0 0 37 11 C 62 2 0 5 16 D 144 21 2 1 1 E 51 7 5 0 1 Significance within each geologic type < 0.00005 0.0017 0.2552 < 0.00005 0.0087 Log-linear model test statistic = 221.03; df = 16; p < 0.0000005. Analysis of environmental co-variables All environmental variables except K (p = 0.060) and Average Summer Precipitation (p = 0.067) demonstrated significant correlation with the ordination diagram. Maximumr2 valuesforsignificantvariablesranged from 0.0246-0.3204. For those varying within sites, pH, - 10 cm soil temperature, Mg, Ca, surface soil temperature, and S were most positively correlated with Axis 1 scores, while vegetation height and P were negatively correlated (Fig. 4). N, disturbance level, % organic matter, and hummock height were positively correlated with Axis 2 scores, while soil moisture was negatively correlated. Of the factors varying only between sites, site perimeter/ area ratio was positively correlated with Axis 1 scores, while site area, air temperature, and site perimeter were negatively correlated. E-W coordinate location was positively correlated with Axis 1 but negatively correlated with Axis 2 scores, while precipitation, growing season, ClusterA(n=100quadrats)ClusterB(n=88)ClusterC(n=85)ClusterD(n=169)ClusterE(n=64) SpeciesFreq.SpeciesFreq.SpeciesFreq.SpeciesFreq.SpeciesFreq. Campanulaaparinoides64.00%Thelypterispalustris93.18%Carexstricta82.35%Asterpuniceus90.30%Muhlenbergiaglomerata87.50% Impatienscapensis64.00%Impatienscapensis63.64%Asterpuniceus82.35%Lycopusamericanus88.48%Violanephrophylla54.69% Asterpuniceus61.00%Polygonumsagittatum61.36%Thelypterispalustris80.00%Muhlenbergiaglomerata81.82%Lycopusamericanus51.56% Carexprairea59.00%Asterpuniceus50.00%Pycnanthemumvirginianum51.77%Violanephrophylla78.79%Parnassiaglauca48.44% Lycopusamericanus48.00%Onocleasensibilis48.86%Solidagoaltissima50.59%Lysimachiaquadriflora72.73%Carexprairea45.31% Carexstricta41.00%Triadenumfraseri44.32%Lycopusamericanus45.88%Carexprairea66.67%Eleochariselliptica42.19% Pileafontana40.00%Carexstricta43.18%Campanulaaparinoides41.18%Carexstricta63.64%Lysimachiaquadriflora39.06% Muhlenbergiaglomerata39.00%Campanulaaparinoides42.05%Asterumbellatus36.47%Pycnanthemumvirginianum60.61%Scirpusvalidus35.94% Calthapalustris34.00%Eupatoriummaculatum31.82%Violanephrophylla35.29%Carexinterior44.24%Carexinterior32.81% Eupatoriummaculatum32.00%Typhalatifolia31.82%Helianthusgrosserratus34.12%Campanulaaparinoides40.00%Rhynchosporacapillacea32.81% Lysimachiathyrsiflora25.00%Lycopusamericanus26.14%Onocleasensibilis29.41%Thelypterispalustris38.18%Eupatoriumperfoliatum31.25% Cardaminebulbosa24.00%Scutellariagalericulata23.86%Muhlenbergiaglomerata23.53%Eleochariselliptica35.15%Lobeliakalmii29.69% Equisetumfluviatile24.00%Pileafontana21.59%Polygonumsagittatum22.35%Eupatoriummaculatum34.55%Eupatoriummaculatum28.12% Violanephrophylla23.00%Boehmeriacylindrica21.59%Eupatoriummaculatum18.82%Lythrumalatum30.91%Carexhystericina28.12% Galiumlabradoricum21.00%ImmatureMuhlenbergia20.45%Fragariavirginica17.65%Salixcandida27.88%Asterumbellatus26.56% Thelypterispalustris19.00%Carexlacustris18.18%Geumaleppicum17.65%Lobeliasiphilitica26.06%Carexstricta26.56% Epilobiumleptophyllum19.00%Asterumbellatus15.91%Polemoniumreptans16.47%Sphenopholisintermedia24.24%Juncusnodosus26.56% Leerziaoryzoides19.00%Carexlanuginosa12.50%Poapratensis15.29%Asterumbellatus20.61%Glyceriastriata25.00% Typhalatifolia18.00%Bidenscoronata12.50%Saxifragapensylvanica15.29%Calthapalustris19.39%Scirpusamericana23.44% Sphenopholisintermedia17.00%Leerziaoryzoides12.50%Carexinterior14.12%Parnassiaglauca19.39%Equisetumarvense20.31% Polygonumsagittatum17.00%Calthapalustris14.12%Scleriaverticillata20.31% Table 2. Most frequent taxa (n = 20 or 21) in each of the five compositional clusters. 72 - Vascular plant compositional gradients within and between Iowa fens - 777 and N-S coordinate location were negatively correlated with Axis 1 but positively correlated with Axis 2. The frequency of the five compositional clusters was found to significantly (p < 0.0000005) vary between the five geologic groups (Table 4). Glacial till sites were over-represented in Cluster D and underrepresented in Clusters A and B (p < 0.00005). Bedrock outcrop sites were over-represented in Cluster D and underrepresented in Clusters B and C (p = 0.0017). Basin sites were over-represented in Clusters A and B and underrepresented in Clusters C, D, and E (p < 0.00005). Eolian Sand sites were over-represented in Clusters B and C and underrepresented in Clusters D and E (p = 0.0087). No significant patterns (p = 0.2552) were noted from the lone Fluvial Sand site sampled. Discussion Major compositional gradients Theseanalysesdemonstratetwostrongcompositional gradients (3 and 2 half-change units) within Iowa fens. The principal axis varies from vegetation supporting few to many calciphiles, while the second axis varies from vegetation supporting few to many hydrophiles. The compositional clusters created to help interpret these patterns are not separated in ordination space, with quadrats from different clusters often being more similar to each other than they are to other quadrats within their own group. The variation along Axis 1 is consistent with the Fig. 3. Box plot of variation within each site along NMDS Axes 1 and 2. Site letter codes correspond to those in Table 1. Fig. 4. Environmental variable biplots for the ordination diagram. 73 778 Nekola, J.C. rich-poor fen gradient noted in most peatland landscapes, including Estonia (Masing 1982), Poland (Wassen et al. 1990a), the eastern Carpathians (Hájek 2002), Switzerland (Bergamini 2001), the southern Alps (Gerdol1990),northwesternEurope(Sjörs1952;Økland 1990; Sjörs & Gunnarsson 2002), Scotland (Charman 1993), boreal Canada (Vitt & Chee 1990), the northern USA (Glaser 1987; Glaser et al. 1981), northeastern USA (Motzkin 1994), and New Zealand (McQueen & Wilson 2000). Similarly, the wet-mesic gradient demonstrated along the Axis 2 has been previously documented from Illinois (Bowles et al. 1996), northern Minnesota (Glaser et al. 1990), Ohio (Stewart 1987), New York (Bernard et al. 1983), Scotland (Charman 1993), and Norway (Singsaas 1989) fens. From a conservation perspective, the most important microsites for rare plants were those that had the highest soil moisture levels. Fourteen rare Iowa plants reached their modal frequencies in Cluster A (moderate pH and cation levels), while 25 rare species reached modal frequencies in Cluster E (high pH and cation levels). This pattern is consistent with previous investigations of Minnesota (Almendinger & Leete 1998a) and Suffolk, England (Fojt & Harding 1995) fens that have documented a positive correlation between soil water levels and rare plant richness. Environmental co-variables As for many other fen systems (e.g. Gerdol 1990; Vitt & Chee 1990; Wassen et al. 1990a; Charman 1993; Motzkin1994;Fojt&Harding1995;McQueen&Wilson 2000; Bergamini et al. 2001; Sjörs & Gunnarsson 2002) the principal axis of compositional variation was significantly correlated with pH, Ca, and Mg. The strong correlationoftheprincipalaxiswithP,vegetationheight, and soil temperature is likely due to the limited solubility of P compounds in areas with high soil/water pH and Ca (Bedford et al. 1999). This causes reduced primary production, fertility, and plant height in Minnesota (Almendinger & Leete 1998a), English (Boyer & Wheeler 1989), Italian alpine (Gerdol 1990) and Dutch (Verhoeven et al. 1990) fens. With increased penetration of solar energy to the ground, areas with less vegetation growth will also have higher soil temperatures. As low vegetation mats support the highest plant diversity and number of rare species in Iowa (twice the number of rare species occur in Cluster E vs. A) and other fen systems (Johnson & Leopold 1994; Wheeler & Shaw 1991; Jensen & Meyer 2001), maintenance of low available P levels will be important to fen conservation. Previous research has also documented that soil moisture, N concentration, and disturbance level will influence fen vegetation. Water levels are one of the most commonly documented correlates with peatland vegetation gradients (Bernard et al. 1983; Stewart 1987; Økland 1989; Singsaas 1989; Glaser et al. 1990; Wassen et al. 1990a; Charman 1993; Bowles et al. 1996). Soil N has been shown to impact vegetation composition in Swiss montane (Bergamini et al. 2001) and Icelandic (Magnússon & Magnússon 1990) fens. Disturbance gradients also underlie some peatland vegetation patterns, with grazing being more intense on drier soils (Bowles et al. 1996). In Iceland these environmental factors are linked as fen sites subjected to higher grazing pressures alsoexhibitthehighestsoilNconcentrations(Magnússon & Magnússon 1990). Climate and geology may also significantly impact peatland ecology as increased precipitation and temperature lead to increased leaching rates and decreased soil Ca levels in boreal North America (Glaser 1992), Fennoscandia (Økland 1989), and Minnesota (Almendinger & Leete 1998a) fens. Similarly, in Iowa the frequency of calcareous microsites increase towards the northwest where precipitation and temperature are the lowest. A linkage between fen soil and aquifer chemistry has been previously demonstrated in Minnesota (Almendinger & Leete 1998b). Not surprisingly, Iowa sites originating from the most base-rich aquifers (e.g. bedrock, till) supported the richest fen vegetation. However, the linkage between site configuration (area, perimiter/area ratio) and vegetation has not been previously documented. While these could be related to edgeeffects, they are more likely related to differential landscape positions for large (basins) vs. small (sideslope) sites, increasing the contribution of cation-rich groundwater in the latter. Within- vs. between-site compositional gradients While most individual sites capture between 25- 50% of total observed compositional variation, this overlap was less evident along Axis 1 as compared to Axis 2. This result is to be expected as Axis 1 environmental co-variables are most strongly related to soil pH, Ca, Mg, and P, which will vary most between sites due to differences in aquifer chemistry (Almendinger & Leete 1998b), while Axis 2 scores are most strongly related to differences in soil water levels, which are largely repeated from the dry margins (lags) to wet centres of each site (Wassen et al. 1990b; Wassen & Barendregt 1992; Gerdol 1993). Thus, water levels vary more within vs. between New York fens (Slack 1994), while in nutrient levels vary more between vs. within Dutch fens (Verhoeven et al. 1990). This differential within- and between-site variation along both axes has important consequences to the study and conservation of Iowa fens. To observe (and protect) 74 - Vascular plant compositional gradients within and between Iowa fens - 779 theprincipalcompositionalgradientfrompoor-richsites, multiple locations within the landscape must be considered. When only single sites are investigated or conserved, gradients associated with water levels gain importance. It is thus not surprising that as previous quantitative vegetation analysis of Midwestern fens have been limited to single or few sites (e.g. van der Valk 1975; Stewart 1987; Wilcox et al. 1986; Stoynoff 1993; McCormac & Schneider 1994; Bowles et al. 1996; Choesin & Boerner 2000) the poor-rich fen gradient has not been previously reported from this landscape. How- ever,thesesingle-sitestudieshavegenerallydocumented the importance of water levels in determining vegetation pattern. Conservation prioritization simply based upon the number of rare species per site will likely lead to an over-representation of calcareous sites, as these areas support the greatest number of rare species. The rare species characteristic of poor and intermediate fens, though fewer in number, may thus be overlooked. Acknowledgements. Gretchen and the late John Brayton suggested that I study diversity patterns within eastern Iowa peat ‘bogs’, and are in large part responsible for my interest in these sites. Bob Moats played a critical role by introducing me to northwestern Iowa fens. Dick Baker, Willard Hawker, John Nehnevaj, Dean Roosa, Arlan Kirchman, Dennis Schlicht and Dave Wendling, provided me with locality information. Fred Nekola provided invaluable assistance in field sampling. Herb Wilson furnished soil survey maps for all counties within the study region. Richard Carlson of the Department of Agronomy at Iowa State University kindly provided climatic data for Iowa. Helge Bruelheide, Paddy Coker, Jim Graves, Robert Peet, Martin Wassen, Thomas Wentworth, Peter White and Susan Wiser made useful comments regarding earlier drafts. The substantial financial support needed for completion of soil analyses was provided by the McElroy Foundation of Waterloo, Iowa. Field work was also funded in part through a National Science Foundation Graduate Fellowship. References Anon. 1995. S-PLUS guide to statistical and mathematical analysis. Version 3.3. Statistical Sciences, Seattle, WA, US. Almendinger, J.E. & Leete, J.H. 1998a. Peat characteristics and groundwater geochemistry of calcareous fens in the Minnesota River basin, U.S.A. Biogeochemistry 43: 17- 41. Almendinger, J.E. & Leete, J.H. 1998b. Regional and local hydrology of calcareous fens in the Minnesota River basin, USA. Wetlands 18: 184-202. Banfield, J.D. & Raftery, A.E. 1992. Model-based Gaussian and non-Gaussian clustering. Biometrics 49: 803-822. Bedford, B.L, Walbridge, M.R. & Aldous, A. 1999. Patterns in nutrient availability and plant diversity of temperate North American wetlands. Ecology 80: 2151-2169. Bellamy, D.J. 1967. Ecological studies on some European mires. Ph.D. Dissertation, University of London. London, UK. Bergamini, A., Peintinger, M., Schmid, B. & Urmi, E. 2001. Effects of management and altitude on bryophyte species diversity and composition in montane calcareous fens. Flora 196: 180-193. Bernard, J.M., Sieschab, F.K. & Gauch, H.G. Jr. 1983. Gradient analysis of the Byron-Bergen swamp, a rich fen in western New York. Vegetatio 53: 85-91. Bowles, M., McBride, J., Stoynoff, N. & Johnson, K. 1996. Temporal changes in vegetation composition and structure in a fire-managed prairie fen. Nat. Areas J. 16(4): 275-288. Boyer, M.L.H. & Wheeler, B.D. 1989. Vegetation patterns in spring-fen calcareous fens: calcite precipitation and constraints on fertility. J. Ecol. 77: 597-609. Burgess, T.M. & Webster, R. 1980. Optimal interpolation and isarithmic mapping of soil properties. II. Block kriging. J. Soil Sci. 31: 333-341. Charman, D.J. 1993. Patterned fens in Scotland: evidence from vegetation and water chemistry. J. Veg. Sci. 4: 543- 552. Choesin, D.V. & Boerner, R.E.J. 2000. Vegetation and ground water alkalinity of Betsch Fen, a remnant preiglacial fen in south central Ohio. Castanea 65: 193-206. Choesin, D.V. & Boerner, R.E.J. 2002. Vegetation boundary detection: a comparison of two approaches applied to field data. Plant Ecol. 158: 85-96. Dahnke, W.C. 1988. Recommended chemical soil test procedures for the North-central Region. North Dakota Agricultural Experiment Station, Bull. No. 499, North Dakota State University, Fargo, ND, US. Equihua, M. 1990. Fuzzy clustering of ecological data. J. Ecol. 78: 519-534. Faith, D.P., Minchin, P.R. & Belbin, L. 1987. Compositional dissimilarity as a robust measure of ecological distance. Vegetatio 69: 57-68. Fojt, W. & Harding, M. 1995. Thirty years of change in the vegetation communities of three valley mires in Suffolk, England. J. Appl. Ecol. 32: 561-577. Gerdol, R. 1990. Vegetation patterns and nutrient status of two mixed mires in the southern Alps. J. Veg. Sci. 1: 663-668. Gerdol, R. 1993. The vegetation of wetlands in the southern Carnian alps (Italy). Gortania Atti del Museo Friuland di Storia Naturale 15: 67-107. Glaser, P.H. 1987. The ecology of patterned boreal peatlands of northern Minnesota: a community profile. USDI Fish and Wildlife Service, Biological Report 85 (7.14), Washington, DC, US. Glaser, P.H. 1992. Raised bogs in eastern North America – regional controls for species richness and floristic assemblages. J. Ecol. 80: 535-554. Glaser, P.H., Wheeler, G.A., Gorham, E. & Wright, H.E. 1981. The patterned mires of the Red Lake peatland, northern Minnesota: vegetation, water chemistry, and landforms. J. Ecol. 69: 575-599. Glaser, P.H., Janssens, J.A. & Siegel, D.I. 1990. The response of vegetation to chemical and hydrological gradients in 75 780 Nekola, J.C. the Lost River peatland, northern Minnesota. J. Ecol. 78: 1021-1048. Gunnarsson, U., Rydin, H. & Sjörs, H. 2000. Diversity and pH changes after 50 years on the boreal mire Skattlösbergs Stormosse, central Sweden. J. Veg. Sci. 11: 277-286. Hájek, M. 2002. The class Scheuchzerio-Caricetea fuscae in the western Carpathians: indirect gradient analysis, species groups and their relation to phytosociological classification. Biologia (Bratislava) 57: 461-469. Hartigan, J.A. 1975. Clustering algorithms. John Wiley Press, New York, NY, US. Hudson, B. 1994. Soil organic matter and available water capacity. J. Soil Water Cons. 49: 189-193. Jensen, K. & Meyer, C. 2001. Effects of light competition and litter on the performance of Viola palustris and on species composition and diversity of an abandoned fen meadow. Plant Ecol. 155: 169-181. Johnson, A.M. & Leopold, D.J. 1994. Vascular plant species richness and rarity across a minerotrophic gradient in the wetlands of St. Lawrence County, New York, USA. Biodivers. Conserv. 3: 606-627. Magnússon, B. & Magnússon, S.H. 1990. Studies in the grazing of a drained lowland fen in Iceland. I. The responses of the vegetation to livestock grazing. Búvísindi 4: 87-108. Masing, V. 1982. The plant cover of Estonian bogs: a structural analysis. in: Masing, V. (ed.) Peatland ecosystems, pp. 50-92. Academy of Sciences of the Estonian S.S.R.Valgus, Tallinn, EE. McCormac, J.S. & Schneider, G.J. 1994. Floristic diversity of a disturbed western Ohio fen. Rhodora 96: 327-353. McQueen, A.A.M. & Wilson, J.B. 2000. Vegetation and environment of a New Zealand raised bog. J. Veg. Sci. 11: 547- 554. Minchin, P.R. 1987. An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio 69: 89- 108. Minchin, P.R. 1990. DECODA software. Australian National University, Canberra, AU. Mooney, E.P. & O’Connell, M. 1990. The phytosociology and ecology of the aquatic and wetland plant communities of the lower Corrib basin, County Galway. Proc. R. Irish Acad. 90b: 57-97. Motzkin, G. 1994. Calcareous fens of western New England and adjacent New York state. Rhodora 96: 44-68. Nekola, J.C. 1994. The environment and vascular flora of northeastern Iowa fen communities. Rhodora 96: 121- 169. Økland, R.H. 1990. Regional variation in SE Fennoscandian mire vegetation. Nord. J. Bot. 10: 285-310. Prior, J.C. 1991. Landforms of Iowa. University of Iowa Press, Iowa City, IA, US. Singsaas, S. 1989. Classification and ordination of the mire vegetation of Stormyra near Tynset, S Norway. Nord. J. Bot. 9: 413-423. Sjörs, H. 1952. On the relation between vegetation and electrolytes in North Swedish mire waters. Oikos 2: 241-258. Sjörs, H. & Gunnarsson, U. 2002. Calcium and pH in north and central Swedish mires. J. Ecol. 90: 650-657. Slack, N.G. 1994. Can one tell the mire type from the Bryophytes alone? J. Hattori Bot. Lab. 75: 149-159. Sokal, R.R. & Rohlf, F.J. 1981. Biometry. W.H. Freeman and Co., San Francisco, CA, US. Southall, E.J., Dale, M.P. & Kent, M. 2003. Floristic variation and willow carr development within a southwest England wetland. Appl. Veg. Sci. 6: 63-72. Stammel, B., Kiehl, K. & Pfadenhauer, J. 2003. Alternative management on fens: Response of vegetation to grazing and mowing. Appl. Veg. Sci. 6: 245-254. Stewart, D.J. 1987. The vegetation and flora of Gott fen (Ohio), with a comparison of community classifications. M.Sc. Thesis, Miami University, Oxford, OH, US. Stoynoff, N.A. 1993. A quantitative analysis of the vegetation of Bluff Spring Fen Nature Preserve. Trans. Illinois Acad. Sci. 86: 93-110. Thompson, C.A., Bettis, E.A. III & Baker, R.G. 1992. Geology of Iowa fens. J. Iowa Acad. Sci. 99: 53-59. van der Valk, A.G. 1975. Floristic composition and structure of a fen in northwest Iowa. Proc. Iowa Acad. Sci. 82: 113- 118. Vanderpuye, A.W., Elvebakk, A. & Nilsen, L. 2002. Plant communities along environmental gradients of high-arctic mires in Sassendalen, Svalbard. J. Veg. Sci. 13: 875-884. Verhoeven, J.T.A., Maltby, E. & Schmitz, M.B. 1990. Nitrogen and Phosphorus mineralization in fens and bogs. J. Ecol. 78: 713-726. Vitt, D.H. & Chee, W.L. 1990. The relationships of vegetation to surface water chemistry and peat chemistry in fens of Alberta, Canada. Vegetatio 89: 87-106. Wassen, M.J. & Barendregt, A. 1992. Topographic position and water chemistry of fens in a Dutch river plain. J. Veg. Sci. 3: 447-456. Wassen, M.J., Barendregt, A., Palczynski, A., de Smidt, J.T. & de Mars, H. 1990a. The relationship between fen vegetation gradients, groundwater flow and flooding in an undrained valley mire at Biebrza, Poland. J. Ecol. 78: 1106-1122. Wassen, M.J., Barendregt, A., Schot, P.P. & Beltman, B. 1990b. Dependency of local mesotrophic fens on a regional groundwater flow system in a poldered river plain in the Netherlands. Landscape Ecol. 5: 21-38. Wheeler, B.D. & Shaw, S.C. 1991. Above-ground crop mass and species richness of the principal types of herbaceous rich-fen vegetation of lowland England and Wales. J. Ecol. 79: 285-301. Wilcox, D.A., Shedlock, R.J. & Hendrickson, W.H. 1986. Hydrology, water chemistry, and ecological relations in the raised mound of Cowles Bog. J. Ecol. 74: 103-117. Zar, J.H. 1984. Biostatistical analysis. Prentice-Hall, Inc., Englewood Cliffs, NJ, US. Received 17 November 2003; Accepted 16 August 2004. Co-ordinating Editor: H. Bruelheide. For App. 1, see JVS/AVS Electronic Archives; www.opuluspress.se/pub/archives/index.htm 76 © 2003 Blackwell Publishing Ltd. http://www.blackwellpublishing.com/journals/ddi 55 BIODIVERSITY RESEARCH Diversity and Distributions (2003) 9, 55–71 Blackwell Science, Ltd Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America JEFFREY C. NEKOLA Department of Natural and Applied Sciences, University of Wisconsin — Green Bay, Green Bay, Wisconsin 54311, U.S.A., E-mail: nekolaj@uwgb.edu Abstract: Previous ordination studies of land snail community composition have been limited to four or fewer habitat types from sites separated by no more than 300 km. To investigate the nature of large-scale patterns, North American land snail assemblages at 421 sites, representing 26 habitat types and covering a 1400 × 800 km area, were ordinated using global, nonmetric multi-dimensional scaling (NMDS). These data were then subjected to model-based cluster analysis and kmeans clustering to identify the main compositional groups and most important environmental covariables. Six primary compositional groups were identified. Three of these largely represent upland forest and rock outcrop sites, while the remaining largely represent either lowland forest, lowland grassland or upland grassland habitats. The geographical location and moisture level of sites also influences community composition. A strong compositional difference exists between sites having duff vs. turf soil surface layers. Only 8% of sites were improperly classified when soil surface architecture was used as the sole predictor variable. Fully 43% of taxa exhibited significant preferences towards one of these surface types, while only 15% of relatively common (10 + occurrence) taxa showed no preferences. Twelve groups of closely related taxa within the same genus had members that favoured different surface types, indicating that differential selection pressures have existed over evolutionary time scales. While turf faunas appeared unaffected by anthropogenic disturbance, duff faunas were strongly impacted, suggesting that their conservation will require protection of soil surface architecture. Key words. Cluster analysis, community ecology, conservation biology, landscape pattern, multidimensional scaling, North America, soil architecture, terrestrial gastropods. INTRODUCTION Land snails are regarded typically as generalist herbivores, fungivores and detritivores (Burch & Pearce, 1990) that exhibit weak levels of intraspecific competition (Cain, 1983; Cowie & Jones, 1987; Smallridge & Kirby, 1988; Barker & Mayhill, 1999). As land snail communities can consume less than 0.5% of annual litter input per year (Mason, 1970), some speculate that few resources, beyond CaCO3 (Boycott, 1934) and appropriate resting site availability (Pearce, 1997), will limit distribution. This concept is supported by high levels of microsympatry in land snail communities, where 13–35 (representing up to 50% of the regional fauna) co-occurring taxa can be found at < 1 m2 grains (Schmid, 1966; Cameron & Morgan-Huws, 1975; Nekola & Smith, 1999; Cameron, 2002). However, at landscape scales, land snail population size and faunistic composition has been suggested to vary with habitat and vegetation types (e.g. Burch, 1956; Wäreborn, 1970; Van Es & Boag, 1981; Young & Evans, 1991; Stamol, 1991; Stamol, 1993; Ports, 1996; Theler, 1997; Nekola, 2002). Habitat preferences for individual species have also been discussed (without supporting empirical data) at subcontinental scales 77 56 J. C. Nekola © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 in both western Europe (Kerney & Cameron, 1979) and eastern North America (Hubricht, 1985). Ordination techniques have documented significant species turnover along local environmental gradients in western Europe (e.g. Tattersfield, 1990; Magnin et al., 1995; Hermida et al., 2000; Ondina & Mato, 2001) and Pacific Island (e.g. Cowie et al., 1995; Barker & Mayhill, 1999) faunas. Unfortunately, these (and all other published snail ordination) studies have been conducted only at limited ecological (< four sampled habitat types) and geographical (maximum separation of no more than 300 km) scales. The following study addresses these concerns by using global nonmetric multi-dimensional scaling (NMDS) ordination and model-based cluster analysis to analyse land snail composition patterns within 26 habitat types across a 1400-km extent of central North America. These data will be used to address: (1) if land snail community composition predictably varies across this subcontinental region; and (2) what environmental and geographical factors underlie any such patterns. This study represents not only the first use of NMDS in the analysis of land snail communities, but also represents the first time North American faunas have been subjected to ordination. METHODS Study region Land snail faunas were sampled across a 1400 × 800 km area centred on the western portion of the Great Lakes basin in eastern and central North America (Fig. 1). This area covers a wide range of bedrock, climate and vegetation types. Both Palaeozoic sedimentary and Precambrian igneous bedrock outcrops in the region. One of the more prominent sedimentary exposures is the Niagaran Escarpment, a band of Silurian limestones and dolomites extending from western New York state to north-eastern Iowa. Outcrops along the western Lake Superior shore typically represent late-Precambrian mafic igneous rocks associated with the Keewenawan midcontinental rift system (Anderson, 1983). Average Fig. 1 Map of the study region showing the location of the 443 sample sites. 78 Great Lakes land snail community patterns 57 © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 minimum winter temperatures range from −25 °C in northern Minnesota to −10 °C in western New York State. Average maximum summer temperatures range from 25 °C along the western Lake Superior shore to 30 °C in south-eastern Iowa. The average length of the 0 °C growing season varies from 100 to 110 days in northern Minnesota and Michigan to 180–190 days in southern Iowa, southern Ontario, and western New York state. In areas adjacent to or downwind from (east of ) the Great Lakes (especially the Lower Peninsula of Michigan, southern Ontario, and western New York State), the climate tends to be buffered over that normally experienced in the continental interior, being warmer in the winter, cooler in the summer and having a longer growing season with more constant precipitation (Eichenlaub, 1979). Matrix vegetation varies from tallgrass prairie in the west to deciduous forest in the east to mixed boreal–hardwood forest in the north (Barbour & Billings, 1988). Study sites Four hundred and forty-three sites (Fig. 1) were surveyed across the range of habitats known to support diverse land snail assemblages (Nekola, 1999). The 26 habitats surveyed were broadly grouped into five major categories: rock outcrops, upland forests, lowland forests, upland grasslands and lowland grasslands (Table 1). While sites generally represent undisturbed examples of their respective habitats, an effort was also made to sample some (25 rock outcrop, 12 Table 1 Distribution of samples among surveyed habitat types. Habitat descriptions can be found in Nekola (1999) Group Habitat type Sites sampled Geographic range Rock outcrop Carbonate cliff 129 Illinois, Iowa, Minnesota, Michigan, Ontario, New York, Wisconsin Lakeshore carbonate ledge 23 Michigan, Ontario, Wisconsin Algific talus slope 27 Illinois, Iowa Igneous cliff 72 Michigan, Minnesota Sandstone/quartzite cliff 5 Wisconsin Shale cliff 3 New York Upland forest Oak–hickory forest 2 Wisconsin Maple–basswood forest 3 Wisconsin Hemlock–birch forest 1 Wisconsin Lakeshore forest 16 Michigan, Wisconsin Rocky woodland 26 Iowa, Michigan, Ontario, Wisconsin Lowland forest Floodplain forest 2 Wisconsin Black ash swamp 6 Wisconsin Tamarack swamp 33 Minnesota, Michigan, Ontario, Wisconsin White cedar swamp 16 Michigan, Ontario, Wisconsin Shrub-carr 2 Wisconsin Upland grassland Tallgrass prairie 1 Iowa Sand dune 1 Wisconsin Bedrock glade 13 Iowa Alvar 6 Michigan, Wisconsin Igneous shoreline 4 Michigan Successional old field 4 Wisconsin Lowland grassland Sedge meadow 5 Michigan, Wisconsin Fen 29 Iowa, Michigan, New York, Wisconsin Calcareous meadow 7 Michigan, Wisconsin Cobble beach 7 Michigan, Wisconsin 79 58 J. C. Nekola © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 upland forest, nine lowland forest, four upland grassland and eight lowland grassland) that had been disturbed anthropogenically by grazing, logging, recreational/urban development or bedrock/ soil removal. Examples of such sites include fieldedge stone piles, abandoned agricultural fields, abandoned building foundations, old quarries, pastures, road verges and exploited forests. Field methods Documentation of terrestrial gastropods from each site was accomplished by hand collection of larger shells and litter sampling for smaller taxa from representative 100–1000 m2 areas. Soil litter sampling was primarily used as it provides the most complete assessment of site faunas (Oggier et al., 1998). As suggested by Emberton et al. (1996), collections were made at places of high micromollusc density, with a constant volume of soil litter (approximately 4 L) being gathered from each site. For woodland sites, sampling was concentrated: (1) along the base of rocks or trees; (2) on soil covered bedrock ledges; and/or (3) at other places found to have an abundance of shells. For grassland sites, samples consisted of: (1) small blocks (c. 125 cm3 ) of turf; (2) loose soil and leaf litter accumulations under or adjacent to shrubs, cobbles, boulders and/or hummocks; and (3) other locations observed to have an abundance of shells. The latitude–longitude location of each sample was determined using either USGS 7.5 minute topographic maps or a hand-held GPS. To minimize bias from use of polar-coordinates, these locations were converted subsequently to Cartesian UTM Zone 16 coordinates using . The presence or absence of anthropogenic disturbance and soil surface architecture (duff vs. turf) was also recorded from each site. For purposes of this study ‘duff’ soils represent sites where the organic horizon was deep (> 4 cm) and subtended by a friable upper A horizon consisting largely of humus and mineral soil. ‘Turf’ soils represent sites where the organic horizon is thin (< 4 cm) and immediately subtended by an upper A horizon firmly bound together by living plant roots. While many habitats only supported a single soil architecture type (e.g. all carbonate cliffs were duff, and all bedrock glades were turf ), some (such as white cedar swamps) could possess either turf or duff surface layers, depending upon individual site conditions. Thus, habitat type could not be used as a surrogate for soil surface architecture. Laboratory procedures Samples were dried slowly and completely in either a low-temperature soil oven (c. 80–95 °C) or in a greenhouse. Dried samples were then soaked in water for 3–24 h, and subjected to careful but vigorous water disaggregation through a standard sieve series (ASTME 3/8′ (9.5 mm), #10 (2.0 mm), #20 (0.85), and #40 (0.425 mm) mesh screens). These fractions were then dried and passed again through the same sieve series, and hand-picked against a neutral-brown background. All shells and shell fragments were removed. All identifiable shells from each site were assigned to species (or subspecies) using the author’s reference collection and the Hubricht Collection at the Field Museum of Natural History (FMNH). Identification of some additional specimens representing Holarctic taxa more common in western Europe were verified by Robert Cameron of the University of Sheffield, UK. All specimens have been catalogued and are housed in the author’s reference collection at the University of Wisconsin — Green Bay. Nomenclature generally follows that of Hubricht (1985), with updates and corrections by Frest (1990, 1991) and Nekola (2002). Statistical procedures Ordination Species lists were determined for each sample. Sites with four or fewer taxa were excluded from further analysis, as such samples can bias results and obscure compositional trends. The remaining sites were subjected to global nonmetric multidimensional scaling (NMDS) using  (Minchin, 1990). NMDS was used as it makes no assumptions regarding the underlying nature of species distributions along compositional gradients. As such, NMDS is the most robust form of ordination for detection of ecological patterns (Minchin, 1987). To ordinate sites, a matrix of dissimilarity coefficients was calculated between all pairwise 80 Great Lakes land snail community patterns 59 © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 combinations of sites using the Czekanowski (Bray–Curtis) index (Faith et al., 1987). All species (including the most rarely encountered) were considered. NMDS in one to four dimensions was then performed, with 200 maximum iterations, a stress ratio stopping value of 0.9999, and a small stress stopping value of 0.01. Output was scaled in half-change units, so that an interpoint distance of 1.0 will correspond, on average, to a 50% turnover in species composition. Because a given NMDS run may locate a local (rather than the global) stress minimum, multiple NMDS runs must be conducted on a given set of data from different initial random starting points to assess the stability of an individual solution (Minchin, 1987). For this ordination,  used a total of 20 random starting configurations. Solutions in each of the four dimensions were compared using a Procrustes transformation to identify those that were statistically identical. The number of unique solutions, and number of runs which fell into each, was then calculated across each of the four dimensions (Minchin, 1990). The modal solution of 20 runs was identified, and was considered a global optimum when it was achieved in at least 50% of starts. Identification of compositional groups The chosen optimal NMDS solution was then subjected to model-based cluster analysis (Banfield & Raftery, 1992) to identify the number of compositional groups most supported by the data. Clustering was performed on the selected ordination output, rather than raw data, as ordination results are more robust and less susceptible to sampling or other inadvertent errors (Equihua, 1990). A sum-of-squares model was chosen, as it assumes that clusters will be spherical in ordination space, making them maximally compact and similar in composition. The approximate weight of evidence for k clusters (AWEk) was calculated via the S + MCLUST algorithm (Statistical Sciences, 1995) for k = 1 to n − 1 clusters (where n = the total number of ordinated sites). The larger the AWEk, the more evidence exists for that number of clusters. After the optimum number of clusters was determined, kmeans iterative relocation (Hartigan, 1975) was used to assign each site to a cluster. Kmeans clustering was chosen as it operates under the same sum-ofsquares criteria used for AWEk calculation. Ordination interpretation The number of occurrences (and frequency) of each species within each kmeans cluster was calculated. Species frequencies between clusters were compared using a Spearman’s rank correlation. The 10 most frequent taxa, taxa reaching modal frequency and species richness for each cluster were calculated. The frequency of the five major habitat groups between the compositional clusters was analysed using a contingency table. As predicted values were sparse (< 5) in more than one-fifth of cells, log-linear modelling was used to estimate significance (Zar, 1984). The maximum correlation vectors for the four recorded environmental variables (UTM E coordinate, UTM N coordinate, soil surface type, presence of anthropogenic disturbance) was calculated by . The significance of each was estimated through Monte-Carlo simulations using 1000 replications. Discriminant analysis was used to help further describe the impact of soil surface type and anthropogenic disturbance on site position in ordination space. Three tests were conducted: (1) effect of soil surface architecture (duff vs. turf) and the effect of anthropogenic disturbance separately for (2) duff and (3) turf soils. Lastly, the number of duff and turf sites containing and lacking each species was calculated. The significance of observed differences in these ratios between duff and turf sites was estimated using log-linear modelling. As this test was repeated for each species, a Bonferroni correction was used to adjust the significance threshold. This conservative adjustment was used so that only the most robust deviations would be used for data interpretation. RESULTS Site ordination One hundred and eight terrestrial gastropod taxa were identified from the 443 inventoried sites (Appendix I). Twenty-two sites were species poor (four or fewer taxa) and removed from further analysis. These included eight igneous cliffs, three lakeshore forests, three tamarack wetlands and single shale cliff, oak–hickory forest, maple– basswood forest, floodplain forest, sand dune, old field, sedge meadow and cobble beach sites. 81 60 J. C. Nekola © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 NMDS of the remaining 421 sites demonstrated that the only stable solution occurred along two axes of variation, where one was achieved in 50% of starts. The minimum stress configuration of this solution was 0.197412. In other dimensions, the most stable solution(s) were achieved in five (one dimension), three (three dimensions) and one (four dimensions) runs (Table 2). Identification and description of compositional clusters Visual observation of the chosen NMDS ordination solution demonstrated apparent natural clustering in faunal composition, with at least one well-defined group existing in the lowercentre of the diagram (Fig. 2). AWEk analysis demonstrated that the maximum score (2131.5) was achieved at the 53rd cluster. As this result provides too many groups to be useful for generalization of faunistic trends, AWEk scores from k = 1–10 were calculated, along with the percentage increase in AWEk from k to k + 1 clusters (Table 3). These data demonstrate that over 50% of maximum AWEk was achieved by the 6th cluster. The percentage increase in AWEk fell by almost 50% for cluster 7 (7.3%), and decreased steadily to the 3.4% range by cluster 10. Based on this, the optimal number of clusters was set at six (Fig. 3), even though it does not represent maximum AWEk. Contingency table analysis (Table 4) demonstrates that habitat representation significantly (P < 0.00005) varies between the six nonoverlapping kmeans clusters. Clusters A–C were equally (P = 0.3778) represented by rock outcrop and upland forest sites, while cluster D was dominated by lowland forests, cluster E by lowland grasslands and cluster F by upland grasslands. The 10 most frequent taxa also varied greatly, Table 2 NMDS summary statistics from an ordination of 421 sites with five or more taxa Dimensions Stress level Runs achieving minimum stress Unique solutions Number of runs in modal solution 1 0.339103 20 11 5 2 0.197412 20 6 10 3 0.147405 18 16 3 4 0.116912 18 18 1 Fig. 2 NMDS ordination of 421 sites with land snail richness of 5 or more, showing distribution of the five main habitat groups. Units are scales in halfchange units, such that a distance of 1 represents a 50% turnover in fauna composition. Table 3 Approximate weight of evidence for k clusters (AWEk) in land snail ordination based on a sum-of-squares model No. of clusters AWEk % Change from AWEk to AWEk−1 1 0 — 2 298.3 — 3 767.8 157.4 4 919.4 19.7 5 1097.8 19.4 6 1244.0 13.3 7 1335.2 7.3 8 1411.4 5.7 9 1456.7 3.2 10 1508.9 3.6 53 2131.5 — 82 Great Lakes land snail community patterns 61 © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 with approximately 50% turnover occurring between even the most similar groups (Table 5). Spearman’s rank correlations of species occurrence frequencies indicated that clusters D and E were most similar (0.831), while clusters A and F were the most different (0.316). Six species possessed modal occurrence frequencies in cluster A, 25 in cluster B, 32 in cluster C, five in cluster D, 20 in cluster E and 20 in cluster F (Table 6). Forty-one total taxa were encountered in cluster A, 78 in cluster B, 87 in cluster C, 58 in cluster D, 60 in cluster E and 60 in cluster F (Appendix I). Analysis of environmental co-variables Monte Carlo testing of the maximum correlation vectors for the four recorded environmental variables (Fig. 4) demonstrated that all were highly significant (P < 0.0005), having maximum rFig. 3 NMDS ordination of 421 sites with land snail richness of five or more. Letters represent each of the six compositional clusters assigned via Kmeans Clustering. Table 4 Contingency table analysis of main habitat groups vs. compositional clusters, with species richness within each cluster Compositional cluster Habitat group Total sites Species richness1 2 3 4 5 A 57 6 2 4 0 69 41 B 100 20 2 1 0 123 78 C 86 9 4 3 0 102 87 D 2 2 37 2 11 54 58 E 0 1 10 2 34 47 60 F 5 5 0 15 1 26 60 Comparison Log-likelihood ratio statistic d.f. P Entire table 451.898 20 < 0.00005 Clusters A,B,C 8.593 8 0.3778 Clusters D,E,F 112.053 8 < 0.00005 Fig. 4 Environmental biplot for NMDS ordination. The direction of each vector represents the angle of maximum correlation, while the length represents the strength of the correlation. 83 62J.C.Nekola ©2003BlackwellPublishingLtd,DiversityandDistributions,9,55–71 Table 5 Ten most frequent taxa in each of the six compositional clusters Rank order Cluster A Cluster B Cluster C Cluster D Cluster E Cluster F 1 Discus catskillensis (82.61%) Discus catskillensis (93.50%) Punctum vitreum (89.22%) Carychium exiguum (88.89%) Gastrocopta tappaniana (93.62%) Hawaiia minuscula (57.69%) 2 Nesovitrea binneyana (82.61%) Punctum minutissimum (91.06%) Gastrocopta contracta (88.24%) Striatura milium (83.33%) Carychium exiguum (85.11%) Vallonia costata (46.15%) 3 Zonitoides arboreus (82.61%) Zonitoides arboreus (90.24%) Carychium exile (88.24%) Nesovitrea electrina (81.48%) Nesovitrea electrina (85.11%) Cochlicopa lubrica (46.15%) 4 Vertigo cristata (72.46%) Strobilops labyrinthica (87.80%) Vertigo gouldi (87.25%) Strobilops labyrinthica (72.22%) Vertigo elatior (74.47%) Helicodiscus parallelus (46.15%) 5 Striatura milium (71.01%) Anguispira alternata (84.55%) Anguispira alternata (85.29%) Striatura exigua (72.22%) Euconulus alderi (70.21%) Gastrocopta contracta (42.31%) 6 Punctum minutissimum (62.32%) Vertigo gouldi (83.74%) Strobilops labyrinthica (79.41%) Zonitoides arboreus (68.52%) Hawaiia minuscula (59.57%) Gastrocopta similis (42.31%) 7 Zoogenetes harpa (53.62%) Columella simplex (83.74%) Gastrocopta holzingeri (78.43%) Vertigo elatior (59.26%) Gastrocopta contracta (57.45%) Punctum vitreum (42.31%) 8 Euconulus fulvus (49.28%) Striatura milium (73.98%) Hawaiia minuscula (78.43%) Punctum minutissimum (59.26%) Oxylama retusa (55.32%) Gastrocopta holzingeri (38.46%) 9 Vertigo paradoxa (49.28%) Helicodiscus shimeki (66.67%) Gastrocopta pentodon (74.51%) Gastrocopta tappaniana (53.70%) Stenotrema leai (55.32%) Vallonia pulchella (38.46%) 10 Striatura exigua (46.38%) Euconulus fulvus (65.85%) Zonitoides arboreus (74.51%) Euconulus alderi (48.15%) Deroceras spp. (55.32%) Vertigo pygmaea (38.46%) 84 GreatLakeslandsnailcommunitypatterns63 ©2003BlackwellPublishingLtd,DiversityandDistributions,9,55–71 Table 6 Taxa reaching modal frequencies in each compositional cluster Cluster A Cluster B Cluster C Cluster D Cluster E Cluster F Nesovitrea binneyana Carychium nannodes Allogona profunda Carychium exiguum Catinella avara Catinella ‘vermeta’ Vertigo cristata Cochlicopa morseana Anguispira alternata Planogyra asteriscus Catinella exile Cepaea nemoralis Vertigo modesta modesta Columella simplex Carychium exile Striatura exigua Cochlicopa lubricella Cochlicopa lubrica Vertigo modesta parietalis Discus catskillensis Catinella ‘gelida’ Striatura milium Discus cronkhitei Gastrocopta armifera Vertigo paradoxa Discus patulus Deroceras spp. Vertigo nylanderi Euconulus alderi Gastrocopta procera Zoogenetes harpa Euconulus fulvus Discus macclintockii Gastrocopta tappaniana Gastrocopta rogersensis Euconulus polygyratus Gastrocopta contracta Hawaiia n.sp. Gastrocopta similis Glyphyalinia rhoadsi Gastrocopta corticaria Helicodiscus n.sp. Glyphyalinia wheatleyi Helicodiscus shimeki Gastrocopta holzingeri Nesovitrea electrina Helicodiscus inermis Mesomphix cupreus Gastrocopta pentodon Oxyloma peoriensis Helicodiscus parallelus Mesomphix inornatus Glyphyalinia indentata Oxyloma retusa Helicodiscus singleyanus Oxychylus draparnaudi Guppya sterkii Pomatiopsis lapidaria Pupilla muscorum Paravitrea multidentata Haplotrema concavum Punctum n.sp. Pupoides albilabris Punctum minutissimum Hawaiia minuscula Stenotrema leai Succinea indiana Striatura ferrea Hendersonia occulta Strobilops affinis Vallonia costata Strobilops labyrinthica Mesodon clausus Triodopsis multilineata Vallonia excentrica Triodopsis albolabris Mesodon pennsylvanicus Vertigo elatior Vallonia parvula Triodopsis denotata Mesodon thyroidus Vertigo milium Vallonia pulchella Triodopsis tridentata Oxychylus cellarius Vertigo morsei Vertigo pygmaea Vallonia gracilicosta Punctum vitreum Vertigo ovata Zonitoides nitidus Vertigo n.sp. Stenotrema barbatum Vertigo bollesiana Stenotrema fraternum Vertigo hubrichti Strobilops aenea Vitrina limpida Succinea ovalis Zonitoides arboreus Trichia striolata Triodopsis alleni Triodopsis fosteri Vallonia perspectiva Vertigo gouldi Vertigo meramecensis Vertigo tridentata Zonitoides limatulus 85 64 J. C. Nekola © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 values ranging from 0.2111 to 0.7909 (Table 7). In duff sites, only UTM N coordinate and anthropogenic disturbance were found to correlate significantly (P < 0.0005) with the ordination diagram (Table 7). Northern sites tended to occur in the lower left of this group (maximum r = 0.8409), while disturbed sites tended to occur further to the right (maximum r = 0.4812; Fig. 5). In turf sites, only UTM N and E coordinates were found to correlate significantly (P < 0.0005) with the ordination diagram (Table 7). In this group, more northern (maximum r = 0.6690) and eastern (maximum r = 0.5067) sites tended to occur to the lower left (Fig. 5). Discriminant analysis demonstrated that the location of duff and turf sites in ordination space differs significantly (P < 0.0005) (Table 8), with duff sites being essentially limited to the upper left half of the diagram, and turf sites being found largely in the lower right (Fig. 6). The classification summary for this analysis indicates that only 33 of the 421 sites (7.8%) were classified improperly when soil surface type was used as the sole predictor variable. Discriminant analyses conducted separately on duff and turf sites demonstrated that disturbed duff sites were significantly (P < 0.0005) shifted to the right of undisturbed ones (Fig. 7). The Table 7 Two-dimensional correlation statistics for environmental variables in land snail ordination space Variable Maximum r Angle to first axis P Entire ordination UTM E-W coordinate 0.2111 129.4 < 0.0005 UTM N-S coordinate 0.7909 148.9 < 0.0005 Soil surface type 0.7843 52.1 < 0.0005 Disturbance presence 0.2829 1.7 < 0.0005 Duff soil sites only UTM E-W coordinate 0.1020 100.4 0.190 UTM N-S coordinate 0.8409 134.7 < 0.0005 Disturbance presence Turf soil sites only 0.4812 31.9 < 0.0005 UTM E-W coordinate 0.5067 150.4 < 0.0005 UTM N-S coordinate 0.6690 162.0 < 0.0005 Disturbance presence 0.0671 1.3 0.820 Fig. 5 Environmental biplots for duff and turf sites. The direction of each vector represents the angle of maximum correlation, while the length represents the strength of the correlation. 86 Great Lakes land snail community patterns 65 © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 classification summary for this analysis indicates that only 50 of the 295 duff sites (16.9%) were improperly classified when anthropogenic disturbance was used as the sole predictor variable. However, no significant differences (P = 0.758) were noted in the location of disturbed vs. undisturbed turf sites (Table 8). Faunistic turnover between duff and turf soils As differences in occurrence frequency between duff and turf sites were analysed for all 108 species, the significance threshold was lowered using a Bonferroni correction to P = 0.000463. Species with P-values ranging from 0.05 to 0.000463 were considered to possess statistically nonsignificant trends in their response to soil surface architecture. Species with P-values exceeding 0.05 were considered generalists. Thirty-six species demonstrated no significant differences in their occurrence frequencies between duff and turf soils (Appendix I). Sixteen of these (Cochlicopa lubricella, Deroceras spp., Discus cronkhitei, Gastrocopta armifera, Gastrocopta contracta, Haplotrema concavum, Hawaiia Table 8 Summary statistics for discriminant analysis of substrate and disturbance comparisons in land snail community ordination Factor Comparison Duff vs. turf (all sites) Disturbed vs. pristine (duff sites only) Disturbed vs. pristine (turf sites only) Canonical correlation 0.784 0.482 0.067 Eigenvalue 1.596 0.302 0.005 Likelihood ratio 0.385 0.768 0.996 Approximate F 322.74 43.915 0.278 Number d.f. 2 2 2 Density d.f. 417 291 123 P < 0.0005 < 0.0005 0.758 Fig. 7 Location of disturbed and undisturbed sites in the ordination diagram for duff and turf sites. Fig. 6 Location of duff and turf sites within the ordination diagram. 87 66 J. C. Nekola © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 minuscula, Helicodiscus parallelus, Helicodiscus singleyanus, Planogyra asteriscus, Striatura ferrea, Striatura milium, Triodopsis multilineata, Vallonia costata, Vertigo pygmaea, Vitrina limpida) were found in 10 or more sites, and clearly represent generalists. However, the remaining 20 (Carychium nannodes, Cepaea nemoralis, Discus patulus, Glyphyalinia wheatleyi, Helicodiscus inermis, Mesodon pennsylvanicus, Mesomphix cupreus, Mesomphix inornatus, Oxychylus cellarius, Oxychylus draparnaudi, Oxyloma peoriensis, Pomatiopsis lapidaria, Pupilla muscorum, Succinea indiana, Trichia striolata, Triodopsis denotata, Triodopsis fosteri, Vallonia excentrica, Vertigo modesta parietalis, Zonitoides limatulus) are known from fewer locations, making Type 2 errors a significant concern. Additional data will be needed to adequately assess the response of these species to duff vs. turf soils. Eighteen species (Cochlicopa lubrica, Cochlicopa morseana, Discus macclintockii, Mesodon clausus clausus, Mesodon thyroidus, Punctum minutissimum, Punctum vitreum, Stenotrema barbatum, Strobilops aenea, Strobilops labyrinthica, Succinea ovalis, Triodopsis albolabris, Triodopsis alleni, Triodopsis tridentata, Vallonia perspectiva, Vertigo meramecensis, Vertigo modesta modesta, Vertigo tridentata) nonsignificantly favoured duff sites. Another eight (Gastrocopta procera, Gastrocopta rogersensis, Gastrocopta similis, Hawaiia n.sp., Helicodiscus n.sp., Striatura exigua, Vallonia pulchella, Zonitoides nitidus) nonsignificantly favoured turf sites. While a number of these (e.g. Discus macclintockii, Gastrocopta procera, Gastrocopta rogersensis, Vertigo meramecensis) demonstrated very strong absolute preferences, their few total occurrences in combination with the conservative Bonferroni correction prevented them from exhibiting significant responses. The remaining 46 species demonstrated clear and significant soil surface preferences. Twentyeight species (Allogona profunda, Anguispira alternata, Carychium exile, Catinella ‘gelida’, Columella simplex, Discus catskillensis, Euconulus fulvus, Euconulus polygyratus, Gastrocopta corticaria, Gastrocopta holzingeri, Gastrocopta pentodon, Glyphyalinia indentata, Glyphyalinia rhoadsi, Guppya sterkii, Helicodiscus shimeki, Hendersonia occulta, Nesovitrea binneyana, Paravitrea multidentata, Stenotrema fraternum, Vallonia gracilicosta, Vertigo bollesiana, Vertigo cristata, Vertigo gouldi, Vertigo hubrichti, Vertigo n.sp., Vertigo paradoxa, Zonitoides arboreus, Zoogenetes harpa) favoured duff soils, while another 18 (Carychium exiguum, Catinella avara, Catinella exile, Catinella ‘vermeta’, Euconulus alderi, Gastrocopta tappaniana, Nesovitrea electrina, Oxyloma retusa, Punctum n.sp., Pupoides albilabris, Stenotrema leai leai, Strobilops affinis, Vallonia parvula, Vertigo elatior, Vertigo milium, Vertigo morsei, Vertigo nylanderi, Vertigo ovata) favoured turf soils. DISCUSSION These data demonstrate clearly that at large environmental and spatial scales most land snail species possess pronounced ecological preferences. They thus represent a paradox, being generalists at small scales, yet responding to specific environmental factors at larger ones. At large scales, species tend to congregate into six major compositional clusters related to habitat type, soil surface architecture, geography, moisture levels and presence of anthropogenic disturbance. Habitat type The six compositional clusters significantly differ in their habitat representations. Clusters A–D generally consist of forested sites while clusters E–F generally consist of grasslands. These results are in agreement with previous studies from other regions, including north-western Spain (Ondina & Mato, 2001), southern France (Magnin et al., 1995), western Switzerland (Baur et al., 1996), Croatia (Stamol, 1991, 1993), Hungary (Bába, 1989) and north-eastern Nevada (Ports, 1996). The contrast between open-ground and forest faunas is not limited to terrestrial gastropods. Other soil invertebrate groups that demonstrate this pattern include fungus-eating microarthropods (Branquart et al., 1995), carabid beetles (McCracken, 1994), terrestrial amphipods (Taylor et al., 1995) and collembola (Greenslade, 1997). In an ordination of global earthworm communities, Lavelle et al. (1995) demonstrated that open-ground and forest assemblages were distinct from the warm-tropics to the arctic. The distinction between forest and grassland faunas thus appears to be a general driving factor in soil biota community composition. 88 Great Lakes land snail community patterns 67 © 2003 Blackwell Publishing Ltd, Diversity and Distributions, 9, 55–71 Soil surface architecture The greater similarity of most lowland forest faunas to lowland grasslands, as opposed to upland forests and rock outcrops (Fig. 2), suggests additional factors underlie observed land snail composition patterns. The potential importance of soil surface architecture is implied as many lowland forest sites (e.g. tamarack, white cedar and most black ash swamp forests), and all lowland grasslands, possess turf soils. Only 8% of sites were improperly classified when soil surface type was used as the sole predictor variable (Table 7). Even this rate may be exaggerated, as most misclassifications were limited to two specific instances. First, even though having turf soils, igneous shoreline habitats had faunas almost identical to surrounding rock outcrop sites. Snails in this habitat, however, were largely restricted to friable accumulations of organic matter under stunted white cedar trees. Secondly, almost all duff sites with faunas similar to upland grasslands had experienced severe levels of anthropogenic disturbance. Striking differences exist between the species of duff and turf sites: 43% of taxa significantly favoured one soil surface type over the other (even with use of a conservative Bonferronicorrected significance threshold), while only 15% of frequent taxa (10 + occurrences) showed no preference. Inspection of these data indicate that for eight groups of closely related taxa within the same genus (24 total), one or more significantly favour duff soils, while the other(s) significantly favour turf (Table 9). In another four groups (10 additional taxa), one or more taxa significantly favour one of these soil types, while the other(s) exhibit a nonsignificant preference (Table 9). These 12 groups represent a wide variety of phylogenetic stocks (representing nine families: Carychiidae, Helicarionidae, Polygyridae, Punctidae, Pupillidae, Strobilopsidae, Succineidae, Valloniidae, Zonitidae), shell shapes (five wider than tall, five taller than wide and two equally tall as wide), and maximum shell dimensions (0.8 mm– 12 mm). The presence of so many pairs of closely related duff- and turf-specialist taxa across such a wide range of phylogenies, shell shapes and dimensions suggests that very strong selective pressures between these soil surface types have extended over evolutionary time scales. Table9Closelyrelatedintergenericspeciespairsthatdemonstratesignificantdifferencesand/ornonsignificanttrendsintheirsoilsurfacepreferences.A significancelevelof**representsthosespecieswhichpossessP<0.000463;*representthosespecieswhere0.0004633.00 30 40 Numberoftrees 20 10 0 Radial growth rate (mm per year) 0 0.34 0.67 1.00 1.34 1.67 2.00 2.34 2.67 >3.00 30 40 Numberoftrees 20 10 0 Numberoftrees Age estimate (years) 0 160 320 480 640 800 960 30 a b c d FFiigguurree 11 Age estimates and growth rates of trees on cliffs. Frequency distributions of estimated tree ages (a, b) and radial growth rates (c, d) for mature trees growing on cliffs in the United States (a, c, nǃ97) and Europe (b, d, nǃ127). Molecular basis of triclosan activity Triclosan (5-chloro-2-(2,4-dichlorophenoxy) phenol) has been used for more than 30 years as a general antibacterial and antifungal agent, and is found in formulations as diverse as toothpastes, cosmetics, antiseptic soaps, carpets, plastic kitchenware and toys. It has recently been suggested that triclosan blocks lipid biosynthesis by specifically inhibiting the enzyme enoyl–acyl carrier protein reductase (ENR)1 . We have carried out a structural analysis and inhibition experiments on a complex of ENR from the bacterium Escherichia coli with triclosan and NAD+ . We find that triclosan acts as a sitedirected, very potent inhibitor of the enzyme by mimicking its natural substrate. ENR catalyses the final, regulatory step in the fatty-acid synthase cycle: the reduction of a carbon–carbon double bond in an enoyl moiety that is covalently linked to an acyl carrier protein. ENR has been identified as a target for the diazaborines2 , a family of antibacterial agents, and many triclosanresistant strains of E. coli have been found to be resistant to diazaborines1 . The structure of the E.coli ENR complexed with NAD+ and triclosan (triclosan was provided by Ciba Speciality Chemicals and Coalite Chemicals) was determined to 2.2 Å resolution. The electron density map of the complex is of high quality (Fig. 1a) and reveals the mode of binding of the triclosan adjacent to the nicotinamide ring of the nucleotide cofactor in the enzyme’s active site. The phenol ring of triclosan forms a face-to-face interaction with the 95 Theoretical relation between δ18 O of land snail shell carbonate and latitude. The red horizontal lines marked A-E denote our five proposed transects to calibrate land snail δ18 Oshell systematics Section III: Spatio-Temporal Ecology Ecological patterns are not solely based upon the local, contemporaneous environment. Due to the lags inherent in complex systems, these patterns are also often impacted by drivers separated by distance in either space or time. As a result, simply knowing local conditions is often not enough to accurately predict what patterns will develop. Understanding how ecological systems change over space and time, and how spatial and temporal contingencies effect current patterns has been a long-standing focus of my research efforts. These investigations help determine the nature of spatial and temporal heterogeneities in ecological systems, and the extent to which these systems are in equilibrium with the local environment. Additionally, there is no better way to determine the potential impacts of global environmental change than to empirically study how past changes have impacted ecological systems. Land snails are excellent for this purpose as they are found across all of Earth’s terrestrial biomes and represent one of the most diverse groups of multicellular life. Their shells also represent one of the most common fossils in the temperate terrestrial geological record. As a result, this group has the potential to be as important in generating paleoclimate proxies for terrestrial systems as foraminifera are in the deep sea. Along with Jeffery Pigati of the US Geological Survey and Jason Rech of Miami University, our team first documented through a series of highly cited papers that many land snail phylogenetic groups have shells that are in equilibrium with environmental 14 C and thus can be used to accurately date terrestrial sediments that lack traditional organic carbon sources (e.g., wood, peat). We have now turned our efforts to O-isotopes. Our current NSF grant (running through 2018) will document climatic, environmental and biological controls on the oxygen isotopic composition of land snail shells. We anticipate that this work will lead to a subsequent proposal which, if funded, will establish through field surveys across North America the fidelity of shell Oisotopes ratios in the documentation of modern and prior 96 regional climatic conditions. This work has the potential to revolutionize terrestrial paleoclimatic proxies given that the same shells can theoretically provide not only useful chronometric but also environmental data. Additionally, along with Michal Horsák and Jan Divíšek I am exploring macroevolutionary process in Holarctic land snails using modern macroecology of these species in combination with their current molecular phylogenetics. These data are being applied to Late Glacial Maximum climate models to identify reasons for allopatric replacement within the boreal land snail fauna between three general regions: Europe, Beringia, and North America. Preliminary results suggest that geographic isolation of these regions during LGM times (climatic between Europe and Beringia; ice sheet between Beringia and North America) may well explain modern observed turnover patterns. Representative Publications [number of citations as of October 27, 2017] Chytrý, M., M. Horsák, V. Syrovátka, J. Danihelka, N. Ermakov, D.A. German, M. Hájek, O. Hájek, P. Hájková, V. Horsáková, M. Kočí, S. Kubešová, P. Lustyk, J.C. Nekola, Z. Preislerová, P. Resl & M. Valachovi. 2017. Refugial ecosystems in central Asia as indicators of biodiversity change during the Pleistocene–Holocene transition. Ecological Indicators. 76:357-367. [4] Bajc, A.F., P.F. Karrow, C.H. Yansa, B.B. Curry, J.C. Nekola, K. Seymour & G.L. Mackie. 2015. Geology and paleoecology of a Middle Wisconsin fossil occurrence in Zorra Township, southwestern Ontario, Canada. Canadian Journal of Earth Sciences. 52:386– 404. [4] Horsák, M., M. Chytrý, P. Hájková, M. Hájek, J. Danihelka, V. Horsáková, N. Ermakov, D.A. German, M. Kočí, P. Lustyk, J.C. Nekola, Z. Preislerová & M. Valachovič. 2015. European glacial relict snails and plants: environmental context of their modern refugial occurrence in southern Siberia. Boreas. 44:638-657. [14] Pigati, J.S., J.P. McGeehin, D.R. Muhs, D.C. Grimley & J.C. Nekola. 2015. Radiocarbon dating loess deposits in the Mississippi Valley using terrestrial gastropod shells (Polygyridae, Helicinidae, Discidae). Aeolian Research. 16:25-33. [5] Rakovan, M.T., J.A. Rech, J.S. Pigati, J.C. Nekola & G.C. Wiles. 2013. An evaluation of Mesodon and other large terrestrial gastropod shells for dating late Holocene and historic alluvium in the midwestern USA. Geomorphology. 193:47-56. [11] Horsák, M., M. Hájek, D. Spitale, P. Hájková, D. Dítě & J.C. Nekola. 2012. The age of islandlike habitats impacts habitat specialist species richness. Ecology. 93:1106-1114. [46] Rech, J.A., J.C. Nekola & J.S. Pigati. 2012. Radiocarbon ages of terrestrial gastropods extend duration of ice-free conditions at the Two Creeks forest bed, Wisconsin, USA. Quaternary Research. 77:289-292. [11] Pigati, J.S., D.M. Miller, S.A. Mahan, J.A. Bright, J.C. Nekola & J.B. Paces. 2011. Chronology, sedimentology, and microfauna of ground-water discharge deposits in the central Mojave Desert, Valley Wells, California. Geological Society of America Bulletin. 123:2224-2239. [24] 97 Rech, J.A., J.S. Pigati, S.B. Lehmann, C.N. McGimpsey, D.A. Grimley & J.C. Nekola. 2011. Assessing open-system behavior of 14-C in terrestrial gastropod shells. Radiocarbon. 53:325-335. [17] Pigati, J.S., J.A. Rech & J.C. Nekola. 2010. Radiocarbon dating of small terrestrial gastropod shells in North America. Quaternary Geochronology. 5:519-532. [94] Nekola, J.C. & C.E. Kraft. 2002. Spatial constraint of peatland butterfly occurrences within a heterogeneous landscape. Oecologia. 130:62-71. [21] Kraft, C.E., P.J. Sullivan, A.Y. Karatayev, L.E. Burlakova, J.C. Nekola, L.E. Johnson & D.K. Padilla. 2002. Landscape patterns of an aquatic invader: assessing dispersal extent from spatial distributions. Ecological Applications. 12:749-759. [68] Nekola, J.C. 1999. Paleorefugia and neorefugia: the influence of colonization history on community pattern and process. Ecology. 80: 2459-2473. [121] 98 Research Paper Radiocarbon dating of small terrestrial gastropod shells in North America Jeffrey S. Pigati a,*, Jason A. Rech b , Jeffrey C. Nekola c a U.S. Geological Survey, Denver Federal Center, Box 25046, MS-980, Denver CO 80225, USA b Department of Geology, Miami University, Oxford, OH 45056, USA c Department of Biology, University of New Mexico, Albuquerque, NM 87131, USA a r t i c l e i n f o Article history: Received 26 May 2009 Received in revised form 20 January 2010 Accepted 21 January 2010 Available online 29 January 2010 Keywords: Radiocarbon Land snails Limestone effect Chronology Quaternary a b s t r a c t Fossil shells of small terrestrial gastropods are commonly preserved in wetland, alluvial, loess, and glacial deposits, as well as in sediments at many archeological sites. These shells are composed largely of aragonite (CaCO3) and potentially could be used for radiocarbon dating, but they must meet two criteria before their 14 C ages can be considered to be reliable: (1) when gastropods are alive, the 14 C activity of their shells must be in equilibrium with the 14 C activity of the atmosphere, and (2) after burial, their shells must behave as closed systems with respect to carbon. To evaluate the first criterion, we conducted a comprehensive examination of the 14 C content of the most common small terrestrial gastropods in North America, including 247 AMS measurements of modern shell material (3749 individual shells) from 46 different species. The modern gastropods that we analyzed were all collected from habitats on carbonate terrain and, therefore, the data presented here represent worst-case scenarios. In sum, w78% of the shell aliquots that we analyzed did not contain dead carbon from limestone or other carbonate rocks even though it was readily available at all sites, 12% of the aliquots contained between 5 and 10% dead carbon, and a few (3% of the total) contained more than 10%. These results are significantly lower than the 20–30% dead carbon that has been reported previously for larger taxa living in carbonate terrain. For the second criterion, we report a case study from the American Midwest in which we analyzed fossil shells of small terrestrial gastropods (7 taxa; 18 AMS measurements; 173 individual shells) recovered from late-Pleistocene sediments. The fossil shells yielded 14 C ages that were statistically indistinguishable from 14 C ages of well-preserved plant macrofossils from the same stratum. Although just one site, these results suggest that small terrestrial gastropod shells may behave as closed systems with respect to carbon over geologic timescales. More work on this subject is needed, but if our case study site is representative of other sites, then fossil shells of some small terrestrial gastropods, including at least five common genera, Catinella, Columella, Discus, Gastrocopta, and Succinea, should yield reliable 14 C ages, regardless of the local geologic substrate. Published by Elsevier B.V. 1. Introduction Gastropods are one of the most successful animal groups on Earth, with at least 70,000 extant species occupying terrestrial, marine, and freshwater habitats. Globally, terrestrial gastropods encompass at least 35,000 species (Barker, 2001), span 4–5 orders of magnitude in shell volume, and represent a variety of trophic levels, including polyphagous detritivores, herbivores, omnivores, and carnivores (Kerney and Cameron, 1979; Burch and Pearce, 1990). They are so exceptionally diverse in their appearance, ecology, and physiology that determining their phylogenetic relationships from conchological and/or anatomical characteristics remains difficult and controversial (e.g., Ponder and Lindberg, 1997 and references therein). It is clear, however, that the preference for terrestrial habitats of North American gastropods developed independently in three of six basal clades (Neritomorpha, Caenogastropoda, and Heterobranchia), with the informal group Pulmonata representing more than 99% of the continental fauna. Of the Pulmonata, the most common size class1 in both modern and fossil assemblages are individuals with adult shells that are <10 mm in maximum dimension (Nekola, 2005) (Fig. 1). * Corresponding author. Tel.: þ1 303 236 7870; fax: þ1 303 236 5349. E-mail address: jpigati@usgs.gov (J.S. Pigati). 1 Size classes of gastropods are categorized by the maximum shell dimension (length or diameter) as follows: large (>20 mm), medium (10–20 mm), small (5– 10 mm), minute (2–5 mm), and micro (<2 mm). Although the size of the gastropods targeted in this study range from small to micro, for simplicity, we refer to them collectively as ‘‘small’’. Contents lists available at ScienceDirect Quaternary Geochronology journal homepage: www.elsevier.com/locate/quageo 1871-1014/$ – see front matter Published by Elsevier B.V. doi:10.1016/j.quageo.2010.01.001 Quaternary Geochronology 5 (2010) 519–532 99 Today, small terrestrial gastropods occupy and thrive in incredibly diverse habitats, from marshes, wet meadows, and grasslands to upland forests and tundra. Species are known from all continents, save Antarctica, and occupy almost all climate regimes except hyperarid deserts and the high Arctic. Their distribution in the fossil record is equally diverse. Gastropod shells are commonly preserved in wetland, alluvial, loess, and glacial deposits, as well as within sediments at archeological sites worldwide (e.g., Evans, 1972). But even though their distribution is widespread and their aragonitic shells contain w12% by weight carbon, terrestrial gastropods are often avoided for 14 C dating because many taxa incorporate 14 C-deficient (or ‘‘dead’’) carbon from limestone and other carbonate rocks when building their shells. This phenomenon, referred to as the ‘‘Limestone Problem’’ by Goodfriend and Stipp (1983), can cause 14 C ages of gastropod shells to be as much as w3000 yrs too old. Despite the Limestone Problem, geochronologists have continued to investigate the possibility of using terrestrial gastropod shells for 14 C dating because of their widespread occurrence and potential for dating Quaternary sediments. Most 14 C studies of gastropod shells have found that gastropods consistently incorporate dead carbon from limestone in their shells when it is available (Frye and Willman, 1960; Leighton, 1960; Rubin et al., 1963; Tamers, 1970; Evin et al., 1980; Goodfriend and Hood, 1983; Goodfriend and Stipp, 1983; Goslar and Pazdur, 1985; Yates, 1986; Goodfriend, 1987; Zhou et al., 1999; Quarta et al., 2007; Romaniello et al., 2008). These studies, however, were generally limited to a few individual gastropods collected from a small number of sites, and were biased toward large taxa and warm climates. Brennan and Quade (1997) analyzed a number of small terrestrial gastropod taxa and found that small shells generally yielded reliable 14 C ages for late-Pleistocene paleowetland deposits in the American Southwest. Pigati et al. (2004) followed by measuring the 14 C activities of a suite of small gastropods living in alluvium dominated by Paleozoic carbonate rocks in Arizona and Nevada and found that while some of the small gastropods did incorporate dead carbon from limestone when building their shells, others did not. Based in part on these initial positive results, small terrestrial gastropod shells have been used recently to date Quaternary wetland and lacustrine deposits in the Americas (e.g., Pedone and Rivera, 2003; Placzek et al., 2006; Pigati et al., 2009). However, it is unclear if the results obtained from modern gastropods collected from a limited number of sites in the American Southwest can be extrapolated to all geologic, ecologic, and climatic environments. Moreover, it is not known if results for one taxonomic level (family, genus, or species) can be extrapolated to other members of the Fig. 1. Photographs of select small terrestrial gastropods included in this study (1 mm bar in each panel for scale). (a) Cochlicopa lubricella, (b) Columella columella alticola, (c) Discus macclintockii, (d) Euconulus fulvus, (e) Hawaiia miniscula, (f) Hendersonia occulta, (g) Punctum minutissimum, (h) Pupilla muscorum, (i) Strobilops labyrinthica, (j) Vallonia gracilicosta, (k) Vertigo elatior, and (l) Zonitoides arboreus. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532520 100 same level living elsewhere, or even between individuals living within the same population. Here we report the results of a comprehensive analysis of the Limestone Problem for small terrestrial gastropods from 163 localities in North America (Fig. 2). All samples that we analyzed were collected from habitats on carbonate terrain and, therefore, the data reported here represent worst-case scenarios. In addition, we measured the 14 C activities of a number of fossil shells recovered from well-dated sediments at a late-Pleistocene site in the American Midwest as a case study to determine if the shells remain closed systems with respect to carbon over geologic timescales. Positive results for both tests for a particular taxon would allow us to consider 14 C ages derived from fossil shells of that taxon to be reliable, regardless of the local geologic substrate. 2. Shell carbonate and 14 C dating All materials (organic and inorganic) that yield reliable 14 C ages have two common characteristics. First, the initial 14 C activity of the material – a plant, for example – was in equilibrium with atmospheric 14 C at the time that it was alive. In other words, the 14 C activity of a plant that lived T yrs ago was the same as the 14 C activity of the atmosphere T yrs ago (after accounting for isotopic fractionation). Second, after death, the material behaved as a closed system; carbon was neither added to nor removed from the sample material. If both of these criteria are met, then the measured 14 C activity is a function of only two parameters: the initial 14 C activity of the atmosphere and the amount of time elapsed since the death of the organism. The measured 14 C activity and the 14 C age of the material are related by the familiar decay equation A ¼ AoeÀlt (1) where A and Ao are the measured and initial 14 C activities of the material, respectively, l is the decay constant, and t is the time elapsed since the death of the organism. Conventional radiocarbon ages assume that the atmospheric 14 C activity is invariant through time (i.e., Ao ¼ 1). Radiocarbon ages can be converted to calendar year ages to account for temporal variations in the 14 C activity of the atmosphere (Reimer et al., 2009). Fig. 2. Locations of modern localities (red dots) and the fossil locality at the Oxford East outcrops in southwestern Ohio (star). Modern localities and collection information are listed in Tables S1 and S2, respectively. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article). J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 521 101 2.1. Sources of shell carbon In order to evaluate the validity of a 14 C age of a given sample, the 14 C contents of the original sources of the carbon and their contribution to the total carbon content must be known. Carbon in gastropod shell carbonate originates from as many as four different sources: atmospheric CO2, food, water, and carbonate rocks. Gastropods incorporate atmospheric CO2 in their shell carbonate via respiration. Respired CO2 is introduced to the bicarbonate pool in the gastropod’s hemolymph and passed along to the extrapallial fluid, from which the shell carbonate is ultimately precipitated (Wilbur, 1972). Estimates of the contribution of atmospheric CO2 to gastropod shell carbonate vary between negligible (Stott, 2002), 16–48% (Romaniello et al., 2008), and 30– 60% (Goodfriend and Hood, 1983). Carbon from food sources (e.g., living plants, fungi, organic detritus) is incorporated into the extrapallial fluid through two mechanisms: direct digestion and breakdown of urea. When gastropods consume and digest food, carbon is introduced to the hemolymph and passed along to the extrapallial fluid in the same manner as atmospheric CO2. There it mixes with atmospheric carbon before becoming incorporated in the shell carbonate (Wilbur, 1972). Carbon derived from urea takes a more indirect pathway. Urea that is not expelled by the gastropod breaks down into CO2 and NH3 via a urease reaction (Stott, 2002). The resulting CO2 is then reintroduced directly to the extrapallial fluid and ultimately incorporated into the shell carbonate. Estimates of the amount of carbon derived from plants, either directly or indirectly through urea, vary between 25 and 40% (Goodfriend and Hood, 1983), 36–73% (Romaniello et al., 2008), and w100% (Stott, 2002). Terrestrial gastropods ingest water from multiple sources, including dew, soil moisture, standing water, and precipitation, all of which contain some amount of dissolved inorganic carbon (DIC). Water is taken up through the foot of the gastropod by contact rehydration (Balakrishnan and Yapp, 2004) and introduced to the hemolymph before being passed on to the extrapallial fluid. Pigati et al. (2004) found that aqueous carbon sources account for w10% of the shell carbon for one species of Catinella, but it is not known if this value is constant across the entire Succineidae family. To our knowledge, data for other terrestrial taxa do not exist. Finally, some terrestrial gastropods are able to scrape carbonate rocks (limestone, dolomite, soil carbonate), and ingest the powder or granules which then dissolve in their stomach acid to produce CO2. As before, the dead carbon from the rocks is introduced to the hemolymph, passed on to the extrapallial fluid, and ultimately incorporated in the shell carbonate. Dead carbon from limestone can account for up to w30% of the total carbon in shells of large terrestrial gastropods. (Goodfriend and Stipp, 1983). 2.2. Effects of carbon sources on 14 C ages of shell carbonate In most environments, the 14 C activities of live plants are in equilibrium with atmospheric carbon. Gastropods that obtain their shell carbon from live plants and the air, therefore, should yield reliable 14 C ages, assuming they behave as closed systems after burial (Fig. 3). Gastropods that consume organic detritus (i.e., decaying plant litter) typically do not pose a significant problem for 14 C dating because the time between plant death, its incorporation into decomposition products, and consumption by gastropods is usually quite short, on the order of a few yrs. The 14 C activity of water that is available for consumption by terrestrial gastropods (e.g., dew, standing water, precipitation) is at or near equilibrium with atmospheric 14 C and, therefore, water is unlikely to introduce a significant error to 14 C ages of terrestrial gastropod shells. Exceptions include gastropods living directly adjacent to springs that discharge waters from deeply-circulating carbonate aquifers, lakes or rivers with significant hard water effects, or in active volcanic areas where 14 C-deficient CO2 in surface waters may be abundant (e.g., Riggs, 1984; Grosjean, 1994). 14 C ages of gastropods living in such areas should be evaluated carefully. The incorporation of 14 C from limestone and other carbonate rocks can present a significant problem for 14 C dating of terrestrial gastropod shells. The 14 C activity of atmospheric carbon, plants, and water consumed by gastropods is essentially the same, w100% modern carbon (pMC). In contrast, because most carbonate rocks are of pre-Quaternary age, their 14 C activity is typically 0 pMC. Thus, for 14 C dating, the magnitude of the potential error introduced by carbonate rocks is a direct function of the amount of carbon from rocks that is incorporated in the gastropod shell (Fig. 3). Unfortunately, a simple correction that accounts for the incorporation of 14 C-deficient carbon in gastropod shells is not possible because we cannot know a priori how much of the shell carbon was derived from carbonate rocks versus other sources. Thus, to be confident in 14 C ages derived from terrestrial gastropod shells, it is imperative that we identify and avoid taxa that incorporate dead carbon from rocks altogether. 2.3. Open-system behavior Even if some terrestrial gastropods consistently manage to avoid the Limestone Problem regardless of the local geologic substrate or environmental conditions, there is another hurdle that must be overcome before we can confidently use their shells for 14 C dating. That is, gastropod shells must remain closed systems with respect to carbon after burial. For reliable 14 C dating, the pool of carbon atoms measured during the 14 C dating process must consist solely of carbon atoms that originally resided in the shell. Thus, following burial, shells must resist the addition or exchange of 14 C atoms with the local environment. Shells that exhibit open-system behavior typically yield 14 C ages that are too young, and the magnitude of the error depends upon the degree of such behavior (Fig. 3). Previous work has suggested that 14 C ages from small terrestrial gastropod shells recovered from fossil deposits in arid environments may be reliable back to at least w30,000 14 C yrs B.P., but a small degree of open-system behavior appears to compromise 14 C dates obtained from shells older than this (Pigati et al., 2009). In -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 0 5 10 15 20 25 30 35 40 True 14 C age (ka) eurtmorfnoitaiveD41 )ak(egaC Fig. 3. Modeled deviation from the true 14 C age for four scenarios: (1) closed-system behavior and no dead carbon (thick solid line), (2) closed-system behavior and 10% dead carbon (dashed line), (3) open-system behavior equivalent to 1% modern carbon contamination and 10% dead carbon (thin solid line), and (4) open-system behavior equivalent to 1% modern carbon contamination and no dead carbon (dotted line). J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532522 102 this study, we analyzed fossil shells from the American Midwest because of their abundance in Quaternary deposits in the region, the presence of multiple calcareous substrates (Paleozoic limestone, calcareous till, and loess), and the humid climate (annual precipitation in southwestern Ohio is w100 cm yrÀ1 ). If fossil shells exhibit even a small degree of open-system behavior in arid environments, it may be exacerbated and, therefore, more easily detected in humid environments where interaction between shells and DIC in ground water is more prevalent. 3. Methods 3.1. Live gastropods Previous ecological sampling by one of us (JCN) has resulted in an extensive collection of modern terrestrial gastropods from North America, constituting w250 taxa and over 470,000 individuals from more than 1000 modern environments. Gastropods were collected at each site from a representative 100–1000 m2 area by hand collection of larger taxa and litter sampling of smaller taxa, which provides the most complete assessment of site faunas (Oggier et al., 1998; Cameron and Pokryszko, 2005). Collections were made at places of high mollusc density, such as loosely compacted leaf litter lying on top of highly compacted damp soil or humus (Emberton et al., 1996). Litter was removed by hand and sieved by shaking, tapping, or other agitation in the field using a shallow sieve (ASTME #10; 2.0 mm mesh) nested inside a second sieve (ASTME #30; 0.6 mm mesh). The process was continued for 15–60 min during which time 50–500 mg of material was collected and retained. Gastropods and detritus were dried at room temperature in the laboratory and then hand-picked against a neutral background. All shells, shell fragments, and slug plates were removed and identifiable material was assigned to species using JCN’s reference collection. Nomenclature generally follows that of Hubricht (1985) with updates and corrections by Nekola (2004). We selected 247 aliquots of shell material (3749 individual shells) from 163 sites across the United States and southern Canada for 14 C analysis. Nearly all of the specimens that we chose for analysis were collected live, but at a few sites, only recently-dead gastropods were available, which were identified by a translucent appearance or the retention of color in the shells. Shells of small gastropods that are dead for more than a year or so become increasingly white and opaque with time (J. Nekola, unpublished data), and were excluded from our study. In the fossil record, species-level identification of fossil shells is possible for most small terrestrial gastropods and, therefore, the results of our investigation of modern gastropods can be applied directly to the fossil record. An exception is the Succineidae family, which is composed of three genera (Catinella, Oxyloma, and Succinea) that are difficult to differentiate in modern faunas, let alone the geologic record (Fig. 4). Their simple shells exhibit few diagnostic characteristics and, therefore, species-level identification is based on soft-body reproductive organ morphology, which is rarely preserved in the fossil record. This presents a significant problem for geochronologists; that is, can we be confident in 14 C ages derived from shells from any taxon within the Succineidae family, or do we need to target a specific genus or species? To address this issue, we measured the 14 C activity of 100 aliquots of gastropod shell material (802 individual shells) for twelve species of the Succineidae family to determine the level of identification (i.e., family, genus, or species) required to apply our results to the fossil record. We prepared aliquots of modern shell material for 14 C analysis at the University of Arizona Desert Laboratory (JSP) and Miami University (JAR). We selected multiple shells at random for X-ray diffraction (XRD) analysis using a Siemens Model D-500 diffractometer to verify that only shell aragonite remained prior to preparation for 14 C analysis. There was no evidence of primary or secondary calcite in any of the shells that we analyzed. When possible, shells were broken, the adhering soft parts were removed using forceps, and the shells were treated with 6% NaOCl for 18– 24 h at room temperature to remove all remnants of organic matter. Shells were not powdered during pretreatment to minimize the potential for adsorption of atmospheric 14 C (Samos, 1949). We selectively dissolved some of the shells by briefly introducing dilute HCl to remove secondary carbonate (dust) from primary shell material. Shells were washed repeatedly in ASTM Type 1, 18.2 MU (hereafter ‘‘ultrapure’’) water, sonicated for a few seconds to remove adhered solution, washed again with ultrapure water, and dried in a vacuum oven overnight at w70 C. Shell aragonite was converted to CO2 using 100% H3PO4 under vacuum at either 50 or 75 C until the reaction was visibly complete (w1 h). Water, SOx, NOx, and halide species were removed using passive Cu and Ag traps held at w600 C and the resulting CO2 was split into two aliquots. One aliquot was converted to graphite by catalytic reduction of CO (modified after Slota et al., 1987) and submitted to the Arizona-NSF Accelerator Mass Spectrometry (AMS) facility for 14 C analysis. The second aliquot was submitted for d13 C analysis in order to correct the measured 14 C activity of the shell carbonate for isotopic fractionation. 14 C data for modern gastropods are presented as D14 C values in per mil (Stuiver and Polach, 1977; Reimer et al., 2004) and analytical uncertainties are reported at the 2s (95%) confidence level. d13 C values are given in the usual delta (d) notation as the per mil deviation from the VPDB standard. Analytical uncertainties for d13 C measurements are less than 0.1& based on repeated measurements of standards. We also measured the 14 C content of several gastropod bodies for comparison to their corresponding shells. The bodies were treated with 10% HCl for 15–30 min to remove any remnants of the carbonate shell, rinsed repeatedly, and dried in a vacuum oven at Fig. 4. Photographs of the three genera of the Succineidae family: Catinella (left panels), Oxyloma (middle panels), and Succinea (right panels); all are w10 mm in length. The simple shells of the three genera contain few diagnostic characteristics and, therefore, species-level identification is based on soft-body reproductive organ morphology, which is rarely preserved in the fossil record. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 523 103 w70 C. The dried bodies were placed in 6 mm quartz tubes along with w100 mg of cupric oxide (CuO) and a small piece (1 mm  5 mm) of silver foil, all of which were pre-combusted at 900 C for 4–6 h. The tube was evacuated, sealed with a glassblower’s torch, and combusted offline at 900 C. The resulting CO2 gas was isolated and converted to graphite as above. 3.2. Modeling of 14 C values of gastropod diets To quantify the amount of carbon in a gastropod shell that was derived from limestone or other carbonate rocks, it is necessary to compare the measured D14 C values of shells with D14 C values of the gastropod diet. D14 C values of live plants consumed by gastropods are identical to D14 C values of the atmosphere (Fig. 5a), which were calculated using 14 C data averaged over the Northern Hemisphere (Hua, 2004). We assigned a 5& uncertainty to the atmospheric values to account for short-term, regional variations (Hsueh et al., 2007) and small changes in atmospheric 14 C values that occur at a given site during the short (usually annual) lifespan of the small gastropods (Manning et al., 1990; Meijer et al., 1995). Plant detritus from previous years’ primary production has a higher 14 C activity than live plants because of the 14 C ‘‘bomb-spike’’ (Hua, 2004). Simply comparing the D14 C values of shells with the D14 C of the atmosphere or live plants during the year that the gastropod was alive, therefore, would ignore the potential impact of detritus in the gastropod diet. We estimated the amount of detritus from a given year that was available for consumption using a wide range of carbon turnover rates (0.2–0.002 yrÀ1 ; Fig. 5b), which are applicable to O horizons and the upper few centimeters of A horizons in which small terrestrial gastropods typically live (Guadinski et al., 2000; Torn et al., 2005; Brovkin et al., 2008). We intentionally chose a wide range of carbon turnover rates, which are equivalent to mean residence times for carbon of 5–500 yrs, to encompass the potentially wide range of detritus ages at the 163 localities included in our study. We then used Monte Carlo simulation to generate 10,000 values for the D14 C of the gastropod diets for each of the past 13 yrs (the time span of ourcollections), which allowed two factors to varyrandomly:thecarbonturnover rate(0.2–0.002yrÀ1 ) andtheage of the detritus fraction included in the diet (range of 0–1000 yrs). We also ran simulations in which we let the age of the detritus vary up to 10,000 yrs, but the results did not change significantly. We took the average and standard deviation of the 10,000 generated values as the gastropod diet D14 C value for each year of collection between 1996 and 2008 (Fig. 5c). All uncertainties are reported at the 2s (95%) confidence level. As expected, the modeled D14 C value of the gastropod diet for a given year is slightly higher than the atmospheric D14 C value for the same year. Uncertainties associated with the modeled values are relatively large, on the order of w35%, because of the large range of detritus D14 C values that could be present at a given site. Shells with D14 C values that are lower than the modeled D14 C value of the gastropod diet during the year in which the gastropod was alive indicate the presence of dead carbon from limestone or other carbonate rocks. When applicable, the difference between the D14 C values was converted into 14 C yrs to estimate the ‘‘limestone effect’’, which represents the potential error introduced by the incorporation of dead carbon in the shells. The magnitude of the limestone effect should be considered a maximum value because the calculation assumes that all of the dead carbon came from carbonate rocks, rather than older (but not infinitely-aged) organic matter. Because of the uncertainties associated with modeling the D14 C of the gastropod diet, we are unable to discern limestone effects smaller than w300 14 C yrs. 3.3. Fossil gastropods Fossil gastropod shells were collected from glacial deposits at the Oxford East glacial outcrops in southwestern Ohio as a case study to determine if the shells remain closed systems with respect to carbon over geologic timescales. These outcrops contain a series of glacial diamictons from the Miami Lobe of the Laurentide Ice 0 50 100 150 200 250 1980 1985 1990 1995 2000 2005 2010 Year 41 )‰(erehpsomtafoC 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 0 200 400 600 800 1000 Time after plant death (years) elbaliavasutirtedfonoitcarF noitpmusnocrof 0.002 yr-1 0.2 yr-1 0.05 yr-1 0.02 yr-1 0.01 yr-1 0 50 100 150 200 250 1993 1998 2003 2008 Date of collection 41 )‰(teiddoportsagfoCΔΔ a b c Fig. 5. Modeling results for (a) the 14 C activity of the gastropod diet using measured atmospheric values for the northern hemisphere (after Hua, 2004), (b) carbon turnover rates (CTRs) ranging from 0.2 to 0.002 yrÀ1 , and (c) Monte Carlo simulation to generate estimates of the D14 C values of gastropod diets for each of the past 13 yrs. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532524 104 Sheet that are separated by thin (3–5 cm) units of calcareous organic-rich silt that contain gastropods, plant macrofossils, and rooted tree stumps. AMS 14 C dating of plant macrofossils and in situ tree stumps has shown that the age of this unit is between w20,100 and 21,400 14 C yrs B.P. (Eckberg et al., 1993; Lowell, 1995). Gastropod taxa identified previously from this unit include Columella columella, Discus cronkhitei, Euconulus fulvus, Hendersonia occulta, Pupilla muscorum, Vertigo elatior, and multiple Succineidae taxa (Dell, 1991). Gastropod-bearing sediment was collected and placed in deionized water with a deflocculant for several days to soften the sediment enough to pass through a 0.5 mm sieve. A few samples were placed in an ultrasonic bath for w1 h to further disaggregate the sediment. Fossil shells were hand-picked from the retained fraction, placed in a beaker of ultrapure water, subjected to an ultrasonic bath for a few seconds, and then repeatedly dunked in a second beaker of ultrapure water to remove sediment that adhered to the shell surface or was lodged within the shell itself. The recovered shells were broken and examined under a dissecting microscope to ensure that the interior whorls were free of secondary carbonate and detritus. Fossil shells that were free of detritus were then processed for 14 C in the same manner as the modern specimens, including random selection of shells for XRD analysis. None of the fossil shells that we analyzed contained measurable quantities of either primary or secondary calcite. Organic samples, which included bark, charcoal, plant fragments, and wood, were subjected to a standard acid–base–acid (ABA) chemical pretreatment with 1N HCl (1 h at 60 C), 1N NaOH (18–24 h at 60 C), and 1N HCl again (2–4 h at 60 C) before combustion at 900 C in the presence of cupric oxide and silver foil. The resulting CO2 was purified and converted to graphite in the same manner as above. Conventional radiocarbon ages are reported in 14 C yrs and, after calibration, in calendar yrs. For calibration, we used the IntCal09.14C dataset (CALIB 6.0.0, Stuiver and Reimer, 1993; Reimer et al., 2009). 4. Results 4.1. Modern shells A few aliquots (11 of 247, or 4.5% of the total) yielded D14 C values that were higher than the modeled dietary D14 C values of the corresponding year of collection, which indicates these individuals consumed unusually high amounts of bomb-spike carbon (Table S1, Fig. 6). Data from these shells were excluded from further analysis. For the remaining 236 aliquots of gastropod shells from the 46 different species that we analyzed, w78% did not contain any dead carbon from limestone or other carbonate rocks even though it was readily available at all sites, w12% contained between 5 and 10% dead carbon, and a few (3% of the total) contained more than 10% (Table S1, Fig. 6). D14 C values for all taxa ranged from À97.5 to 158.4&, and limestone effects averaged only w180 14 C yrs. Dead carbon was not detected in the shells of at least 23 different species, including (number of shells in parentheses) Catinella avara (99), Catinella gelida (66), Catinella vermeta (39), Cochlicopa lubricella (17), Cochlicopa morseana (16), Columella columella (94), Discus catskillensis (40), Discus cronkhitei (47), Discus macclintockii (5), Euconulus alderi (71), Euconulus polygyratus (57), Gastrocopta pentodon (131), Nesovitrea binneyana (46), Punctum minutissimum (542), Pupilla hebes (7), Strobilops affinis (36), Succinea bakeri (27), Succinea grosvernori (1), Succinea n. sp. ‘Minnesota A’ (1), Succinea ovalis (50), Succinea strigata (14), Vertigo hubrichti (144), and Vertigo modesta (73). D14 C values were most negative (i.e., contained the most dead carbon) for shells from the Pupilla and Vallonia genera. Maximum limestone effects for these genera ranged from 780 Æ 310 14 C yrs for P. muscorum to 1590 Æ 280 14 C yrs for P. sonorana, and from 1010 Æ 380 14 C yrs for Vallonia perspectiva to 1500 Æ 270 14 C yrs for V. cyclophorella, respectively (Table 1). The only other species that exhibited a limestone effect that was greater than 1000 14 C yrs was H. occulta (1210 Æ 250 14 C yrs). These results can be applied directly to the fossil record if it is possible to identify the taxa to the species-level based on shell morphology. For the Succineidae family, the data must be evaluated at the family or genus level to be applicable. Taking the Succineidae family as a whole, 85% of the shell aliquots that we analyzed did not contain measurable amounts of dead carbon; the remaining aliquots contained an average of 5.2% dead carbon. Within the family, D14 C values for the genus Catinella ranged from 63.7 to 147.0&, Oxyloma values ranged from À0.8 to 135.0&, and Succineavalues ranged from 16.1 to 147.8&. Members of the Catinella genus incorporate little, if any, dead carbon from limestone or other carbonate rocks in their shells; 32 of 33 aliquots (97%) of Catinella shells did not contain measurable amounts of dead carbon. The remaining sample (DL-170) contained only a very small amount of dead carbon, equivalent to a limestone effect of 320 Æ 310 14 C yrs. Similarly, 36 of 39 aliquots (92%) of Succinea shells did not contain dead carbon. The remaining three aliquots were all Succinea indiana; limestone effects for these shells ranged from 430 Æ 370 to 610 Æ 380 14 C yrs. For Oxyloma, 15 of 24 aliquots (63%) did not contain measurable amounts of dead carbon. Limestone effects for the remaining samples ranged between 250 Æ 210 and 670 Æ 240 14 C yrs. 4.2. Modern gastropod bodies We also measured the 14 C activity of the bodies of eight gastropods to determine the magnitude of the offset between the body carbon and shell carbonate (Table 2). Ideally, we would have preferred to measure the 14 C content of the extrapallial fluid to comparewith the shell carbonate todetermine if carbon isotopes are fractionated when the shells are formed, but this was not feasible because the gastropod bodies were simply too small. Regardless, the measured D14 C values of gastropod body carbon ranged from 42 to 101&, similar to the D14 C values of shell carbonate from the same sites, which ranged from 43 to 133&. However, we did not observe a clear relation between the 14 C activity of gastropod body carbon and shell carbonate. D14 C values of bodies of Oxyloma retusa collected from Maquokata River Mounds, Iowa were indistinguishable from the D14 C values of their corresponding shells (D14 Cbody ¼ 42 Æ 6&, D14 Cshell ¼ 43 Æ 10&), as were values for S. ovalis from Dave Pepin Homestead, Minnesota (D14 Cbody ¼ 101 Æ8&, D14 Cshell ¼ 110 Æ 8&). In contrast, D14 C values for bodies of S. ovalis were significantly lower than their shells from Brewer Boat Ramp, Maine (D14 Cbody ¼ 55 Æ 6&, D14 Cshell ¼ 82 Æ 5&) and Zippel Bay State Park, Minnesota (D14 Cbody ¼ 84 Æ 6&, D14 Cshell ¼ 133 Æ 12&). The reason(s) for this difference is unclear. 4.3. Fossil shells Well-preserved fossil organic material (bark, plant macrofossils, and wood) recovered from sediments at the Oxford East outcrops yielded calibrated ages that ranged from 24.60 Æ 0.40 to 25.28 Æ 0.55 ka, with an average of 24.93 Æ 0.30 ka (n ¼ 5; Table 3, Fig. 7). Gastropod shells recovered from the same stratigraphic unit yielded ages that ranged from 23.92 Æ 0.66 to 25.81 Æ 0.94 ka, and averaged 24.73 Æ 0.44 ka (n ¼ 18). Average ages of six of the seven fossil taxa were indistinguishable from the average age of the organic matter: Discus shimeki (24.80 Æ 0.17 ka; n ¼ 3), P. muscorum (24.34 Æ 0.32 ka; n ¼ 3), Vallonia gracilicosta (24.20 Æ 0.40 ka; n ¼ 2), Vertigo hannai (24.90 Æ 0.77 ka; n ¼ 1), V. modesta J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 525 105 (24.86 Æ 0.32 ka; n ¼ 3), and Succineidae (25.29 Æ 0.54 ka; n ¼ 3). The average age of H. occulta (24.34 Æ 0.15 ka; n ¼ 3) was slightly younger than the organic ages. 5. Discussion 5.1. Small terrestrial gastropods and the limestone problem Approximately 78% of the modern shells that we analyzed did not contain any dead carbon from limestone or other carbonate rocks even though it was readily available at all sites, w12% of the aliquots contained between 5 and 10% dead carbon, and a few (3% of the total) contained more than 10%. Even at the high end, the amount of dead carbon in the small shells is significantly less than the 20–30% dead carbon that has been previously reported for larger taxa (e.g., Goodfriend and Stipp, 1983). If we extrapolate our results to the fossil record and assume that the shells behave as closed systems with respect to carbon over geologic timescales, then small terrestrial gastropod shells should provide accurate 14 C ages w78% of the time and ages that are a Fig. 6. Shell carbonate D14 C values compared to modeled dietary D14 C values for modern gastropods. Data points that fall on the solid black line in each panel represent gastropods that obtained their carbon from live plants and the atmosphere. Data points that fall below the solid line indicate that dead carbon from limestone or other carbonate rocks was incorporated during shell construction. The magnitude of this phenomenon, called the ‘‘limestone effect’’, depends upon the amount of shell carbon that was derived from rocks as shown by the dashed lines. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532526 106 within w1000 14 C yrs of the true age w97% of the time. Shells from at least 23 different species did not contain dead carbon, and therefore should yield reliable 14 C ages if the modern shell data can be applied directly to the fossil record. For the Succineidae family as a whole, 85% of the shell aliquots that we analyzed did not contain measurable amounts of dead carbon; the remaining aliquots contained an average of 5.2% dead carbon, equivalent to a limestone effect of 425 14 C yrs. At the genus level, shells of the genus Catinella should yield reliable 14 C ages w97% of the time, again assuming closed-system behavior, and ages that are within w300 14 C yrs of the true age every time. Similarly, Succinea shells should yield reliable 14 C ages w92% of the time and ages that are within w600 14 C yrs of the true age every time. Results for the genus Oxyloma suggest that some caution should be used when evaluating 14 C ages derived from these shells. AMS results for the Oxyloma shells show that nearly 1 in 3 aliquots contained at least some dead carbon. Although the amount was relatively minor, <7% of the total, dead carbon was present in Oxyloma shells more frequently than in either Catinella or Succinea shells. b Fig. 6. (continued). J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 527 107 5.2. Ca-limiting hypothesis Large gastropod shells (>20 mm in maximum dimension) routinely contain 20–30% dead carbon when living in habitats on carbonate terrain (e.g., Evin et al.,1980; Goodfriend and Stipp,1983), whereas the small shells measured in this study rarely contained more than w10%. We speculate that calcium may hold the clues to determining the reasons for the difference. Gastropod shell carbonate (aragonite) is composed of three elements – calcium, carbon, and oxygen. The latter two elements are readily available in the environments in which gastropods live and, therefore, cannot be considered as possible limiting factors for shell construction. In most settings, however, calcium is present inplants and water in lowconcentrations (typically parts per million). If small terrestrial gastropods can acquire enough calcium from their‘‘normal’’diet (plants, detritus, and water), then they may not have to consume carbonate rocks to supplement theircalcium intakewhen buildingtheir shells.Largertaxamay find it more difficult to obtain enough calcium from these sources without turning to carbonate rocks when they are available. Our results support this hypothesis, but only on a gross scale. There is clearly a significant difference in the amount of dead carbon incorporated in the shells of large taxa previously studied and the small taxa studied here. However, shell size alone is not the only factor to consider when evaluating the results within the small body size class. For example, in the present study, we did not observe a significant correlation between shell size and measured D14 C values (R2 ¼ 0.015). The largest taxon that we included in our analysis, H. occulta, averaged 15.6 mg per shell and contained approximately the same amount of dead carbon as P. muscorum and Vallonia shells, which averaged only 1.4 and 0.7 mg per shell, respectively. Similarly, we did not find a clear correlation between shell size and measured D14 C values even within a single family. For Succineidae, Catinella shells were generally the smallest, averaging 1.1 mg per shell and contained the least amount of dead carbon, and Table 1 Summary of 14 C results for modern gastropod shells. Limestone effecta (14 C yrs) Family Genus Species Aliquots Shells Negligibleb Maximumc Cochlicopidae Cochlicopa Cochlicopa lubrica 3 17 53% 650 Æ 390 Cochlicopa lubricella 3 17 100% – Cochlicopa morseana 4 16 100% – Discidae Discus Discus catskillensis 6 40 100% – Discus cronkhitei 6 47 100% – Discus macclintockii 3 5 100% – Discus shimeki 7 33 97% 430 Æ 240 Helicarionidae Euconulus Euconulus alderi 3 71 100% – Euconulus dentatus 3 34 68% 670 Æ 290 Euconulus fulvus 8 107 57% 730 Æ 300 Euconulus polygyratus 3 57 100% – Helicinidae Hendersonia Hendersonia occulta 13 13 46% 1210 Æ 250 Punctidae Punctum Punctum minutissimum 3 542 100% – Pupillidae Columella Columella columella 3 94 100% – Gastrocopta Gastrocopta pentodon 3 131 100% – Gastrocopta tappaniana 3 105 92% 540 Æ 360 Pupilla Pupilla blandi 9 90 67% 1000 Æ 270 Pupilla hebes 1 7 100% – Pupilla muscorum 8 80 38% 780 Æ 310 Pupilla sonorana 2 29 66% 1590 Æ 280 Vertigo Vertigo elatior 3 113 38% 500 Æ 280 Vertigo hannai 3 123 65% 280 Æ 270 Vertigo hubrichti 3 144 100% – Vertigo modesta 3 73 100% – Vertigo paradoxa 3 124 74% 380 Æ 240 Strobilopsidae Strobilops Strobilops affinis 3 36 100% – Strobilops labyrinthica 3 43 65% 520 Æ 320 Succineidae Catinella Catinella avara 9 99 100% – Catinella exile 15 316 97% 320 Æ 310 Catinella gelida 9 66 100% – Catinella vermeta 4 39 100% – Oxyloma Oxyloma retusa 17 136 65% 540 Æ 360 Oxyloma verrilli 7 47 57% 670 Æ 240 Succinea Succinea bakeri 8 27 100% – Succinea grosvernori 1 1 100% – Succinea indiana 4 6 17% 610 Æ 380 Succinea n.sp. ‘Minnesota A’ 1 1 100% – Succinea ovalis 19 50 100% – Succinea strigata 6 14 100% – Valloniidae Vallonia Vallonia cyclophorella 6 129 95% 1500 Æ 270 Vallonia gracilicosta 7 155 60% 1370 Æ 270 Vallonia perspectiva 5 221 68% 1010 Æ 380 Zonitidae Hawaiia Hawaiia minuscula 3 152 29% 340 Æ 270 Nesovitrea Nesovitrea binneyana 3 46 100% – Nesovitrea electrina 4 33 73% 710 Æ 240 Zonitoides Zonitoides arboreus 4 20 85% 530 Æ 240 a Defined as the theoretical difference between the measured and true 14 C ages for gastropods that incorporate the same amount of dead carbon in their shells as the aliquots measured here. These values are based on the difference between the modeled diet and shell carbonate D14 C values and converted into 14 C yrs. b Percent of shells measured by AMS that did not contain dead carbon from limestone or other carbonate rocks (i.e., the D14 C values for the shells were statistically indistinguishable from the modeled diet D14 C value). c Maximum limestone effect for a given taxon (given in 14 C yrs). Uncertainties are given at the 2s (95%) confidence level. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532528 108 Oxyloma shells averaged 2.1 mg per shell and contained the most dead carbon. Succinea shells were the largest, averaging 5.8 mg per shell, but were between Catinella and Oxyloma in terms of the amount of dead carbon in their shells (Table S1). The results presented here suggest that the Limestone Problem for small terrestrial gastropods is often negligible and always much less than the 20–30% dead carbon for larger taxa. However, there are additional factors that apparently influence the dietary intake of carbonate rocks of small terrestrial gastropods living side by side, which may include opportunistic behavior, variations in microhabitats, and the dietary needs or wants of individual gastropods. Table 2 14 C results for modern gastropod bodies and corresponding shells. Lab # AA # Taxon Sitea Lat ( N) Long ( W) Mass (mg) d13 C (vpdb) F14 Cb Shell D14 C Atmos D14 C Diet D14 C Limestone Effectc (14 C yrs) Bodies MU-109 80177 Succinea ovalis 1 44.819 68.723 4.22 À24.9 1.0547 Æ 0.0064 55 Æ 6 70 Æ 5 96 Æ 33 330 Æ 280 MU-114 80181 Succinea ovalis 2 48.410 94.819 8.64 À24.8 1.0996 Æ 0.0082 101 Æ 8 80 Æ 5 107 Æ 36 50 Æ 300 MU-108 80176 Succinea ovalis 3 48.906 96.027 3.79 À25.1 1.0946 Æ 0.0062 96 Æ 6 80 Æ 5 107 Æ 36 90 Æ 300 MU-111 80178 Succinea ovalis 4 47.874 96.422 11.05 À26.0 1.0877 Æ 0.0062 89 Æ 6 80 Æ 5 107 Æ 36 150 Æ 300 MU-117 80183 Oxyloma retusa 5 42.559 90.713 9.04 À26.1 1.0419 Æ 0.0064 42 Æ 6 52 Æ 5 73 Æ 25 250 Æ 210 MU-115 80182 Succinea ovalis 6 50.264 66.411 12.88 À25.0 1.0485 Æ 0.0072 49 Æ 7 57 Æ 5 79 Æ 27 250 Æ 230 MU-112 80179 Succinea strigata 7 64.858 147.862 4.01 À24.2 1.0740 Æ 0.0060 74 Æ 6 54 Æ 5 75 Æ 26 0 Æ 210 MU-113 80180 Succinea ovalis 8 48.866 94.843 6.88 À23.5 1.0833 Æ 0.0060 84 Æ 6 80 Æ 5 107 Æ 36 180 Æ 300 Shells MU-162 80928 Succinea ovalis 1 44.819 68.723 8.80 À10.0 1.0817 Æ 0.0048 82 Æ 5 70 Æ 5 96 Æ 33 110 Æ 270 MU-128 80194 Succinea ovalis 2 48.410 94.819 11.72 À9.8 1.1090 Æ 0.0084 110 Æ 8 80 Æ 5 107 Æ 36 0 Æ 300 MU-179 80942 Oxyloma retusa 5 42.559 90.713 13.08 À11.0 1.0425 Æ 0.0100 43 Æ 10 52 Æ 5 73 Æ 25 250 Æ 220 MU-141 80910 Succinea ovalis 8 48.866 94.843 8.91 À9.6 1.1316 Æ 0.0116 133 Æ 12 80 Æ 5 107 Æ 36 0 Æ 300 Uncertainties are given at the 2s (95%) confidence level. a Key to sites: 1 ¼ Brewer Boat Ramp, Maine; 2 ¼ Dave Pepin Homestead, Minnesota; 3 ¼ Duxby, Minnesota; 4 ¼ Huot Forest WMA, Minnesota; 5 ¼ Maquokata River Mounds, Iowa; 6 ¼ September Islands, Quebec; 7 ¼ University of Alaska – Fairbanks; 8 ¼ Zippel Bay State Park, Minnesota. b F14 C values are derived from the measured 14 C activity, corrected for fractionation, and account for decay that occurred between the time of collection and the AMS measurement. c Defined in Table 1. Table 3 14 C results for the Oxford East outcrops. Lab # AA # Taxon Na Mass (mg) d13 C (vpdb) F14 C 14 C age (ka) Calendar age (ka)b Pc Organics MU-212 82584 bark – 4.47 À24.6 0.0731 Æ 0.0026 21.02 Æ 0.28 25.15 Æ 0.34 1.00 MU-213 82585 bark – 3.28 À25.1 0.0767 Æ 0.0025 20.62 Æ 0.27 24.64 Æ 0.37 1.00 MU-214 82586 twig – 3.80 À24.9 0.0721 Æ 0.0028 21.13 Æ 0.32 25.28 Æ 0.55 1.00 MU-211 82583 wood – 3.82 À22.6 0.0743 Æ 0.0042 20.88 Æ 0.45 24.97 Æ 0.63 1.00 MU-210 82582 wood chip – 4.29 À23.6 0.0772 Æ 0.0025 20.58 Æ 0.26 24.60 Æ 0.40 1.00 Average 24.93 Æ 0.30 Gastropod shells MU-194 82567 Discus shimeki 2 9.20 À6.4 0.0758 Æ 0.0052 20.72 Æ 0.55 24.73 Æ 0.75 1.00 MU-195 82568 Discus shimeki 2 10.94 À6.9 0.0762 Æ 0.0052 20.68 Æ 0.54 24.67 Æ 0.74 1.00 MU-196 82569 Discus shimeki 2 14.99 À6.4 0.0745 Æ 0.0053 20.87 Æ 0.57 25.00 Æ 0.79 1.00 Average 24.80 Æ 0.17 MU-188 82561 Hendersonia occulta 1 12.41 À6.4 0.0805 Æ 0.0051 20.24 Æ 0.51 24.18 Æ 0.67 1.00 MU-189 82562 Hendersonia occulta 1 10.29 À6.1 0.0787 Æ 0.0055 20.42 Æ 0.56 24.37 Æ 0.68 1.00 MU-190 82563 Hendersonia occulta 1 18.42 À7.1 0.0777 Æ 0.0055 20.52 Æ 0.57 24.47 Æ 0.70 1.00 Average 24.34 Æ 0.15 MU-199 82572 Pupilla muscorum 8 10.86 À6.5 0.0760 Æ 0.0052 20.71 Æ 0.55 24.71 Æ 0.74 1.00 MU-200 82573 Pupilla muscorum 9 11.88 À6.4 0.0805 Æ 0.0052 20.24 Æ 0.52 24.18 Æ 0.68 1.00 MU-201 82574 Pupilla muscorum 10 14.13 À6.4 0.0809 Æ 0.0051 20.20 Æ 0.51 24.13 Æ 0.67 1.00 Average 24.34 Æ 0.32 MU-191 82564 Succineidae 8 10.58 À5.3 0.0712 Æ 0.0052 21.23 Æ 0.59 25.33 Æ 0.80 1.00 MU-192 82565 Succineidae 4 9.32 À5.5 0.0689 Æ 0.0056 21.49 Æ 0.65 25.81 Æ 0.94 1.00 MU-193 82566 Succineidae 15 9.53 À5.8 0.0759 Æ 0.0055 20.71 Æ 0.58 24.73 Æ 0.78 1.00 Average 25.29 Æ 0.54 MU-202 82575 Vallonia gracilicosta 15 10.72 À6.0 0.0776 Æ 0.0058 20.53 Æ 0.60 24.49 Æ 0.76 1.00 MU-203 82576 Vallonia gracilicosta 15 11.08 À6.2 0.0824 Æ 0.0058 20.05 Æ 0.56 23.92 Æ 0.66 0.98 Average 24.20 Æ 0.40 MU-205 82577 Vertigo hannai 30 10.46 À6.5 0.0750 Æ 0.0052 20.81 Æ 0.55 24.90 Æ 0.77 1.00 MU-206 82578 Vertigo modesta 15 11.47 À6.7 0.0751 Æ 0.0052 20.80 Æ 0.55 24.88 Æ 0.77 1.00 MU-207 82579 Vertigo modesta 15 14.14 À7.0 0.0771 Æ 0.0053 20.59 Æ 0.55 24.54 Æ 0.71 1.00 MU-208 82580 Vertigo modesta 20 15.28 À6.7 0.0729 Æ 0.0053 21.04 Æ 0.58 25.18 Æ 0.77 1.00 Average 24.86 Æ 0.32 Uncertainties for the raw and calibrated 14 C ages are given at the 2s (95%) confidence level. a Number of shells per aliquot. b Calibrated ages were calculated using CALIB v. 6.0.0, IntCal09.14C dataset; limit 50.0 calendar ka B P. Calibrated ages are reported as the midpoint of the calibrated range. Uncertainties are reported as the difference between the midpoint and either the upper or lower limit of the calibrated age range, whichever is greater. Multiple ages are reported when the probability of a calibrated age range exceeds 0.05. c P ¼ probability of the calibrated age falling within the reported range as calculated by CALIB. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 529 109 5.3. Small terrestrial gastropods and open-system behavior The present study includes only a single fossil locality, the Oxford East outcrops of southwestern Ohio, which we present here as a case study. Average ages of organic materials and fossil gastropod shells from the Oxford East outcrops were statistically indistinguishable; 24.93 Æ 0.30 ka for the organics and 24.73 Æ 0.44 ka for the shells. Six of the seven taxa that we analyzed (D. shimeki, P. muscorum, Succineidae, V. gracilicosta, V. hannai, and V. modesta) yielded average 14 C ages that are indistinguishable from the organic ages; the seventh (H. occulta) yielded ages that were slightly younger than the organic ages. The range of ages of the shell material, 1.8 ka, is significantly larger than the range of ages of the organics, 0.8 ka. If the dispersion of shell ages was related to opensystem behavior, then we would expect the ages to be systematically younger than the organic matter ages, which they are not. It may be that the small number of organic samples fails to adequately capture the full range of time represented by the sampled stratum. More work needs to be done on this subject, including additional comparisons of shell and organic ages from other sites, but the results from the Oxford East outcrops suggest that small terrestrial gastropod shells may behave as closed systems with respect to carbon in the American Midwest for at least the past w25 ka. 6. Summary and conclusions Fossil shells of small terrestrial gastropods are commonly preserved in Quaternary sediment across North America, including loess,wetland, glacial,andalluvial deposits, aswellas in sedimentsat many archeological sites. Their aragonitic shells contain w12% by weight carbon, and therefore contain sufficient carbon for 14 C dating. However, terrestrial gastropod shells in carbonate terrains are often avoided for 14 C dating because large taxa are known to incorporate dead carbon from limestone or other carbonate rocks when building their shells, which can cause their 14 C ages to be up to 3000 yrs too old. Previous studies suggested that small terrestrial gastropod shells may yield reliable 14 C agesin arid environments, but a systematic and comprehensive analysis was needed before ages derived from their shells could be considered reliable outside of the Desert Southwest. To this end, we measured the 14 C activity of 247 aliquots of modern shell material (3749 individual shells) from 163 localities across North America. Approximately 78% of the aliquots did not contain measurable amounts of dead carbon even though limestone orother carbonate rocks were readilyavailable at all sites, w12 of the aliquots contained between 5 and 10% dead carbon, and the remaining few (3% of the total) contained more than 10%. The average Limestone Effect for these samples was only w180 14 C yrs, which is significantly less than the 2000–3000 14 C yrs that previous researchers found for larger taxa. Assuming that the small gastropod shells behave as closed systems with respect to carbon after burial, they should yield reliable 14 C ages w78% of the time, and ages that are within w1000 yrs of the true age w97% of the time, regardless of the taxon analyzed, local bedrock type, climate, or environmental conditions. If fossil shells can be identified to the species level, then at least 23 different species should yield reliable 14 C ages if the modern shell data can be applied directly to the fossil record. The terrestrial gastropod family Succineidae is one of the most common gastropod taxa in North America. Unlike the other Fig. 7. Photograph of the section at the Oxford East outcrops and the calibrated ages obtained from the organic material and fossil gastropod shells. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532530 110 gastropods studied here, our 14 C data for Succineidae must be evaluated at the genus or even family level because species-level identification is based on soft-part morphology, which is rarely preserved in the fossil record. Based on the data from modern shells, the Succineidae family as a whole should yield reliable 14 C ages w85% of the time, and ages that are within w700 14 C yrs every time. At the genus level, Catinella should yield reliable 14 C ages w97% of the time, again assuming closed-system behavior, and ages that are within w300 14 C yrs of the true age every time. Similarly, Succinea shells should yield reliable 14 C ages w92% of the time and ages that are within w600 14 C yrs of the true age every time. Caution should be used when evaluating shells of the genus Oxyloma, however, as nearly 1 in 3 aliquots contained dead carbon, equivalent to a limestone effect of up to w700 14 C yrs. Fossil shells of small terrestrial gastropods recovered from welldated, late-Pleistocene sediments in the Midwest yielded ages that were statistically indistinguishable from ages obtained from wellpreserved plant macrofossils (wood, bark, plant remains). Although just one site, these results suggest that small terrestrial gastropod shells may behave as closed systems with respect to carbon over geologic timescales. More work on this subject is needed, but if our case study site is representative of other sites, then fossil shells of some small terrestrial gastropods, including at least five common genera, Catinella, Columella, Discus, Gastrocopta, and Succinea, should yield reliable 14 C ages, regardless of the local geologic substrate. Fossil shells of these and other small terrestrial gastropods are common in a wide range of Quaternary deposits in North America and, therefore, our results may have broad chronologic applications to Quaternary geology and New World archeology. Acknowledgments We thank K. Gotter for converting carbon dioxide samples to graphite at the JAR’s lab at Miami University and J. Quade and the University of Arizona Desert Laboratory for access to their vacuum extraction systems. We also thank J. Rosenbaum and D. Van Sistine of the U.S. Geological Survey for their help with Figs. 1 and 2, respectively. This manuscript benefited from constructive reviews from E. Ellis, G. Hodgins, D. Muhs, M. Reheis, J. Southon, and two anonymous reviewers. This project was funded by the National Science Foundation Sedimentary Geology and Paleobiology Competition, award #EAR 0614840. Additional support was provided to JSP by the U.S. Geological Survey’s Mendenhall postdoctoral research program. Appendix. Supporting information Supplementary data associated with this article can be found in the online version at doi:10.1016/j.quageo.2010.01.001. Editorial Handling by: A. Hogg References Balakrishnan, M., Yapp, C.J., 2004. Flux balance models for the oxygen and carbon isotope compositions of land snail shells. Geochimica et Cosmochimica Acta 68, 2007–2024. Barker, G.M., 2001. Gastropods on land: phylogeny, diversity, and adaptive morphology. In: Barker, G.M. (Ed.), Biology of Terrestrial Molluscs. CABI Publishing, Oxon, U.K, pp. 1–146. Brennan, R., Quade, J., 1997. Reliable Late-Pleistocene stratigraphic ages and shorter groundwater travel times from 14 C in fossil snails from the southern Great Basin. Quaternary Research 47, 329–336. Brovkin, V., Cherkinsky, A., Goryachkin, S., 2008. Estimating soil carbon turnover using radiocarbon data: a case study for European Russia. Ecological Modelling 216, 178–187. Burch, J.B., Pearce, T.A., 1990. Terrestrial Gastropoda. Soil Biology Guide. John Wiley & Sons, New York. Cameron, R.A.D., Pokryszko, B.M., 2005. Estimating the species richness and composition of land mollusc communities. Journal of Conchology 38, 529–547. Dell, A.M., 1991. Reconstruction of late Pleistocene paleoecology in southwestern Ohio from nonmarine gastropod assemblages. Unpublished M.S., University of Cincinnati, 97 p. Eckberg, M.P., Lowell, T.V., Stuckenrath, R., 1993. Late Wisconsin glacial advance and retreat patterns in southwestern Ohio, USA. Boreas 22, 189–204. Emberton, K.C., Pearce, T.A., Randalana, R., 1996. Quantitatively sampling land-snail species richness in Madagascan rainforests. Malacologia 38, 203–212. Evans, J.G., 1972. Land Snails in Archaeology. Seminar Press, New York, NY. Evin, J., Marechal, J., Pachiaudi, C., 1980. Conditions involved in dating terrestrial shells. Radiocarbon 22, 545–555. Frye, J.C., Willman, H.B., 1960. Classification of the Wisconsinan stage in the Lake Michigan glacial lobe. Illinois Geological Survey Circular 285, 1–16. Goodfriend, G.A., 1987. Radiocarbon age anomalies in shell carbonate of land snails from semi-arid areas. Radiocarbon 29, 159–167. Goodfriend, G.A., Hood, D.G., 1983. Carbon isotope analysis of land snail shells: implications for carbon sources and radiocarbon dating. Radiocarbon 25, 810–830. Goodfriend, G.A., Stipp, J.J., 1983. Limestone and the problem of radiocarbon dating of land-snail shell carbonate. Geology 11, 575–577. Goslar, T., Pazdur, M.F., 1985. Contamination studies on mollusk shell samples. Radiocarbon 27, 33–42. Grosjean, M., 1994. Paleohydrology of the Laguna Lejı´a (North Chilean Altiplano) and climatic implications for late-glacial times. Palaeogeography, Palaeoclimatology, Palaeoecology 109, 89–100. Guadinski, J.B., Trumbore, S.E., Davidson, E.E., Zheng, S., 2000. Soil carbon cycling in a temperate forest: radiocarbon-based estimates of residence times, sequestration rates, and partitioning of fluxes. Biogeochemistry 51, 33–69. Hsueh, D.Y., Karkauer, N.Y., Randerson, J.T., Xu, X., Trumbore, S.E., Southon, J.R., 2007. Regional patterns of radiocarbon and fossil fuel-derived CO2 in surface air across North America. Geophysical Research Letters 34, L02816. doi:10.1029/ 2006GL027032. Hua, Q., 2004. Review of tropospheric bomb 14 C data for carbon cycle modeling and age calibration purposes. Radiocarbon 46, 1273–1298. Hubricht, L., 1985. The distributions of the native land mollusks of the eastern United States. Fieldiana: Zoology 24, 1–191. Kerney, M., Cameron, R.A.D., 1979. Field Guide to the Land Snails of Britain and Northwestern Europe. Collins Press, London. Leighton, M.M., 1960. The classification of the Wisconsin Glacial stage of north central United States. Journal of Geology 68, 529–552. Lowell, T.V., 1995. The Application of Radiocarbon Age Estimates to the Dating of Glacial Sequences: an Example from the Miami Sublobe, Ohio, USA. Quaternary Science Reviews 4, 85–99. Manning, M.R., Lowe, D.C., Melhuish, W.H., Sparks, R.J., Wallace, G., Brenninkmeijer, C.A.M., McGill, R.C., 1990. The use of radiocarbon measurements in atmospheric studies. Radiocarbon 32, 37–58. Meijer, H.A.J., van der Plicht, J., Gislefoss, J.S., Nydal, R., 1995. Comparing long term atmospheric 14 C and 3 H records near Groningen, the Netherlands with Fruholmen, Norway and Izan˜a, Canary Islands 14 C stations. Radiocarbon 37, 39–50. Nekola, J.C., 2004. Terrestrial gastropod fauna of northeastern Wisconsin and the southern Upper Peninsula of Michigan. American Malacological Bulletin 18, 21–44. Nekola, J.C., 2005. Geographic variation in richness and shell size of eastern North American land snail communities. Records of the Western Australian Museum Supplement 68, 39–51. Oggier, P., Zschokke, S., Baur, B., 1998. A comparison of three methods for assessing the gastropod community in dry grasslands. Pedobiologia 42, 348–357. Pedone, V., Rivera, K., 2003. Groundwater-discharge deposits in Fenner Wash, eastern Mojave Desert. Geological Society of America Abstracts with Programs 35, 257. Pigati, J.S., Bright, J.E., Shanahan, T.M., Mahan, S.A., 2009. Late Pleistocene Paleohydrology Near the Boundary of the Sonoran and Chihuahuan Deserts, Southeastern Arizona, USA. Quaternary Science Reviews 28, 286–300. Pigati, J.S., Quade, J., Shanahan, T.M., Haynes Jr., C.V., 2004. Radiocarbon dating of minute gastropods and new constraints on the timing of spring-discharge deposits in southern Arizona, USA. Palaeogeography, Palaeoclimatology, Palaeoecology 204, 33–45. Placzek, C., Quade, J., Patchett, P.J., 2006. Geochronology and stratigraphy of late Pleistocene lake cycles on the southern Bolivian Altiplano: implications for causes of tropical climate change. Geological Society of America Bulletin 118, 515–532. Ponder, W.F., Lindberg, D.R., 1997. Towards a phylogeny of gastropod molluscs: an analysis using morphological characters. Zoological Journal of the Linnean Society 119, 83–265. Quarta, G., Romaniello, L., D’Elia, M., Mastronuzzi, G., Calcagnile, L., 2007. Radiocarbon age anomalies in pre- and post-bomb land snails from the coastal Mediterranean basin. Radiocarbon 49, 817–826. Reimer, P., Baillie, M., Bard, E., Bayliss, A., Beck, J., Blackwell, P., Ramsey, C.B., Buck, C., Burr, G., Edwards, R., Friedrich, M., Grootes, P., Guilderson, T., Hajdas, I., Heaton, T., Hogg, A., Hughen, K., Kaiser, K., Kromer, B., McCormac, F., Manning, S., Reimer, R., Richards, D., Southon, J., Talamo, S., Turney, C., van der Plicht, J., Weyhenmeyer, C., 2009. IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal B.P. Radiocarbon 51, 1111–1150. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532 531 111 Reimer, P.J., Brown, T.A., Reimer, R.W., 2004. Discussion: reporting and calibration of post-bomb 14 C data. Radiocarbon 46, 1299–1304. Riggs, A.C., 1984. Major carbon-14 deficiency in modern snail shells from southern Nevada springs. Science 224, 58–61. Romaniello, L., Quarta, G., Mastronuzzi, G., D’Elia, M., Calcagnile, L., 2008. 14 C age anomalies in modern land snails shell carbonate from Southern Italy. Quaternary Geochronology 3, 68–75. Rubin, M., Likins, R.C., Berry, E.G., 1963. On the validity of radiocarbon dates from snail shells. Journal of Geology 71, 84–89. Samos, G., 1949. Some observations on exchange of CO2 between BaCO3 and CO2 gas. Science 110, 663–665. Slota, P.J., Jull, A.J.T., Linick, T.W., Toolin, L.J., 1987. Preparation of small samples for 14 C accelerator targets by catalytic reduction of CO. Radiocarbon 29, 303–306. Stott, L.D., 2002. The influence of diet on the d13 C of shell carbon in the pulmonate snail Helix aspersa. Earth and Planetary Science Letters 195, 249–259. Stuiver, M., Polach, H.A., 1977. Reporting of 14 C data. Radiocarbon 19, 355–363. Stuiver, M., Reimer, P.J., 1993. Extended 14 C database and revised CALIB radiocarbon calibration program. Radiocarbon 35, 215–230. Tamers, M.A., 1970. Validity of radiocarbon dates on terrestrial snail shells. American Antiquity 35, 94–100. Torn, M.S., Vitousek, P.M., Trumbore, S.E., 2005. The influence of nutrient availability on soil organic matter turnover estimated by incubations and radiocarbon modeling. Ecosystems 8, 352–372. Wilbur, K.M., 1972. Shell formation in mollusks. In: Florkin, M., Scheer, B.T. (Eds.), Chemical Zoology, Mollusca, vol. 7. Academic Press, New York, pp. 103–145. Yates, T., 1986. Studies of non-marine mollusks for the selection of shell samples for radiocarbon dating. Radiocarbon 28, 457–463. Zhou, W., Head, W.J., Wang, F., Donahue, D.J., Jull, A.J.T., 1999. The reliability of AMS radiocarbon dating of shells from China. Radiocarbon 41, 17–24. J.S. Pigati et al. / Quaternary Geochronology 5 (2010) 519–532532 112 Abstract The occurrence of ten butterfly taxa (Clossiana eunomia dawsonii, Clossiana freija, Clossiana frigga, Clossiana titania, Coenonympha inornata, Erebia discoidalis, Incisalia augustinus, Lycaena dorcas, Lycaena epixanthe, Oeneis jutta) was analyzed within three acid peatland habitat types from the Lake Superior drainage basin of northwestern Wisconsin. Both first(nearest-neighbor spatial analysis) and second-order (Ripley’s K) spatial point process statistics were used to identify the extents over which each distribution pattern significantly deviated from random expectations. Versions of these tests were used that identified significant spatial pattern uncorrelated to habitat location and habitat preference. These analyses documented non-random occurrence patterns in seven species. Deviations from random were largely confined to two extents: <50 km and 70–100+ km. The majority of non-random patterns at <50 km extents were examples of aggregation, while the majority of non-random patterns noted at the 70–100+ km scale were examples of segregation. These results demonstrate that even for winged animals inside a limited landscape, spatially constrained processes can be important determinants of distribution. It is likely that metapopulation dynamics and dispersal limitation help explain why aggregation is dominant at small scales. The mechanisms underlying the predominance of segregation at large scales are less clear, but may be related to migration history and/or weak environmental gradients. Keywords Butterflies · Metapopulation · Nearest-neighbor spatial analysis · Ripley’s K · Spatial pattern Introduction Species occurrence patterns have often been described as a simple reflection of species niche requirements and underlying environmental conditions (MacArthur 1972; Tilman 1988). This idea requires the assumption that dispersal never limits distribution at regional scales, giving all species access to all potential habitats (Krebs 1985). Saur (1988) equated this concept with Beijerinck’s Law, which states that “everything is everywhere but the environment selects”. In such situations, species occurrences should not have significant spatial pattern beyond that exhibited by the physical environment. However, the spatial relationship of populations, independent of the physical environment, may also influence occurrence patterns. Supply-side ecology (Roughgarden et al. 1987) and the mass effect (Shmida and Ellner 1984) demonstrate how proximity to potential source populations can positively influence occurrence frequency. Additionally, analyses of distance decay (Nekola and White 1999) and competitive co-equivalency (Shmida and Ellner 1984) demonstrate how habitat isolation can negatively influence population frequency. For these reasons, patch size and isolation can correlate as significantly to patch occupancy as does the physical environment (Moilanen and Hanski 1998). When such spatial factors are at work, non-random spatial occurrence patterns should be evident independent of the physical environment. Identification of such patterns is an important first step in the recognition of situations where spatially constrained processes are (or have been) important (Diggle 1983). How frequently do populations deviate from random occurrence patterns, independent of their environment? One way of assessing this question is to compare distributions across a well-defined taxonomic group within a particular habitat in a given landscape. This allows for at least partial control of the environmental, geographic, temporal, and phylogenetic templates upon which distributions have developed. Butterflies of northern Wisconsin acid peatlands represent such a system. Recent and historical investigaJ.C. Nekola (✉) Department of Natural and Applied Sciences, University of Wisconsin-Green Bay, 2420 Nicolet Drive, Green Bay, WI 54311, USA e-mail: nekolaj@gbms01.uwgb.edu Tel.: +1-920-4652937 C.E. Kraft Department of Natural Resources, Cornell University, Ithaca, NY, USA Oecologia (2002) 130:53–61 DOI 10.1007/s004420100782 Jeffrey C. Nekola · Clifford E. Kraft Spatial constraint of peatland butterfly occurrences within a heterogeneous landscape Received: 21 March 2000 / Accepted: 19 July 2001 / Published online: 1 September 2001 © Springer-Verlag 2001 113 tions have documented ten species [Clossiana eunomia dawsonii (Nymphalidae), Clossiana freija (Nymphalidae), Clossiana frigga (Nymphalidae), Clossiana titania (Nymphalidae), Coenonympha inornata (Satyridae), Erebia discoidalis (Satyridae), Incisalia augustinus (Lycaenidae), Lycaena dorcas (Lycaenidae), Lycaena epixanthe (Lycaenidae), Oeneis jutta (Satyridae)], which live out their entire life cycles in these habitats (Masters 1971a, b, 1972; Ferge and Kuehn 1976; Kuehn 1983; Swengel 1995; Nekola 1998). Within the Lake Superior drainage basin of northwestern Wisconsin, three types of acid peatland habitats exist: muskeg, kettlehole, and coastal. Muskeg sites are dominated by open black spruce-cottongrass-wiregrass savanna, are relatively dry (except in the proximity of moats or lakes), and usually have an elevation similar to the surrounding uplands. Kettlehole peatlands are generally wetter, commonly contain floating sphagnum-leatherleaf mats, and are typically found fringing lakes or in kettlehole depressions. Coastal wetlands are limited to estuaries along the Lake Superior coast. Considerable similarity exists within the physical environment, habitat size, and flora of each peatland type, leading to the presence of very similar habitats throughout the region. However, subtle environmental and vegetation differences exist between peatland types, producing slightly different butterfly faunas within each (Nekola 1998). The following paper analyzes the occurrence patterns of these butterflies within these habitats in order to identify the frequency of non-random distributions that are uncorrelated with environment, and the spatial extents at which the majority of any such non-random patterns occur. These results identify potentially important spatial processes, and can be used to guide conservation strategy. Materials and methods Site selection Identification of all high quality peatlands within or adjacent to the Lake Superior drainage basin (Fig. 1) was accomplished by a fly-over of the study area on 20 May 1996 in a small aircraft. Approximately 5 h of flight time were required to survey the entire region at an altitude of 3,000–5,000 m in north-south transects running approximately 10 km apart. Other high-quality sites were identified through review of the Biological Conservation Database and discussions with the staff of the Wisconsin Department of Natural Resources Bureau of Endangered Resources. While an attempt was made to inventory all high-quality sites, some (like the Kakagon Sloughs) were not visited due to difficulty in political and/or physical access. As few such sites existed, and as they were spread across the study region, we assume that their absence from this analysis has not biased results. Field sampling The latitude-longitude location of each surveyed site was determined through digitization of USGS 7.5 min topographic quads using the ATLAS DRAW software package. The centroids for these sites were then converted to Zone 15 UTM coordinates using PC-ARCINFO. The peatland habitat type represented by each site (muskeg, kettlehole, or coastal) was also recorded. Muskeg peatlands (37 to- 54 Fig. 1 Distribution of surveyed peatland habitats in northwestern Wisconsin. The shaded area in the south of the region represents areas outside of the Lake Superior drainage basin Fig. 2 A Occurrence pattern of Clossiana eunomia dawsoni within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations 114 tal inventoried sites) were concentrated along the southern divide basin boundary in the western and eastern sides of the study area and were absent from the pitted outwash plain in the central region. Kettlehole peatlands (23 sites) were generally concentrated in the central outwash plain that lacked muskeg sites, although a few were also scattered throughout the eastern half. Coastal peatlands (10 sites) were limited to the Lake Superior shore in the Bayfield Peninsula and the Apostle Islands (Fig. 1). Each site was visited 4–7 times during the 1996 flight season. Surveys were only conducted during sunny, dry weather when temperatures were 18°C or higher. Surveys were made at weekly intervals early in the flight season when a number of closely spaced adult emergences occur. Later in the season, when species emergence periods were more temporally separated and individual populations were in flight for longer periods of time, the survey interval increased to 3 weeks. Each site visit lasted for 15–90 min, depending upon site size, and all encountered peatland butterfly taxa were recorded. Voucher specimens are housed in the first author’s collection at the University of Wisconsin-Green Bay. Nomenclature follows that of Miller and Brown (1981). Cumulative presence-absence species lists for all peatland taxa from each site were compiled from site visit lists combined with any additional collections archived at the Milwaukee Public Museum and element occurrence records maintained by the Wisconsin Bureau of Endangered Resources. As weather and logistical constraints prevented visitation of all sites during a given flight time, only sites inventoried during the flight duration for a given species were used for statistical analysis. Flight times for a given species were determined by noting the earliest and latest dates of adult encounters in the study region during 1996. Statistical analyses Two general approaches exist to analyze spatial point patterns: first-order statistics consider the mean distances between points, while second-order statistics consider the variance in these distances. One of the most commonly used first-order statistics is nearest-neighbor analysis (Diggle 1983; Manly 1991), while one of the most commonly used second-order statistics is Ripley’s K analysis (Ripley 1977; Haase 1995). As little has been written to suggest whether first- or second-order statistics are more effective at identifying non-random patterns, both were used. Both methods, as commonly employed, generate random null models from the assumption that occurrence is possible anywhere within the study region (e.g. Diggle 1983; Manly 1991). This is not appropriate for acid peatland butterflies, which cannot reside outside of peatland habitats. If peatlands are non-randomly distributed (as is the case in northwestern Wisconsin), it will be impossible to determine from such null models whether a non-random occurrence pattern for a given species is due to habitat distribution or other spatially constrained processes. Additionally, the physical environment may lead to significant non-random pattern when differences exist between habitats (as is also the case for peatland butterflies in our study). To more clearly test for non-random distributions uncorrelated to such environmental pattern, the firstand second-order statistics used in this study constrained random null models to actual peatland locations and to the actual occurrence frequencies of each species within each peatland type. Nearest-neighbor spatial analysis (NNSA; Davis et al. 2000) was chosen as the first-order statistic. This method is related to the nearest-neighbor method described by Manly (1991), in which the average distance between 1st, 2nd, 3rd to n–1 nearest neighbors (where n equals the total number of occurrences of that species in the landscape) is calculated and compared to expected values generated from random distributions. Unlike Manly’s test, the NNSA method generates random expectations based upon the subsampling of a larger finite set of points within a landscape. This permits factoring out spatial pattern caused solely by habitat distribution. This method also constrains random subsamples to be drawn at the same frequencies for which taxa were observed in that given habitat type. This allows any spatial pattern caused by habitat differences to be factored out (Davis et al. 2000). NNSA null expectations were based on 5,000 Monte Carlo simulations. Two-tailed 95% confidence intervals were generated from this process, and plotted versus observed nearest-neighbor distances (Figs. 2, 3, 4, 5, 6, 7, 8, 9, 10, 11). From these, the total 55 Fig. 3 A Occurrence pattern of Clossiana freija within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations Fig. 4 Occurrence pattern of Clossiana frigga within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time 115 range of distances were determined over which observed distributions deviated from random. Ripley’s K (Ripley 1977) was chosen as the second-order statistic. This test evaluates the expected number of points within a distance t of an arbitrary point in the area being evaluated (Haase 1995). If occurrences are randomly distributed, the expected value of K(t)=πt2. The unbiased estimate of K(t) for an observed spatial point pattern is: where n is the number of points in the landscape area A; uij is the distance between events i and j; It(u), the counter variable, equals 1 if u<=t and 0 if u>t; wij is the proportion of the circumference of a circle centered at the event i with radius uij lying within A; and the summation is over all pairs of a given species occurrence (Ripley 1977). Random expectations for K(t) were based upon a random subsample of n peatlands in the landscape, with n representing the number of species occurrences, as is outlined in Kraft et al. (in press). Randomly chosen sites were selected according to the actual proportion of sites colonized by each species within each peatland type. In this way, spatial pattern due to habitat distribution and preference were factored out. S-Plus was used for estimating Ripley’s K (Venables and Ripley 1997). Null expectations were based on 5,000 Monte Carlo simulations. Two-tailed 95% confidence intervals were generated and plotted against observed values (Figs. 2, 3, 4, 5, 6, 7, 8, 9, 10, 11). From these, the total range of distances was determined over which observed deviated from random. Because our null expectations are based upon randomly selected subsets of given peatland sites, the expectation that K(t) will approximate πt2 is not valid. Thus, the boundary issues considered by Haase (1995) do not apply to our analysis, as occurrences are constrained to actual peatland locations. Any such boundary-based biases should be accounted for by our use of Monte Carlo simulations to estimate the 95% confidence interval for K(t). Results were summarized across all species via histograms documenting the total number of significant aggregations and segregations at each analyzed distance for both methods (ca. 0–150 km for NNSA, 0–91 km for Ripley’s K). 56 Fig. 5 A Occurrence pattern of Clossiana titania within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations Fig. 6 A Occurrence pattern of Coenonympha inornata within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations Fig. 7 Occurrence pattern of Erebia discoidalis within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time 116 Results Two taxa (Clossiana frigga and Erebia discoidalis) were only twice encountered. Their populations were limited to muskeg sites in the west and east. Although their distributions have been mapped for sake of completeness (Figs. 4, 7), their few occurrences precluded further statistical analysis. Significant deviations from random were noted in the occurrences of seven of the remaining eight taxa. The only species which did not significantly deviate from random in either test was Coenonympha inornata, which was located at nine scattered sites (4 muskeg, 1 kettlehole, 4 coastal; Fig. 7). Clossiana eunomia dawsonii was located at 34 sites (14 muskeg, 14 kettlehole, 4 coastal), with the majority being located in the western half of the study region (Fig. 2). NNSA identified significant aggregation in these occurrences from the 6th–14th (min. P<0.0005; 21–44 km), 25rd–27th (min. P=0.008; 73–77 km), and 29th (P=0.021; 85 km) nearest neighbors. Ripley’s K identified significant clustering (min. P<0.0005) in these occurrences from 1–48 km and 67–87 km extents. Clossiana freija was located at 12 muskeg sites, with populations being restricted to the far west and east (Fig. 3). NNSA identified significant aggregation over the 1st–2nd (min. P=0.012; 5–11 km) and 8th (P=0.017; 47 km) nearest neighbors. Significant segregation was noted at the 11th (P=0.001; 150 km) nearest neighbor. Ripley’s K identified significant aggregation from 4–55 km (min. P<0.0005), and significant segregation from 69–91 km (min. P<0.0005). Clossiana titania was located at 6 muskeg sites in the far west (Fig. 5). NNSA identified significant aggre- 57 Fig. 8 A Occurrence pattern of Incisalia augustinus within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations Fig. 9 A Occurrence pattern of Lycaena dorcas within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations 117 gation (minimum p<0.0005) across all neighbors (5– 15 km). Too few occurrences were present to allow calculation of Ripley’s K. Incisalia augustinus was located at 39 sites (30 muskeg, 8 kettlehole, 1 coastal), which were most frequently encountered in the west and east (Fig. 8). NNSA identified significant segregation (min. P<0.0005) from the 22nd–38th nearest neighbors (81–133 km). Ripley’s K identified significant aggregation in occurrence at 5 km (P=0.023) and from 10–44 km (min. P<0.0005), and significant segregation from 61–91 km (min. P<0.0005). Lycaena dorcas was located at 13 muskeg sites, with the majority of populations being found in the eastern half (Fig. 9). NNSA identified significant (P=0.003) segregation at the 3rd nearest neighbor (40 km). Ripley’s K identified significant aggregation in its occurrence from 16–60 km (min. P<0.0005), and significant segregation from 81–82 km (min. P=0.013) and 86–91 km (min. P=0.006). Lycaena epixanthe was located at 23 sites (7 muskeg, 7 kettlehole, 9 coastal), with populations sporadically occurring throughout the entire region (Fig. 10). No significant deviations from random were noted by NNSA. However, Ripley’s K identified significant segregation (min. P<0.0005) in its occurrences from 3 to 21 km. Oeneis jutta was located at 18 sites (15 muskeg, 3 kettlehole), with populations being noted in the west and south central (Fig. 11). NNSA identified significant aggregation from the 3rd–5th (minimum p=0.009; 15–23 km) nearest neighbors. Ripley’s K identified significant aggregation from 5–8 km (min. P=0.008), 10 km (P=0.022), 13 km (P=0.022), 16 km (P=0.024), and from 18–49 km (min. P=0.001). Comparison of these results across species (Fig. 12) show that significant deviation from random was most apparent at three extents. At <20 km, NNSA identified non-random aggregation for two taxa (C. freija, C. titania), while Ripley’s K identified significant aggregation 58 Fig. 10 A Occurrence pattern of Lycaena epixanthe within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations Fig. 11 A Occurrence pattern of Oeneis jutta within the study region. Black circles represent occupied sites; open circles represent sites vacant during flight time. B Black line shows distance to nearest neighbor for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo NNSA simulations. C Black line shows K(t) for observed distribution of occupied sites; dashed lines show upper and lower 95% confidence intervals from 5,000 Monte Carlo Ripley’s K simulations 118 within each peatland type, possessing similar (if not identical) microhabitats, vascular plant, and bryophyte species. Additionally, given the small total extent of the sample region (ca. 150 × 75 km), it is unlikely that strong climatic gradients exist. Thus, we believe that the patterns of peatland butterfly occurrence are likely related to other factors. Assuming that peatlands within a given type have approximately similar environments for these butterflies, the existence of non-random occurrence patterns independent of habitat location and preference would be unlikely if dispersal allowed equal access to all suitable sites. The frequent presence of significant non-random occurrence pattern in these species suggests that within this landscape spatial limitation of butterfly dispersal is common. Inference of spatial process from spatial pattern The local clustering of occurrences is not surprising as many ecological processes show some degree of spatial autocorrelation over relatively small extents (Shmida and Ellner 1984; Burrough 1986; Okubo and Levin 1989). In particular, limitations on butterfly movement contribute to patchy occurrences in a number of European taxa, including Clossiana (Boloria) aquilonaris (Mousson et al. 1999), C. (Proclossiana) eunomia (Nève et al. 1996a), Hesperia comma (Hill et al. 1996), and Melitaea cinxia (Hanski et al. 1994). When dispersal limitation is coupled with other aspects of metapopulation dynamics, local aggregations of butterfly occurrence, independent of the physical environment, are often evident. For instance, M. cinixa occurrences on the main Åland island in southwestern Finland demonstrate what visually appears to be strong aggregation over 1–10 km extents among appropriate habitat patches (Moilanen and Hanski 1998). Occurrences of H. comma in the North Downs, Surrey, UK, also appear to be aggregated over 2–6 km extents within appropriate sites (Hill et al. 1996). The metapopulations of all five edaphically-restricted butterfly taxa of Sierra Nevada (California) serpentine barrens also demonstrate apparent spatial aggregation in their occurrences among known sites (Gervais and Shapiro 1999). Much of the previous work on occurrence patterns in Lepidoptera has focused on the factors predicting habitat occupancy or vacancy (Ehrlich and Murphy 1987; Thomas et al. 1992; Hanski et al. 1994). While these studies have elucidated the processes influencing individual patches, none have investigated the spatial patterns in occurrence generated by these processes across an entire landscape. Analyzing such patterns may prove a useful tool for identifying those species whose distribution is influenced by spatial processes (Hanski 1999). For instance, C. eunomia is known to exist in discrete metapopulations within Belgium and France, with individuals capable of traversing up to 8 km of unfavorable habitat within a season (Nève et al. 1996b). Our analyses demonstrate strong aggregation for this species at relatively small (<50 km) scales. A previous investigation on 59 Fig. 12 Frequency of non-random aggregation and segregation patterns from 0–100 km extents for both NNSA and Ripley’s K analyses in four taxa (C. eunomia dawsonii, C. freija, I. augustinus, O. jutta). Additionally, Ripley’s K identified significant segregation over this range for L. epixanthe. From 20–50 km, NNSA identified aggregation for three taxa (C. eunomia dawsonii, C. freija, O. jutta), while Ripley’s K identified aggregation for five taxa (C. eunomia dawsonii, C. freija, I. augustinus, L. dorcas, O. jutta). Significant segregation was identified at this scale by NNSA for L. dorcas. Few non-random occurrence patterns were identified by either method from 50–70 km. However, at larger scales both methods identified significant aggregation in C. eunomia dawsonii. Significant segregation at this scale was identified by both methods for C. freija and I. augustinus, and by Ripley’s K for L. dorcas. Discussion Perhaps the most obvious and important finding from these analyses is the simple realization that in northwestern Wisconsin, across a broad range of spatial scales (0–90 km), most acid-peatland butterflies appear to possess non-random occurrences independent of habitat location and preference. As we did not sample physical environmental variables from these sites (e.g. soil and water chemistry, local climate, amount and type of plant cover), it is possible that some of these distribution patterns may be related to such factors. If present, however, these factors must be cryptic, as habitats appeared to be homogeneous 119 Coenonympha tullia (often lumped with C. inornata) distribution in Northumberland, UK, documented that occurrences were more strongly related to habitat quality than geographic factors (Dennis and Eales 1999). In northwestern Wisconsin, this species was the only one that exhibited random occurrence patterns at all scales. These examples, in conjunction with our analysis of spatial distribution pattern, illustrate how analyses of spatial pattern can reflect potential processes underlying organism distribution. While Real and McElhaney (1996) note that spatial pattern analysis, alone, cannot distinguish between all processes capable of producing a given pattern, we believe that the analysis of spatial pattern provides an instructive first step that can help improve field experiments or modeling efforts used to evaluate the processes affecting distribution. It is not clear for the majority of the fauna why aggregation is essentially limited to <50 km extents. This is especially interesting given the fact that many of these species appear to share similar niches: in the region Clossiana freija, C. eunomia, C. titania, L. dorcas, and L. epixanthe all are obligate cranberry consumers, while Coenonympha inornata, E. discoidalis, and O. jutta appear limited to cottongrass and/or wiregrass (Nekola 1998). While these species might thus be expected to demonstrate some form of competitive exclusion, they instead appear to show largely coincident occurrence patterns at regional scales. Such correspondence between occurrence patterns is likely related to factors that have equally influenced all taxa, such as habitat and host plant distribution. If related to cryptic environmental gradients, the location of colonies may be expected to remain static over time. However, if shifting metapopulation dynamics are responsible, the location of occurrence clusters may change over ecological time scales within the landscape. Another enigma is the presence of significant smallscale segregation in occurrences of L. epixanthe. While it is possible that colonies may have become segregated due to competitive interactions, such processes are generally thought to be limited to small extents within sites (Shmida and Ellner 1984). It is difficult to envision how such mechanisms could operate at distances up to 20 km. A number of other spatial mechanisms may underlie larger scale (70–150 km) segregated spatial distributions. Migration history may provide one explanation. For instance, migration of C. eunomia dawsonii out of Minnesota (beyond the study area boundaries) could help account for its higher occurrence frequencies in the west. Likewise, the absence of species from the center (e.g., C. freija, L. dorcas, O. jutta) could be due to dual migration pathways from peatland-rich landscapes to the west and east. However, these patterns could also be related to cryptic soil and water chemistry or climate gradients that our statistical methods could not control for. In this event, identification of such cryptic gradients would be a useful result, given their difficulty in direct detection. Teasing apart the historical and environmental influences on distribution at large scales may prove difficult due to covariation between these factors. Population genetics analyses may provide one possible avenue to distinguish between them. If historical factors are responsible for the observed large-scale pattern, it seems likely that significant genetic differences may occur across the study region due to differences in initial population source pools. However, if dispersal is not limiting and contemporaneous environmental gradients are responsible for the observed pattern, differences in allele frequency should be diminished from more thorough genetic mixing. Conservation implications Relatively few populations (e.g., <25 stations) of these taxa (except I. augustinus) have been previously reported from Wisconsin (Ebner 1970; Masters 1971a, b, 1972; Ferge and Kuehn 1976; Kuehn 1983; Swengel 1995). As acid peatlands are frequent in the northern Wisconsin landscape, it has been generally assumed that these species were simply undercollected and would eventually be found within most appropriate sites (Wisconsin Rare Butterfly Working List 1996). As such, none of these species have been granted endangered or threatened species protection within the state (Wisconsin Endangered Species List 1999). The existence of complex spatial structure in the occurrence of these taxa suggests that such expectations may be false. The aggregation of most inventoried taxa indicates that, at some scale, high-quality sites will be unoccupied. Modifications to sites within limited areas of occurrence would have profound implications to the long-term survival of those species across the entire landscape. For example, C. freija and C. titania are essentially restricted to a small cluster of sites in the far west of the study region. Unfortunately, peatlands in this same area were observed to be undergoing the heaviest “all terrain vehicle” abuse within the entire region. Thus, even though appropriate habitat for these species exists throughout the study region, the correspondence of high levels of disturbance in the region where most colonies occur suggests that their continued existence within this landscape may be in jeopardy. In this region, the vulnerability of peatland butterfly species appears to be generally unrelated to habitat frequency. Such patterns may be common at range edges where increased environmental stress and patch isolation lead to lower habitat occupancy rates (Thomas et al. 1992). When species distribution does not simply mimic habitat abundance, determination of population vulnerability to extirpation can only be assessed by conducting geographically systematic inventories and documenting the actual distribution of taxa. Additionally, repeated monitoring of distribution may be necessary if shifting metapopulations exist. Comparison of first- and second-order spatial point statistics Given the different assumptions and algorithms underlying these methods, the general robustness of the results 60 120 suggests that the underlying patterns are quite strong. Only one species (L. dorcas) exhibited a major discrepancy in outcomes, with NNSA identifying segregation in the same 30–40 km extent where Ripley’s K showed strong aggregation. Based upon visual inspection of this taxon’s occurrences, the Ripley’s K result appeared more accurate. Perhaps NNSA was unable to identify aggregation for this species as occurrences were concentrated on a region of widely separated sites. Comparing results across species, Ripley’s K identified more non-random distributions at a given extent, and identified larger regions for non-random distribution, as compared to NNSA. It would thus appear that NNSA is the more conservative of the two tests. However, visual inspection of maps suggests that NNSA might be overly conservative, as it failed to identify areas of apparent aggregation or segregation at some scales (e.g., C. freija, I. augustinus, L. epixanthe, O. jutta). A more rigorous comparison of these methods would help to better determine their relative strengths and weaknesses. Acknowledgements The computer code used for conducting the NNSA analyses was graciously provided by Jennifer Davis. Advice on the implementation of Ripley’s K using S-Plus was provided by Patrick Sullivan. Eric North helped in field sampling and Heather Voster collected data from the Milwaukee Public Museum collections. Patricia Schneider piloted the 20 May air reconnaissance of the study region. Peter White and Michael Draney read and made useful comments regarding earlier drafts. This research was funded through a grant from the Lake Superior Coastal Wetland Inventory which was administered by the Wisconsin Department of Natural Resources and the U.S. Environmental Protection Agency. References Burrough PA (1986) Principles of geographical information systems for land resources assessment. Oxford University Press, New York Davis JH, Howe RW, Davis GJ (2000) A multi-scale spatial analysis method for point data. Landscape Ecol 15:99–114 Dennis RLH, Eales HT (1999) Probability of site occupancy in the large heath butterfly Coenonympha tullia determined from geographical and ecological data. Biol Conserv 87:295–301 Diggle PJ (1983) The analysis of spatial point patterns. Academic Press, New York Ebner JA (1970) Butterflies of Wisconsin. Milwaukee Public Museum Popular Science Handbook 12, Milwaukee, Wisconsin Ehrlich PR, Murphy DD (1987) Conservation lessons from longterm studies of checkerspot butterflies. Conserv Biol 1:122–131 Ferge LA, Kuehn RM (1976) First records of Boloria frigga (Nymphalidae) in Wisconsin. J Lepid Soc 30:233–234 Gervais BR, Shapiro AM (1999) Distribution of edaphic-endemic butterflies in the Sierra Nevada of California. Glob Ecol Biogeogr Lett 8:151–162 Haase P (1995) Spatial pattern analysis in ecology based on Ripley’s K-function: introduction and methods of edge correction. J Veg Sci 6:575–582 Hanski I (1999) Habitat connectivity, habitat continuity, and metapopulations in dynamic landscapes. Oikos 87:209–219 Hanski I, Kuussaari M, Nieminen M (1994) Metapopulation structure and migration in the butterfly Melitaea cinxia. Ecology 75:747–762 Hill JK, Thomas CD, Lewis OT (1996) Effects of habitat patch size and isolation on dispersal by Hesperia comma butterflies: implications for metapopulation structure. J Anim Ecol 65: 725–735 Kraft CE, Sullivan PJ, Karatayev AY, Burlakova LE, Nekola JC, Johnson LE, Padilla DK (in press) Landscape patterns of zebra mussel inland lake invasions: evaluating dispersal from distribution. J Appl Ecol Krebs CJ (1985) Ecology. Harper and Row, New York Kuehn RM (1983) New Wisconsin butterfly records. J Lepid Soc 37:228–235 MacArthur RH (1972) Geographical ecology. Harper and Row, New York Manly BFJ (1991) Randomization and Monte Carlo methods in biology. Chapman and Hall, New York Masters JH (1971a) Ecological and distributional notes on Erebia discoidalis (Satyridae) in the north central states. J Res Lepid 9:11–16 Masters JH (1971b) First records of Boloria eunomia (Nymphalidae) in Wisconsin. J Lepid Soc 25:149 Masters JH (1972) Critical comments on James A. Ebner’s “Butterflies of Wisconsin.” In: Keuhn RM, Masters JH (eds) Wisconsin butterflies. Mid-Continent Lepidoptera Series 59, pp 5-11 Miller LD, Brown FM (1981) A catalogue/checklist of the butterflies of America north of Mexico. Lepidopterist’s Society Memoir 2 Moilanen A, Hanski I (1998) Metapopulation dynamics: effects of habitat quality and landscape structure. Ecology 79:2503– 2515 Mousson L, Nève G, Baguette M (1999) Metapopulation structure and conservation of the Cranberry Fritillary Boloria aquilonaris (Lepidoptera, Nymphalidae) in Belgium. Biol Conserv 87:285–293 Nekola JC (1998) Butterfly (Lepidoptera: Lycaeniidae, Nymphalidae, and Satyridae) faunas of three peatland habitat types in the Lake Superior drainage basin of Wisconsin. Great Lakes Entomol 31:27–37 Nekola JC, White PS (1999) The distance decay of similarity in bioigeography and ecology. J Biogeogr 26:867-878 Nève G, Mousson L, Baguette M (1996a) Adult dispersal and genetic structure of butterfly populations in a fragmented landscape. Acta Oecol 17:621–626 Nève G, Barascud B, Hughes R, Aubert J, Descimon H, Lebrun P, Baguette M (1996b) Dispersal, colonization power and metapopulation structure in the vulnerable butterfly Proclossiana eunomia (Lepidoptera: Nymphalidae). J Appl Ecol 33:14–22 Okubo A, Levin SA (1989) A theoretical framework for data analysis of wind dispersal of seeds and pollen. Ecology 70:329– 338 Real LA, McElhany P (1996) Spatial pattern and process in plantpathogen interactions. Ecology 77:1011–1025 Ripley BD (1977) Modeling spatial patterns. J R Stat Soc Ser B 39:172–212 Roughgarden J, Gaines SD, Pacala SW (1987) Supply side ecology: the role of physical transport processes. Br Ecol Soc Symp 27:491–518 Saur JD (1988) Plant migration. University of California Press, Berkeley Shmida A, Ellner S (1984) Coexistence of plant species with similar niches. Vegetatio 58: 29–55 Swengel AB (1995) Butterflies of Northwestern Wisconsin. Madison Audubon Society, Madison Thomas CD, Thomas JA, Warren MS (1992) Distributions of occupied and vacant butterfly habitats in fragmented landscapes. Oecologia 92:563–567 Tilman D (1988) Plant strategies and the dynamics and structure of plant communities. Princeton University Press, Princeton Venables WN, Ripley BD (1997) Modern applied statistics with S-Plus. Springer, Berlin Heidelberg New York Wisconsin Endangered Species List (1999) Wisconsin Department of Natural Resources, Madison Wisconsin Rare Butterfly Working List (1996) Bureau of Endangered Resources, Wisconsin Department of Natural Resources, Madison 61 121 2459 CONCEPTS & SYNTHESIS EMPHASIZING NEW IDEAS TO STIMULATE RESEARCH IN ECOLOGY Ecology, 80(8), 1999, pp. 2459–2473 ᭧ 1999 by the Ecological Society of America PALEOREFUGIA AND NEOREFUGIA: THE INFLUENCE OF COLONIZATION HISTORY ON COMMUNITY PATTERN AND PROCESS JEFFREY C. NEKOLA1 Curriculum in Ecology, 229 Wilson Hall, CB 3275, University of North Carolina, Chapel Hill, North Carolina 27514 USA Abstract. Two types of biological refugia (habitats that support populations not able to live elsewhere in a landscape) can be defined from relative refugium age as compared to surrounding matrix age; paleorefugia are now-fragmented relicts of a formerly widespread matrix community, whereas neorefugia have formed more recently than the matrix. This difference should make extinction a relatively more important process in determining species occurrence in paleorefugia, whereas immigration should be relatively more important in neorefugia. Based on these differences, a series of eight a priori predictions relating to the diversity and distribution patterns for the biota of such sites can be generated: (1) the slope of the species–area relationship, and amount of variance explained by it, should be greater in paleorefugia as compared to neorefugia; (2) the negative relationship between habitat isolation and species richness should be stronger in neorefugia as compared to paleorefugia; (3) species richness should be expected to decrease over time in paleorefugia, but to increase over time in neorefugia; (4) the inverse correlation between site distance and community similarity (distance decay) should be stronger in neorefugia as compared to paleorefugia; (5) neorefugia should be enriched in highly vagile species relative to paleorefugia, whereas paleorefugia should be rich in less vagile species relative to neorefugia; (6) geographic factors should be more important predictors of species occurrence for neorefugia than for paleorefugia; (7) paleorefuge sites should possess more and stronger correlations between community composition and environmental covariables (such as soil chemistry, climate, etc.) as compared to neorefuge sites; and (8) the number of competitive co-equivalents held within a system of neorefugia should be greater then the number held within a series of paleorefugia. The most readily testable predictions (numbers 1, 2, 4, and 5) were evaluated by comparing species-richness and community-composition patterns within two northeastern Iowa refugia: algific talus slopes (paleorefugia) and fens (neorefugia). Results from these tests were consistent with predictions. These results illustrate that colonization history may influence contemporaneous species diversity and community-composition patterns. They also suggest that (1) equilibrium has yet to be achieved in the example systems after 5000– 10 000 yr, (2) the ecological-biogeographic debate centered around the mutual exclusivity of vicariance and dispersal is intrinsically flawed, and (3) optimum reserve-design strategies for biodiversity protection within paleorefuge and neorefuge systems will differ. Key words: biodiversity; biogeography, historical; colonization history, influence on community pattern; conservation biology; dispersal; fens; Iowa (USA), northeastern; island biogeography; refugia, paleorefugia vs. neorefugia; species richness; talus slopes, algific; vicariance. INTRODUCTION Past environmental changes have caused repeated fluctuations in species and habitat distributions that can Manuscript received 6 February 1998; revised 10 November 1998; accepted 27 November 1998; final version received 21 December 1998. 1 Present address: Department of Natural Applied Sciences, University of Wisconsin, 2420 Nicolet Drive, Green Bay, Wisconsin 54311 USA. E-mail: nekolaj@uwgb.edu lead to formation of island-like habitats and isolated populations through two contrasting pathways. They may represent remnants of once-more-widespread distributions that have become fragmented, or they may represent the de novo development of habitats or populations in a landscape where they were previously absent. Borrowing terminology used for endemic species (Stebbins and Major 1965, Krukeberg and Rabinowitz 1985), I refer to habitats that have been colonized through these processes, respectively, as ‘‘paleo- 122 2460 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis FIG. 1. Diagrammatic representation of the development of paleorefugia and neorefugia. At time period A, environmental conditions are conducive to the existence of a particular habitat (shaded). After critical environmental change (time period B), this habitat has retreated completely from the focal landscape (inner box) in the case of neorefugia but persists as isolated patches in the case of paleorefugia. By time period C, the environment has changed again, allowing the habitat to exist again within portions of its original range. At time period C, paleorefugia are found within essentially the same microenvironmentally buffered patches present at time period B. Neorefugia, in contrast, have formed as appropriate microhabitats develop in isolation from the main habitat area. refugia’’ and ‘‘neorefugia.’’ More specifically, paleorefugia represent habitats that are older than the surrounding biological matrix, whereas neorefugia represent habitats that are younger than the surrounding biological matrix. Such habitats are considered refugia as they support communities or populations unable to survive elsewhere in the landscape (i.e., Pielou 1979). As these terms reflect the relative age of a refugium to its matrix, a paleorefugium in one landscape may be of more recent origin than a neorefugium in another. It is also possible that a particular habitat type that developed as a paleorefugium in one landscape may be a neorefugium elsewhere. In this paper, I ask how such paleorefugia and neorefugia would be expected to differ in their contemporaneous species richness, species distribution, and community-composition patterns. I develop eight predictions about these differences and then investigate the four most easily testable ones with data collected from two types of insular habitats that co-occur in northeastern Iowa. THE INFLUENCE OF EXTINCTION AND IMMIGRATION The relative importance of extinction and immigration processes varies greatly between paleorefugia and neorefugia. First, consider the development of paleorefugia sites. Before environmental change, a given community is continuous over a wide geographic extent, harboring a diverse and well-mixed biota potentially near equilibrium with the environment. Following environmental change, the habitat becomes restricted to a few discrete patches in the landscape (Fig. 1). As the community becomes fragmented, habitat size decreases and isolation between fragments increases, leading to increased extinction rates on individual fragments (MacArthur and Wilson 1967, Diamond 1975, Bierrgaard et al. 1992). The richness of individual sites falls, with larger fragments retaining more taxa than smaller ones (Diamond 1975). Although some recolonization from nearby sites undoubtedly occurs (the rescue effect of Brown and Kodric-Brown [1977]), this should not greatly influence large-scale compositional pattern as nearby sites should have been originally colonized from similar species pools. In contrast, neorefugia sites are created through the development of novel environmental conditions in a landscape (compare time periods B to C in Fig. 1). Consequently, the composition of neorefugia must be a product of immigration. Factors that effect immigration, such as habitat isolation or age, should therefore exert influence on community composition and species richness (Carlquist 1974, Diamond 1975, Bush and Whittaker 1993). Although extinction certainly occurs on these sites as newly recruited populations disappear, this process should be of limited importance as populations can only become extinct if a successful immigration has previously occurred, and as incomplete dispersal between sites will limit rates of competitive exclusion (Shmida and Ellner 1984). The differential contributions of extinction and immigration have been empirically documented for various insular habitats. Fernald (1925) hypothesized (incorrectly) about the role played by unglaciated nuntacks (reputed paleorefugia) in the development of eastern North American species ranges. Diamond (1973, 1975) noted that the species richness of land-bridge islands (paleorefugia) is primarily determined by extinction rates. Brown (1971) also discussed the role of extinction in shaping rodent faunas of Great Basin montane and alpine communities, which became frag- 123 December 1999 2461NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis mented following the close of Wisconsinan glaciation (paleorefugia). The importance of immigration in the development of the biota of young volcanic islands (neorefugia) has been shown on Krakatau, Long, and Ritter islands in the eastern Pacific (Diamond 1975, Bush and Whittaker 1993). Similar patterns were demonstrated by Simberloff and Wilson (1969) in their manipulative experiments of mangrove islands. Research on the importance of immigration in shaping community composition and structure has also been referred to as ‘‘supply-side ecology’’ (Roughgarden et al. 1987). However, none of these previous works have attempted to compare the contemporaneous ecological patterns of isolated communities that possess different colonization histories and that have, therefore, been exposed to different levels of extinction and immigra- tion. PREDICTIONS ABOUT SPECIES DISTRIBUTION, SPECIES RICHNESS, AND COMMUNITY COMPOSITION My primary assumption is that across a large range of taxonomic groups, extinction will most influence the species composition of paleorefugia, whereas immigration will most influence the species composition of neorefugia. From this assumption, I have developed eight a priori expectations regarding differential species richness, distribution, and community-composition patterns: (1) All else being equal, the slope of the species– area relationship, and the amount of variance explained by it, should be greater in paleorefugia than in neorefugia. Paleorefugia can be expected to have steeper species–area curves than neorefugia, because for a given-size island the paleorefuge will be approaching equilibrium number through species loss, whereas the neorefuge will be approaching this number through species accretion. Paleorefugia were originally more connected, and presumably once harbored similar numbers of species per unit area. Consequently, differences in species richness should be largely a function of postfragmentation extinction rates. Such rates are believed to be strongly correlated with habitat size (e.g., Diamond 1975, Bierrgaard et al. 1992). The species richness of neorefugia, however, should also be a function of immigration processes, which are related less to habitat size than to other factors like habitat age and isolation. Consequently, paleorefugia can be expected to have less variance around the species–area curve than neorefugia. This prediction may be violated for neorefugia that have remained isolated over evolutionary time scales, as strong species–area relationships may be generated on island habitats through speciation (Carlquist 1974). (2) All else being equal, the negative relationship between habitat isolation and species richness should be stronger in neorefugia than in paleorefugia. I expect this as propagule movement (and hence immigration rate) is largely a function of barrier width (Okubo and Levin 1989). Thus, less-isolated neorefugia should have access to a larger propagule pool than more-isolated ones. This result would not be anticipated from paleorefugia as, over limited geographic extents, sites were likely colonized from similar species pools. As such, contemporaneous site isolation should not strongly affect richness. Isolation could influence paleorefugia richness when recolonization from nearby sites (Brown and Kodric-Brown 1977) is important. (3) Species richness should decrease over ecological time scales in paleorefugia, while increasing over similar time scales in neorefugia. Extinction, or ‘‘relaxation’’ (sensu Diamond 1975), should eliminate taxa from paleorefugia, as is typical for land-bridge islands. Such taxa may include those that are not well adapted to local site conditions, that are competitively inferior (e.g., Tilman 1988), or that exhibit area sensitivity (McDonald and Brown 1992). However, as neorefugia form de novo out of a hostile landscape, they will initially be void of species. Over time their richness should increase as taxa from the surrounding species pools immigrate onto sites (Zobel 1997). It is also possible that over time the species richness of neorefugia and paleorefugia will converge. As time increases, immigration between neorefugia should become more and more complete, thereby increasing the importance of extinction in the structuring of site biotas and strengthening the species–area relationship. On paleorefugia, the stochastic loss of species may eventually lower site richness to levels anticipated for neorefugia. It is not clear what factors will most strongly influence this rate of convergence. (4) The inverse correlation between site distance and community similarity (distance decay) should be stronger for neorefugia than for paleorefugia. A fundamental principle of geography is the inverse correlation between similarity and intersample distance (Tobler 1970). In the geographical literature this decrease in similarity has been termed ‘‘distance decay’’ and been applied to phenomena as diverse as human communication networks, migration patterns (Bennett and Gade 1979, Fotherinham 1981), and spatial interpolation (Burrough 1986). The rate of distance decay among natural communities has been shown to vary with the dispersal strategy of organisms and with the degree of landscape fragmentation (Nekola and White, in press). I expect that the rate of compositional distance decay for paleorefugia should primarily reflect pre-fragmentation rates as long as no strong environmental gradients are present in the current landscape. Consequently, in most cases the rates of compositional distance decay for paleorefugia should be low within regions. However, because immigration is required for development of neorefugia biotas, sites across a region may be colonized through different species pools when source areas or migration corridors differ. As a result, similarity 124 2462 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis of neorefugia communities should be more strongly correlated with intersample distance. (5) Neorefugia should be enriched in highly vagile species relative to paleorefugia, whereas paleorefugia should be enriched in less vagile species relative to neorefugia. I expect that species with good dispersal abilities will be more frequently represented in neorefugia, as colonization on such sites is only possible through the crossing of dispersal barriers–effectively eliminating poor dispersers while increasing the relative importance of good dispersers in the neorefugia species pool. Poor dispersers not only can be a component of paleorefugia biotas (as sites were colonized prior to fragmentation), but may be favored over time on these sites if they lose a smaller percentage of propagules beyond site boundaries (Carlquist 1974), and have larger, more competitively fit progeny (Harper 1977). These mechanisms have been suggested by Carlquist (1974) to explain the higher frequency of poorly dispersing plant and animal species on remote oceanic islands. (6) Geographic factors (such as distance to nearest population) should be more important predictors of species occurrence on neorefugia over ecological time scales. All paleorefuge species (theoretically) had similar access to sites prior to fragmentation. For this reason, contemporaneous geographic factors should be of relatively limited importance in predicting current species distribution. Those environmental factors that directly affect the survival of populations should be of pronounced importance in predicting species occurrence on these sites due to the prolonged competitive sorting that has occurred there. However, the distance to contemporaneous propagule sources should be an important predictor of species occurrence on neorefugia, as the probability of colonization will increase with decreasing dispersal distance. Additionally, direct environmental factors may be less important in predicting species occurrence in neorefugia due to increased habitat breadths that result from limited competitive sorting on these sites and incomplete immigration of potential competitors (Cox and Ricklefs 1977). (7) The higher levels of competitive sorting in paleorefugia should lead to a stronger tie between environmental and compositional gradients (Peet and Christensen 1988) than in neorefugia. Although community composition within paleorefugia should initially contain more and stronger correlations with environmental covariables than neorefugia, convergence can be expected over ecological time. (8) The number of competitive co-equivalents held within an archipelago of neorefugia should be greater than that held within a similar set of paleorefugia. The more intense competitive sorting within and the lower levels of pre-fragmentation isolation between paleorefugia should have resulted in similar initial compositions and a strongly competitive environment. As a consequence, there should have been little opportunity for establishment of equivalents with varying fitnesses (Shmida and Ellner 1984). However, the always-isolated nature of neorefugia should allow for less complete intersite dispersal, making it easier for less fit coequivalents to persist in the landscape. As is true for many ecological processes that operate over large spatial and temporal extents, collection of data to test these hypotheses may be difficult, or in some cases impossible. And, even if appropriate data exist, it may be equally difficult to control for confounding factors (Brown 1995). This should not imply, however, that hypotheses regarding differential processes between paleorefugia and neorefugia are inherently unfalsifiable. By comparing current ecological patterns between sites within the same geographic region, confounding effects of the local environment, climatic history, and migration history can be at least partially controlled. The most easily tested hypotheses are those that concern directly quantifiable variables such as current species richness, community composition, habitat size, habitat isolation, and habitat location. Consistency with Hypothesis 1 can be easily tested as it is based on a comparison of species richness to habitat area. Hypothesis 2, which is based on species richness and siteisolation data, can also be readily assessed. Hypothesis 4 can also be easily tested, as it only requires community-composition and geographic-location data. Lastly, Hypothesis 5 can be easily assessed by investigation of the dispersal strategies exhibited for paleorefuge and neorefuge species. Hypothesis 7 can also be relatively easily tested, as it is based on analysis of how quickly and predictably composition changes over a given amount of environmental-gradient space. Testing of this question is slightly more complex, however, as the documentation of these patterns requires use of multivariate data-reduction procedures such as ordination. It is possible that use of different data-reduction techniques could produce different outcomes. Previous knowledge of which environmental variables are critical in the system may also be necessary. Hypotheses 6 will be more difficult to test. To assess the effect of geography on species occurrence, it is first necessary to control for variation in distribution related to environmental gradients. Although in some cases it may be relatively easy to quantify this through use of parametric or nonparametric regression techniques (Austin et al. 1990), such analyses will be problematic when environmental variables demonstrate strong spatial autocorrelation or when habitats are clustered in a landscape. In such situations it is difficult to know what the null expectation will be for the effect of geography on species occurrence, independent of environmental variation. While these problems are likely tractable through use of Monte Carlo or other randomization 125 December 1999 2463NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis FIG. 2. Cross-section view of a Blanding Formation algific talus slope showing major geomorphological features (adapted from Frest [1991: Fig. 2]). A complete system has small sinks in the forested upland; vertical mechanical karst fissures in the Hopkington Formation; ice reservoirs in mostly horizontal mechanical karst in the thin-bedded Blanding Formation; underlying relatively impervious Tete des Mortes, Mosalem, and/or Maquoketa formations. techniques (Hilborn and Mangel 1997), the exact form of these tests have yet to be formalized. Hypotheses 3 and 8 will both be hard to test. Determination of absolute habitat age will be difficult, if not impossible, to measure in most cases. Even when it is possible to age habitats through 14C dating of organically rich basal sediments (peatlands, for example), it would be difficult or impossible to document total species richness over this time period. Hypothesis 8 is difficult to address as it requires determination of which species represent co-competitors. It is not clear how one could adequately assess this without knowing, for a given set of species, all interactions between all potential limiting environmental variables. Based on these factors, a priori Hypotheses 1, 2, 4, and 5 appear to be the most easily testable. To provide a preliminary assessment for these hypothesized differences, I have used a paleorefugia system (algific talus slopes; Frest 1982) and neorefugia system (fens; Nekola 1994) that co-occur in northeastern Iowa to examine four specific predictions: (1) the species–area relationship is steeper and stronger in algific slopes as compared to fens; (2) the species–isolation relationship is stronger in fens as compared to algific talus slopes; (3) distance-decay patterns are more important in fens as compared to algific talus slopes; and (4) the occurrence frequency of good and poor dispersers varies between algific talus slopes and fens. METHODS Study sites Algific talus slopes.—Algific talus slopes (‘‘algific’’ coined from the Greek root ‘‘algos’’ and translates to ‘‘cold producing’’) occur within a deciduous-forest matrix on steep, usually north-facing, carbonate talus slopes where cold air flows out of ice-filled caves (Frest 1981; Fig. 2). These caves developed through ice wedging associated with periglacial activity (Hedges 1972), and have maintained a below-freezing mean annual temperature through thermal buffering provided by overlying cliff-forming carbonates and subtending talus slopes (Frest 1981, 1982). Algific talus slopes are characterized by an unique buffered microclimate where soil temperatures rarely exceed 15Њ C in the summer (Frest 1982). Approximately 300 algific talus slope sites have been located across a 10-county region of northeastern Iowa (USA) in exposures of Prairie-duChien, Galena, Blanding, or Hopkington Formation carbonates (Fig. 3; Frest 1981, 1982, 1983, 1984, 1986a, b, 1987). A limited number of sites are also known from adjacent portions of Minnesota, Wisconsin, and Illinois, USA (Frest 1991). Algific talus slope communities range from densely forested to moderately open. Characteristic woody species include Acer saccharum, A. spicatum, Abies balsamea, Betula lutea, B. papyrifera, Fraxinus americana, Juglans cinerea, Pinus strobus, Salix bebbiana, Sambucus pubens, Taxus canadensis, and Tilia americana. The ground layer is typically dominated by a dense bryophyte cover that supports large populations of pteridophytes such as Cystopteris fragilis, Gymnocarpium robertianum, and Equisetum scirpoides. Woody plant cover decreases and bryophyte/pteridophyte cover increases in general with increasing intensity of cold-air seepage. Algific talus slopes harbor populations of over 60 vascular plant species that are disjunct in Iowa from wet woods and bogs in northern or western boreal forests, including Adoxa moschatellina, Carex media, Cornus canadensis, Linnaea borealis, Mertensia paniculata, Poa paludigena, Pyrola asarifolia, Rhamnus alnifolia, Ribes hudsonianum, Streptopus roseus, and Viola renifolia (Thorne 1964, Frest 1982). At least three near-endemic vascular plant taxa (Aconitum noveboracense, Chrysosplenium iowense, and Sullivantia renifolia) occur on these sites. Algific talus slopes also harbor populations of at least eight 126 2464 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis FIG. 3. Distribution of algific talus slopes and major bedrock units in northeastern Iowa, USA. land snail taxa once thought to have become extinct at the end of the Wisconsinan (Frest 1991). Relict mites in the family Rhagidiidae have also been discovered on these sites (V. Ruzicka, personal communication). These taxa were components of a full-glacial community associated with quasi-maritime climates found south of the ice margin (Frest and Dickson 1986). Populations became restricted to algific talus slopes when this climatic regime disappeared following the retreat of glacial ice from the continental interior (Frest 1981). Due to the microclimatic buffering provided by coldair seepage, which mimics the full-glacial regional climate, populations were able to persist on algific talus slopes through the Hypsithermal to the cooler and wetter conditions of modern times. Algific talus slopes thus represent paleorefugia whose biota predates the surrounding deciduous forest matrix. Fens.—In northeastern Iowa, fens occur within a tallgrass-prairie matrix at sites of local groundwater discharge from pre-Illinoian tills, bedrock, aeolian sands, fluvial sands, or oxbow-meander cutoffs (Nekola 1994, Fig. 4). Sites typically occur as hillside terraces, or more rarely as flat expanses in low swales or mounds. The constant issue of groundwater at fen sites creates a buffered habitat somewhat similar to algific talus slopes, with soil temperatures being cooler in the summer, warmer in the winter, and with more constant soil moisture than is otherwise found within the surrounding landscape. Fens are scattered across a 30county region where at least 2333 sites occurred as recently as 50 yr ago and where 160 sites remain extant (Fig. 5). The bulk of presettlement and extant fens are found on the Iowan Erosional Surface, which was formed during the Wisconsinan through intense periglacial erosion that removed older tills, created a stepped landscape surface, and prevented the accumulation of deep loess deposits (Hallberg et al. 1978, Prior 1991). Similar fen habitats are known from northwestern Iowa (Anderson 1943), Illinois (Moran 1981), Minnesota (Coffin and Pfannmuller 1988), South Dakota (Ode 1985), and Wisconsin (Curtis 1959). In northeastern Iowa, fen habitats are dominated by a dense sedge turf, which includes Carex buxbaumii, C. interior, C. lanuginosa, C. lasiocarpa, C. prairea, C. stricta, C. suberecta, and C. tetanica. In discharge zones Scirpus validus, Typha angustifolia, and T. la- 127 December 1999 2465NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis FIG. 4. Cross section of the five geologic features that give rise to fen formation in northeastern Iowa: (A) pre-Illinoian till sites, with groundwater seepage emanating from sand and gravel lenses interbedded between till units; (B) bedrock sites, with groundwater seepage emanating from fractured carbonate, sandstone, or shale units; (C) aeolian sand dune sites, with groundwater seepage emanating from the contact between sand and an impervious till or paleosol; (D) fluvial sand sites, with groundwater seepage emanating from the base of fluvial or outwash terraces; and (E) oxbow sites, which form through hydric succession of cutoff oxbow meanders. tifolia are often dominant. Scattered clumps of low shrubs are also frequent, and include Cornus stolonifera, Salix bebbiana, S. discolor, S. petiolaris, and S. rigida. In areas of highest soil saturation, vascular-plant growth often becomes stunted and sparse, with bryophytes becoming dominant. Approximately 50% of the vascular-plant species of these sites are regionally rare (Nekola 1994). Over 80 of these species are disjunct from open peatlands in boreal or northeastern North American, including Aster junciformis, Betula pumila, Carex sterilis, Epilobium strictum, Eriophorum angustifolium, Galium labradoricum, Gentiana procera, Lobelia kalmii, Menyanthes trifoliata, Mimulus glabratus var. fremontii, Parnassia glauca, Rhynchospora capillacea, Salix candida, S. pedicellaris, Solidago uliginosa, Triglochin maritimum, and T. palustris. Northeastern Iowa fens also harbor populations of 19 rare Iowa butterfly and skipper species (Nekola 1994), and seven rare Iowa land snail species (Frest 1990). 14 C dates of basal peat layers indicate that northeastern Iowa fens initiated peat development following the end of Hypsithermal warming, less than 6000 yr ago (Thompson 1992). Although a few sites possess basal peat dates extending to 10 000 yr ago (Thompson and Bettis 1994), the presence of oxidized layers within these beds (Hall 1971, Van Zant and Hallberg 1976) indicate that deposition was not continuous, and that during the Hypsithermal even these sites dried out. Thus, although fens may have occurred at modern locations in the late-glacial landscape, the geological record indicates that they did not maintain environmental buffering during the Hypsithermal. Because of this, isolated populations of boreal taxa occurring within these sites cannot represent late Pleistocene or early Holocene relicts, but are instead relatively recent immigrants. In northeastern Iowa, therefore, fens represent neorefugia that are younger than their surrounding tallgrass-prairie matrix. Data sets Species richness.—The ecological distance between fen and algific talus slope habitats and the surrounding matrix habitats is not so great as to preclude occurrence of matrix taxa. Such species can colonize sites via short-range dispersal or mass effect (Shmida and Ellner 1984), independent of the colonization history of a given site. As such, inclusion of these ubiquitous or waif species could seriously swamp any potential ecological differences caused by differential colonization histories. By limiting analysis to only those species restricted to fens or algific talus slopes, this source of noise can be filtered out. For this reason, observations of species composition and richness for fen and algific talus slope sites was limited to the universe of vascularplant taxa restricted (Ͼ80% of known northeastern Iowa occurrences) to either of these habitats. In total, 80 restricted vascular-plant species were identified from fens and 51 from algific slopes, based on occurrence records reported in Frest (1982), Howe et al. (1984), and Nekola (1990, 1994), as well as field data collected during this study. The presence or absence of each of these taxa was subsequently noted for all 160 extant northeastern Iowa fens and all 76 extant algific talus slopes in the Blanding Formation outcrop region. Habitat size.—Sizes of fen sites were estimated through digitization of soil pedons (named variously 128 2466 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis FIG. 5. Presettlement and 1991 distribution of northeastern Iowa fens within major landform boundaries, based on Prior (1991). 129 December 1999 2467NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis as Palms Muck, Houghton Muck, Muck, or Peat soils) associated with fen communities on USDA Soil Conservation Service soil maps. Sizes of algific talus slope sites were determined through digitization of site boundaries recorded from field observations on 7.5minute U.S. Geological Survey topographic maps. Isolation.—To quantify the isolation of sites, a weighted summation was calculated for the distance between each site and all others, assuming that the influence of surrounding sites exponentially decreases as intersite distance increases: n Ϫcd I ϭ e͸iϭ1 where I ϭ isolation index, n ϭ number of other sites in a landscape, d ϭ distance between the reference site and site i, and c ϭ a constant modifying the rate of exponential decay The smaller the value of this index for a site, the less it is expected to interact with surrounding sites. This index differs from a summed addition of site distances as it down-weights the importance of distant areas to the isolation of a given site. Justification of this form of weighting is based upon theoretical (Okubo and Levin 1989) and quantitative (Preston 1962, Nekola and White, in press) studies that have shown that dispersal ability and community similarity tend to exhibit an exponential decay with increasing intersample distance. No a priori rules exist to suggest the appropriate value of c, the constant that modifies the level of downweighting of distant sites. To allow a robust test of the effect of isolation on species richness, isolation-index scores were calculated over five different values of c (1.0, 0.5, 0.25, 0.1, and 0.05), corresponding to five different neighborhood radii (roughly equivalent to 3, 6, 12, 30, and 60 km) that demarcate those sites that most strongly contribute to isolation values. Sites that fall outside these neighborhoods contribute very little to the overall isolation-index values for a given loca- tion. Statistical analysis Prediction 1: The species–area relationship will be steeper and stronger in algific slopes as compared to fens.—The slope of the vascular plant species–area relationship was estimated by multiple least-squares linear regression after both the species-richness and habitat-area values had been log transformed. A double log transform was used as these residuals were more homoscedastic than those generated from untransformed richness vs. log-transformed area. To allow inclusion of sites with zero species, a one was added to the species richness of each site prior to log transformation. Tests for differences between the slopes and intercepts of the two species–area relationships were obtained by analysis of a binary variable representing habitat type that was added into the model following methods outlined in Kleinbaum et al. (1988). The form of this regression model was ln(species richness ϩ 1) ϭ ␤0 ϩ ␤1(ln(habitat size)) ϩ ␤2(habitat type) ϩ ␤3(habitat type ϫ ln(habitat size)) ϩ ␧ (1) where ␤0 ϭ intercept for algific slopes, ␤1 ϭ slope for algific slopes, ␤2 ϭ difference in intercept between algific slopes and fens, and ␤3 ϭ difference in slope between algific slopes and fens. As fens can be significantly larger than algific talus slopes, comparison of species–area slopes and intercepts calculated from the entire data set may lead to inaccurate conclusions (Conner and McCoy 1979). To compensate, the regression analysis was repeated using only those sites falling within the habitat-size overlap between algific talus slopes and fens. The relative strength of the species–area relationship was determined through comparison of the correlation coefficients generated through separate regressions of log-transformed richness vs. log-transformed area for each habitat. The significance of observed differences between these r values was quantified using the correlation-coefficient homogeneity test (Sokal and Rohlf 1981). Prediction 2: Species–isolation relationships will be stronger in fens as compared to algific talus slopes.— The significance and amount of extra variance in vascular-plant species richness explained by site isolation was estimated by separately adding each of the five isolation-index values into the previously defined species–area model. The form of this regression model, for each of the isolation-index values, was ln(species richness ϩ 1) ϭ ␤0 ϩ ␤1(ln(habitat size)) ϩ ␤2(isolation index) ϩ ␧ (2) where ␤0 ϭ intercept, ␤1 ϭ slope of habitat-size relationship, and ␤2 ϭ slope of isolation relationship. The amount of additional variance each of the isolationindex values explained was determined through subtraction of r2 values generated from these models from the r2 values observed from the species-richness vs. habitat-size regression. Prediction 3: Distance-decay patterns will be more pronounced in fens as compared to algific talus slopes.—Distance-decay rates in the restricted vascular-plant floras were estimated in both habitats using the procedures outlined in Nekola and White (in press). These analyses were conducted at five different sample grain sizes: single sites, plus grid cells of 10, 20, 40, and 80 km on a side laid over the study region. However, 80 ϫ 80 km grid cells were not analyzed for algific talus slopes because of their more limited extent. The flora of a given cell was determined by summing together the floras of all sites whose centroids fell within the boundaries of that cell. The centroid location of each cell was set as the average location of all sites 130 2468 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis FIG. 6. Species–area curves (Eq. 1). Log-transformed restricted-species richness as a function of log-transformed habitat size for northeastern Iowa algific talus slope and fen habitats. Each group of three curves shows the position of best-fit regression line and 95% confidence interval around the mean of each. found within that quadrat. Floristic similarity between samples was calculated using Jaccard’s index, and distance was calculated as the aerial separation between cell centroids. The amount of variance in similarity explained by distance was estimated through a linear regression of natural-log-transformed similarity vs. untransformed distance (Nekola and White, in press). The significance of the similarity–distance relationship was estimated by Mantel matrix randomization tests (RT software package, Manley 1992). All Mantel P values were generated using 10 000 replications, as suggested by Jackson and Somer (1989) for biological-data sets. Prediction 4: The proportional frequency of species with good or poor dispersal abilities will vary between the algific talus slope and fen flora.—Each restricted vascular-plant species found in either of the two habitat types was assigned to one of four dispersal classes expected to differ in their dispersal abilities: (1) macroscopic (Ͼ0.1 mm), non-plumose seeds in dry fruits or capsules; (2) fleshy berries, fruits, or nuts; (3) microscopic (Ͻ0.01 mm) seeds or spores; and (4) plumose seeds. The frequency of these four categories in the flora of each habitat was determined, and differences tested using the Pearson chi-square statistic. RESULTS Prediction 1: The species–area relationship will be steeper and stronger in algific slopes as compared to fens Both habitats demonstrate a significant positive relationship between species richness and habitat size (Fig. 6, Table 1), with both slope and intercept for algific talus slopes being significantly (P Ͻ 0.0005) larger than those reported for fens. When this analysis was repeated for only those sites within the range of habitat-size overlap, the slope and intercept were again found to be significantly larger in algific slopes (P Ͻ 0.017 and P Ͻ 0.0005, respectively). In both analyses, small algific talus slopes had fewer vascular-plant species than fens of identical size, whereas large algific talus slopes had greater richness than fens of identical size. The amount of variance explained in these regressions also differed greatly (Table 2). Log-transformed habitat size accounted for 32.6% of observed variance in algific talus slope species richness but accounted for only 6.3% of the variance in fen species richness. This 131 December 1999 2469NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis TABLE 1. Results of multiple-regression analysis of the effect of habitat area and habitat type on restricted vascular-plant species richness in northeastern Iowa fen and algific talus slope habitats (see Eq. 1). Restricted species are those present only in fens or algific talus slopes. Habitat N Value 1 SE t P Algific talus slope sites 76 Intercept, ␤0 Slope, ␤1 2.335 0.542 0.102 0.090 22.881 6.003 Ͻ0.0005 Ͻ0.0005 Difference between algific slopes and fen sites 160 Intercept difference, ␤2 Slope difference, ␤3 Ϫ0.476 Ϫ0.428 0.115 0.097 Ϫ4.139 Ϫ4.421 Ͻ0.0005 Ͻ0.0005 Notes: N ϭ no. of observations. There were 160 extant fens and 76 extant algific talus slopes in the Blanding Formation outcrop region. TABLE 3. Summary statistics for restricted vascular-plant species richness vs. area and isolation in northeastern Iowa fen and algific talus slope habitats. Exponen- tial decay coefficient Algific talus slopes P Additional r2 Fens P Additional r2 c ϭ 1.0 c ϭ 0.5 c ϭ 0.25 c ϭ 0.1 c ϭ 0.05 0.150 0.266 0.497 0.802 0.778 0.019 0.012 0.005 0.001 0.001 0.027 0.008 0.003 0.003 0.008 0.029 0.041 0.053 0.050 0.041 TABLE 2. Significance and r2 for the restricted vascular-plant species richness vs. habitat-area relationship in northeastern Iowa fen and algific talus slope habitats. Habitat No. of observations P r2 Fens Algific talus slopes 160 76 0.001 0.000 0.063 0.326 Note: Significance of the correlation-coefficient homogeneity test: P ϭ 0.006. difference was found to be highly significant (P ϭ 0.006). Prediction 2: Species–isolation relationships will be stronger in fens as compared to algific talus slopes Isolation was found to be a significant additional predictor of vascular-plant species richness in fens for each of the calculated isolation-index values (P ranging from 0.003–0.027; Table 3). The amount of additional variance explained by isolation in fen habitats varied between 0.026 and 0.053, or between 46% and 84% of the variance explained by log-transformed habitat area. However, for algific talus slopes, isolation was never a significant additional predictor of species richness at any of the isolation-index scores (P ranging between 0.150 and 0.802). Prediction 3: Distance–decay patterns will be more pronounced in fens as compared to algific talus slopes In fens, significant (P ϭ 0.0001) distance–decay relationships were observed between the restricted species floras at all grain sizes (Table 4). At the largest grain (80 ϫ 80 km regions), this relationship explained over one third of the variance in regional restricted vascular-plant species composition (Table 5). At smaller grains, distance accounted for Ͻ5% of observed variance in floristic similarity. In algific slopes, only the distance–decay relationship at the 20 ϫ 20 km sample grain size was found to be significant at the 0.05 level. At this scale, 23% of the observed variance in vascularplant assemblage similarity was accounted for by intersample distance. Prediction 4: The proportional frequency of species with good or poor dispersal abilities will vary between the algific talus slope and fen flora The frequencies of dispersal classes differed significantly (P Ͻ 0.0005) between the two habitats (Table 6). Species with berries, fleshy fruits, or nuts were ϳ15 times more frequent in algific slopes than fens, whereas species with plumose seeds were ϳ7 times more frequent in fens than algific slopes. DISCUSSION Distinguishing among alternative explanations In each case the results obtained agree with the a priori hypotheses, which predicted differences between the ecological patterns of paleorefugia and neorefugia and suggest that extinction and immigration have played differing roles in algific talus slope and fen communities (Table 7). However, it is impossible to control for all confounding factors in a natural experiment such as this. By explicitly stating what some of these factors may be, the preliminary nature of these investigations, and potential avenues for future research, may be made more clear. Two potential confounders stem from the different source pools for the flora of these two habitats, and from the different landscape dynamics of the primary disturbances that effect them. Although the flora of each habitat is a subset of the boreal wetland flora, the species of algific slopes have greater affinity with conifer swamps while the species of fens have greater 132 2470 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis TABLE 4. P values for the distance–decay relationships of the restricted vascular-plant floras of northeastern Iowa fen and algific talus slope habitats at differing grain sizes. Habitat Grain size (km) 80 40 20 10 Individual sites Fen Algific slope 0.0002 ··· 0.0001 0.6245 0.0001 0.0213 0.0001 0.2832 0.0001 0.2339 Note: P values were calculated using Mantel tests on distance and similarity matrices for comparisons between each region or site. TABLE 5. The r2 values for the distance–decay relationships of the restricted vascular-plant floras of northeastern Iowa fen and algific talus slope habitats at differing grain sizes. Habitat Grain size (km) 80 40 20 10 Individual sites Fen Algific slope 0.335 ··· 0.014 0.082 0.044 0.226 0.034 0.005 0.027 0.001 Note: The r2 values were calculated using linear regression on log-transformed data. TABLE 6. Frequency of four dispersal classes among restricted vascular-plant species in northeastern Iowa fen and algific slopes. The numbers in parentheses represent the percentage of the total number of species recorded in each habitat contained in each dispersal class. Habitat Dispersal class† 1 2 3 4 Fen Algific slope 42 (52.5%) 26 (51.0%) 1 (1.2%) 10 (19.6%) 17 (21.3%) 13 (25.4%) 20 (25.0%) 2 (3.9%) Note: Chi-square statistic for differences in frequencies for all dispersal classes ϭ 20.998; P Ͻ 0.0005. † Dispersal classes: 1 ϭ macroscopic, non-plumose seeds in dry capules; 2 ϭ berry/fleshy fruits; 3 ϭ spore or microscopic seeds; and 4 ϭ plumose seeds. similarity with open peatlands (e.g., Curtis 1959). Differences in the composition, coexistence patterns, and dispersal strategies of species in these pools could have contributed to, or created, observed differences. Additionally, algific slopes are occasionally impacted by small-scale talus dislodgement from game trails or individual tree-falls, whereas fens are more often affected by larger-scale events such as fires and regional droughts. Consequently, algific slopes seem more likely than fens to be in a state of successional quasiequilibrium (Shugart 1984, Turner et al. 1994). This difference could potentially account for the stronger species–area relationship in algific slopes, or for the predominance of vagile species in fens. Additional research will be necessary to tease apart the effects of these and other factors from effects of colonization history. For example, it would be instructive to extend these analyses of algific slope and fen vascular-plant floras to additional taxonomic groups found on the same sites that possess varying dispersal abilities, such as bryophytes, birds, lepidoptera, or terrestrial gastropods. Such analyses would help document whether the patterns observed for vascular plants in this investigation are general in nature, and whether species groups with different dispersal abilities will respond to identical levels of geographic isolation and patterns of habitat creation in the same way. It is possible that poorly dispersing taxa (snails, for example) may be more influenced by extinction and that morevagile species (lepidoptera, for example) may be more influenced by immigration processes on the same sites. Important additional insights may also be garnered if paleorefugia and neorefugia are studied over a more diverse set of environmental situations and landscapes. For instance, comparing identical habitats that have formed at different times in different landscapes (creating a neorefugia in one and a paleorefugia in the other) may be a good way to control for the confounding effects of species source pools or disturbance dynamics. It may also be interesting to study the contrast between these refugia types across a series of landscapes in which sites vary from highly to less isolated. Neorefugia and paleorefugia of very ancient origin may also demonstrate different biogeographic patterns. For instance, some ancient oceanic islands represent neorefugia (like the Hawaiian archipelago, which formed through volcanic eruptions) whereas others represent paleorefugia (like New Zealand and New Caledonia, which were created through fragmentation of continental land masses). As time has been sufficient for speciation to occur, a strong species–area relationship should develop for both. However, the fact that dispersal was originally more important for ancient oceanic neorefugia should allow for stronger species–isolation relationships, and different dispersal-strategy frequencies in their contemporaneous biota as compared to ancient oceanic paleorefugia. Hypotheses regarding paleorefugia and neorefugia are also amenable to some forms of manipulative experimentation for those processes that operate over limited spatial and temporal extents, or to computer simulations for largerscale patterns. The future research priorities regarding paleorefugia and neorefugia should include more taxa, more habitats, more landscapes, and more times of origin. Through such additional work a clearer understanding of the role of colonization history on contemporaneous ecological pattern can be achieved. Implications for ecology, biogeography, and conservation The central question of this paper is whether the historical contingency of habitat formation can have a long-lasting impact on community patterns and processes. The answer to this question has important implications for ecological debates surrounding equilibrium vs. non-equilibrium dynamics, vicariance vs. long-range dispersal, and the delineation of optimum reserve designs for maintenance of biological diversity. 133 December 1999 2471NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis TABLE 7. Summary outcomes of analysis: observed and predicted differences between algific talus slopes (paleorefugia) and fens (neorefugia) in northeastern Iowa. Factor Response Neorefugia Paleorefugia Result† Relative steepness and strength of species–area relationship Less Greater Y Strength of species–isolation relationship High Low Y Change in richness over time Increasing Decreasing ··· Strength of distance-decay relations High Low Y Dispersal ability of species in habitat Good Poor Y Relative importance of geography in predicting species occurrence Strong Weak ··· Strength of community–environment correlation Weak Strong ··· Relative number of co-competitors in landscape More Less ··· † Y, relationship found statistically significant and in agreement with a priori predictions; ···, not addressed in this analysis. Equilibrium vs. non-equilibrium dynamics.—Although much classical ecological theory (e.g., MacArthur and Wilson 1967, MacArthur 1972, Tilman 1988) assumes existence of equilibrium conditions, the data presented above suggest that differences in colonization history may still imprint biotas 5000–10 000 yr following habitat development. If ‘‘transient dynamics’’ continue to exist after such time periods, equilibrium conditions cannot be assumed to dominate biogeographic pattern. These results support paleoecological work (Davis 1984) that suggests that disequilibrium between climate and species range may persist over relatively long temporal extents. Vicariance vs. long-distance dispersal.—The relative importance of vicariance vs. long-distance dispersal in the development of biogeographic pattern has been debated since the times of Wallace, Schimper, Engler, and Hooker (Cain 1944). Many ecologists view dispersal as so efficient that no barriers to movement exist for any species within continental areas, making biogeographic pattern simply the result of physiological limitations and competitive interactions (e.g., Krebs 1985). Saur (1988) has summarized this belief as Beijerinck’s Law: everything is everywhere but the environment selects. However, some cladistic biogeographers have stated that dispersal either does not occur between isolated habitats, or is so unpredictable that it cannot be studied. For these researchers, only the process of vicariance can parsimoniously explain disjunct distributions (e.g., Humphries and Parenti 1986, Brudin 1988). Analysis of paleorefugia and neorefugia suggest that such absolute positions may be fundamentally flawed as both vicariance and dispersal can operate contemporaneously in the same landscape. Instead of asking which of these processes is responsible for observed biogeographic patterns, it seems more productive to determine how much each has contributed to that pattern. If additional research confirms generality to the patterns attributed above, it may be possible to use similar analyses in additional landscapes to document the occurrence of neorefugia and paleorefugia, and thus the frequency of vicariance and long-distance dispersal. Conservation of biological diversity.—These preliminary analyses suggest that algific talus slopes are floristically more self-contained than fens, with immigration playing a relatively minor role in the determination of species occurrence. Protection of the vascular flora of these habitats should thus center on those sites that harbor maximum diversity and the internal site dynamics that are responsible for that diversity. However, these preliminary analyses also suggest that fens are not as self-contained, with immigration from surrounding sites strongly influencing vascularplant composition. Diversity in these communities varies conversely with isolation, and community composition is strongly influenced by the proximity of source areas. To ensure adequate protection of the processes that gave rise to site diversity in such systems, reserves may have to be chosen to protect not only internal site dynamics, but also the immigration potential between sites. This will be possible only if many sites are protected across the landscape, including those that may not currently support diverse floras. Consequently, neorefugia reserve design may require an integrated, regional strategy. Unfortunately, the reality of fen destruction in Iowa may make such goals unrealistic. Since European settlement, 93% of northeastern Iowa fens have been destroyed (Nekola 1994; Fig. 5). Even if all remaining sites could be saved, immigration rates cannot be maintained at presettlement levels. The ultimate effect of this will be severely lowered species-immigration rates compared to presettlement times, and low probabilities that species will be able to reestablish themselves following local extinction. If extinction rates are high, strategies to guard fen biodiversity may require not only protection of all remaining sites, but also the physical transport of propagules between sites to reestablish presettlement immigration levels. If this approach is eventually deemed necessary, it would require creation of stochastic immigration models to guide land managers in their attempts to simulate ‘natural’ immigration between sites. ACKNOWLEDGMENTS This research was supported by a National Science Foundation Graduate Fellowship and by the McElroy Foundation of Waterloo, Iowa. I wish to thank John and Gretchen Bray- 134 2472 Ecology, Vol. 80, No. 8JEFFREY C. NEKOLA Concepts&Synthesis ton, Bob Moats, Richard Baker, Willard Hawker, Louise McEarchen, John Nehnevaj, Dean Roosa, Arlan Kirchman, Dennis Schlicht, Bill Thomas, Robert Thomson, Dave Wendling, and Herb Wilson for providing me with locality information for extant fen sites. Terrence Frest was instrumental in instructing me in the geology and ecology of algific talus slopes, and initially located many of the Blanding outcrop sites inventoried in this analysis. Peter White, Michael Palmer, Susan Wiser, James Hazen Graves, Robert Howe, Clifford Kraft, Ed Connor, Paul Keddy, three anonymous reviewers, and especially Robert K. Peet read and made useful comments regarding earlier drafts of this manuscript. LITERATURE CITED Anderson, W. A. 1943. A fen in northwestern Iowa. American Midland Naturalist 29:787–791. Austin, M. P., A. O. Nicholls, and C. R. Margules. 1990. Measurement of the realized qualitative niche: environmental niches of five Eucalyptus species. Ecological Monographs 60:161–177. Bennett, D. G., and O. Gade. 1979. Geographic perspectives in migration research. Department of Geography Studies in Geography, Number 12. University of North Carolina, Chapel Hill, North Carolina, USA. Bierrgaard, R. O., Jr., T. E. Lovejoy, V. Kapos, A. Augusto dos Santos, and R. W. Hutchings. 1992. The biological dynamics of tropical rainforest fragments. BioScience 42: 859–866. Brown, J. H. 1971. Mammals on mountaintops: non-equilibrium insular biogeography. American Naturalist 105: 476–478. . 1995. Macroecology. University of Chicago Press, Chicago, Illinois, USA. Brown, J. H., and A. Kodric-Brown. 1977. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58:445–449. Brudin, L. Z. 1988. Phylogenetic biogeography. Pages 343– 370 in A. A. Meyers and P. S. Giller, editors. Analytical biogeography. Chapman & Hall, New York, New York, USA. Burrough, P. A. 1986. Principles of geographical information systems for land resources assessment. Oxford University Press, New York, New York, USA. Bush, M. B., and R. J. Whittaker. 1993. Non-equilibrium in island theory of Krakatau. Journal of Biogeography 20: 453–457. Cain, S. A. 1944. Foundations of plant geography. Harper & Row, New York, New York, USA. Carlquist, S. 1974. Island biology. Columbia University Press, New York, New York, USA. Coffin, B., and L. Pfannmuller. 1988. Minnesota’s endangered flora and fauna. University of Minnesota Press, Minneapolis, Minnesota, USA. Conner, E., and E. D. McCoy. 1979. The statistics and biology of the species–area relationship. American Naturalist 113:791–833. Cox, G. W., and R. E. Ricklefs. 1977. Species diversity, ecological release, and community structuring in Caribbean land bird faunas. Oikos 29:60–66. Curtis, J. T. 1959. The vegetation of Wisconsin. University of Wisconsin Press, Madison, Wisconsin, USA. Davis, M. B. 1984. Quaternary history and the stability of forest communities. Pages 132–151 in D. C. West, H. H. Shugart, and D. B. Botkin, editors. Forest succession. Springer-Verlag, New York, New York, USA. Diamond, J. M. 1973. Distributional ecology of New Guinea birds. Science 179:759–769. . 1975. The island dilemma: lessons of modern biogeographic studies for the design of nature reserves. Biological Conservation 7:129–146. Fernald, M. L. 1925. Persistence of plants in unglaciated North America. Memoirs of the American Academy of Arts and Sciences 15:241–342. Fotherinham, A. S. 1981. Spatial structure and distance-decay parameters. Annals of the Association of American Geographers 71:425–436. Frest, T. J. 1981. Final report, project SE-1-2 (Iowa Pleistocene Snail). Iowa Conservation Commission, Des Moines, Iowa, USA. . 1982. Final report, project SE-1-4 (Iowa Pleistocene Snail). Iowa Conservation Commission, Des Moines, Iowa, USA. . 1983. Final report (contract number 30181-1259) northern Driftless Area survey. U.S. Fish and Wildlife Service, Fort Snelling, Minnesota, USA. . 1984. Final report, project SE-1-6 (Iowa Pleistocene Snail). Iowa Conservation Commission, Des Moines, Iowa, USA. . 1986a. Final report, project SE-1-6 number 2 (Iowa Pleistocene Snail). Iowa Department of Natural Resources, Des Moines, Iowa, USA. . 1986b. Final report, project SE-1-7 (Iowa Pleistocene Snail). Iowa Department of Natural Resources, Des Moines, Iowa, USA. . 1987. Final report, project SE-1-8 (Iowa Pleistocene Snail). Iowa Department of Natural Resources, Des Moines, Iowa, USA. . 1990. Final report (contract number 65-2454). Field survey of Iowa spring fens. Iowa Department of Natural Resources, Des Moines, Iowa, USA. . 1991. Summary status reports on eight species of candidate land snails from the Driftless Area (Paleozoic Plateau), Upper Midwest. Final report (contract number 301-01366). U.S. Fish and Wildlife Service, Fort Snelling, Minnesota, USA. Frest, T. J., and J. R. Dickson. 1986. Land snails (Pleistocene-recent) of the Loess Hills: a preliminary survey. Proceedings of the Iowa Academy of Science 93:139–157. Hall, S. A. 1971. Paleoecological interpretation of bison, mollusks, and pollen from the Hughes peat bed, Linn County, Iowa. Thesis. University of Iowa, Iowa City, Iowa, USA. Hallberg, G. R., T. E. Fenton, G. A. Miller, and A. J. Lutenegger. 1978. The Iowan Erosion Surface: an old story, an important lesson, and some new wrinkles. Pages 2.1 to 2.94 in R. Anderson, editor. Geology of East-central Iowa, 42nd annual tri-state geological field conference. Iowa Geological Survey, Iowa City, Iowa, USA. Harper, J. L. 1977. Population biology of plants. Academic Press, New York, New York, USA. Hedges, J. 1972. Expanded joints and other periglacial phenomena along the Niagara Escarpment. Biuletyn Peryglacjalny 21:87–126. Hilborn, R., and M. Mangel. 1997. The ecological detective. Princeton Monographs in Population Biology number 28. Princeton University Press, Princeton, New Jersey, USA. Howe, R. W., M. J. Huston, W. P. Pusateri, R. H. Laushman, and W. E. Schennum. 1984. An inventory of significant natural areas in Iowa. Iowa Conservation Commission, Des Moines, Iowa, USA. Humphries, C. J., and L. R. Parenti. 1986. Cladistic biogeography. Clarendon Press, Oxford, UK. Jackson D. A., and K. M. Somer. 1989. Are probability estimates from the permutation model of Mantel’s test stable? Canadian Journal of Zoology 67:766–769. Kleinbaum, D. G., L. L. Kupper, and K. E. Muller. 1988. Applied regression analysis and other multivariate methods. PWS-Kent Publishing, Boston, Massachusetts, USA. Krebs, C. J. 1985. Ecology. Harper & Row, New York, New York, USA. Krukeberg, A. R., and D. Rabinowitz. 1985. Biological as- 135 December 1999 2473NEOREFUGIA AND PALEOREFUGIA Concepts&Synthesis pects of endemism in higher plants. Annual Review of Ecology and Systematics 16:447–479. Manley, B. F. J. 1992. RT—a program for randomization testing. Version 1.02. The Centre for Applications of Statistics and Mathematics, University of Otago, Dunedin, New Zealand. MacArthur, R. H. 1972. Geographical ecology. Harper & Row, New York, New York, USA. MacArthur, R. H., and E. O. Wilson. 1967. The theory of island biogeography. Princeton University Press, Princeton, New Jersey, USA. McDonald, K. A., and J. H. Brown. 1992. Using montane mammals to model extinctions due to global change. Conservation Biology 6:409–415. Moran, R. C. 1981. Prairie fens in northeastern Illinois: floristic composition and disturbance. Pages 164–168 in R. Stuckey and K. J. Reece, editors. Proceedings of the sixth North American Prairie Conference. The Ohio State University, Columbus, Ohio, USA. Nekola, J. C. 1990. Rare Iowa plant notes from the R. V. Drexler Herbarium. Proceedings of the Iowa Academy of Science 97:55–73. . 1994. The environment and flora of northeastern Iowa fen communities. Rhodora 96:121–169. Nekola, J. C., and P. S. White. In press.The distance decay of similarity in ecology and evolution. Journal of Bioge- ography. Ode, D. J. 1985. Rediscovered plants of fen habitats of South Dakota. Proceedings of the South Dakota Academy of Science 64:217. Okubo, A., and S. A. Levin. 1989. A theoretical framework for data analysis of wind dispersal of seeds and pollen. Ecology 70:329–338. Peet, R. K., and N. L. Christensen. 1988. Changes in species diversity during secondary forest succession on the North Carolina Piedmont. Pages 233–245 in H. J. During, M. J. A. Werger, and J. H. Willems, editors. Diversity and pattern in plant communities. SPB Academic Publishing, The Hague, The Netherlands. Pielou, E. C. 1979. Biogeography. John Wiley & Sons, New York, New York, USA. Preston, F. W. 1962. The canonical distribution of commonness and rarity: part II. Ecology 43:410–432. Prior, J. C. 1991. Landforms of Iowa. University of Iowa Press, Iowa City, Iowa, USA. Roughgarden, J., S. D. Gaines, and S. W. Pacala. 1987. Supply-side ecology: the role of physical transport processes. Symposium of the British Ecological Society 27:491–518. Saur, J. D. 1988. Plant migration. University of California Press, Berkeley, California, USA. Shmida, A., and S. Ellner. 1984. Coexistence of plant species with similar niches. Vegetatio 58:29–55. Shugart, H. 1984. A theory of forest dynamics. SpringerVerlag, New York, New York, USA. Simberloff, D. S., and E. O. Wilson. 1969. Experimental zoogeography of islands: a model for insular conservation. Ecology 50:296–314. Sokal, R. R., and F. J. Rohlf. 1981. Biometry. W. H. Freeman, San Francisco, California, USA. Stebbins, G. L., and J. Major. 1965. Endemism and speciation in the California flora. Ecological Monographs 35:1– 35. Thompson, C. A., and E. A. Bettis III. 1994. Age and development of Iowa fens. Proceedings of the Iowa Academy of Science 101:73–77. Thorne, R. F. 1964. Relict nature of the flora of White Pine Hollow Forest Reserve, Dubuque County, Iowa. State University of Iowa Studies in Natural History 20:1–33. Tilman, D. 1988. Plant strategies and the dynamics and structure of plant communities. Princeton University Press, Princeton, New Jersey, USA. Tobler, W. R. 1970. A computer movie simulating urban growth in the Detroit region. Economic Geography 46:234– 240. Turner, M. G., W. H. Romme, and R. H. Gardner. 1994. Landscape disturbance models and the long-term dynamics of natural areas. Natural Areas Journal 14:3–11. Van Zant, K. L., and G. R. Hallberg. 1976. A late-glacial pollen sequence from northeastern Iowa: Sumner Bog revisited. Iowa Geological Survey Technical Informational Bulletin Series number 3. Zobel, M. 1997. The relative role of species pools in determining plant species richness: an alternative explanation of species coexistence. Trends in Ecology & Evolution 12: 266–269. 136 Section IV: Biogeography and Macroecology The ecological processes that most influence global biodiversity operate at spatial and temporal scales too large for typical manipulative experiments. To understand the impact of such drivers requires observational studies based on multiple sites that encompass large extents and/or represent natural experiments. Identification of general mechanism in biogeography and macroecology requires that pattern and process be investigated over a wide range of organisms representing many different body plans, trophic states, dispersal strategies and reproductive modes. In spite of this, the great majority of work is limited to vertebrates, vascular plants, and a few anthropods (e.g. lepidoptera). Land snails serve as an important potential alternative group because of their very poor active but often excellent passive dispersal abilities and hermaphroditism with common instances of unipartental reproduction. As a result, this group displays a number of unique patterns. Companion articles published in 2014 document that continental (as well as most regional) faunas display strongly bimodal body-size distributions, as opposed to the typical rightskewed pattern. Additionally, they display strong scale-dependence, with site faunas tending to be dominated by small (<5.7 mm³) species while regional/continental faunas are characterized by large (>90.5 mm³) taxa. This pattern is more pronounced at lower latitudes and appears driven by a strong inverse correlation between dispersal ability and body size, with small species tending to have much larger ranges and lower levels of allopatric replacement across a given geographic gradient Prior work has also shown that land snail communities demonstrate high levels of microsympatry with up to 25% of a regional fauna co-existing within 0.04 m² areas. My best known contribution in this field is my 1999 paper with Peter White which popularized Distance Decay analyses to the field of Ecology. This work is now considered by many to be a classic in the field with this pattern being one of the major statistics used to test Macroecological hypotheses. I am also interested in documenting the nature and causes of spatial heterogeneity in occurrence patterns, the spread of organisms across landscapes, and continental gradients of body size variation within and between communities. And as intimated in Section III, I am using macroecological pattern to inform about macroevolutionary process in Holarctic land snails. Representative Publications [number of citations as of October 27, 2017] 1. Edited Volumes Cameron, R.A.D., J.C. Nekola, B.M. Pokryszko & F.E. Wells. 2005. Pattern and process in land mollusc diversity. Records of the Western Australian Museum, Supplement #68. 137 2. Peer-reviewed Articles Nekola, J.C. 2014. North American terrestrial gastropods through either end of a spyglass. Journal of Molluscan Studies. 80:238-248. [7] Nekola, J.C. 2013. Biodiversity refuges. pp. 141-148 in: MacLeod, N., D.J. Archibald & P. Levin (eds), Grzimek's Animal Life Encyclopedia: Extinction. Gale Publishing, Farmington Hills, Michigan. [6] Nekola, J.C., G.M. Barker, R.A.D. Cameron & B.M. Pokryszko. 2013. Latitudinal variation of body size in land snail populations and communities. pp. 62-82 in: F. Smith & K. Lyons (eds), Global Patterns of Body Size. University of Chicago Press. [12] Nekola, J.C. 2005. Geographic variation in richness, evenness, and shell size of eastern North American land snail communities. Records of the Western Australian Museum. Supplement 68:39-51. [39] Nekola, J.C. 2003. Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Diversity and Distribution. 9:55-71. [52] Nekola, J.C. & P.S. White. 1999. Distance decay of similarity in biogeography and ecology. Journal of Biogeography. 26:867-878. [1069] 138 North American terrestrial gastropods through each end of a spyglass Jeffrey C. Nekola Biology Department, University of New Mexico, Albuquerque, NM 87131, USA Correspondence: J. C. Nekola; e-mail: jnekola@unm.edu (Received 30 December 2013; accepted 3 March 2014) ABSTRACT Some suggest that because of scale independence major biodiversity metrics can be estimated at large scales from analysis of a well chosen suite of individual sites. Others have attempted to estimate individual site patterns from analysis of the continental pool. But does such cross-scale extrapolation work? This issue is addressed for the North American terrestrial gastropod fauna by comparison of family representation, species richness and body-size patterns across site to continental scales. These data demonstrate profound differences: while the continental fauna is dominated by large body-size families such as the Polygyridae, Helminthoglyptidae, Oreohelicidae, Succineidae and Urocoptidae, average site faunas are most frequently represented by small body-sized families like the Vertiginidae, Gastrodontidae, Oxychilidae, Euconulidae, Punctidae, Valloniidae, Strobilopsidae and Ellobiidae. Species richness within sites tends to be 2–7 times smaller than random draws of individuals of the same number from regional or continental pools, indicating the potential for strong bias in the construction of site faunas. And, while the body-size spectrum for average site faunas is strongly right-skewed, the continental pool is strongly left-skewed. Thus, although taxa with biovolumes .16 mm3 dominate the continental fauna (79.4% of total), they make up only a small average fraction (4.1%) of individual site species lists. Within most regions, site faunas are overrepresented in species with biovolumes ,4 mm3 and underrepresented in species with biovolumes .128 mm3 as compared with the regional pool. As a result, assumptions of self-similarity between observational scales in terrestrial gastropods are inappropriate. INTRODUCTION The search for universal patterns and simplifying assumptions is common practice among natural scientists, especially those working in complex systems such as biology. One common form taken by this search is the identification of patterns that are scale-independent, i.e. patterns that remain largely similar across all observational scales. Such relationships are often referred to as expressing ‘self-similarity’, in which large scale datasets are made up of smaller scale components possessing the same basic properties. In the realm of fractal geometry, for instance, self-similarity across multiple scales has been noted in crystals (Mandelbrot, 1983), clouds (Hentschel & Procaccia, 1984), fluids (Nittman, Daccord & Stanley, 1985) and water drainage networks (Milne, 1988). This assumption also underlies most common algorithms generating fractal artwork, lending aesthetic regularity to the generated images. However, the assumption of self-similarity has been found to be less useful in biological systems in which the major drivers are themselves scale dependent (Wiens, 1989). For instance, none of the vegetation patterns analysed by Palmer (1988) demonstrated scale independence and self-similarity. Yet, the siren-call of self-similarity has been especially strong in biodiversity analyses as its presence would allow insights regarding patterns and parameters present at one scale to be accurately applied to adjacent scales without the need for empirical observations. Direct measurement of various diversity metrics at tractable (usually single site) scales could thus theoretically be used to generate expectations at regional to global scales where empirical observation may be logistically impossible. Finlay et al. (2006), for instance, claimed that selfsimilarity in various species composition, richness and bodysize metrics allows for accurate extrapolation of insect diversity patterns at Hilbre Island and Monks Woods Nature Reserves in England to the entire globe. Is it really the case that biodiversity patterns appear the same when peering through each end of a spyglass? Can observation of only a few sites allow for accurate estimation of regional to global scale patterns? And, is it also possible that knowledge of regional to global scale patterns informs accurately about relationships within individual sites? These questions will be empirically tested through an analysis of North American terrestrial gastropod species composition, body-size and abundance patterns across site, regional and continental scales. # The Author 2014. Published by Oxford University Press on behalf of The Malacological Society of London, all rights reserved Journal of The Malacological Society of London Molluscan Studies Journal of Molluscan Studies (2014) 80: 238–248. doi:10.1093/mollus/eyu028 Advance Access publication date: 2 June 2014 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 139 MATERIAL AND METHODS Faunistic enumeration Site scale. Species lists and abundances were considered from 1,574 sites ranging from the Alaskan North Slope and southern California coast east to central Quebec, the Atlantic seaboard and the Florida Keys (Table 1; Fig. 1). These were sampled according to methods presented by Nekola (2010), with individual sample areas being no more than 1,000 m2 . An attempt was made to sample sites from across the entire range of habitats within a given biogeographic region and included 416 bedrock outcrops, 455 upland forests, 335 lowland forests, 126 upland grasslands/shrublands and 242 lowland grasslands/shrublands. Regional scale. Using the maps of Pilsbry (1948), Burch (1962) and observed community composition patterns as a guide, the continent was subdivided into 21 biogeographic regions with boundaries being set to enclose areas of similar faunistic composition and to also allow each species to occur within at least one region (Fig. 1). The total faunas for each region was then determined through the county-scale distribution maps of Hubricht (1985) for eastern North America (using updates provided by Nekola & Coles, 2010 and J.C. Nekola, unpubl.), with Pilsbry (1948) and various regional lists (e.g. Metcalf & Smartt, 1997; Roth & Sadeghian, 2003; Forsyth, 2004) being used for western North America. The number of survey sites per region ranged from 8 (southern Appalachians) to 252 (western Niagaran Escarpment). Continental scale. The total North American fauna was based on Nekola (in press). Only species having native or naturalized populations in North America north of Mexico were considered; species only occurring south of the USA boundary are considered to belong to the Central American fauna. The initial point of departure for this dataset is all terrestrial gastropods listed by Turgeon et al. (1998). To this were added all subsequently described species as determined via the Zoological Record. Additionally, all species listed by Hubricht (1985)—but not by Turgeon et al. (1998)—were included, as these represent dead shells limited to drift along the south Texas Gulf Coast that could easily have been sourced from local extant colonies. The list was also expanded to include all subspecies-level entities of Pilsbry (1948) that appear, based on their unique shell features, ranges and/or ecological preferences, to represent valid species-level entities. Finally, 14 undescribed new species encountered by the author during field sampling across North America have also been included, representing six Vertigo plus single representatives of Columella, Daedalochila, Glyphyalinia, Hawaiia, Helicodiscus, Paravitrea, Punctum and Succinea. Taxonomy and body size Family assignments for all species were based on those of Bouchet & Rocroi (2005), with the placement of genera into these families generally following Schileyko (2006 and previous works). Generic level assignments generally followed Turgeon et al. (1998). Shell/ body-size dimensions for each taxon were determined from either the published literature, or in the case of undescribed taxa from lots held in the Nekola collection. Calculation of body volume (in mm3 ) was then determined for each species using the formulae presented by McClain & Nekola (2008) and Nekola (in press). Statistical analyses All analyses were conducted in the R Statistical Environment with scripts being available upon request. Family-level composition. The proportional representation of each family within each site was calculated by dividing the total number of species present within each family into the total species richness of that site. This process was repeated across all sites, with the mean proportion for each family then being calculated across all sites. This process was also repeated within each biogeographic region. The proportion of each family within the continental fauna was calculated by dividing the number of species within each family into the continental species richness. Table 1. Statistical summary for the continental and regional datasets of North American terrestrial gastropods. Region No. of sites Regional richness Site richness No. of individuals Reported Encountered Median Maximum Total Median/site Whole continent 1,574 1,204 460 12 39 697,778 237 Churchill 23 17 15 6 12 9,303 261 Central Manitoba 24 53 41 10 21 11,014 288 Laurentian Plateau 48 80 40 8 18 9,729 131 Northwestern Minnesota 192 63 59 15 27 142,305 463 Western Superior Uplands 82 55 44 10 18 5,025 47 Eastern Niagaran Escarpment 73 149 86 16 31 26,804 305 Western Niagaran Escarpment 252 98 84 15 34 92,778 216 New England 194 133 82 13 34 84,661 217 Upper Mississippi Valley 131 103 90 20 39 105,075 587 Ozarks 35 149 88 20 34 20,601 255 Central Appalachians 22 246 86 16 35 7002 219 Southern Appalachians 8 278 42 12 33 2,181 210 Carolinas 38 131 62 8 31 9,298 132 Gulf Coast 37 156 34 3 11 2,260 43 Peninsular Florida 59 130 97 12 28 33,848 301 Southern Plains 14 154 58 12 23 3,379 128 Southern Rockies 161 308 113 7 21 72,116 230 California 30 271 51 5 10 8,219 99 Great Basin 39 76 45 7 13 12,635 206 Pacific Northwest 48 187 64 9 17 13,895 197 Alaskan Interior 64 33 32 7 16 25,644 216 NORTH AMERICA LAND SNAIL BIODIVERSITY 239 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 140 This process was also repeated at the regional level. All proportions were multiplied by 100 to give percentage representation. Statistical significance of the difference between site and regional scale representation was calculated using a sign test for all families occurring in at least 10 regions. The number of cases in which mean site percentage was less than the regional percentage was counted, as was the number of cases in which regional percentage was less than mean site percentage. Under a null expectation of no relationship, these numbers should be equal. The P-value for deviations from this null was calculated using the binomial distribution. Because the test was repeated on 18 different families, the P ¼ 0.05 significance threshold was adjusted using a Bonferroni correction to P ¼ 0.0028. Site species richness. The ratio between observed and random expected richness was calculated. Expected random richness was determined by generating a vector representing the species identities of all encountered individuals across either the entire continental dataset or a respective region. For each site, n random individuals were selected from this vector without replacement, where n ¼ the number of all identified individuals from a given site. The number of unique species in the random draw was then reported, with this process being repeated 1,000 times. The mean number of unique species across all random samples was then recorded. This process was repeated across all sites. Observed site richness was then divided into mean random richness. A box plot was then generated showing the variation in observed: random richness across all sites in the continent or respective region. Statistical significance for differences in observed vs randomized richness was estimated using the paired Wilcoxon rank-sum test. Body-size spectra. Observed body-size spectra were determined across all sites by counting the number of species falling within each of 41 log2 volume classes with bins starting at 22.5 and increasing by 0.5 log2 units to 18.0. The first bin value ranged from 22.5 to 22.0 ( 0.18 to 0.25 mm3 ) and the last bin from 17.5– 18.0 (185 363.8 to 262 144 mm3 ). The proportion of the total site fauna found within each bin was then determined by dividing the observed bin richness by total site richness. Random expectations were then generated by randomly drawing n species, without replacement, from the continental or respective regional species pool, where n ¼ number of recorded species from that site. The proportion that each bin represented in the random draw was then calculated. This procedure was repeated 1,000 times, with the mean proportion across all randomizations being reported across all 41 bins. This process was repeated for each site. From this, body-size spectra graphs were generated showing across all bins the mean observed site representation and the minimum and maximum mean random values as drawn from the continental or respective regional pool. Regional scale patterns were illustrated through three latitudinal transects: western (Alaskan Interior, Pacific Northwest, Great Basin and southern Rockies), central (Churchill, Northwestern Minnesota, Upper Mississippi Valley and Ozarks), and eastern (Laurentian Plateau, New England, Carolinas, Peninsular Florida). Figure 1. Location of the 1,574 sample sites and 21 biogeographic regions in North America. Regional codes are: A, Churchill; B, Central Manitoba; C, Laurentian Plateau; D, Northwestern Minnesota; E, Lake Superior Uplands; F, Western Niagaran Escarpment; G, Eastern Niagaran Escarpment; H, New England; I, Upper Mississippi Valley; J, Ozarks; K, Central Appalachians; L, Southern Appalachians; M, Carolinas; N, Gulf Coast; O, Peninsular Florida; P, Southern Plains; Q, Southern Rockies; R, California; S, Great Basin; T, Pacific Northwest; U, Alaskan Interior. J. C. NEKOLA 240 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 141 RESULTS In all 1,204 species-level taxa are reported from North America, partitioned among 170 genera and 51 families (Nekola, in press). A total of 460 species (38% of the continental total) and 697,778 individuals were recorded from the 1,574 sites (Table 1). Regional faunistic richness ranged from 17 (Churchill) to 308 (Southern Rockies), with encountered species ranging from 15 (Churchill) to 113 (Southern Rockies). Median site richness ranged from 3 (Gulf Coast) to 20 (Upper Mississippi Valley and Ozarks), while maximum richness ranged from 10 (California) to 39 (Upper Mississippi Valley). The total number of encountered individuals ranged from 2,181 (Southern Appalachians) to 142,305 (Northwestern Minnesota), while the median number of individuals encountered per site ranged from 43 (Gulf Coast) to 587 (Upper Mississippi Valley). Family-level composition For the entire continental fauna (Table 2), the ten most frequently represented families were the Polygyridae (22.51%), Helminthoglyptidae (14.78%), Vertiginidae (8.64%), Oreohelicidae (6.98%), Pristilomatidae (5.40%), Oxychilidae (4.07%), Succineidae (4.07%), Gastrodontidae (2.91%), Arionidae (2.82%) and Urocoptidae (2.58%). This differs greatly from average site faunas, which were most frequently represented by the Vertiginidae (29.93%), Gastrodontidae (11.19%), Oxychilidae (7.60%), Euconulidae (7.09%), Punctidae (5.44%), Discidae (5.40%), Valloniidae (4.88%), Strobilopsidae (3.23%), Polygyridae (3.20%) and Ellobiidae (2.97%). Regional species pools demonstrated variability in familylevel representation, with the most common family ranging from the Vertiginidae (Churchill, Central Manitoba, Lake Superior Uplands, Laurentian Plateau, New England, Northwestern Minnesota, Eastern Niagaran Escarpment, Western Niagaran Escarpment, Upper Mississippi Valley and Alaskan Interior) to the Polygyridae (Southern Plains, Ozarks, Southern Appalachians, Central Appalachians, Carolinas, Gulf Coast and Peninsular Florida), Oreohelicidae (Pacific Northwest and Great Basin) and Helminthoglyptidae (California and Southern Rockies). The sign test demonstrates that five families show significant (and another three marginally significant) differences in their representation between site and regional scales (Table 3). The Euconulidae, Punctidae and Vertiginidae all exhibited significantly (Bonferroni adjusted P , 0.0028) more instances of mean site fauna proportion exceeding the respective regional species pool proportion than would be expected at random. The Gastrodontidae, Oxychilidae and Strobilopsidae exhibited this pattern to only a marginal (0.05 , P , 0.0028) degree. However, the Polygyridae and Succineidae demonstrated significantly more instances of mean site fauna proportion being less than the respective regional pool proportion. Ten families (Agriolimacidae, Cochlicopidae, Discidae, Ellobiidae, Haplotrematidae, Helicodiscidae, Pristilomatidae, Pupillidae, Valloniidae and Vitrinidae) exhibited similar proportional representations between average site faunas and the regional species pool. Because of limited occurrence (,10 regions occupied), statistical power was too low to conduct this test on 25 families. However, among these it should be noted that in all four regions supporting members of the small body-size Thysanophoridae, percent representation within site faunas always exceeded the proportion in respective regional species pools, while in the large body-size Helicidae, Helminthoglyptidae, Humboldtianidae, Hygromiidae, Monadeniidae and Oreohelicidae site fauna representation was always less than the respective regional pool. This was also the case in 75% of the regions supporting large body-size Orthalicidae. Site richness Observed site richness values were approximately seven times smaller than random draws taken from the entire continental dataset (Fig. 2). The paired Wilcoxon rank-sum test demonstrated that this difference was highly significant (P , 0.0000001). At the regional scale, observed richness generally ranged from 1.85 times smaller than random draws (Lake Superior Uplands) to 6.3 times smaller (Southern Rockies). Observed site faunas were 2.4 and 3.6 times less rich than average random draws from the regional individual pool for Churchill and the Alaskan Interior, respectively. This number ranged from 2.7 to 3.1 for Central Manitoba, the Laurentian Plateau, New England, Northwestern Minnesota, the Eastern and Western Niagaran Escarpments and the Upper Mississippi Valley. In the Southern Plains, Ozarks and Central Appalachians, individual site faunas were 2.6–2.2 times smaller than random draws from their respective regional individual pools. This number ranged from 3.4–3.9 in the Carolinas, Gulf Coast and Peninsular Florida, and 3.1–4.7 in the Pacific Northwest, Great Basin and California. The paired Wilcoxon rank-sum test demonstrated that these differences were all highly significant (P , 0.0000027 with a Bonferroni-adjusted P-value threshold of 0.00238). The main exception was the Southern Appalachians, which had observed richness ranging between c. 1.3 times larger to 4.2 times smaller than the size of a random draw from the regional individual pool (median ¼ 1.5 times smaller). While sites were generally less rich than random draws from the entire regional individual pool, this difference was only marginally significant (P ¼ 0.007813). Body-size spectra Observed mean body-size spectra within sites differed strongly from random draws made from the continental species pool (Fig. 3). While observed site faunas were strongly right-skewed with a mode at the 5th log2 body-size bin (0.71–1 mm3 ), random draws of the same richness from the continental species pool were strongly left-skewed, with modal values ranging from the 21st–24th log2 body-size bin (181–512 mm3 ). Site faunas were greatly enriched in species with biovolumes of 5.7 mm3 or less, and greatly underrepresented in species with biovolumes of 90.5 mm3 or greater. These differences are highly significant, with the site-scale mode for species with biovolumes ,5.7 mm3 being more than three times larger than the maximum observed randomized mean at that same size class, and the minimum score of the randomized means for species with biovolumes .90.5 mm3 being at least five times larger than observed scores for these same body-size classes. Species with biovolumes of 5.7– 90.5 mm3 were represented at approximately equal proportions within observed and randomized site faunas. The enrichment of small species and impoverishment of large species at site scales was also present within most biogeographic regions (Fig. 4). Among the 12 illustrated relationships in Figure 4, the Pacific Northwest, Great Basin, Southern Rockies, Upper Mississippi Valley, Ozarks, Laurentian Plateau, New England and Carolinas all strongly demonstrate this pattern. It is also more weakly exhibited in Northwestern Minnesota and Peninsular Florida. Very similar observed vs randomized bodysize spectra were noted for the Alaskan Interior and Churchill. DISCUSSION These analyses clearly show that family composition, body-size spectra and species richness are not self-similar and scale independent among North American terrestrial gastropods. While the continental fauna is dominated by families with large-sized individuals like the Polygyridae and Helminthoglyptidae, NORTH AMERICA LAND SNAIL BIODIVERSITY 241 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 142 Table2.ComparisonofterrestrialgastropodfamilypercentrepresentationacrossobservationalscalesinNorthAmerica. FamilyEntireContinentChurchillCentralManitobaLakeSuperior Uplands LaurentianPlateauNewEnglandNWMinnesotaENiagaraWNiagaraUpperMississippiSouthernPlainsOzarks R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S% Polygyridae22.513.205.460.3412.780.501.590.0312.083.988.162.7810.686.4832.4718.5429.5318.40 Helminthoglyptidae14.780.28 Vertiginidae8.6429.9364.7168.1135.8531.5627.2720.9626.2536.9315.7930.9331.7524.4615.4521.8620.4128.0621.3631.5812.9926.5515.4423.49 Oreohelicidae6.980.37 Pristilomatidae5.402.661.89–1.820.612.260.473.183.062.692.693.062.661.944.562.608.506.716.40 Oxychilidae4.077.603.7712.057.2711.785.0014.269.779.173.189.537.389.008.165.114.855.124.555.736.719.11 Succineidae4.072.7511.7711.1713.213.345.461.057.501.716.021.967.944.326.711.639.183.6210.682.247.149.006.710.52 Gastrodontidae2.9111.195.668.029.0921.716.257.537.5219.706.3511.927.3816.636.1212.615.835.663.255.815.379.75 Arionidae2.82–3.64–11.25–6.02–3.18–6.04–4.08–3.88– Urocoptidae2.580.181.95– Discidae2.245.405.88–3.7713.085.4612.703.758.662.263.734.766.704.037.904.087.184.855.141.300.993.362.36 Helicodiscidae2.082.763.77–3.643.741.250.983.013.543.181.762.013.933.064.483.885.955.205.772.694.68 Phylomicidae1.74–5.26–4.70–1.02–2.91–1.95–4.03– Haplotrematidae1.330.440.750.090.670.291.020.070.971.130.671.70 Orthalicidae1.330.232.601.130.671.35 Monadeniidae1.330.03 Pupillidae1.161.885.882.711.89–1.250.281.500.413.180.081.340.922.040.051.941.770.653.730.672.50 Valloniidae1.084.889.434.477.276.578.756.714.514.669.526.634.703.477.143.735.835.083.901.562.691.19 Euconulidae0.917.095.8813.783.7710.255.465.332.5012.793.017.253.187.232.695.924.086.333.885.411.953.052.694.40 Ellobiidae0.912.973.773.733.642.122.50–1.503.683.186.332.014.452.045.081.943.981.300.480.672.21 Punctidae0.835.443.776.991.824.942.502.741.507.244.766.392.016.063.065.732.915.130.652.441.344.94 Humboldtianidae0.830.00 Helicinidae0.750.631.820.151.021.650.971.733.251.130.671.08 Binneyidae0.75– Subulinidae0.660.280.65– Helicidae0.660.053.750.143.760.034.030.591.02–2.60–1.34– Hygromiidae0.660.023.75–1.500.142.690.06 Megomphicidae0.660.01 Veronicellidae0.66– Agriolimacidae0.501.305.884.241.891.961.820.103.750.151.500.963.181.692.012.562.041.731.943.270.65–1.340.45 Strobilopsidae0.423.231.891.101.824.601.25–3.014.173.184.162.013.973.064.892.914.242.604.452.694.44 Thysanophoridae0.420.971.30– Truncatellidae0.420.130.65– Limacidae0.42–2.50–3.01–3.36–1.02–1.30–2.01– Spiraxidae0.330.132.600.51 Pleurodontidae0.330.06 Cochlicopidae0.251.993.772.375.462.542.501.772.261.243.184.052.012.843.063.672.911.251.430.22 Vitrinidae0.171.431.891.101.820.761.255.360.750.151.591.650.671.231.020.420.970.11 Pomatiidae0.170.12 Pomatiopsidae0.170.070.75–0.670.051.020.141.940.180.670.83 Ferrussaciidae0.170.00 Charopidae0.170.09 Sagdidae0.080.07 Oleacinidae0.080.07 Bradybaenidae0.080.03 Streptaxidae0.080.02 Helicarionidae0.080.01 Cepolidae0.080.00 Cochlicellidae0.08– Milacidae0.08–1.25– Testacellidae0.08–1.25–0.67– J. C. NEKOLA 242 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 143 FamilySouthern Appalachians Central Appalachians CarolinasGulfCoastPeninsularFloridaAlaskanInteriorPacificNorthwestGreatBasinCaliforniaSouthernRockies R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S%R%S% Polygyridae30.2212.8023.1711.0725.194.1630.138.3114.6212.709.634.348.862.9217.861.93 Helminthoglyptidae1.07–38.7512.1224.350.51 Vertiginidae6.1217.739.7623.4914.5036.8312.8241.3613.0817.5345.4659.0912.3025.6619.7419.837.7522.2312.6634.17 Oreohelicidae14.971.1028.957.470.37–12.341.46 Pristilomatidae12.9514.397.327.341.531.833.210.521.543.275.885.141.32–2.955.980.334.13 Oxychilidae12.5913.8210.9811.359.1618.208.9714.204.624.816.063.313.216.433.951.392.210.901.625.75 Succineidae1.80–4.891.227.63–8.330.396.931.309.099.946.422.0313.163.593.691.942.270.57 Gastrodontidae8.2717.9210.1610.876.119.997.6917.303.082.541.609.982.632.480.7411.310.657.21 Arionidae2.85–2.29–14.44–7.75–0.33– Urocoptidae2.313.478.440.50 Discidae5.400.382.851.412.290.361.28–6.063.553.216.522.636.941.482.640.654.13 Helicodiscidae5.04–4.473.552.292.982.560.343.081.540.540.230.370.330.651.12 Phylomicidae4.68–4.89–3.05–3.21–1.54–0.33– Haplotrematidae0.721.910.411.940.761.400.64–1.603.044.065.56 Orthalicidae0.36–1.28–5.393.831.620.49 Monadeniidae1.070.195.901.11 Pupillidae0.36–0.810.380.760.330.64–1.541.973.035.771.070.846.5814.950.370.673.907.31 Valloniidae1.08–2.030.642.290.641.281.681.544.316.061.342.144.176.5811.820.742.392.278.87 Euconulidae1.806.832.036.733.055.363.216.953.855.429.0914.271.078.632.636.680.744.200.656.95 Ellobiidae2.164.242.444.652.291.081.280.250.770.571.600.830.370.330.330.04 Punctidae1.448.461.637.773.056.141.282.430.770.246.060.881.6011.541.324.972.2120.870.972.87 Humboldtianidae3.250.04 Helicinidae0.72–0.410.520.64–2.314.77 Binneyidae4.28–0.37– Subulinidae0.41–1.530.291.28–6.157.170.37–0.33– Helicidae1.630.441.53–2.56–1.54–1.60–1.32–2.580.480.65– Hygromiidae2.29–0.64–0.77–0.37– Megomphicidae1.600.351.85– Veronicellidae6.15– Agriolimacidae0.72–0.811.311.530.910.640.340.770.133.031.342.670.312.631.091.111.220.970.31 Strobilopsidae1.08–1.633.192.299.092.565.612.312.86 Thysanophoridae1.544.311.076.191.324.680.654.78 Truncatellidae2.313.400.74– Limacidae0.72–1.63–1.53–1.92–2.14–1.32–1.48–0.65– Spiraxidae0.76–0.640.340.773.04 Pleurodontidae3.081.62 Cochlicopidae1.080.950.811.280.760.423.030.131.070.121.320.200.37–0.332.41 Vitrinidae0.41–3.030.390.541.151.3213.650.372.810.333.99 Pomatiidae1.543.34 Pomatiopsidae0.72–0.810.850.76–0.64– Ferrussaciidae0.41–1.540.07 Charopidae0.541.201.320.260.330.45 Sagdidae0.771.77 Oleacinidae0.771.93 Bradybaenidae0.770.69 Streptaxidae0.64–0.770.62 Helicarionidae0.770.33 Cepolidae0.770.07 Cochlicellidae0.76–0.37– Milacidae0.41–0.54–0.37–0.33– Testacellidae0.54–0.37– R%,thepercentageofspeciesinthecontinental/regionalfaunaoccurringwithinthatfamily.S%,thepercentagerepresentationofthatfamilywithinanaveragesitewithinthatregion.Dashesindicatefamiliesthatarepresentintherespectivefaunabuthavenotyetbeenencounteredwithinsites. NORTH AMERICA LAND SNAIL BIODIVERSITY 243 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 144 typical site faunas are dominated by families with small-sized individuals like the Vertiginidae and Gastrodontidae. Likewise, while the continental fauna demonstrates a left-skewed body-size spectrum dominated by species with biovolumes .90.5 mm3 , average site faunas are right-skewed and dominated by species with biovolumes ,5.7 mm3 . Random draws representing the same number of individuals from the entire continental dataset contained on average at least seven times more species than the actual number observed at that site. These discrepancies were generally also observed between sites and their respective biogeographic regions. Typically, families with small-sized individuals (Euconulidae, Gastrodontidae, Oxychilidae, Punctidae, Strobilopsidae, Thysanophoridae and Vertiginidae) were overrepresented in typical site faunas as compared to their respective regional species pool, while families with large-sized individuals (Helminthoglypidae, Monadeniidae, Oreohelicidae, Polygyridae and Succineidae) were underrepresented. This bias is demonstrated across most regions (e.g. Pacific Northwest, Great Basin, Southern Rockies, New England and Carolinas). However, both arctic regions tended to show little discrepancy between typical site body-size spectra and the regional pool. Site richness also generally ranged from two to six times smaller than would be expected from random draws from the respective regional individual pool. Even in the case of the Southern Appalachians, where the eight sampled sites were only collected across a 20-km extent and only 42 total species were encountered (15% of the entire regional fauna) the majority of sites were still found to have fewer species than would be expected from a random draw of all encountered individuals. As more samples are made across a larger extent within this region and more of this regional fauna is encountered, the distribution of observed to random ratio scores will be lowered and made to fall in line with other regions. These analyses indicate that site faunas are not a simple random sample drawn from the continental or regional pool. Rather, important biases are present. The most important of these are not only the predictable filtering of species along environmental gradients (principally soil architecture, moisture and acidity; Nekola, 2003, 2010) but also purely geographic filtering caused by dispersal limitation and other forms of spatial constraint (Nekola & White, 1999; Hubbell, 2001). As a result, it is not possible to extrapolate accurately these macroecological metrics across scales for North American terrestrial gastropods. Rather, the fauna looks considerably different depending upon whether one is looking down from continental scales, or up from site scales. From the continental-scale perspective, the fauna is generally dominated by families with largesized individuals. However, from the site scale the fauna is dominated by a series of families with small body-sized individuals. Documenting the faunas of a few individual sites will thus not allow for an accurate portrayal of the continental fauna and knowledge of the continental fauna informs little about what one would expect within individual sites. The lack of self-similarity within the North American terrestrial gastropod fauna appears to be principally caused by a strong inverse correlation between body and range size. In this fauna, many small species (especially within the Euconulidae, Gastrodotidae, Oxychilidae, Punctidae, Strobilopsidae and Table 3. Comparison of family representation at regional vs site scales for terrestrial gastropods in North America. A. 10 or more regions B. ,10 regions Family R . S S . R P-value Family R . S S . R Agriolimacidae 14 5 0.063568 Bradybaenidae 1 0 Cochlicopidae 11 5 0.210114 Cepolidae 1 0 Discidae 6 12 0.237885 Charopidae 1 2 Ellobiidae 11 8 1.0 Ferrussaciidae 1 0 Euconulidae 1 20 0.000021 Helicarionidae 1 0 Gastrodontidae 3 16 0.004425 Helicidae 5 0 Haplotrematidae 3 7 0.343750 Helicinidae 3 5 Helicodiscidae 7 9 0.803619 Helminthoglyptidae 2 0 Oxychilidae 5 15 0.041389 Humboldtianidae 1 0 Polygyridae 16 0 0.000031 Hygromiidae 2 0 Pristilomatidae 7 9 0.803619 Megomphicidae 1 0 Punctidae 2 18 0.000402 Monadeniidae 2 0 Pupillidae 10 7 0.629059 Oleacinidae 0 1 Strobilopsidae 1 12 0.003418 Oreohelicidae 3 0 Succineidae 17 2 0.000729 Orthalicidae 3 1 Valloniidae 12 7 0.359283 Pleurodontidae 1 0 Vertiginidae 3 18 0.001490 Pomatiidae 0 1 Vitrinidae 6 7 1.0 Pomatiopsidae 3 2 Sagdidae 0 1 Spiraxidae 2 1 Streptaxidae 1 0 Subulinidae 2 1 Thysanophoridae 0 4 Truncatellidae 0 1 Urocoptidae 1 1 R . S represents the number of regions in which family representation in the regional pool exceeded the representation on an average site. S . R represents the number of regions in which family representation within an average site exceeded the representation in the regional pool. Assuming the null of R . S ¼ S . R, P-values for families present in at least ten regions are based on the bionomial sign test. Families with P , 0.0028 have been highlighted in bold font; families with 0.05 , P , 0.0028 have been highlighted in italic font. J. C. NEKOLA 244 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 145 Vertiginidae) tend to have component species harboring extensive ranges. Examples include Euconulus fulvus (Mu¨ ller, 1774), E. alderi Gray, 1840, Punctum minutissumum (I. Lea, 1841) and Vertigo arthuri von Martens, 1884, which range across the entire continent north of 388N, while also extending south along the crest of the Rockies to the Mexican border. Gastrocopta pellucida (Pfeiffer, 1841), Glyphyalinia umbilicata (Singley, in Cockerell, 1893), Hawaiia miniscula (A. Binney, 1840) and Straitura meridionalis (Pilsbry & Ferriss, 1906) likewise extend across much of North America south of 358N. And, both Strobilops labyrinthica (Say, 1817) and Strobilops aenea Pilsbry, 1926 range across most of the southeastern quarter of the continent. A number of genera with large body-sized species, however, demonstrate significant levels of allopatric replacement and local endemism. These are especially pronounced within the Polygyridae (e.g. Ashmunella, Daedalochila, Stenotrema, Inflectarius, Patera, Triodopsis, Trilobopsis and Vespericola), Helminthoglyptidae (e.g. Helminthoglypta, Micrarionta and Sonorella), Oreohelicidae (e.g. Oreohelix and Radiocentrum), Urocoptidae (e.g. Holospira and Coelostemma), Monadeniidae (Monadena) and Humboldtianidae (Humboldtiana), with it being uncommon for species within these genera to be sympatric at the site scale. It seems likely that the combined effects of body size on passive dispersal and uniparental reproduction rates help explain this counter-intuitive pattern. Because terrestrial gastropods are among the poorest active dispersers known among terrestrial animals, with individuals moving perhaps no more than 1–10 m over their lifetime (Schilthuizen & Lombaerts, 1994) and populations being unable actively to cross barriers of only 100–1,000 m (Baur, 1988; Schilthuizen & Lombaerts, 1994), Figure 2. Box plot showing distribution of observed site richness vs randomized expectations from the regional/continental individual pool of terrestrial gastropods in North America. The site vs continental-scale comparison is identified by grey fill. Figure 3. Site vs continental body-size spectra for terrestrial gastropods of North America as represented by 41 log2 body-size classes. The solid line represents the observed average for individual sites. The upper and lower dashed lines represent the maximum and minimum mean scores (respectively) observed for that size class across 1,000 random samples drawn per site from the continental pool. The range between these extremes has been filled with grey. NORTH AMERICA LAND SNAIL BIODIVERSITY 245 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 146 passive dispersal takes precedence in determining population movement and species range size. Small snails/slugs are much more easily moved via passive vectors than larger ones, because small-sized individuals are less likely to be pulled off by gravity or fluid mechanics during movement. Additionally, a number of North American genera with species of small body-size contain some that are capable of uniparental reproduction (Pokryszko, 1987; Bulman, 1990). For these, movement of only a single unmated individual is required to found a new population, greatly increasing the effectiveness of long-range dispersal events in causing range expansion. DNA sequence data illustrate, for instance, that members of the genus Balea have been repeatedly carried across 9,000 km of open sea in the eastern Atlantic Ocean by migrating birds (Gittenberger et al., 2006). Species within the Vertigo gouldii group in North America may possess ranges exceeding 5,000 km in extent even though their modern Figure 4. Site vs regional body-size spectra of terrestrial gastropods in North America, as represented by 41 log2 body-size classes across twelve selected biogeographic regions. Panels are arranged in three columns (western, central and eastern North America) from most northern (top) to most southern (bottom). In each panel, the solid line represents the observed average for individual sites in that region. The upper and lower dashed lines represent the maximum and minimum mean scores (respectively) observed for that size class across 1,000 random samples drawn per site from the respective regional pool. The range between these extremes has been filled with grey. J. C. NEKOLA 246 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 147 ranges were covered by continental ice as recently as 12,000 yr BP (Nekola, Coles & Bergthorsson, 2009). This is not a universal pattern, however, with many Pristilomatidae and Helicodiscidae exhibiting local endemism in the eastern USA (e.g. Helicodiscus, Paravitrea, Pilsbryna and Polygyriscus). Comparison of the breeding systems and ecology between large-range vs small-range small body-sized species will undoubtedly document important differences. For instance, many locally endemic Helicodiscus, Paravitrea and Polygyriscus species tend to be restricted to highly insular rock talus habitats, perhaps making it easier for their populations to become isolated and speciate. Because similar analyses have not yet been performed in other regions, it cannot be conclusively stated whether strong scaledependence is a general phenomenon or is unique to North America terrestrial gastropods. Because a strong inverse correlation between body and range size (Pokryszko & Cameron, 2005; Cameron, Pokryszko & Horsa´ k, 2010, 2012) in combination with an overabundance of small individuals and large species (Nekola et al., 2013) exists in the European fauna, it seems likely that scale-dependence also occurs there. Preliminary investigations of the New Zealand fauna (Nekola et al., 2013), however, show apparent strong coupling in species richness and individual abundance across the body-size spectrum. Given that this fauna, as well as that of eastern Australia (Stanisic et al., 2007, 2010), also harbours many narrow-range endemics of small body size (especially in the Charopidae and Punctidae), self-similarity within these faunas is possible. Yet, the commonness in western Australia of large body-size camaenids demonstrating strong alloptaric replacement with smaller species demonstrating extensive ranges (Solem, 1988; Cameron, 1992) suggests that at least parts of this continental fauna may behave in a fashion similar to that in North America. Lastly, it is important to point out that at least some of the ‘self-similar’ patterns reported by Finlay et al. (2006) are almost certainly not rooted in ecological mechanisms. For instance, S-shaped species-abundance distributions along a rankabundance axis and power-law species-area relationships have both been shown to be common expectations across a wide variety of complex systems spanning not only the physical, biological and social sciences but also the arts (Nekola & Brown, 2007). As a result, the similarity of these patterns across scales only means that these systems have remained complex across these scales—and no more. Given that complex systems typically demonstrate nesting, with a given system being made up of complex parts, while at the same time being a component of even larger-scale complex systems (Brown, 1994; West, 2006), these findings are to be expected and are likely mathematically trivial. Because of scale-dependence, no easy mathematical short-cut exists for documenting the North American terrestrial gastropod fauna. It is not only impossible to look at a few sites and be able accurately to assemble continental or most regional scale biodiversity patterns, but it is also equally impossible to deduce the makeup of site faunas given knowledge of the continental or regional pool. This fauna can only be accurately known by looking at many sites spread across all biogeographic regions and then summarizing these data to allow for expression of emergent properties. The only way this can be accurately accomplished is thus through extensive, time consuming and expensive fieldwork. ACKNOWLEDGEMENTS The generation of the continental and region-scale datasets was made possible by the invaluable help of Gabe Martinez, Sarah Martinez, and Khayman Mondragon of the Albuquerque Institute for Math and Science. The generation of site-scale data from across North America is due to the generous support of a large number of organizations and people. While most have been mentioned in previous papers, I would especially like to thank for recent aid Ken Hotopp of Appalachian Conservation Biology for fieldwork in the middle Appalachians, the Bailey-Matthew Shell Museum for work in Peninsular Florida, the Valhalla Wilderness Society for fieldwork in southeastern British Columbia and adjacent areas in Idaho, Charles Drost of the USGS for fieldwork in the Channel Islands and along the California coast. I would also like especially to thank Michal Horsa´ k and Brian Coles for helping keep the lights on in my lab though their generous personal donations. This work would simply have been impossible without their assistance. REFERENCES BAUR, B. 1988. Microgeographical variation in shell size of the land snail Chondrina clienta. Biological Journal of the Linnean Society, 35: 247–259. BOUCHET, P. & ROCROI, J.-P. 2005. Classification and nomenclator of gastropod families. Malacologia, 47: 1–397. BROWN, J.H. 1994. Complex ecological systems. In: Complexity: metaphors, models, and reality (G.A. Cowan, D. Pines & D. Melzer, eds), pp. 419–449. Santa Fe Institute Studies in the Science of Complexity, Proceedings Volume XVIII. Addison-Wesley, Reading. BULMAN, K. 1990. Life history of Carychium tridentatum (Risso, 1826) (Gastropoda: Pulmonata: Ellobiidae) in the laboratory. Journal of Conchology, 33: 321–333. BURCH, J.B. 1962. The eastern land snails. W. C. Brown Co., Dubuque, IA. CAMERON, R.A.D. 1992. Land snail faunas of the Napier and Oscar Ranges, Western Australia: diversity, distribution and speciation. Biological Journal of the Linnean Society, 45: 271–286. CAMERON, R.A.D., POKRYSZKO, B.M. & HORSA´ K, M. 2010. Land snail faunas in Polish forests: patterns of richness and composition in a post-glacial landscape. Malacologia, 53: 77–134. CAMERON, R.A.D., POKRYSZKO, B.M. & HORSA´ K, M. 2012. Forest snail faunas from Crimea (Ukraine), an isolated and incomplete Pleistocene refugium. Biological Journal of the Linnean Society, 109: 424–433. FINLAY, B.L., THOMAS, J.A., MCGAVIN, G.C., FENCHEL, T. & CLARKE, R.T. 2006. Self-similar patterns of nature: insect diversity at local to global scales. Proceedings of the Royal Society B, 273: 1935–1941. FORSYTH, R.G. 2004. Land snails of British Columbia. Royal British Columbia Museum handbook. Royal British Columbia Museum, Victoria. GITTENBERGER, E., GROENENBERG, D.S.J., KOKSHOORN, B. & PREECE, R.C. 2006. Molecular trails from hitch-hiking snails. Nature, 439: 409. HENTSCHEL, H.G.E. & PROCACCIA, I. 1984. Relative diffusion in turbulent media: the fractal dimension of clouds. Physics Review A, 29: 1461–1470. HUBBELL, S.P. 2001. The unified neutral theory of biodiversity and biogeography. Monographs in Population Biology 32. Princeton University Press, Princeton, NJ. HUBRICHT, L. 1985. The distributions of the native land mollusks of the eastern United States. Fieldiana, New Series, 24: 1–191. MCCLAIN, C. & NEKOLA, J.C. 2008. The role of local-scale on terrestrial and deep-sea gastropod body size distributions across multiple scales. Evolutionary Ecology Research, 10: 129–146. MANDELBROT, B. 1983. The fractal geometry of nature. W.H. Freeman, San Francisco. METCALF, A.L. & SMARTT, R.A. 1997. Land snails of New Mexico. New Mexico Museum of Natural History and Science, Albuquerque, New Mexico. MILNE, B.T. 1988. Measuring the fractal geometry of landscapes. Applied Mathematics and Computation, 27: 67–79. NEKOLA, J.C. 2003. Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Diversity and Distribution, 9: 55–71. NEKOLA, J.C. 2010. Acidophilic terrestrial gastropod communities of North America. Journal of Molluscan Studies, 76: 144–156. NORTH AMERICA LAND SNAIL BIODIVERSITY 247 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 148 NEKOLA, J.C. in press. Overview of the North American terrestrial gastropod fauna. American Malacological Bulletin. NEKOLA, J.C., BARKER, G.M., CAMERON, R.A.D. & POKRYSZKO, B.M. 2013. Latitudinal variation of body size in land snail populations and communities. In: Global patterns of body size (F. Smith & K. Lyons, eds), pp. 62–82. University of Chicago Press, Chicago. NEKOLA, J.C. & BROWN, J.H. 2007. The wealth of species: ecological communities, complex systems, and the legacy of Frank Preston. Ecology Letters, 10: 188–196. NEKOLA, J.C. & COLES, B.F. 2010. Pupillid land snails of eastern North America. American Malacological Bulletin, 28: 29–57. NEKOLA, J.C., COLES, B.F. & BERGTHORSSON, B. 2009. Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Molecular Phylogenetics and Evolution, 53: 1010–1024. NEKOLA, J.C. & WHITE, P.S. 1999. Distance decay of similarity in biogeography and ecology. Journal of Biogeography, 26: 867–878. NITTMAN, J., DACCORD, G. & STANLEY, H.E. 1985. Fractal growth of viscous fingers: quantitative characterization of fluid instability phenomenon. Nature, 314: 141–144. PALMER, M.W. 1988. Fractal geometry: a tool for describing spatial patterns in plant communities. Vegetatio, 75: 91–102. PILSBRY, H.A. 1948. Land Mollusca of North America (north of Mexico). Academy of Natural Sciences of Philadelphia Monograph, 1(1)–2(2). POKRYSZKO, B.M. 1987. On aphally in the Vertiginidae (Gastrocopta: Pulmonata: Orthurethra). Journal of Conchology, 32: 365–375. POKRYSZKO, B.M. & CAMERON, R.A.D. 2005. Geographical variation in the composition and richness of forest snail faunas in northern Europe. Records of the Western Australian Museum, Supplement, 68: 115–132. ROTH, B. & SADEGHIAN, P.S. 2003. Checklist of the land snails and slugs of California. Santa Barbara Museum of Natural History Contributions in Science, 3: 1–81. SCHILEYKO, A.A. 2006. Treatise on recent terrestrial pulmonate molluscs, Part 13. Helicidae, Pleurodontidae, Polygyridae, Ammonitellidae, Oreohelicidae, Thysanophoridae. Ruthenica, Suppl. 2: 1765–1906. SCHILTHUIZEN, M. & LOMBAERTS, M. 1994. Population structure and levels of gene flow in the Mediterranean land snail Albinaria corrugata (Pulmonata: Clausiliidae). Evolution, 48: 577–586. SOLEM, A. 1988. Maximum in the minimum: biogeography of land snail from the Ningbing Ranges and Jeremiah Hills, north-east Kimberly, Western Australia. Journal of the Malacological Society of Australia, 9: 59–113. STANISIC, J., CAMERON, R.A.D., POKRYSZKO, B.M. & NEKOLA, J.C. 2007. Forest snail faunas from S.E. Queensland and N.E. New South Wales (Australia): patterns of local and regional richness and differentiation. Malacologia, 49: 445–462. STANISIC, J., SHEA, M., POTTER, D. & GRIFFITHS, O. 2010. Australian land snails Volume 1: a field guide to eastern Australian species. Queensland Museum, Brisbane. TURGEON, D.D., QUINN, J.F., BOGAN, A.E., COAN, E.V., HOCHBERG, F.G., LYONS, W.G., MIKKELSEN, P., NEVES, R.J., ROPER, C.F.E., ROSENBERG, G., ROTH, B., SCHELTEMA, A., THOMPSON, F.G., VECCHIONE, M. & WILLIAMS, J.D. 1998. Common and scientific names of aquatic invertebrates from the United States and Canada: mollusks. Edn 2. American Fisheries Society Special Publication 26. Bethesda, MD. WEST, B.J. 2006. Thoughts on modeling complexity. Complexity, 11: 33–43. WIENS, J.A. 1989. Spatial scaling in ecology. Functional Ecology, 3: 385–397. J. C. NEKOLA 248 atUniversityofNewMexicoonSeptember9,2014http://mollus.oxfordjournals.org/Downloadedfrom 149 Journal of Biogeography, 26, 867–878 The distance decay of similarity in biogeography and ecology Jeffrey C. Nekola∗ and Peter S. White Curriculum in Ecology, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599, U.S.A. Abstract Aim Our aim was to understand how similarity changes with distance in biological communities, to use the distance decay perspective as quantitative technique to describe biogeographic pattern, and to explore whether growth form, dispersal type, rarity, or support affected the rate of distance decay in similarity. Location North American spruce-fir forests, Appalachian montane spruce-fir forests. Methods We estimated rates of distance decay through regression of log-transformed compositional similarity against distance for pairwise comparisons of thirty-four white spruce plots and twenty-six black spruce plots distributed from eastern Canada to Alaska, six regional floras along the crest of the Appalachians, and six regional floras along the east–west extent of the boreal forest. Results Similarity decreased significantly with distance, with the most linear models relating the log of similarity to untransformed distance. The rate of similarity decay was 1.5–1.9 times higher for vascular plants than for bryophytes. The rate of distance decay was highest for berry-fruited and nut-bearing species (1.7 times higher than plumose-seeded species and 1.9 times higher than microseeded/spore species) and 2.1 times higher for herbs than woody plants. There was no distance decay for rare species, while species of intermediate frequency had 2.0 times higher distance decay rates than common species. The rate of distance decay was 2.7 times higher for floras from the fragmented Appalachians than for floras from the contiguous boreal forest. Main conclusions The distance decay of similarity can be caused by either a decrease in environmental similarity with distance (e.g. climatic gradients) or by limits to dispersal and niche width differences among taxa. Regardless of cause, the distance decay of similarity provides a simple descriptor of how biological diversity is distributed and therefore has consequences for conservation strategy. Keywords Similarity, spatial dependence, distance decay, biological diversity, boreal forest. INTRODUCTION analyses of spatial autocorrelation and eventually to the field of geostatistics, a field that has grown tremendously in the last The similarity between two observations often decreases or decade (Cressie, 1993). decays as the distance between them increases, a pattern long While ecologists and biogeographers have rarely referred to recognized in the geographical literature and once called ‘the distance decay per se, a negative relationship between distance first law of geography’ (Tobler, 1970). That early interest in and similarity is implicit in several ecological and evolutionary the distance decay of similarity led subsequently to formal phenomena. For example, species turnover along spatial environmental gradients produces a decrease of similarity with ∗ Present address: Department of Natural and Applied Science, distance (e.g. Whittaker, 1975; Cody, 1985). Mass effect (Shmida University of Wisconsin-Green Bay, Green Bay, Wisconsin 52311, U.S.A. & Ellner, 1984), source-sink dynamics in metapopulations email: nekolaj@uwgb.edu (Hanski & Gilpin, 1991), and supply side ecologyCorresponding author: Peter S. White, Department of Biology, (Roughgarden, Gaines & Pacala, 1987) have stressed theCB#3280, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599, U.S.A. email: pswhite@unc.edu importance of dispersal and therefore distance in explanations © 1999 150 868 J. C. Nekola and P. S. White of population dynamics and community composition (Palmer, DISTANCE DECAY OF BIOLOGICAL SIMILARITY: CONCEPTUAL FRAMEWORK1988). Studies in island biogeography have found a decrease in percent species saturation of oceanic or habitat islands as a function of their distance from a source pool of immigrants Distance decay and geostatistics (Vuilleumier, 1970; Kadmon & Pulliam, 1993), producing a Our approach to the distance decay of similarity is parallel to, decrease in the similarity of species lists with distance. Dispersal, but methodologically different from, geostatistical approaches migration, and gene flow produce geographic patterns in the to the analysis of spatial autocorrelation (Cressie, 1993). It is distribution of genetic variation and in the distribution of now standard practice to use semi-variograms to characterize species that result from vicariance and adaptive radiations the relationship of variance of a parameter to intersample (Meyers & Giller, 1988; Briggs, 1991), leading to spatially distance or lag (Ripley, 1988), a critical step in the optimal autocorrelated distributional patterns that were described in interpolation technique called kriging (Burrough, 1986; Webster the early biogeographic literature under the age and area and & Oliver, 1990). To construct a semi-variogram, variance is centre of origin hypotheses (Cain, 1944; Willis, 1922). calculated by comparing values of the parameter at two These examples suggest two very different causes of locations of known distance apart. The statistic we model here, biological distance decay. On the one hand, a decay of similarity similarity, is a parallel concept to variance in the sense that it results from a simple decrease in environmental similarity results from comparison of two observations from different with distance (e.g. movement along a topographic or climatic locations. In our case the observations consist of species lists; gradient). In this case, the underlying explanation is competitive similarity is a complex function of the occurrence of many sorting of species with different physiological abilities; species (Faith, Minchin & Belbin, 1987). By contrast, variance composition, at least after sufficient time has elapsed, is is calculated for univariate factors (e.g. soil pH, rainfall, or predictable from environment. We term this source of biological abundance of a single species). distance decay the niche difference model. Variance and similarity both measure differences between The second cause of distance decay is that the spatial samples and can be modelled as a function of intersample configuration (the size and isolation of habitats), spatial context distance. In this paper, we use linear regression to model that (the nature of the matrix surrounding the habitats), and time relationship. The intercept of this model is similar to the nugget influence species and gene movement across landscapes. effect (residual intersample variance as distance decreases to Variation in distance decay rates among landscapes occurs in zero) in semi-variograms. The shape of the similarity-distance this case because different landscapes have different resistances relationship can be examined for evidence of a range and sill to the movement of organisms. Variation in distance decay which, in semi-variograms are, respectively, the distance at rates among different species groups will also arise if the groups which variance becomes independent of further increases in have different dispersal abilities—in essence, different species distance and the amount of variance at that distance. Regression will perceive a given spatial configuration of habitats differently. of similarity against distance has the advantage in our case of Variation in landscape resistance to movement and in dispersal producing a direct estimate of the rate of distance decay and abilities suggest an interaction between the spatial template thus the rate of composition change through space. One of our and time: the greater the resistance to movement or the less purposes is to ask whether species with different dispersal types vagile the organism, the greater the amount of time required or growth forms show different rates of compositional change. for the organism to cross a given spatial template. As a result, the nature of the spatial configuration of habitats will affect how long the effects of historic events persists in terms of Scale dependence and distance decay distributional patterns of species. When time or dispersal are limiting, current environmental conditions will not fully explain Grain, extent and related geostatistical concepts Observations of distance decay will show scale dependence.species distributions or local composition. We term this source of distance decay the model of temporal and spatial constraint: Ecologists have divided scale into grain and extent (Allen & Hoekstra, 1991; Weins, 1989). Grain, which is similar to thehistory and the nature of the spatial template explain species distributions. At large scales of space and time, the imprint of concept of support or resolution in the geostatistical literature (Odland, 1988; Webster & Oliver, 1990; Cressie, 1993), is thepast events has been termed historic biogeography (Schluter & Ricklefs, 1993). contiguous area over which a single observation is made (e.g. quadrat, plot, or pixel). In this sense it is a special case ofIn this paper we explore distance decay as a descriptor of compositional variation. We first discuss scale dependence in support, which is defined as an n-dimensional volume within which a linear averaged value of a regionalized value maythe detection of distance decay and develop a conceptual framework for distance decay. We then examine distance decay be computed (Olea, 1991). Spatial extent, called area in the geostatistical literature (Cressie, 1993), is the space over whichrates in boreal and montane spruce-fir forests, comparing the rate of similarity decay in several species groups (vascular observations are made. The grain size or support of samples will effect observedplants, mosses, species abundance classes, growth form classes, and dispersal classes) across the same set of plots. Finally, similarities, a phenomenon known as the support effect in the geostatistical literature. Consider two areas with identicalwe discuss the implications of distance decay perspective for community composition, biogeographical pattern, and species lists. If a small quadrat is sampled from each area, species-area relationships dictate that these quadrats will haveconservation. © Blackwell Science Ltd 199926867–878 151 Distance decay of similarity 869 only a subset of the possible species and will contain identical species lists only a portion of the time. If we increase grain size, however, species lists will become more complete until, at a grain size that is equivalent to the size of the study areas, all possible species will be contained and the observed similarity will equal the real similarity. If two regions have a similarity greater than zero, small samples, by chance, can have either higher or lower similarity than the real similarity. The distribution of sample scores will vary with the distribution of individuals among species, the spatial pattern of species and individuals, the sampling design, and that component of the species-area relationship that we term the species-grain relationship (Palmer & White, 1994). In general, we expect that the variation in observed similarities among samples will increase as grain size decreases, assuming that the real similarity is a positive value. This effect of change in support or grain size on sample statistics has been well discussed in the geographical literature as the modifiable areal unit problem (Openshaw, 1983). It has generally been found that variation increases as support or grain size decreases. Although distance decay proposes that similarity decreases with extent, natural periodicities in the environment may cause a lack of distance decay at some scales of observation (Fig. 1). Geostatisticians have treated similar patterns of variance as the problem of non-stationarity (Webster & Oliver, 1990). For example, if extent is only large enough to include a mountain crest and its adjacent valley, distance decay will likely occur as a result of the strong environmental gradient that is present. However, if extent is increased to include the next mountain crest, similarity of community composition will likely increase in relation to the first mountain summit. If extent is increased further still to include a large number of mountain crests and Figure 1 The influence of sample grain and extent on the valleys, distance decay will likely re-emerge from either climatic observation of distance decay. A, A hypothetical landscape with gradients or the presence of dispersal barriers between places spatial periodicities in similarity controlled by strong environmental with similar environments. gradients (elevation contours); B, the distance decay of similarity as The detection of distance decay requires that the variation detected by samples of differing grain and extent. due to small grain size or support be less than the variation due to spatial extent. Increasing the grain size or support of a the spatial template) and two characteristics of the organisms sample will decrease variation, and may allow detection of under study (niche-breadth and dispersal ability). The effects weak distance decay trends. Alternatively, increasing extent of each of these factors are discussed below. while holding sample grain or support constant may allow identification of distance decay trends because the variation Environmental distance due to distance will be larger than the noise present from small For a fixed spatial distance, the steeper the rate of environmental grain size or support. The strongest distance decay relationships change, the more rapid the turnover in species composition should be observed when the variation due to sampling error and the higher the expected rate of distance decay. Ecologists is minimized (with large sample grains or support) and the often have assumed that composition is solely a function of variation due to distance decay is maximized (with large sample environment (e.g. MacArthur, 1972; Tilman, 1988). For the extents). rate of distance decay to be solely controlled by environmental distance, all species must have access to all appropriate habitats and time must be sufficient to allow competitive interactionsA conceptual model for variation in distance decay to sort species as a function of environment. In essence, thisrates summarizes Beijerinck’s Law: everything is everywhere but the The rate of distance decay will vary with the strength of environment selects (Saur, 1988). Niche characteristics and environmental gradients and will differ among taxa and competition determine composition. between landscapes of different spatial configurations and histories. More formally, we suggest that the rate of decay in The spatial template Characteristics of the spatial template (habitat size, isolation,biological similarity will be correlated with two characteristics of the environment (environmental distance and the nature of and the nature of the matrix) will effect the rate of distance © Blackwell Science Ltd 199926867–878 152 870 J. C. Nekola and P. S. White decay, all else being equal. The more isolated the habitat, the The second data set consisted of regional vascular plant lists for the boreal forest and Appalachian montane spruce-fir. Whileless efficient dispersal will be. Some species will not locate all available habitats (Huffaker, 1958; Shmida & Ellner, 1984), the boreal forest exists in a relatively continuous band from Alaska to Newfoundland, montane spruce-fir is restricted toleading to lower levels of species overlap among sites and higher rates of distance decay. Factors which would effect the highest peaks, with the areas dominated by spruce and fir separated from one another as a series of island-like clustersdistance decay rates are: the unfavourability of the matrix (which would affect rates of movement), the favourability of (Cogbill & White, 1991). Species lists for six boreal and six Appalachian spruce-fir forest regions were compiled from thethe habitat patches (which would affect size, reproduction, and persistence of the populations), and the presence of corridors literature and corrected to a common nomenclatural base (Crandall, 1958; Curtis, 1959; Grigal, 1968; Lakela, 1965; LaRoi,or stepping stone connections. In a sense, these measure the resistant to movement of the landscape. The percolation 1967; Maycock, 1961; Maycock & Curtis, 1960; McIntosh & Hurley, 1964; Nicholson, Holway & Scott, 1969; Ohmann &coefficient is an example of a measure of the affect of spatial characteristics on movement. This coefficient represents the Ream, 1971; Oosting & Billings, 1951; Ramseur, 1960; Siccama, 1974; White, 1982; White & Miller, 1988). Similarity waspermeability of a given landscape to unidirectional organism movement (Gardner et al., 1989). calculated between all pairwise comparisons of these regions within both groups. This data set has a larger support (each list is a regional flora) than the one from LaRoi (1967).Niche breadth and overlap The wider the average niche breadth and overlap for a group We addressed several questions using these data sets: Does community similarity decrease with distance and, if so, whatof organisms, the lower the rate of distance decay which will occur over a fixed amount of environmental distance. Realized statistical model best describes this relationship? Do isolated habitats (Appalachian mountain summits and slopes) have aniche breadth and overlap may decrease over successional time as species sort along gradients (Werner & Platt, 1976; higher rate of distance decay than more contiguous ones (boreal forest)? Do data sets with larger grain size or support (theChristensen & Peet, 1984). regional floristic lists), because they have more complete species lists, have higher overall similarities and lower distance decayDispersal ability Highly vagile species will be less affected by barriers and will rates than data sets with smaller grain size or support (boreal forest plot data)? Do wind-dispersed species (bryophytes andbe able to locate most or all of the appropriate habitats in a landscape (Bush & Whittaker, 1991; Dzwonko & Loster, 1992) wind-dispersed vascular plants) have higher rates of distance decay than less vagile species? Do taller and dominant speciesand such dispersal may compensate for low persistence rates (e.g. as suggested by the source-sink population model and the (which may have broader environmental tolerances) have lower distance decay rates than smaller and understory species?rescue effect, Brown & Kodric-Brown, 1977). More vagile taxa should therefore show lower rates of distance decay over a We addressed the questions about dispersal, dominance, and frequency by using data from a single set of plots, thus holdinggiven landscape. Species with poorer vagilities will be unable to cross barriers, leading to their absence from potential sites spatial configuration and environmental distance constant for all groups compared. The distance decay rates for bryophytes(Weddell, 1991), contributing to higher distance decay rates. and vascular plants were compared with data from both white spruce and black spruce plots. Comparisons among vascular DISTANCE DECAY IN SPRUCE-FIR FORESTS plant classes were made with data from white spruce plots only. The following paragraphs describe the questions addressed withThe boreal and montane spruce-fir forests of eastern and northern North America have similar structures over large white spruce plots in more detail. areas (Cogbill & White, 1991). Two data sets were constructed from published studies of these forests. In the first, the vascular plants and bryophytes of the North American boreal forests The contribution of species in different frequency classes to distance decay ratewere analysed from data in LaRoi (1967) and LaRoi & Stringer (1976). These papers present species lists for nine hectare plots Species were divided into three classes: species found on one to three plots (139 species), four to fourteen plots (eighty-four(300×300 m) distributed from Newfoundland to Alaska in two community types: upland white spruce forests (thirty-four plots species), and fifteen to thirty-four plots (twenty-nine species). We also summed the last two classes to produce a fourth class,containing 252 vascular plant species and 118 bryophyte species) and lowland black spruce forests (twenty-six plots containing those species found on four to thirty-four plots (113 species). Common species are shared by many plots and therefore195 vascular plant species and seventy-nine bryophyte species). Similarity and distances were calculated for all pairwise decrease the overall rate of distance decay (in the extreme, for species found in all plots, intersample similarity is always 1.0combinations of plots within these two habitat types (a total of 561 values for white spruce stands and 351 for black spruce and is independent of distance). Species found in only one plot would lower the similarity between any two plots but wouldstands). These data sets were chosen because they allow us to examine distance decay for relatively homogeneous forests not contribute to a spatial trend in similarity at this scale (similarity based on these species would be zero at all distances).(each of the habitats is dominated by the same tree species throughout) in relatively unbroken topography across a large We therefore hypothesized that species with intermediate frequency would show the steepest rate of distance decay.geographic extent. © Blackwell Science Ltd 199926867–878 153 Distance decay of similarity 871 Variation in distance decay rate among growth forms were assembled (we used non-zero values because we were interested in the slope of the distance decay relationship overCasual observation suggested that trees in the boreal forest have wider geographic distributions than plants with other those distances that have a measured similarity—presumably similarity falls to zero at some distance and then isgrowth forms. This suggests wider environmental tolerances and a lower distance decay rate. We therefore investigated independent of distance thereafter); (2) after testing the difference in mean similarity for the two data subsets,whether there were systematic differences in distance decay rates for plants with different growth forms. Species were similarity values within each subset were rescaled to a common mean (we sought to test the difference in slopesdivided into five classes: trees (nineteen species), tall shrubs (twenty-two species), low shrubs (thirty-five species), subshrubs of the distance decay relationships independent of differences in the means); (3) for each pair of sites, the two similarity(sixteen species), and herbs (160 species). We also compared the bryophyte data to the data from these vascular plant growth values (one for each of the two data subsets to be compared) were randomly reassigned to the two data subsets; (4) afterform classes. this randomization had been carried out for all pairs of sites, linear regression was used to determine the slope ofVariation in distance decay rate among dispersal classes the distance decay relationship for each of the randomizedMore vagile species should have lower distance decay rates subsets; (5) step four was repeated to producethan less vagile species. Species were divided into three major 10,000 values for the slope of the distance decay relationshipand five total dispersal classes: wind dispersed species (a total for randomized subsets; (6) finally, the absolute differenceof sixty-three species, comprised of two subcategories, twentybetween the slopes in each of the 10,000 randomizationseight spore/microseed species and thirty-five plumose seed was compared to the difference between the observed slopesspecies), berry-fruited/nut-bearing species (sixty-four species), based on the original data (and thus the true assignmentsand species with other dispersal types (125 species). We also to data subset categories) and we calculated how many timescompared the bryophyte data to the data from these vascular (out of 10,000 trials) randomization produced a differenceplant dispersal classes. in slope between the subsets that was at least as great as the observed difference in slope based on the original data. Similarity index If correct assignment to data categories is significant (if the We used Jaccard’s Index (Mueller-Dombois & Ellenberg, 1974) categories are indeed different in slope), then the observed due to its simplicity, its widespread use, its accessibility through difference should exceed the randomized differences a high common statistical packages, and its use of presence/absence percentage of the time. We used this percentage as the P data. We chose a presence-absence metric because it represents value for our tests. a more conservative measure of community similarity than ones based on species abundance data which are more sensitive to disturbance and local environmental differences. RESULTS Data analysis Distance decay in spruce-fir forests For the largest empirical data sets (white spruce and black spruce stands of the boreal forest), similarity and distance were In all data sets except for rare species in white spruce stands, similarity decreased significantly with distance (Figs 2–6, Tablecalculated and then log-transformed. Linear regression was used to calculate distance decay rates or all four combinations 1). The most linear relationship and the most heteroscedastic regression residuals were found in models that used the logof these transformed and untransformed variables. Inspection of the plots and regression residuals were used to determine transformation of similarity and untransformed distance. A semi-log model was thus used in all of the regression modelswhich formulation most closely approximated a linear relationship over most intersample distances. presented here. This implies an exponential rate of distance decay. For the boreal forest plot data, the rate of similarityBecause the values in the matrices (561 for white spruce stands, 351 for black spruce stands) are not independent, decay ranged from −0.17 (black spruce bryophytes; units are in ln (similarity) per 1000 km of distance) to −0.55 (whiterandomization tests were used to establish the significance of regression models and to make comparisons of distance decay spruce vascular plants of intermediate frequency) and r2 values ranged from 0.24 (black spruce bryophytes) to 0.71 (commonrates among various subsets of the data. Mantel tests with 10,000 replications (Jackson & Somer, 1989) were used to vascular plants on white spruce plots). For the regional floras, the rate of similarity decay ranged −0.25 (boreal forest, r2 =estimate significance of each model (Manly, 1991). The comparison of distance decay rates among various 0.54) and −0.67 (Appalachian montane forest, r2 =0.80). All of these relationships are fit well by a straight line in log-linearsubsets of the data was done with a randomization test developed by Blake Weiler of the Ecology Curriculum at the space (Figs 2–6), with no obvious sill (i.e. no distance at which the data imply a change of slope for the regression). Our closestUniversity of North Carolina at Chapel Hill. Our objective was to compare the slopes of the distance decay relationship observations are not close enough for a detailed analysis of the intercept value for the regressions, which is analogous tobetween two data subjects (e.g. trees v. herbs) across all plots. The method followed this outline: (1) all non-zero the nugget effect in semi-variograms, although we make a few comments on that parameter below.similarities based on the two data subsets to be compared © Blackwell Science Ltd 199926867–878 154 872 J. C. Nekola and P. S. White Figure 2 The log of similarity plotted against distance for all pairwise comparisons of plots sampled across the boreal forest based on data in LaRoi (1967) and LaRoi & Stringer (1976). A, Vascular plants in thrityfour white spruce stands; B, bryophytes in thirty-four white spruce stands; C, vascular plants in twenty-six black spruce stands; D, bryophytes in twenty-six black spruce stands. Figure 3 The log of similarity plotted against distance for species frequency classes (white spruce stands only; data from LaRoi, 1967). A, Vascular plants found in four or more plots; B, vascular plants found in fifteen to thirty-four plots; C, vascular plants found in four to fourteen plots; D, vascular plants found in three or fewer plots. The effect of frequency, growth form, dispersal type, distance decay rates for regional floras (Fig. 6) with those for plot data (Fig. 2). Both include data taken over the geographicand habitat isolation on distance decay rates extent of boreal forest, but the floristic lists include rare species that are at their range limits at each location, whereas the plotsSpecies frequency When rare species (those occurring in three or less of the 34 are less likely to sample species that are very rare in the landscape as a whole. The fuller regional species lists (andwhite spruce plots) were eliminated from the data set, the rate of distance decay was unaffected (Fig. 3A; Table 1), but larger support) of the floristic data set lowers the observed rate of similarity loss per unit distance. In a sense, the larger supportsimilarities were generally higher, the intercept was closer to unity (complete similarity at a distance of 0), and the r2 value reduces the observed rarity of species. Thus, the rate of distance decay for boreal forest plot data was 1.5 times higher thanwas higher. Thus, rare species lower overall similarity but do not affect the rate of distance decay. As might be expected that for boreal forest floristic lists across the same geographical extent (Table 1).from this analysis, the rare species themselves showed no decay of similarity as a function of distance (Fig. 3B). The species occurring in four or more plots were divided Growth form In the white spruce plots, herbs had 1.7–2.1 times higherinto two groups: common species (fifteen or more plots) and intermediate species (four to fourteen plots). The common distance decay rate than woody plant classes (P < 0.005–0.0001; Tables 1 and 2, Fig. 4). The higher distance decay rate for thespecies had a higher intercept value than intermediate species and a rate of distance decay (−0.27) that was only one half smallest vascular plants was reversed for the contrast between vascular plants and bryophytes: herbs had 1.6 times the ratesas great as the rate for intermediate species (−0.55) (Fig. 3C–D; Table 1). for bryophytes in the same plots (Fig. 2, Table 1). Interestingly, bryophytes had similar rates of distance decay to tall woodyThe influence of rare species can also be seen by comparing © Blackwell Science Ltd 199926867–878 155 Distance decay of similarity 873 Figure 4 The log of similarity plotted against distance for growth form classes (white spruce stands only; data from LaRoi, 1967). A, Trees; B, tall shrubs; C, low shrubs; D, subshrubs; E, herbs. Figure 5 The log of similarity plotted against distance for dispersal classes (white spruce stands only; data from LaRoi, 1967). A, Wind-dispersed species; B, species with plumose seeds; C, species with spores or microscopic seeds; D, species with nuts or fleshy berries; E, species with other types of seeds. plants on the same plots; they also had similar rates to wind- Dispersal type For vascular plants on white spruce plots, berry-fruited anddispersed vascular plants (see below). Vascular plants as a group had 1.5 (white spruce plots) and 1.9 (black spruce plots) nut-bearing species had nearly twice the rate of distance decay shown by spore/microseed and plumose seed species (−0.43 v.times higher distance decay rates than bryophytes of the same plots (comparisons significant at the P < 0.001 level; Table 1). −0.23 and −0.25, P < 0.001; Tables 1 and 2, Fig. 5B, C, and © Blackwell Science Ltd 199926867–878 156 874 J. C. Nekola and P. S. White Figure 6 The log of similarity plotted against distance for all pairwise comparisons of vascular plant floras from six regions across the northern boreal forest [data modified from Curtis (1959), Grigal (1968), Lakela (1965), LaRoi (1967), Maycock (1961), Maycock & Curtis (1960) and Ohmann & Ream (1971)] and six areas along the crest of the Appalachian mountains [data modified from Crandall (1958), LaRoi (1967), McIntosh & Hurley (1964), Nicholson, Holway & Scott (1969), Oosting & Billings (1951), Ramseur (1960), White (1982)]. A, Boreal spruce-fir forests; B, Appalachian spruce-fir forests. Table 1 Regression statistics for the decay of similarity with distance in spruce-fir data sets. The sample size (n) is the number of similarity values and distances in each matrix. The number of plots or areas are as follows: white spruce plots, 34; black spruce plots, 26; boreal spruce-fir areas, 6; Appalachian montane spruce-fir areas, 6. For the three data subsets of the white spruce plots, the sample size is sometimes lower than 561 because of the removal of plot pairs with a similarity of zero. Intercepts are in units of ln(similarity). Slope values are in units of ln(similarity) per 1000 km distance. Data set Species n Intercept Slope r2 P White spruce vascular plants 252 561 −0.75 −0.37 0.67 0.0001 White spruce bryophytes 118 561 −0.87 −0.25 0.53 0.0001 Black spruce vascular plants 195 351 −0.88 −0.33 0.57 0.0001 Black spruce bryophytes 70 351 −0.98 −0.17 0.24 0.0001 Data subsets (white spruce plots only) Frequency class: 4–34 plots 113 561 −0.57 −0.37 0.71 0.0001 15–34 plots 29 561 −0.24 −0.27 0.59 0.0001 4–14 plots 84 467 −1.04 −0.55 0.66 0.0001 1–3 plots 139 63 −2.40 −0.006 0.00 n.s. Growth form: Trees 19 555 −0.51 −0.19 0.30 0.0001 Tall shrubs 22 271 −1.21 −0.20 0.17 0.0001 Low shrubs 35 472 −1.09 −0.21 0.20 0.0001 Subshrubs 16 528 −0.66 −0.28 0.45 0.0001 Herbs 160 554 −0.84 −0.40 0.52 0.0001 Dispersal class: Wind 63 485 −1.14 −0.29 0.38 0.0001 Plumose 35 329 −1.09 −0.25 0.31 0.0001 Spore/microscopic 28 408 −1.02 −0.23 0.28 0.0001 Fleshy fruits, nuts 64 532 −0.62 −0.43 0.62 0.0001 Other 125 561 −0.84 −0.27 0.48 0.0001 Boreal forest regions 250 15 −0.76 −0.25 0.54 0.0039 Appalachian spruce-fir 226 15 −0.18 −0.67 0.80 0.0018 © Blackwell Science Ltd 199926867–878 157 Distance decay of similarity 875 Table 2 Significance levels for differences in slopes for the growth form and dispersal classes. A. Growth form Tall shrubs Low shrubs Subshrubs Herbs Trees 0.05 0.05 0.0001 0.0001 Tall shrubs (0.18) 0.01 0.005 Low shrubs 0.005 0.0001 Subshrubs 0.0001 B. Dispersal class Plumose Spore/micro. Fleshy Other Wind (0.0649) 0.0001 0.0001 0.004 Plumose (0.645) 0.0001 (0.79) Spore/microseeded species 0.0001 (0.306) Fleshy fruited, nut-bearing 0.0001 species E). Berry-fruited and nut-bearing species had the highest r2 change in similarity per unit distance. This is attractive because it suggests that the processes underlying the distance decayvalues among the dispersal types (Table 1). These data suggest that wind-dispersed species are both more widely dispersed relationship vary only with environmental distance, spatial configuration, and organism attributes but not with distanceand more variable in their occurrences. Bryophytes had similar distance decay rates to spore/microseeded and plumose seeded independent of these factors. Hubbell (1995) has produced a simulation model of species colonization and extinction thatvascular plants on the same plots (Table 1). results in just such a decrease in similarity with distance. Habitat isolation Appalachian forest communities had a rate of similarity decay Variation in distance decay rates 2.7 times that observed for the continuous northern boreal forest (Fig. 6, Table 1). The rate of distance decay varied among species groups in expected ways: on the same plots, spore/microseeded and plumose species had lower rates of distance decay than larger DISCUSSION fruited species. Trees, other woody plants, wind dispersed, and more frequent species raised overall similarity and lowered the Form of the distance decay model rate of distance decay, while herbaceous plants, heavy fruited, and less frequent species lowered the rate. Further, the moreIn the geographic literature, a distance effect on the interaction between human populations has been modelled with a power fragmented montane forests dominated by the same genera of trees as the boreal forest had much higher rates of distanceor gravity function (Bennett & Gade, 1979): decay than the more contiguous boreal forest itself. We will I=A∗d−c consider alternative explanations for these patterns. While it can be suggested that due to their small propagulewhere I is interaction, A is a constant, d is distance, and c is the friction coefficient (correlated with technologies for size bryophytes should be more vagile and possess wider ranges along geographic gradients than vascular plant species (Saur,communication and barriers to movement over a given distance). This model implies an ever decreasing effect of a 1988), other explanations are reasonable. The smaller size of bryophytes compared to vascular plants may allow persistenceunit distance as distance becomes larger and a linear relationship only under double log transformation. in microsites not available to vascular plants, thus increasing geographic range. Bryophytes may have broader physiologicalThe finding that a semi-log model showed the most linear relationships suggests similarity could in fact be related to tolerances and higher persistence rates because they can cycle rapidly between active and dormant states.distance through an exponential model: In light of these possibilities, it is interesting to note that S=S0e−cd herbs, which are the smallest vascular plants (thus sharing, with bryophytes, environments that have less fluctuating airwhere S is similarity at distance d; S0 is initial similarity or similarity at distance 0; c is a constant for the rate of distance temperatures), showed the highest rates of similarity loss with distance of all groups. By contrast, wind-dispersed vasculardecay; and d is distance. Exponential decay models have previously been used in biogeography and ecology to model plants had very similar rates of distance decay to bryophytes on these plots regardless of plant stature. Tall vascular plantsplant dispersal (Okubo & Levin, 1989) and to describe the relationship between spatial distance and floristic similarity in had as low or lower distance decay rates than bryophytes. In terms of the comparison of wind and non-wind dispersedthe Galapagos (Preston, 1962). The exponential model, in contrast to the power model, implies a constant proportional species, dispersal type could be correlated with other traits that © Blackwell Science Ltd 199926867–878 158 876 J. C. Nekola and P. S. White influence the rate of geographic turnover. For example, wind- Regardless of whether we can evaluate causes, the decay of similarity with distance is a relatively simple descriptor ofdispersed species could have wider niche breadth (as has been argued for early successional species generally), thus showing variation in species composition and has consequences for how we conserve biological diversity. Whether because ofa lower rate of distance decay. On the other hand, winddispersed species often have small seeds and may therefore environmental distance decay or spatial and historical factors, the distance decay paradigm suggests a natural world that is,have a lower probability of establishment from a given dispersal event, particularly in undisturbed forest studies here (i.e. the at some scales, everywhere unique. Distance decay rates describe the rate of turnover that underlies this uniqueness.probability of establishment in late successional stands could be correlated with seed size). It is unclear how this would Simberloff & Abele (1970) argued that the theory of island biogeography was silent on preserve design; that is, this theoryinfluence distance decay; a simple expectation is that it would tend to make similarity more variable at any distance. could not be used to argue whether one large area would have more species than several smaller areas that summed to theThe higher rate of distance decay in the Appalachian montane spruce-fir forests could be due to less efficient dispersal among same total area. They suggested that the critical factor was the degree of overlap (i.e. the similarity) among the smaller areas.fragmented habitats, but it could also be the result of history (the southern and northern Appalachian spruce-fir forest may Shafer (1990) showed that whether one area or several small areas of the same total area had more species or not washave had access to different refugial populations after climatic warming started 15,000 years BP) and/or steeper rate of dependent on two factors: the slope of the log(species)-log(area) relationship and the similarity among the smaller areas. Weenvironmental change with distance. In terms of environmental distance, montane spruce-fir forests have very similar growing add here that if the smaller areas (which are by definition non-contiguous and thus spatially dispersed compared to oneseason temperatures across the wide range of latitudes covered in the data set (a mean July temperature of 17°C is correlated contiguous area) are more different from one another as a function of the distance between them (see also Palmer &with their lower elevational boundaries), but are exposed to increasingly low winter temperatures northward (Cogbill & White, 1994), then the decay of similarity with distance is a central explanation of why some investigators (e.g. StohlgrenWhite, 1991). Sorting out the relative importance of environment and & Quinn, 1991) find that smaller dispersed conservation areas collectively have more species than large conservation areas.niche-based explanations from spatial configuration and dispersal-based ones is a difficult challenge. Explanations of In essence, conservation strategy requires both grain size for population persistence (Shoenwald-Cox, 1983; Newmark, 1987)species composition based on environmental factors and competition are old ones in ecology. However, explanations and spatial extent (for capturing areas dissimilar to one another). The SLOSS (single large or several small) debate canbased on history and dispersal have also been proposed from the beginning. Gleason (1923) suggested that in the Midwest never be resolved because size (grain) and subdivision (extent) maximize different aspects of how biological diversity ispassive dispersal of propagules was the major factor explaining community composition. Palmgren (1926) found that many distributed. The distance decay of similarity, in this sense, provides a metric and a conceptual base for such efforts asplant species of the A˚ land Archipelago, Finland, were absent from seemingly suitable sites, and suggested that dispersal GAP analysis that seek to identify habitats and species left out of existing conservation areas (Davis et al., 1990).limitation was responsible for this pattern. When species do not disperse to all appropriate sites, composition will be constrained by the spatial configuration of sites and dispersal abilities of species. At one extreme (if all REFERENCES species had the same dispersal abilities), community Allen, T.F.H. & Hoekstra, T.W. (1991) Role of heterogeneity in scalingcomposition would be a stochastic subset of the species which of ecological systems under analysis. Ecological heterogeneity (ed.can occupy that habitat. Shmida & Ellner (1984) and Shmida by J. Kolasa and S.T.A. Pickett), pp. 47–68. Ecological Studies 86. & Wilson (1985) have shown how such conditions may lead Springer-Verlag, New York. to the coexistence of co-competitive species in a landscape, Bennett, D.G. & Gade, O. (1979) Geographic perspectives in migration which are allowed to co-exist only because populations are research. University of North Carolina at Chapel Hill Department spatially separated from each other at any given moment. At of Geography Studies in Geography, Number 12. the other extreme (if species have a graded series of dispersal Briggs, J.C. (1991) Historical biogeography: the pedagogical problem. J. Biogeogr. 18, 3–6.abilities), community composition at each location would be Brown, J. & Kodric-Brown, A. (1977) Turnover rates in insulara nested subset (varying with habitat isolation) of the full biogeography: effect of immigration on extinction. Ecology, 58,species list. Hierarchical nesting of species lists has been much 445–449.discussed in the context of habitat area and area-sensitive Burrough, P.A. (1986) Principles of geographical information systems species (Patterson, 1987); similarly, there may be isolationfor land resources assessment. Oxford University Press, New York. sensitive species. Presences explained not by local environment Bush, M.B. & Whittaker, R.J. (1991) Cracked: colonization patterns but by continued dispersal has been termed mass effect (Shmida and hierarchies. J. Biogeogr. 18, 341–356. & Wilson, 1985); here we add that absences can be the result, Cain, S.A. (1944) Foundations of plant geography. Hafner Publishing not of local environment, but of dispersal limitations and of Co., New York. stochastic patterns of original colonizing populations (e.g. the Christensen, N.L. & Peet, R.K. (1984) Secondary forest succession on the North Carolina Piedmont. Forest succession: concepts andimprint of initial refuge competition and migration routes). © Blackwell Science Ltd 199926867–878 159 Distance decay of similarity 877 application (ed. by D.C. West, H.H. Shugart and D.B. Botkin), biogeography. Analytical biogeography (ed. by A.A. Meyers and P.S. Giller), pp. 3–12. Chapman and Hall, New York.pp. 230–245. Springer-Verlag, New York. Cody, M.L. (1985) Towards a theory of continental species diversities. Mueller-Dombois, D. & Ellenberg, H. (1974) Aims and methods of vegetation ecology. John Wiley and Sons, New York.Ecology and evolution of communities (ed. by M.L. Cody and J.M. Diamond), pp. 214–257. Harvard University Press, Cambridge, Mass. Newmark, W.D. (1987) A land-bridge perspective on mammalian extinctions in western North American parks. Nature, 325, 430–432.Cogbill, C.V. & White, P.S. (1991) The latitude-elevation relationship for Appalachian spruce-fir. Vegetatio, 94, 153–175. Nicholson, S., Holway, J.G. & Scott, J.T. (1969) A floristic comparison of undisturbed spruce-fir forests of the Adirondacks with four otherCrandall, D.L. (1958) Ground vegetation patterns of the spruce-fir area of the Great Smoky Mountains National Park. Ecol. Monogr. 28, regions. Atm. Sci. Res. Center, State U. of N.Y. at Albany. Rept. 92. Odland, J. (1988) Spatial autocorrelation. SAGE Publications, Newbury337–360. Cressie, N.A. (1993) Statistics for spatial data. John Wiley & Sons, Park, California. Ohmann, L.F. & Ream, R.R. (1971) Wilderness ecology: virgin plantNew York. Curtis, J.T. (1959) The vegetation of Wisconsin. University of Wisconsin communities of the Boundary Waters Canoe Area. North Central For. Exp. Sta. Rept. NC-63. St. Paul, Minnesota.Press, Madison. Davis, F.W., Stoms, D.M., Estes, J.E., Scepan, J. & Scott, J.M. (1990) Okubo, A. & Levin, S.A. (1989) A theoretical framework for data analysis of wind dispersal of seeds and pollen. Ecology, 70, 329–338.An information systems approach to the preservation of biological diversity. Int. J. Geogr. Info. Syst. 4, 55–78. Olea, R.A. (1991) Geostatistical glossary and multilingual dictionary. Oxford University Press, Oxford, England.Dzwonko, Z. & Loster, S. (1992) Species richness and seed dispersal to secondary woods in southern Poland. J. Biogeogr. 19, 195–204. Oosting, H.J. & Billings, W.D. (1951) A comparison of virgin sprucefir in the northern and southern Appalachian systems. Ecology, 32,Faith, D.P., Minchin, P.R. & Belbin, L. (1987) Compositional dissimilarity as a robust measure of ecological distance. Vegetatio, 69, 84–103. Openshaw, S. (1983) The modifiable areal unit problem. Concepts and57–68. Gardner, R.H., O’Neill, ??, Turner, M.G. & Dale, V.H. (1989) Quan- Techniques in Modern Geography No. 38. Geobooks, Norwich, England.tifying scale-dependent affects of animal movement with simple percolation models. Landscape Ecol. 3, 217–227. Palmer, M.W. (1988) Fractal geometry: a tool for describing spatial pattern of plant communities. Vegetatio, 75, 91–102.Gleason, H.A. (1923) The vegetation history of the middle west. Ann. Assoc. Am. Geogr. 12, 39–85. Palmer, M.W. & White, P.S. (1994) Scale dependence and the speciesarea relationship. Am. Nat. 144, 717–740.Grigal, D.F. (1968) The relationship between plants communities and soils in northeastern Minnesota. PhD Dissertation, University of Palmgren, A. (1926) Chance as an element in plant geography. Proceedings of the International Congress of Plant Sciences, Ithaca, NewMinnesota, Minneapolis. Hanski, I. & Gilpin, M. (1991) Metapopulation dynamics: brief history York (ed. by B.M. Duggar), pp. 591–602. George Banta Publishing Company, Menasha, Wisconsin.and conceptual domain. Biol. J. Linn. Soc. 42, 3–16. Hubbell, S.P. (1995) Towards a theory of biodiversity and biogeography Patterson, B.D. (1987) The principle of nested subsets and its implications for biological conservation. Conserv. Biol. 1, 323–334.on continuous landscapes. Preparing for global change: a midwestern perspective (ed. by G.R. Carmichael, G.E. Folk and J.L. Schnoor), Preston, F.W. (1962) The canonical distribution of commonness and rarity: parts I and II. Ecology, 43, 185–215; 410–432.pp. 173–201. Academic Publishing bv, Amsterdam. Huffaker, C.B. (1958) Experimental studies on predation: dispersion Ramseur, G.S. (1960) The vascular flora of the high mountain communities of the southern Appalachians. J. Elisha Mitchell Sci. Soc.factors and predator-prey oscillations. Hilgardia, 27, 343–383. Jackson, D.A. & Somer, K.M. (1989) Are probability estimates from 76, 82–112. Ripley, B.D. (1988) Statistical inference for spatial processes. Cambridgethe permutation model of Mantel’s test stable? Can. J. Zool. 67, 766–769. University Press, Cambridge. Roughgarden, J., Gaines, S.D. & Pacala, S.W. (1987) Supply sideKadmon, R. & Pulliam, H.R. (1993) Island biogeography: effect of geographical isolation on species composition. Ecology, 74, 977–981. ecology: the role of physical transport processes. Organization of communities: past and present (ed. by J.H.R. Gee and P.S. Giller),Lakela, O. (1965) A flora of northeastern Minnesota. University of Minnesota Press, Minneapolis. pp. 491–518. Blackwell Scientific, Oxford. Saur, J.D. (1988) Plant migration. University of California Press,LaRoi, G.H. (1967) Ecological studies in the boreal spruce-fir forests of the North American taiga. I. Analysis of the vascular flora. Ecol. Berkeley. Schluter, D. & Ricklefs, R.E. (1993) Species diversity: an introductionMonogr. 37, 229–253. LaRoi, G.H. & Stringer, M.H. (1976) Ecological studies in the boreal to the problem. Species diversity in ecological communities (ed. by R.E. Ricklefs and D. Schluter), pp. 1–10. University of Chicago Press,spruce-fir forests of the North American taiga. II. Analysis of the bryophyte flora. Can. J. Bot. 54, 619–643. Chicago. Schoenwald-Cox, C. (1983) Guidelines to management: a beginningMacArthur, R.H. (1972) Geographical ecology. Harper and Row, New York. attempt. Genetics and conservation (ed. by C. Schoenwald-Cox, S.M. Chambers, B. MacBryde and W.L. Thomas), pp. 414–446.Manly, B.F.J. (1991) Randomization and Monte Carlo methods in biology. Chapman and Hall, New York. Benjamin/Cummings Publishing, New York. Shafer, C. (1990) Nature reserves: Island theory and conservationMcIntosh, R.P. & Hurley, R.T. (1964) The spruce-fir forests of the Catskill Mountains. Ecology, 45, 314–326. practice. Smithsonian Institute Press, Washington, DC. Shmida, A. & Ellner, S. (1984) Coexistence of plant species with similarMaycock, P.F. (1961) Botanical studies on Mont St. Hilaire, Rouville County, Quebec. Can. J. Bot. 39, 1293–1325. niches. Vegetatio, 58, 29–55. Shmida, A. & Wilson, M.V. (1985) Biological determinant of speciesMaycock, P.F. & Curtis, J.T. (1960) The phytosociology of boreal conifer-hardwood forests of the Great Lakes region. Ecol. Monogr. diversity. J. Biogeogr. 12, 1–20. Siccama, T.G. (1974) Vegetation, soil, and climate on the Green30, 1–35. Meyers, A.A. & Giller, P.S. (1988) Process, pattern, and scale in Mountains of Vermont. Ecol. Monogr. 44, 325–349. © Blackwell Science Ltd 199926867–878 160 878 J. C. Nekola and P. S. White Simberlof, D.S. & Abele, L.G. (1976) Island biogeography and con- richness in the southern Appalachian high peaks. J. Ecol. 76, 192–199. Whittaker, R.H. (1975) Communities and ecosystems. MacMillan Pub-servation practice. Science, 191, 285–286. lishing, New York.Stohlgren, T.J. & Quinn, J.F. (1991) Evaluating the contribution of Willis, J.C. (1922) Age and area: a study in geographical distributionnorthern California National Park areas to regional biodiversity. and origin of species. Cambridge University Press, Cambridge.Proceedings of the Symposium on Biodiversity of Northwestern California, 226–232. Tilman, D. (1988) Plant strategies and the dynamics and structure of plant communities. Princeton University Press, Princeton. BIOSKETCHES Tobler, W.R. (1970) A computer movie simulating urban growth in the Detroit region. Econ. Geogr. 46, 234–240. Vuilleumier, F. (1970) Insular biogeography in continental regions. 1. Jeffrey C. Nekola received his Ph.D. in Ecology from the The northern Andes of South America. Am. Nat. 104, 378–388. University of North Carolina at Chapel Hill in 1994. His Webster, R. & Oliver, M.A. (1990) Statistical methods in soil and land previous work includes the documentation of the effect of resource survey. Oxford University Press, Oxford. colonization history on community composition and the Weddell, B.J. (1991) Distribution and movement of Columbian ground spatial distribution of rare vascular plant and butterfly squirrels (Spermophilus columbianus (Ord)): are habitat patches like species within and between isolated habitats. He is islands? J. Biogeogr. 18, 385–394. currently investigating species-rich land snail communities Weins, J.A. (1989) Spatial scaling in ecology. Funct. Ecol. 3, 385–397. at scales ranging from 1 to 10,000 m. Werner, P.A. & Platt, W.J. (1976) Ecological relationships of co- Peter S. White is Professor, Department of Biology, occurring Goldenrods (Solidago: Compositae). Am. Nat. 110, 959– University of North Carolina at Chapel Hill and Director 971. of the North Carolina Botanical Garden. His work ranges White, P.S. (1982) The flora of the Great Smoky Mountains National from studies of biogeography, rare species, exotic species Park: an annotated checklist of the vascular plants and a review of invasions, and species richness to vegetation dynamics and previous floristic work. USDI, National Park Service, Southeast disturbance. Regional Office, Res./Resource Manage. Rept. SER-55. White, P.S. & Miller, R.I. (1988) Topographic models of vascular plant © Blackwell Science Ltd 199926867–878 161 Section V: Ecological Theory and Modeling Without the context provided by theory, ecological research may become bogged down in trivial detail. However, pure theory may easily become untethered from reality unless it is constantly informed by real-world observations. For this reason, I have chosen actively pursue theoretical issues while also maintaining an active research program spanning multiple frames of reference across multiple taxa groups. There are a number of biodiversity patterns which appear to be universal across multiple scientific disciplines. For instance, rarity-enriched Species Abundance Distributions, power-law Species Area/Time Relationships, and non-linear Distance Decay patterns appear ubiquitous across a wide array of complex systems, including not only ecology but also meteorology, geoscience, materials science, economics, sociology, and even the arts. Determining the common mechanisms underlying the distribution of trees in tropical forests, the longevity of drink tumblers in restaurants or of song performance frequencies in Cowboy Junkies setlists will help ecologists and others understand what patterns are caused by statistical mechanics rather than disciplinary mechanisms such as birth, death, immigration, or competition for limiting resources. All disciplines must begin this search for such fundamental mechanisms which are both more universal and less explicitly mechanistic than those typically sought. In this way we will be able to define the rigid rules for life’s sonnet, within which there is an infinity of detail. A particular foci for my theoretical interests is consideration of the role played by spatial or temporal distance. In particular, I am interested in applying Tobler’s First Law of Geography: “Everything is related to everything else, but near things are more related to each other” to biogeographic analysis and modeling. For instance, this premise can be used, via gravity models, to predict the spread of organisms. Such models offer an important counterpoint to diffusion models as they assume landscape heterogeneity and do not require organisms to migrate via a moving wavefront. Gravity models provide a much more accurate portrayal of Zebra Mussel movement through small lakes of the upper Midwest USA than are possible via diffusion models. My latest contribution in this general topic area derives this relationship from individual species occurrences and then illustrates that the functional form is related to sample scale, with powerlaw decay dominating within and exponential-decay dominating between communities. It also documents why distance decay is a universal expectation across a wide range of physical, social, artistic and biological systems. There are perhaps a dozen or more other diversity metrics that likewise appear to be universal statistical expectations, with rarity-enriched species abundance distributions and power-law species area/ time relationships serving as classic examples. As a result, the mere existence of these patterns likely say little about specific mechanism (e.g. neutral 162 vs. deterministic community assembly) other than illustrating that the system consists of many quasi-independent parts. Thus, before we attempt to discern underlying mechanism we must: A. determine which patterns are simply cross-disciplinary expectations derived from statistical mechanics and thus say little per se about explicit ecological mechanism; B. determine what parameters of these relationships, if any (e.g. functional form or rate of change), are most likely to reflect differences in underlying ecological mechanism; C. investigate variation in these parameters to determine their utility in differentiating competing ecological models. Representative Publications [number of citations as of October 27, 2017] Nekola, J.C. & B. McGill. 2014. Scale dependency in the functional form of the distance decay relationship. Ecography. 37:309-320. [18] McGill, B. & J.C. Nekola. 2010. Mechanisms in macroecology: AWOL or purloined letter? Towards a pragmatic view of mechanism. Oikos. 119:591-603. [57] Nekola, J.C., A.L. Šizling, A.G. Boyer & D. Storch. 2008. Artifactions in the logtransformation of species abundance distributions. Folia Geobotanica. 43:259-468. [21] Moses, M.E., C. Hou, W.H. Woodruff, G.B. West, J.C. Nekola, W. Zuo & J.H. Brown. 2008. Revisiting a model of ontogenetic growth: estimating model parameters from theory and data. The American Naturalist. 171:632-645. [98] Nekola, J.C. & J.H. Brown. 2007. The wealth of species: ecological communities, complex systems, and the legacy of Frank Preston. Ecology Letters. 10:188-196. [101] Bossenbroek, J.M., C.E. Kraft & J.C. Nekola. 2001. Prediction of long-distance dispersal using gravity models: Zebra Mussel invasion of inland lakes. Ecological Applications. 11:1178-1788. [226] Examples of linear, exponential, and power-law distance decay showing the best-fit regression to the mean and the upper threshold (95% percentile). 163 309 Scale dependency in the functional form of the distance decay relationship Jeffrey C. Nekola and Brian J. McGill J. C. Nekola (jnekola@unm.edu), Dept of Biology, Univ. of New Mexico, Albuquerque, NM 87131, USA. – B. J. McGill, School of Biology and Ecology, Univ. of Maine, 5751 Murray Hall, Orono, ME 04469, USA. We examine a novel mathematical approach which posits that the decay of similarity in community composition with increasing distance (aka distance decay) can be modeled as the sum of individual species joint-probability vs distance relationships. Our model, supported by analyses of these curves from three datasets (North American breeding birds, North American taiga plants, and tropical forest trees), suggest that when sampling grain is large enough to avoid absences due to stochastic sampling effects, and/or sampling extent is large enough to generate species turnover through the deterministic crossing of environmental and/or geographical range limits, species joint-probability over increasing distance will generally exhibit exponential decay. However, at small scales where occurrence is driven more by stochastic sampling effects, species joint-probability curves exhibit a power-law decay form. Lacking a theoretical prediction of how individual species joint-probability relationships combine to generate community distance decay, we also performed a meta-analysis of 26 ecological and 4 human-system datasets, using non-linear regression to mean and quantile non-linear regression at tau  0.95 for linear, exponential, and power-law decay forms. These analyses demonstrate that the functional form of community distance decay – as shown by comparison of AIC ranks – is largely determined by observational scale, with power law decay prevailing within domains where the species pool remains constant, while exponential decay prevails at larger scales over which the species pool varies, paralleling the patterns predicted in our mathematical approach. The decay of compositional similarity with increasing inter-observation distance (aka distance decay) is a ubiquitous pattern found across a wide array of physical, biological, social and other complex systems (Tobler 1970, Nekola and Brown 2007). It has garnered considerable interest within the fields of community ecology (Condit et  al. 2002, Green et  al. 2004, Soininen et  al. 2007), macroecology (McKnight et  al. 2007, Blanchette et  al. 2008, Qian et  al. 2009) and conservation biology (Bell 2003, Steinitz et al. 2005), and has been suggested to shed important insights into underlying fundamental mechanisms (Nekola and White 1999, Hubbell 2001, Tuomisto et al. 2003, Gilbert and Lechowicz 2004). In spite of this interest, controversy exists regarding its expected functional form. While a number of empirical analyses have firmly established exponential decay (Nekola and White 1999, Qian et al. 2005, 2009, Jobe 2007), linear (Blanchette et  al. 2008, Perez-del-Olmo et  al. 2009) and power law (Harte et  al. 1999, Condit et  al. 2002, Green et  al. 2004) forms have also been documented. In spite of this, all theoretical derivations of the distance decay relationship (Harte and Kinzig 1997, Hubbell 2001, Chave and Leigh 2002, Houchmandzadeh 2008, Morlon et  al. 2008) generate power-law or power-law-like forms. Here we begin to develop analytical underpinnings for a theory of community distance decay based on jointprobabilities for individual species. We show that: 1) expectation of exponential or power-law forms for individual species joint-probability curves is a function of sampling scale, and 2) the community distance decay curve is a sum of single species joint-probability curves. Because a theoretical prediction for the shape of community distance decay curves from individual species joint-probability curves is not analytically tractable (except in a few specific cases), we consider this issue through empirical metaanalysis of 26 community datasets collected over both spatial and temporal extents, covering a large range of ecosystems (simulated data to abyssal plain to continental mountains, arctic tundra to tropics), taxa groups (including mammals, birds, lepidoptera, robber flies, land snails, vascular plants, bryophytes), sample grains (100 m2 – subcontinental and daily-yearly) and total extent (3–9000 km and 1–25 yr). Four human system examples are also considered (1 spatial, 3 temporal; Nekola and Brown 2007) to Ecography 37: 309–320, 2014 doi: 10.1111/j.1600-0587.2013.00407.x © 2014 The Authors. Ecography © 2014 Nordic Society Oikos Subject Editor: Jens-Christian Svenning. Accepted 11 October 2013 164 310 identify if general patterns in community distance decay functional forms exist. Individualistic model of distance decay Single species distance decay We begin with consideration with the decay of jointprobability (Palmer 2005) for a single species/agent/event with increasing inter-observation distance. Specifically, if comparing two plots L and M of the same size A at distance d apart (Fig. 1) for species i, then: pi(dLM)  Prob{species i present at M |      species i present at L} (1) or more generally: pi(d)  Prob{species present at any site distance     d from a site where species i is present} (2) While this definition has a clear relationship to the occupancy function used by numerous authors (Harte et al. 1999, He and Legendre 2002), in these previous formulations p has been a function of the plot area, A, rather than the distance, d, from reference plot L. Thus, our approach is the first to explicitly consider spatial structure. It is clear that under biologically realistic conditions where species have finite ranges, pi(d ) needs to fit the following criteria: 1) pi(0)  1; 2) pi()  0; 3) pi(dAB)   pi(dBA) (i.e. isotropy or direction does not matter); 4) pi(d1)   pi(d2) if d1  d2 (i.e. pi monotonically decreasing); 5) pi(d1  d2)  g(p(d1), p(d2)) for some functional form g if d1 and d2 are linear and contiguous as shown in Fig. 1 (i.e. decomposability with distance). Note that 1, 2 and 4 combined can be used to show that 0  pi(d )  1 if 0  d  ∞. There are an infinite number of functional forms that fit these criteria. Three that have been commonly used to analyze community distance decay (Nekola and White 1999, Soininen et  al. 2007) include the linear [pi(d )   1 2 cd], exponential [pi(d )  exp(2cd)], and power law [pi(d )  d2c]. Additionally, Harte (2011) suggests a logarithmic decay [pi(d )  1 2 c*log(d )] form. All contain a single parameter, c, which denotes decay rate. Note that the power law form violates criterion 1, the linear form violates criterion 2, while the logarithmic form violates both. As a result, these functional forms can in practice be used only over limited intervals: [0, dmax] where dmax  d at which pi(d )  0 for the linear, [∈, ∞] for the power law, and [∈, dmax] for the logarithmic. While Palmer (2005) calculated pi(d ) for tropical trees at the La Selva Biological Station in Costa Rica, he did not propose that it take on any particular mathematical form. However, his plot of pi(d) across all species and sites appears to demonstrate an exponential trace. What form should g in criterion 5 take? This function describes how pi should behave when a long distance d is decomposed into two smaller distances d1 and d2 (where d1  d2  d and lie along a straight line; Fig. 1). Clearly it should be commutative and associative. The most obvious, simple forms are g(x,y)  x 1 y or g(x,y)  x*y. However, the additive form violates criterion 4 [e.g. p(6)  p(4  2)  p(4)  1 p(2)  p(4) but p(6)  p(4)] and can easily produce pi   1. The multiplicative form fits well with our intuitive sense of how similarity decays [i.e. pi(d1  d2)   pi(d1)*pi(d2)], and is essentially the law of independent probabilistic events [P(A and B)  P(A)*P(B)]. Note that pi(d1  d2)  pi(d1)pi(d2) can only be satisified by pi(d )  exp(cd ) with criteria 1 and 2, requiring that c  0. A negative exponential form for the decay of joint occurrence probability with increasing distance thus conforms well to criteria 1–5. It is instructive to proceed with a more rigorous examination of the form of g. Considering the straight line connecting L to M to N in Fig. 1, via the law of total probability (Karlin and Taylor 1975): P(present at N|present at L)     P(present at N|present at M)     3 P(present at M|present at L)    1 P(present at N|NOT present at M)    3 P(NOT present at M| present at L) (3) or: pi(d)  pi(d2)pi(d1)  X(d2)(1 2 pi(d1)) (4) where: X(d) is the probability of a species being present at one point conditioned on it NOT being present at another point [note this is not simply 1 2 pi(d)]. More simply: pi(d)  pi(d1)pi(d2)  Y(d1,d2) (5) where Y(d1,d2)  Prob{present at N, absent at M|present at L} or essentially the probability of encountering an ‘occurrence hole’ (sensu Rapoport 1982, Hurlbert and White 2007) within the range of species i. Y  0 when species i is never absent at M if present at L and N. In such cases, we have the multiplicative form for g and pi(d)  exp(2cd). If Y ≠ 0, then g is more complex, and pi(d) is not a simple exponential. Community distance decay Assuming pi(d ) is exponential for one species, then what distance decay form should be expected across a community assemblage? As is common in theoretical ecology, we will work with the Sørenson similarity measure due to its tractable mathematics (Plotkin and Muller-Landau 2002, Morlon et  al. 2008). Let SLM be the number of species found in both plots L and M, while SL is the number of species found in plot L and similarly for SM. We assume approximately S  SL  SM. because systematic changes in richness will lead to bias in parameter estimation (Jobe dMNdLM L M N dLN Figure 1. Heuristic diagram demonstrating the relationship between sample plots and distances used in the theoretical model. 165 311 2007). Then Sørenson (dLM)  Sørenson (L,M)  SLM/average (SL,SM). If we sum the probability that a species is present in M over all species present in L, this gives: S rensonø ( ) ( )L,M p d S S i LMi SL   1 1 ∑ eexp( ) exp( )    c d E c di i i i S 1 ∑ (6) where Ei is the expectation (or averaging) operator taken with respect to i or over all species. We call this the community aggregated distance decay function. This makes the strong assumption that the joint probability decay is independent between all species. A sum of exponentials as found in Eq. 6 can take on a vast variety of shapes, dependent entirely on the distribution of ci values. If we treat ci as a random variable, C, from which the ci are drawn and using the tools for calculating the expected value of a function of a random variable (Lindgren 1976 p. 113) we have: S rø ( ) exp( ) exp( ) ( )d E c d xd f x dxi i   ∫ (7) Where f(x) is the pdf of the random variable describing the distribution of C. A uniform distribution for C across [0, c] yields 1 2 [exp(2c*d ))/(c*d)]. In the limiting case of c having a Dirac delta distribution (i.e. ci is constant with no variation), then Eq. (6) and (7) give community Sørenson similarity as an exponentially decaying function of distance. Assuming C is distributed according to an exponential distribution yields Sør(d )  c/(c  d ) while assuming C is distributed according to a power law yields Sør(d )  G(c)c/(bc*dc). However, as will be shown in the results, C is strongly right-truncated, making a Beta distribution the most likely. While this does not yield an analytically solvable integral, a partial understanding can realized by taking a Taylor-series expansion of Ei exp(2cid). Let C now be the mean value of ci (C  ci – ) and Δi  ci 2 C, with the expansion around C: being the occupancy rate for species i (i.e. the fraction of sites supporting species i) or equivalently the probability that species i is found at site L. Then we have: S rø ( ) exp( ) exp( )d S w c d E w c di i i i i i S      1 1 ∑ (9) We call this the weighted community aggregate distance decay function. This curve should closely approximate the traditional community-level decay of similarity curve which is generated by calculating pairwise similarity between samples (Nekola and White 1999). Model assumptions Two main assumptions underlie our approach. First, to produce an exponential decay of joint-probability over increasing distance for a single species we assumed that Y(d1,d2)  0 or equivalently that a species must occur at M if present at L and N. In effect Y informs about the probability that a sample plot falls inside an occurrence hole within a species range. The value of Y should show strong scale dependence: if sample grain is small with significant sampling error and extent does not exceed the environmental and/or geographic range limits of species within the pool, then it is likely that the main reason for compositional turnover will be sampling effects (Nekola and White 1999). In such instances Y ≠ 0, and because in such situations pi(d) is not exponential, this form will also not provide a good fit to the community data. The expected shape in such cases should initially show a more rapid decrease in joint-probability than the exponential expectation because of occurrence holes, but will then ultimately asymptote  0 because the probability of a given species within the pool is also  0. Although not a rigorous proof, at a minimum the above demonstrates that a more power-law like than exponential form would result from cases where Y ≠ 0. Multiple theoretical distance decay models assuming such small grains and uniform species pools E c d E C d Cd E d d i i i i i i i exp( ) exp( ( ) ) exp( ) ( ) ! ( ) !         ∆ ∆ ∆ 1 3 2 3 2 (( ) ! ( ) ! ∆ ∆i id d4 5 4 5  …       (8) The infinite series inside the brackets is a function of d that alternates in sign, with kn/n! converging to zero for large enough n (where k  Δid). Thus, while the series does converge, it occurs in a way that depends on Δi and d. In particular if Δid is  1 over much of the range of interest for d, the possibility exists for very slow convergence and strong deviations from exp(2Cd). Nothing further can be said about this situation without a detailed knowledge of the distribution and magnitude of Δi relative to d. After examining empirical distributions in the Results, we will present further analysis in the Discussion as to when convergence on exp(2Cd) may be expected. Equation 6 can be expanded for the case where the evaluated species pool is not just the species present at L but includes all taxa in the regional pool. To do this, we redefine S to be the richness of the regional species pool, with wi have all produced power law decay forms (Harte and Kinzig 1997, Chave and Leigh 2002, Houchmandzadeh 2008, Morlon et al. 2008). When sample grain is large relative to the organism/agent/ event (making for guaranteed detection), with extent being large enough to pass beyond species environmental and/or geographic range limits (allowing the community pool to vary across the sample), then the main cause of species turnover will be the exceeding of species distributional limits, minimizing the chance that a species will be absent at M but present at L and N. In these situations Y  0, and the exponential form should represent a good approximation. Note that these conditions may be met at various absolute scales. For instance, if strong local spatial or temporal gradients exist, even over very limited extents the community pool will not be uniform and the exponential form should occur. 166 312 detection error and year-on-year variability, we took average abundance over the years 2003–2007, and used only the 1430 routes that had a quality code reflecting appropriate weather and observer ability for all five years. Thus, a bird was counted as being present at a given route if it appeared at least once in five years. The third dataset was the Barro Colorado Island tree plot containing 225 species (Condit et al. 2000). Every tree  10 cm DBH was recorded within each of the 50 one hectare plots. Distance was recorded between hectare midpoints. Due to their continental extents, we a priori expect the LaRoi and BBS datasets to have individual site compositions drawn from varying species pools. Decay of individual species joint-probabilities in these data should thus generally take on an exponential form. However, because the BCI dataset covers only a 1 km extent, all samples are likely to be drawn from only two nested species pools (young within old forest; Svenning et al. 2004). In this situation, species absences are likely due to stochastic sampling effects and/or disturbance history, with a power-law like decay-form prevailing. In all three cases the community distance decay curve should be the sum of the individual species decay curves, although the fit should be worse for the BCI data because of the greater impact of stochastic sampling errors. While no explicit prediction can be made for the functional form of the community distance decay curve, the heuristic considerations given above suggest that an exponential form may be more likely at large grains and/or extents which sample across varying species pools. We also predict the weighted community aggregate curve should perform better for the La Roi and BBS data since they have large enough extents/strong enough gradients to exceed the limits of individual species, while conversely the BCI data should be better fit by an unweighted curve. To test the individual species portion of the model (prediction 1), for each dataset we fit an exponential decay curve to the each species joint-probability by calculating similarity (1 if the species was present in both plots; 0 of not) and distance between each occupied site and all others. Because calculations were only conducted from occupied sites, we constrained similarity to equal 1 at the origin and fit only a single parameter to the decay function. Other functional forms (e.g. linear, power law, logarithmic) were not fit because of their violations of criteria 1 and/or 2. We then calculated mean similarity across distance bins of 200 km (LaRoi and BBS) or 200 m (BCI). Due to the computational challenges in the 1403 site   600 species BBS data (reflecting a total potential number of ~ 6  108 calculations), prior to model fit we randomly removed absent sites in order to generate occupancy rates of approximately 10%. We then sampled 10 000 individuals from all site by site pairs. We recorded the c and r2 value for each species. To calculate the community aggregated curve (Eq. 6), we predicted values every 200 km or 200 m, took their average, and plotted a line through the data. We calculated the weighted community aggregate curve (Eq. 7) similarly except that each species was weighted by its observed occupancy. Because singleton and doubleton species do not possess enough degrees of freedom to accurately fit an exponential model (singletons are points in space while doubletons can only demonstrate linear patterns), we The second major assumption is that the decay of each species joint-probability is independent of that exhibited by others. This allows summing of individual decay curves to produce Eq. 6. Such an assumption of independence between co-occurring species is one of the foundational bases of modern community ecology (Gleason 1926, Curtis 1955, Whittaker 1975) that accurately reproduces a number of community patterns (McGill 2010, 2011, He and Legendre 2002, Green and Plotkin 2007). Although species pairs can be identified which clearly show non-independent distributions (e.g. red bellied woodpecker and red-headed woodpecker – Root 1989; American beach and its obligate parasite beachdrops – Voss 1985), these interactions represent such a small fraction of all possible pairwise interactions (Barker and Mayhill 1999) that the assumption of independence is a valid approximation. Summary of model predictions 1) Individual species will exhibit exponential decay in their joint-probability with increasing distance if and only if observational scale is large enough to allow a varying species pool and to avoid quasi-stochastic occurrence holes. 2) The Sørenson decay of community similarity is equivalent to the occupancy-weighted community average of the individual pi(d) curves in the regional species pool. 3) The weighted community aggregate curve (Eq. 8) will fit better than the unweighted curve in cases where species are not equally likely across the species pool (or equivalently that the species pool changes over the extent studied), but will fit worse when the species pool is constant across sites. 4) The functional form for community distance decay [Sør(d)] depends on the distribution of ci, the decay rates of for individual joint-probability curves. Unfortunately this distribution has never been studied and we have no empirical knowledge of its shape, making impossible the prediction of community distance decay functional form. Even when all single species joint-probability curves are exponential, community decay curves of innumerable forms can be produced. Methods Test of model predictions Three different datasets were used to test the general accuracy of our approach. The first is the vascular plant data of LaRoi (1967), which was sampled from 34 nine-ha (300  300 m) mature, undisturbed upland white spruce forest plots at relatively equal distances across boreal North America. A total of 220 vascular plant species were encountered, with individual plot richness ranging from 19–56. The second is the North American Breeding Bird Survey (BBS), which is run annually at over 2000 routes across the continental US and southern Canada (Sauer et al. 2011). A 3-min point count is taken every half mile (approximately 0.8 km) for 50 stops along a transect of 25 miles (approximately 40 km). An individual route typically has 50–100 bird species detected and overall almost 600 species of birds have been observed at least once. To avoid issues with 167 313 Preserve, Linn County, Iowa (F. Olsen pers comm.); and seasonal robberfly emergences at the Sevilleta LTER (H. Lease pers. comm.). Representative human-system datasets include: concert setlists for Cowboy Junkies ( www.setlist.com ) and Phish ( www.ihoz.com/ PhishStats.html ) performances; all commercially available garden vegetable varieties in the US and Canada as reported by the Garden Seed Inventory of the Seed Savers Exchange; all ingredients reported for the cuisines of Ethiopia, Hungary, India, Iran, Ireland, Korea, Mexico, Norway, Puerto Rico, and Thailand (Smith 1990). The Cowboy Junkies setlist, Garden Seed Inventory, and global cuisine distance decay analyses were previously reported in Nekola and Brown (2007). For each of these datasets, Jaccard similarity was calculated between each pairwise combination of samples, while distance was calculated as the linear difference in space or time between observations. The Jaccard metric was employed because of its common use in distance decay analyses (Nekola and White 1999). The Jaccard and Sørenson indices differ only slightly in their denominator values, and perform in a highly similar (and for all intents, identical) fashion. These data were then subjected to non-linear regression to the mean using the ‘nls’ routine in R, and quantile regression using the ‘nlrq’ routine in R with tau  0.95. Quantile regression was used because the central tendency may not be as important in the distance decay relationship as the shape of the upper bound (Rocchini and Cade 2008). For each of these regression approaches three different functional forms were fit to the data: Sd  Si 2 cd (linear decay); Sd  Si e2cd (exponential decay) and Sd  Ad2c (power-law decay) where Sd  intersample similarity at distance d; Si  initial similarity or nugget (Nekola and White 1999); A  arbitrary intercept; d  intersample distance; c  rate of similarity change. We did not fit logarithmic decay [Sd  a 2 c(log(d))] because we found it to neither outperform power-law fits in the three intensively studied datasets (LaRoi, BBS, BCI) nor in those data which appeared by eye to most likely to express the logarithmic functional form (Tallgrass Prairie Preserve, Iowa Butterfly County occurrences, Portal rodents, and Cowboy Junkes setlists). AIC values were calculated for each regression and ranked from 1 (least) to 3 (largest) across the three functional forms for each dataset. We then constructed a contingency table for each of the regression methods, showing 1st, 2nd, and 3rd AIC ranks vs linear, exponential, and power-law fits. A Fisher-exact test was conducted on these tables to determine if cell frequency violated null expectations. Results Model test All three model predictions were confirmed across the three datasets. In general, the exponential decay form was found to be a good approximation for taiga vascular plant and continental bird species joint-probabilities with increasing observational distance (Fig. 2). While the vast majority of LaRoi and BBS species were fit by an exponential model also calculated a community aggregate curve in which each species was equally weighted (wi  1/S) after removal of singletons and doubletons. We term this the filtered community aggregate curve. The traditional community distance decay curve was also calculated for each dataset. Because of its use in model formulation, Sørenson similarity was used. Since the total BBS data would represent almost 2 000 000 pairwise comparisons, we limited analysis to 10 000 randomly-drawn pairs. To these data we then fit a linear decay curve using ordinary least squares (OLS) regression and an exponential and power-law decay curves using non-linear least squares (NLS) regression. Note that under the assumption of normal errors, OLS and NLS give maximum likelihood estimates allowing the use of the Akiake information criteria (AIC), which we used to compare the three models. We also plotted a LOESS line on these plots to allow for easy visual comparison of the central tendency in the empirical data vs the community aggregated curves. Meta-analysis of community similarity datasets Lacking a theoretical prediction for the functional form of community distance decay, we also conducted a metaanalysis of 26 ecological and 4 human-system datasets. Two datasets represent simulated community turnover data using COMPAS (Minchin 1987), with composition lists being generated for 50 randomly-spaced sites along a single gradient containing 20 species. Two different turnover scenarios were used: constant turnover where all species possess the same modal abundance, standard deviation, and are equally spaced along the gradient, and lognormal turnover, where species modes, standard deviations, and gradient placement follow a lognormal distribution. The remaining data are empirical, represent examples of both spatial and temporal decay, and attempt to maximize ecosystem range, taxa groups, sample grains, and extents. Information regarding the habitat type, taxa group, maximum extent, sample grain, number of samples, and additional notes on these datasets is found in Table 2. These data include: Galapagos Islands vascular plant flora (Preston 1962); North America upland taiga vascular plants (LaRoi 1967) and bryophytes (LaRoi and Stringer 1976); fen-restricted vascular plants in northeastern Iowa (Nekola 1994); Martha’s Head to Bermuda abyssal plain bivalves (Allen and Sanders 1996); Belize terra firma tropical forest (Bird 1998); entire vascular plant floras for all algific talus slope sites along Buck Creek, Clayton County, Iowa (Nekola 1999); Peruvian and Panamanian terra firma forest (Condit et al. 2002); Barro Colorado Island terra firma forest (Condit et  al. 2000); New Zealand land snails (Barker 2005); eastern North American and Asian vascular plants (Qian et al. 2005); Iowa county butterfly faunas (Schlicht et al. 2007); small rodent and winter annuals plants at Portal, Arizona (Ernest et al. 2009); eastern North American rock outcrop, Atlantic coastal plain acidophile, and southwestern USA montane land snail faunas (Nekola 2010); North American Breeding Bird Survey (Sauer et al. 2011); southern Flint Hills tallgrass prairie vascular plants (McGlinn and Palmer 2011); seasonal butterfly and skipper emergences at the Rock Island 168 314 parameters in the weighted community aggregate curve were adjusted to make them fit the true Sørenson decay curve. Although no community aggregate curve provides a good fit to the empirical BCI community decay data, the unweighted curve is a better fit, confirming prediction 3. Linear, exponential and power law forms were then fit to the empirical Sørenson community distance decay (Fig. 4). Community distance decay in the LaRoi and BBS data are best fit by the exponential form, while the BCI data is best fit by a power law (Table 1). Meta-analyses All 26 ecological and 4 human-system datasets demonstrated a negative relationship between community similarity and with r2  0.9, a noticeably higher frequency of poor fits were noted in the BCI data (Fig. 3). These results confirm prediction 1. It should be noted, however, that six BBS species (1% of total), present at only a handful of sites, demonstrated increasing similarity with distance. These were removed from subsequent analysis. The distribution of ci values was unimodal with a strong right truncation, making exponential and power law distributions for C a poor description of the empirical data (Fig. 3). Visual comparison of the weighted community aggregate vs the LOESS line for the true Sørenson decay (Fig. 4) shows a remarkable convergence in the LaRoi and BBS data, confirming prediction 2. Table 1 also demonstrates that the community weighted aggregate produces exponential decay parameters that are very close to the observed true community decay values. This is quite remarkable given that no 0 5000 0 0.5 1 Sorenson Viburdum edule 0 1000 2000 0 0.5 1 Sorenson d (km) Arctostaphylos rubra 0 1000 2000 3000 0 0.5 1 Sorenson Dickcissel 0 1000 2000 3000 0 0.5 1 Sorenson d (km) Scissor-tail flycatcher 0 500 1000 0 0.5 1 Sorenson Acalypha macrostachya 0 500 1000 0 0.5 1 Sorenson d (m) Adelia trioba Figure 2. Individual species joint-probability decay with distance [pi(d )] curves for arbitrary species. Blue dots are observed similarities. Red line is LOESS smoothed regression. Left column represents two species from the LaRoi dataset. Middle column represents two species from the Breeding Bird Survey dataset. Right column represents two species from the Barro Colorado Island dataset. Note that for a few species (like Viburnum edule) exponential decay is a poor fit. However, as seen in the bottom row of Fig. 3 there only a few such species. –5 0 5 0 50 100 LaRoi Ci Ci Ci #spec –5 0 5 x 10–3x 10–3 0 50 100 150 BBS #spec –0.01 0 0.01 0 50 100 150 BCI #spec 0 0.5 1 0 50 100 150 r2 #spec 0 0.5 1 0 200 400 r2 #spec 0 0.5 1 0 50 100 r2 #spec Figure 3. Results of exponential fits to individual species joint-probability occurrence decay. Top row shows distribution of estimated coefficents, ci. Bottom row shows a histogram of r2 values for these fits. Left column is for LaRoi data, middle column is for the Breeding Bird Survey, and the right column is for the Barro Colorado Island data. 169 315 distance (Table 2). Based on AIC ranks, each of the three functional forms were found to provide a best-fit to at least some of the ecological datasets under both non-linear regression to mean and quantile regression to tau  0.95. However, in all four human-system datasets the power-law model was best fit to the data, followed by the exponential model, with the linear model providing the poorest fit. For ecological systems exponential decay was best fit 14 times to the mean, and 13 times to tau  0.95 (Table 3). Powerlaw forms were best fit in 8 and 6 cases, respectively, while linear was the best fit in 4 and 7 cases. While exponential equations never provided the poorest fits, in almost half of the datasets linear fits performed the most poorly (13 and 10, respectively), while power-law fits performed worst in at least 50% of datasets (13 and 16, respectively). Fisherexact tests on these contingency tables indicate a highly significant (p  0.00002 and p  0.000014) deviation of cell frequencies from a uniform null model. Discussion These analyses clearly demonstrate that the distance decay of community similarity reflects the aggregate of individual species joint-probability decay functions across the species pool. As shown in both large-scale detailed datasets (BBS, LaRoi), when the grain size and/or extent are large enough to generate exponential individual jointprobability decay, the community decay curve is also best fit by this form. This outcome is replicated in the metaanalysis, where large scale samples crossing into multiple species pools also exhibited exponential community distance decay. Examples include: simulated communities along a strong environmental gradient; vascular plant flora of the Galapagos archipelago; vascular plant and bryophyte assemblages across boreal North America; vascular plants of North American and Asian temperate forests; North American Atlantic coastal plain land snails; BBS data; and the adult emergence of both butterflies/skippers and robberflies along a seasonal gradient. 0 2000 4000 6000 0 0.5 1 Distance (km) Sorenson LaRoi(A) (B) (C) 0 2000 4000 6000 8000 0 0.5 1 Distance (km) Sorenson BBS 0 200 400 600 800 1000 0.4 0.6 0.8 1 Distance (km) Sorenson BCI Raw Data LOESS Exponential Power CommAgg CommAgg Wtd CommAgg Filt Figure 4. Empirical distance decay relationships, with each blue dot representing a plot pair. Red solid line shows LOESS fit. Red dash-dotted red line shows an exponential fit. Red dotted line shows a power law fit. The green lines show the community aggregated curves. The solid green line equally considers all species (i.e. Eq. 6). The dotted green line weights species by their occupancy (i.e. Eq. 7). The dash-dotted line includes all species equally but eliminates species found in 2 or fewer sites. (A) LaRoi boreal plant data. (B) North American Breeding Bird Survey (note only 10 000 random pairs were used out of the over 2 000 000 possible). (C) Barro Colorado Island data. Table 1. Summary of results for three datasets. True Community aggregate Weighted community aggregate C AIC C AIC C AIC LaRoi Linear 20.0001 22578.8 20.00012 2125.4 20.0001 2142.4 Exponential 20.00027 22590.4 20.00056 2183.5 20.00032 2185.1 Power 20.286 22355.1 20.482 2147.0 20.353 2149.4 BBS Linear 20.00011 238294 20.00007 2156.0 20.00008 2167.5 Exponential 20.00039 239287 20.00047 2202.3 20.00034 2215.1 Power 20.370 237635 20.484 2211.5 20.409 2199.0 BCI Linear 20.00006 27646 20.00039 231.0 20.00021 243.2 Exponential 20.00009 27647 20.00059 234.0 20.00024 245.4 Power 20.0367 27677 20.202 239.6 20.085 240.9 170 316 Table2.Source,system,samplescale,andregressioninformationfortheanalyzeddatasets.TheupperAICvaluesandranksrepresentstatisticsgeneratedfromnon-linearregressiontomean,whilethe lowervaluesrepresentstatisticsgeneratedfromquantilenon-linearregressionfortau  0.95. No. samples AICvalue/rankorder SourceLocationHabitatTaxagroupMaxExtentGrainNotesLinearExponentialPower COMPASSimulated50Constantturnover23665.7/2 4730.4/2 23956.1/1 3837.2/1 21478.3/3 5430.9/3 COMPASSimulated50Lognormalturnover21317.6/2 7018.5/2 21976.0/1 6937.4/1 2721.0/3 7372.3/3 Preston1962Galapagos Islands Volcanic archipelago Vascular plants 286.2km1.1–4640km2182222.7/2 51.0/1 2224.6/1 51.6/2 2206.7/3 61.5/3 LaRoi1967North America UplandtaigaVascular plants 5687.4km9ha3421212.6/3 1770.0/2 21244.1/1 1729.3/1 21027.2/2 2116.4/3 LaRoiandStringer1976North America UplandtaigaBryophytes5687.4km9ha3421205.1/2 1810.7/2 21221.7/1 1804.4/1 21122.2/3 1919.9/3 Nekola1994Northeastern Iowa FenSpecialist vascular plants 254.0km0.1–50.1ha10291.1/1 291.7/1 291.0/2 290.7/2 289.7/3 287.4/3 AllenandSanders1996Atlantic abyssal plain TotalfaunaProtobranch bivalves 777.5kmca0.1ha21 3000mdepthonly2286.8/1 498.3/1 2274.4/2 501.9/2 2208.7/3 522.6/3 Bird1998BelizeTropicalforestCanopy trees 205.5km1ha302798.2/3 1290.1/3 2810.0/2 1282.6/2 2947.6/1 1068.2/1 Nekola1999BuckCreek valley,Iowa Algifictalus slope Vascular plants 3.1km0.125–1ha7283.1/1 2109.2/1 282.9/2 2109.1/2 277.6/3 2107.5/3 Conditet al.2000Barro Colorado Island TropicalforestTrees   10cmdbh 1km1ha5024167.7/3 213446/3 24168.8/2 213448/2 24199.0/1 213496/1 Conditet al.2002PanamaTropicalforestCanopy trees 110.6km1ha10029045.8/3 38969.4/3 211850/1 37667.5/1 211125/2 37820.6/2 Conditet al.2002PeruTropicalforestCanopy trees 102.2km1ha152275.4/3 2105.6/3 2276.1/2 2107.8/2 2305.8/1 2136.2/1 Barker2005NewZealandTotalfaunaLandsnails1590.9km282460.6/2 1197.3/1 2463.0/1 1205.8/2 2416.5/3 1371.6/3 Qianet al.2005ENorth America TotalfloraVascular plants 1895.3km67000–249000km22547chosen comparisons 264.6/2 2107.2/1 267.0/1 2106.6/2 264.3/3 238.3/3 Qianet al.2005EAsiaTotalfloraVascular plants 1870.6km67000–249000km21840chosen comparisons 261.4/3 293.0/2 276.2/1 2116.9/1 270.9/2 210.2/3 Schlichtet al.2007IowaCounty faunas Butterflies and skippers 502.1km1042–2524km260Limitedtowell sampled(40  taxa) counties 23973.9/3 9840.0/3 23975.8/2 9839.1/2 23982.2/1 9838.4/1 Ernestet al.2009Portal, Arizona Desertscrub and grassland Winter annual plants 16yrSeasonsummaryfor250m217Plotno.22from 1989–2005 10.4/3 264.7/3 10.3/2 265.3/1 9.7/1 265.4/2 Ernestet al.2009Portal, Arizona Desertscrub and grassland Rodents25yrSeasonsummaryfor250m226Plotno.22from 1976–2005 2301.9/3 1227.5/3 2310.9/2 1219.0/2 2326.0/1 1197.8/1 (Continued) 171 317 Table2.(Continued). No. samples AICvalue/rankorder SourceLocationHabitatTaxagroupMaxExtentGrainNotesLinearExponentialPower Nekola2010USAtlantic Coast AcidhabitatsLandsnails2013km10 000m2152147.1/2 70.9/2 2148.5/1 69.9/1 2146.5/3 72.0/3 Nekola2010ENorth America RockoutcropLandsnails3031.710 000m24821269.0/2 6446.9/1 21297.1/1 6462.6/2 21225.6/3 6553.0/3 Nekola2010SWNorth America Montane forest Landsnails841.1km10 000m2142104.9/1 26.4/2 2102.9/2 26.3/1 295.7/3 41.0/3 McGlinnandPalmer2011OklahomaTallgrass prairie Vascular plants 15.6km100m2151Comparisonswithplot no.5only 2358.0/3 7.8/3 2359.8/2 5.6/2 2367.5/1 211.0/1 McGlinnandPalmer2011OklahomaTallgrass prairie Vascular plants 9yrSeasonsummaryfor100m210Plotno.5from 1997–2006 2149.5/3 2156.7/2 2151.0/2 2157.2/1 2155.6/1 2152.2/3 Saueret al.2011North America Totalfauna 2002–2007 Breeding birds 6139.8km~ 16ha~ 37001000drawsfrom allpairwise comparisons 2968.9/2 28687.5/2 21084.8/1 28698.1/1 2878.4/3 28405.5/3 H.Leasepers.comm.Sevielleta LTER DesertscrubRobberflies193d1d262006season2173.9/3 1289.9/3 2231.7/1 1269.6/1 2227.5/2 1275.2/2 F.Olsenpers.comm.RockIsland Preserve, Iowa Tallgrass prairie Butterflies and skippers 174d2–4h381990season2713.1/3 3301.1/3 2741.9/1 3244.2/1 2718.3/2 3277.2/2 NekolaandBrown2007Concert setlists Cowboy Junkies 6856d1show331987to20062894.8/3 24559.7/3 2921.3/2 24657.3/2 21047.1/1 24910.7/1 www.ihoz.com/ PhishStats.html Concert setlists Phish6070d1show181983–20002485.5/3 246.5/3 2511.5/2 251.0/2 2525.5/1 269.9/1 NekolaandBrown2007Commercially available varieties GardenSeed Inventory 23yr1GardenSeedInventory71981to2004240.4/3 2183.4/3 250.4/2 2199.4/2 254.4/1 2231.6/1 NekolaandBrown2007Ingredient lists Global cusines 16500km1ethnicregion102126.1/3 2453.2/3 2128.5/2 2454.7/2 2136.6/1 2456.3/1 172 318 domesticates now feature prominently in new world cuisines – e.g. wheat, cinnamon, anise, chickens and pigs in Mexico. The existence of public and private seed banks has also made it possible for ‘lost’ commercial varieties to be made available after absences of a few decades. And, it is also highly unlikely that either the Cowboy Junkies or Phish would be unable to play any song in their repertoire given more than a few days’ rehearsal. As a result, agents/events within the analyzed human systems tend to be drawn from a uniform pool, with absences being largely the result of occurrence holes or stochastic sampling effects. Such situations should – and do – lead to power-law community distance decay. While we are unable to analytically prove that exponential community distance decay should be expected in situations where individual species joint-probability curves are also largely exponential, consideration of ci values for LaRoi and BBS in conjunction with Eq. 8 (the Taylor series expansion), helps explain this correspondence. Recall that convergence to exp(2Cd) is dependent upon how fast (Δid)n/n! converges to zero as n increases. Our empirical data (Fig. 3) indicate that the distribution of Δi (or equivalently, ci) is strongly truncated on both tails, making the values of ci strongly centered around the mean. As a result ci (or more specifically Δi 5 ci 2 ci) is never larger than a small multiple of ci (e.g. 2–3 times). Second we note that ci is order-of-magnitude close to but slightly larger than 1/d such that cid is slightly greater than 1 (i.e. approximately in the interval [1,10] for most of the values of d in the decay curve graphs). Since, by tail truncation, Δi is never larger than 2ci or 3ci, then Δid also must fall in the range of roughly 1–10. Additionally note that (Δid)n/n! converges on zero quite quickly when Δid is in the range [1,10]. Also note Δid is less than 1 when d is close to zero, which also forces a fast convergence on zero. As a result, the Taylor series expansion necessitates a community distance decay function close to exp(2Cd) when species joint-probability occurrence is exponential and the distribution of ci values is strongly truncated and order of magnitude close to 1/d. While providing a numerically based understanding for the preponderance of observed exponential-like community distance decay, this analysis can not be construed as a general proof, as the end result is strongly dependent upon the form of the joint-probability decay curve and ci distribution properties. These analyses also document two additional important additional insights regarding the nature of ecological communities. First, it is striking how similar the parameterized and predicted coefficients are between the true and weighted aggregate community distance decay function for large-scale data (Fig. 4; Table 1). This correspondence provides strong evidence of individualistic species sorting (Gleason 1926). Second, the different expected forms for distance decay within (power law) vs between (exponential) communities allows for an empirical test of the observational scales under which communities exist. For instance, in Panamanian tropical forest, strong power-law decay apparent at distances of 1 km or less indicate that samples are being drawn from within the same community. However, the exponential decay observed at distances of 5–100 km indicates that at this larger scale different communities are being surveyed. However when community data is sampled at small grains or within a mostly uniform species pool at limited extents (BCI), exponential decay is a poorer describer of jointprobability decay with community distance decay exhibiting a power-law like form. This result was also replicated in the meta-analysis. Examples include: vascular plants within 100 m2 quadrats along a 15.6 km extent in the Tallgrass Prairie Preserve in northeastern Oklahoma; canopy trees within 1 ha tropical forest stands along a 1 km extent at BCI, a 100 km extent in lowland Peru, and a 200 km extent in Belize; seasonal summaries for rodents within a 250 m2 sample over a 25 yr extent in Portal, Arizona. While Iowa county butterfly faunas would appear a violation, with power-law decay being favored even though ~ 25% of the fauna reaching its range limit within the state boundary, it is also important to realize that range-limit species also tend to be rare. Since rare species do not exhibit a distance decay signal (Nekola and White 1999), they will not contribute greatly to the overall shape of the decay function. As a result, the observed power law shape is largely due to the ~ 75% of the fauna which is cosmopolitan. Power-law community distance decay has also previously been shown to describe the variation of vascular plant composition at  10 km extents within Colorado subalpine meadows (Harte et  al. 1999) and for soil microbes over ~ 1000 km extents in Australia (Green et  al. 2004). It is important to note in this latter example that because of their exemplary long-distance dispersal abilities (Green and Bohannan 2006) local microbe communities are likely drawn from a much more extensive and uniform species pool than would be the case for most macroscopic organism groups. The predominance of power-law community distance decay in human-systems is also readily explained using this framework. Because of the immense human capacity to move ideas and goods across large spatial scales, and to store and retrieve data across long time periods, absolute range limits for commercial goods or ideas are likely not crossed at either planetary spatial extents or decadal time spans. For instance, global trade over the last 600 yr has allowed essentially all domesticated crops to be available world wide. As a result, many new world domesticates are now common aspects in old world cuisines – e.g. tomatoes in Italian, chiles in Indian/Ethiopian/Chinese/Thai, while old-world Table 3. Contingency table of AIC rank vs functional form of the distance decay relationship. The smallest AIC value among the three functional forms was assigned a rank of ‘1’, the second smallest a rank of ‘2’, with the largest being assigned a rank of ‘3’. The first number in a cell represents the number of datasets achieving that given rank based on AIC values calculated from non-linear regression to mean, while the second represents the number achieving that rank based on quantile non-linear regression for tau  0.95. Mathematical distance decay form AIC rank Linear Exponential Power law 1 4/7 14/13 8/6 2 9/9 12/13 5/4 3 13/10 0/0 13/16 Fisher exact test for identical cell frequencies: p  0.00001966 for non-linear regression to mean; p  0.00001374 for quantile non-linear regression. 173 319 Green, J. L. and Plotkin, J. B. 2007. A statistical theory for sampling species abundances. – Ecol. Lett. 10: 1037–1045. Green, J. L. et  al. 2004. Spatial scaling of microbial eukaryote diversity. – Nature 432: 747–750. Harte, J. 2011. Maximum entropy and ecology. – Oxford Univ. Press. Harte, J. and Kinzig, A. P. 1997. On the implications of species–area relationships for endemism, spatial turnover, and food web patterns. – Oikos 80: 417–427. Harte, J. et al. 1999. Estimating species–area relationships from plot to landscape scale using species-turnover data. – Oikos 86: 45–54. He, F. and Legendre, P. 2002. Species diversity patterns derived from species–area models. – Ecology 83: 1185–1198. Houchmandzadeh, B. 2008. Neutral clustering in a simple experimental ecological community. – Phys. Rev. Lett. 101: 078103. Hubbell, S. P. 2001. The unified neutral theory of biodiversity and biogeography. Monographs in population biology no. 32. – Princeton Univ. Press. Hurlbert, A. H. and White, E. P. 2007. Ecological correlates of geographical range occupancy in North American birds. – Global Ecol. Biogeogr. 16: 764–773. Jobe, R. T. 2007. Estimating landscape-scale species richness: reconciling frequency- and turnover-based approaches. – Ecology 89: 174–182. Karlin, S. and Taylor, H. M. 1975. A first course in stochastic processes, 2nd ed. – Academic Press. LaRoi, G. H. 1967. Ecological studies in the boreal sprucefir forests of the North American taiga. I. Analysis of the vascular flora. – Ecol. Monogr. 37: 229–253. LaRoi, G. H. and Stringer, M. H. 1976. Ecological studies in the boreal spruce-fir forests of the North American taiga. II. Analysis of the bryophyte flora. – Can. J. Bot. 54: 619–643. Lindgren, B. W. 1976. Statistical theory. – Macmillan Publishing. McGill, B. J. 2010. Towards a unification of unified theories of biodiversity. – Ecol. Lett. 13: 627–642. McGill, B. J. 2011. Linking biodiversity patterns by autocorrelated random sampling. – Am. J. Bot. 98: 481–502. McGlinn, D. J. and Palmer, M. W. 2011. Quantifying the influence of environmental texture on the rate of species turnover: evidence from two habitats. – Plant Ecol. 212: 495–506. McKnight, M. W. et al. 2007. Putting beta-diversity on the map: broad scale congruence and coincidence in the extremes. – PLoS Biol. 5: 2424–2432. Minchin, P. R. 1987. Simulation of multidimensional community patterns: towards a comprehensive model. – Plant Ecol. 71: 145–156. Morlon, H. et al. 2008. A general framework for distance-decay of similarity in ecological communities. – Ecol. Lett. 11: 904–917. Nekola, J. C. 1994. Environment and vascular flora of northeastern Iowa fen communities. – Rhodora 96: 121–169. Nekola, J. C. 1999. Paleorefugia and neorefugia: the influence of colonization history on community pattern and process. – Ecology 80: 2459–2473. Nekola, J. C. 2010. Acidophilic terrestrial gastropod communities of North America. – J. Molluscan Stud. 76: 144–156. Nekola, J. C. and White, P. S. 1999. Distance decay of similarity in biogeography and ecology. – J. Biogeogr. 26: 867–878. Nekola, J. C. and Brown, J. H. 2007. The wealth of species: ecological communities, complex systems, and the legacy of Frank Preston. – Ecol. Lett. 10: 188–196. Palmer, M. W. 2005. Distance decay in an old-growth neotropical forest. – J. Veg. Sci. 16: 161–166. Perez-del-Olmo, A. et al. 2009. Not everything is everywhere: the distance decay of similarity in a marine host–parasite system. – J. Biogeogr. 36: 200–209. Since its popularization within ecological research over a decade ago, there has been a mismatch between theoretical models and empirical results of distance decay analysis. We hope the approach described within this paper helps resolve this conundrum. Specifically, sampling scales associated with small grain, limited extents, and absent/limited environmental gradients are expected to have power-law community distance decay, while sampling scales associated with large grain and strong environmental gradients are expected to have exponential community distance decay. Acknowledgements – The following graciously provided us with their raw datasets for analysis: Gary Barker (New Zealand land snail data); Matt Barthel (Garden Seed Inventory data); Neil Bird (Belize tropical forest data); Jim Brown, Morgan Ernst and Tom Valone (Portal rodent and vascular plant data); Richard Condit (Panama and Peru tropical forest data); Hilary Lease (Sevilleta LTER robberfly data); Frank Olsen (Rock Island Preserve lepidoptera data); Mike Palmer and Dan McGlinn (Tallgrass Prairie Preserve data); Peter White (eastern Asian and North American regional flora data). Peter Minchin provided the COMPAS code for simulating community turnover along environmental gradients. Initial work was funded through a Santa Fe Inst. workshop grant. References Allen, J. A. and Sanders, H. L. 1996. The zoogeography, diversity and origin of the deep-sea protobranch bivalves of the Atlantic: the epilogue. – Prog. Oceanogr. 38: 95–153. Barker, G. M. 2005. The character of the New Zealand land snail fauna and communities: some evolutionary and ecological perspectives. – Rec. West. Aust. Mus. Suppl. 63: 52–102. Barker, G. M. and Mayhill, P. C. 1999. Patterns of diversity and habitat relationships in terrestrial mollusc communities of the Pukeamaru Ecological District, northeastern New Zealand. – J. Biogeogr. 26: 215–238. Bell, G. 2003. The interpretation of biological surveys. – Proc. R. Soc. B 270: 2531–2542. Bird, N. M. 1998. Sustaining the yield: improved timber harvesting practices in Belize, 1992–1998. – Natural Resources Inst., Univ. of Greenwich, NRI Catalog Services, CAB International, Wallingford, UK. Blanchette, C. A. et al. 2008. Biogeographical patterns of rocky intertidal communities along the Pacific coast of North America. – J. Biogeogr. 35: 1593–1607. Chave, J. and Leigh, E. G. Jr 2002. A spatially explicit neutral model of b-diversity in tropical forests. – Theor. Popul. Biol. 62: 153–168. Condit, R. et  al. 2000. Spatial patterns in the distribution of tropical tree species. – Science 288: 1414–1418. Condit, R. et al. 2002. Beta-diversity in tropical forests. – Science 295: 666–669. Curtis, J. T. 1955. A prairie continuum in Wisconsin. – Ecology 36: 558–566. Ernest, S. K. M. et  al. 2009. Long-term monitoring and experimental manipulation of a Chihuahuan Desert ecosystem near Portal, Arizona, USA. – Ecology 90: 1708. Gilbert, B. and Lechowicz, M. J. 2004. Neutrality, niches, and dispersal in a temperate forest understory. – Proc. Natl Acad. Sci. USA 101: 7651–7656. Gleason, H. A. 1926. The individualistic concept of plant association. – Bull. Torrey Bot. Club 53: 7–26. Green, J. L. and Bohannan, B. J. M. 2006. Spatial scaling of microbial biodiversity. – Trends Ecol. Evol. 21: 501–507. 174 320 Schlicht, D. W. et al. 2007. The butterflies of Iowa. – Univ. of Iowa Press. Soininen, J. et al. 2007. The distance decay of similarity in ecological communities. – Ecography 30: 3–12. Smith, J. 1990. The frugal gourmet on our immigrant ancestors. – Harper-Collins. Steinitz, O. et al. 2005. Predicting regional patterns of similarity in species composition for conservation planning. – Conserv. Biol. 19: 1978–1988. Svenning, J.-C. et  al. 2004. Ecological determinism in plant community structure across a tropical forest landscape. – Ecology 85: 2526–2538. Tobler, W. R. 1970. A computer movie simulating urban growth in the Detroit region. – Econ. Geogr. 46: 234–240. Tuomisto, H. et  al. 2003. Dispersal, environment, and floristic varia­tion of western Amazonian forests. – Science 299: 241–244. Voss, E. G. 1985. Michigan flora. – Cranbrook Inst. of Science, Bloomfield Hills, MI, Bulletin 59. Whittaker, R. H. 1975. Communities and ecosystems. – MacMillan Publishing. Plotkin, J. B. and Muller-Landau, H. C. 2002. Sampling the species composition of a landscape. – Ecology 83: 3344–3356. Preston, F. W. 1962. The canonical distribution of commonness and rarity. – Ecology 43: 185–215, 410–432. Qian, H. et al. 2005. Beta diversity of angiosperms in temperate floras of eastern Asia and eastern North America. – Ecol. Lett. 8: 15–22. Qian, H. et al. 2009. The latitudinal gradient of beta diversity in relation to climate and topography for mammals in North America. – Global Ecol. Biogeogr. 18: 111–122. Rapoport, E. H. 1982. Areography: geographical strategies of species. – Pergamon Press. Rocchini, D. and Cade, B. S. 2008. Quantile regression applied to spectral distance decay. – Geosci. Remote Sens. Lett. 5: 640–643. Root, T. 1989. Atlas of wintering North American birds. – Univ. of Chicago Press. Sauer, J. R. et al. 2011. The North American Breeding Bird Survey, results and analysis 1966–2009. Version 3.23.2011. – USGS Patuxent Wildlife Research Center, Laurel, MD. 175 IDEA AND PERSPECTIVE The wealth of species: ecological communities, complex systems and the legacy of Frank Preston Jeffrey C. Nekola* and James H. Brown Department of Biology, University of New Mexico, Albuquerque, NM 87131, USA *Correspondence: E-mail: jnekola@unm.edu. Abstract General statistical patterns in community ecology have attracted considerable recent debate. Difficulties in discriminating among mathematical models and the ecological mechanisms underlying them are likely related to a phenomenon first described by Frank Preston. He noted that the frequency distribution of abundances among species was uncannily similar to the Boltzmann distribution of kinetic energies among gas molecules and the Pareto distribution of incomes among wage earners. We provide additional examples to show that four different Ôdistributions of wealthÕ (species abundance distributions, species–area and species–time relations, and distance decay of compositional similarity) are not unique to ecology, but have analogues in other physical, geological, economic and cultural systems. Because these appear to be general statistical patterns characteristic of many complex dynamical systems they are likely not generated by uniquely ecological mechanistic processes. Keywords Community ecology, competitive sorting, complexity science, distance decay, ecological theory, neutral models, species–abundance distribution, species–area relationship, species–time relationship, statistical mechanics. Ecology Letters (2007) 10: 188–196 Frank Preston (1950) published a little paper on ÔGas Laws and Wealth LawsÕ in The Scientific Monthly. In it he commented on the remarkable similarities between the Boltzmann frequency distribution of molecular kinetic energies in gasses, the Pareto frequency distribution of personal incomes in countries, and the frequency distribution of species abundances in ecological communities. After noting how such distributions might arise, he went on to remark – in comments which would now be politically incorrect in reference to personal incomes – about whether it is Ôwise to try to change the laws of natureÕ. Interest in community Ôdistributions of wealthÕ (DOWs), in particular the species–abundance distribution (SAD), species–area relationship (SAR), species–time relationship (STR), and distance decay of compositional similarity (DD), have preoccupied community ecologists ever since Preston’s seminal papers (Preston 1948, 1962a,b, 1980; see also Williams 1964; MacArthur 1972; May 1975; Nee et al. 1991; Rosenzweig 1995; Nekola & White 1999; Gaston & Blackburn 2000; Hubbell 2001). As new, larger, and more comprehensive data sets have become available, the discipline of community ecology has been enlivened by debate over which of several different conceptual frameworks or formal mathematical models might be necessary or sufficient to explain these patterns. These debates have primarily had two foci. First, authors have argued about the relative roles of ÔcompetitiveÕ (deterministic ecological differences between species) vs. ÔneutralÕ (stochastic processes of birth, death and dispersal) mechanisms in community assembly. This issue has been difficult to resolve as formal models invoking complete neutrality generate very similar outcomes to models that assume unique species-level abiotic requirements or strong biotic interactions (e.g. Bell 2000; Chave et al. 2002; Mouquet & Loreau 2003; Chave 2004; Tilman 2004; Gaston & Chown 2005). The second focus has been about which particular mathematical distribution (and corresponding ecological mechanism) best describes empirical DOWs. This debate has been constrained by the fact that mathematical distributions having different mechanistic interpretations can be very difficult to empirically distinguish. For instance, it has proved extremely difficult to determine whether a particular SAD data set is best fit by a power law, truncated Ecology Letters, (2007) 10: 188–196 doi: 10.1111/j.1461-0248.2006.01003.x Ó 2007 Blackwell Publishing Ltd/CNRS 176 log-normal, or zero sum multinomial (ZSM) distribution (e.g. McGill 2003a; Volkov et al. 2003; Chave 2004; Etienne & Olff 2005), and similarly whether a power law or exponential distribution best fits a given SAR/STR relationship (Loehle 1990; White et al. 2006). May (1975) and Connor & McCoy (1979) made these points decades ago for the SAD and SAR. We regard as healthy the renewed interest in community ecology, and especially the focus on large-scale, multispecies macroecological pattern and process which are often not amenable to testing by manipulative experiments. But we also wonder whether the current framing of questions and tenor of debates are productive. Perhaps, as Preston suggested, DOWs are not the unique provenance of ecology, but rather are common properties of many seemingly disparate systems. If so, it may be best to seek explanations for these patterns that are as general as the systems that exhibit them. To address this issue, we have extended Preston’s examples to a broader selection of data sets spanning a wide range of physical and human economic, social and artistic systems. Physical systems: (i) Yearly precipitation averages for each of 1027 North American sites (see Nekola 2005). These data are based on 1-km resolution global precipitation maps created by the WORLDCLIM Project (http://www.worldclim.org); (ii) Mineral species richness from county to global scales, as reported by the MINDAT data base (http://www.mindat.org). Data represent the total number of minerals for the entire terrestrial globe, for 10 countries (Australia, Canada, Chile, China, India, Mexico, Namibia, Turkey, the UK and the USA), all the USA states (including the District of Columbia), all Canadian provinces/ territories, and two selected counties per state. Areas for each unit were based on data provided with ArcMap 9.1 (ESRI, Redlands, California, USA). Economic systems: 2004 Stock volumes for all publicly traded corporations in the USA, based on data reported by Bloomberg (http://pages.stern. nyu.edu/$adamodar/New_Home_Page/data.html). Social systems: (i) Citation frequencies for all papers catalogued by the Institute for Scientific Information from 1981 to 1997 (http:// physics.bu.edu/$redner/projects/citation/isi.html); (ii) the list of all commercially sold garden vegetable varieties in the USA and Canada from 1981 to 2004 as reported by the Garden Seed Inventory of the Seed Savers Exchange; (iii) the list of all ingredients reported in recipes for 10 global cuisines (Ethiopia, Hungary, India, Iran, Ireland, Korea, Mexico, Norway, Puerto Rico and Thailand) from Smith (1990). Artistic systems: (i) The number of unique words and total word length for the 1863 texts documented in Project Gutenberg as of 2000 (http://www.mine-control.com/ zack/guttenberg/); (ii) concert setlists for 33 approximately evenly spaced Cowboy Junkies performances from 1987 to 2006 (http://setlist.com). Using standard ecological protocols (e.g. ÔoctavesÕ or log2 bin widths for SADs and Jaccard’s similarity in the calculation of compositional similarity between all pairwise combinations of observations for DD analyses), we treated each as if they were ecological data sets. All DOW analogues were calculated from the Cowboy Junkies example, as only this data set provided abundance data across multiple observations. For the remaining, analyses were limited to the DOW equivalents that could be assessed from the given data: SAD analogues for precipitation class frequencies, stock volumes and scientific citations; SAR analogues for mineral species richness and Project Gutenberg texts; STR analogues for North American garden seed offerings; and DD analogues for global cuisine and vegetable seed offerings. These analyses corroborate Preston’s observation that typical DOW patterns are by no means unique to ecological systems. North American precipitation classes, 2004 stock volumes, scientific citation frequency and song frequencies from Cowboy Junkies setlists (Fig. 1) are all similar to truncated log-normal or ZSM distributions with rare events being more frequent than abundant ones. Using code developed by McGill (2003a) the ZSM can be shown to well fit both the precipitation (h ¼ 55.173, m ¼ 0.179) and Cowboy Junkies setlist (h ¼ 89.446, m ¼ 0.240) data. Power law SAR and STR-like relationships are also widespread (Fig. 2), being observed not only for the accumulation of mineral species across space, but also for unique words as a function of book size and for unique vegetable varieties and Cowboy Junkies song performances as a function of time. The fitted relationships explained 46–99% of observed variance. Finally, nonlinear DD relationships are also common (Fig. 3), being evident in cuisine ingredient lists, the garden seed industry, and Cowboy Junkies setlists. For these examples, power law decay models fit best, accounting for 38–88% of observed variance. The convergence between ecological and non-ecological community patterns is not limited to these examples. In his final Ecology paper, Preston (1981) documented rarityenriched ZSM-like SAD analogues for the service life of restaurant drink tumblers, the static fatigue of glass and other materials, and first marriage age for Danish, UK and the USA women. Power law STR analogues have also been noted in comprehensive examination scores for degree candidates in the University of Oslo, Department of Biology and for Norsk Hydro stock prices on the Norwegian market (Ugland et al. 2005). And, given the convergence between the body-size distribution of beetles in Borneo tree canopies and cars in York and Heathrow Airport parking lots (Gaston et al. 1993), it appears such similarities may not be limited to DOWs. These examples caution that the general mechanisms generating many familiar DOW patterns are not unique to ecology. Certainly, the processes of birth, death, Idea and Perspective The wealth of species 189 Ó 2007 Blackwell Publishing Ltd/CNRS 177 immigration and speciation invoked by Hubbell (2001), or the competitive tradeoffs invoked by Tilman (2004) or Chave et al. (2002) could not have generated the ZSM-like SADs for precipitation classes, paper citations, stock volumes, drink tumbler longevity, and marriage ages, or power law SARs/STRs for unique mineral, word, and garden seed occurrences, or comprehensive examination scores. Conversely, it seems unlikely that Cowboy Junkies performances mimic community ecology process, even though they display rarity-enriched SAD, power law STD and nonlinear DD patterns. Preston (1950) concluded that the remarkable convergence between the Boltzmann, Pareto and SAD might be the result of Ôstatistical mechanicsÕ. Even though he could not identify the common underlying mechanisms, he urged scientists to Ôunderstand [this] law, and the causes that bring it aboutÕ. We question whether the recent revival of interest in ecological DOWs has advanced this goal. These congruent patterns seem to beg for some kind of explanation that is both more universal and also less explicitly mechanistic than the kinds typically sought and offered by community ecologists. Over the last few decades investigations in the new interdisciplinary field of complexity science have begun to address such issues. Complexity science developed as natural and social scientists sought to identify and explain common features exhibited by complex dynamical systems in such seemingly disparate fields as quantum physics, computer science, economics, sociology, political science, linguistics, astronomy, geology and meteorology. While a single definition for a Ôcomplex systemÕ remains elusive, a number of common features have been identified (Brown 1994a,b; West 2006a,b): (1) They are composed of many components of many different kinds. (2) These components interact with each other and the extrinsic environment in many different ways and on multiple spatial and temporal scales. (3) These interactions give rise to complex structures and complicated nonlinear dynamics. (4) These structures and dynamics are neither completely stochastic nor entirely deterministic, but instead represent a combination of randomness and order. (5) They contain both positive and negative feedback mechanisms, causing either amplification or damping of temporal and spatial variation, depending on conditions. Figure 1 Species–abundance distribution analogues for non-ecological systems, including North American precipitation classes, stock volumes for all publicly traded US corporations, song performances for Cowboy Junkies setlists, and citation frequencies for scientific papers. All distributions have been binned into log2 ÔoctavesÕ. For precipitation classes and Cowboy Junkies performance frequencies, the best-fit zero sum multinomial distribution has also been calculated using code provided by McGill (2003a). These best-fit curves are presented as the graphed lines in both of these panels. 190 J. C. Nekola and J. H. Brown Idea and Perspective Ó 2007 Blackwell Publishing Ltd/CNRS 178 (6) They are open systems which require exchanges of energy, materials, and/or information from extrinsic sources to maintain highly organized states far from thermodynamic equilibrium. (7) They are historically contingent, so that their present configurations reflect the influence of initial conditions and subsequent perturbations. (8) They are often nested within other complex systems, giving rise to hierarchical organizations that can be approximated by fractal geometry and dynamic scaling laws. Ecological communities clearly demonstrate these features. Even the simplest contain thousands to billions of individuals of tens to thousands of different species, ranging from unicellular procaryotes, protists, and fungi to multicellular plants and animals. These individuals and species interact with each other and their extrinsic abiotic environment across multiple spatial and temporal scales. These relationships are often inherently nonlinear, ranging from Michaelis–Menten curves for nutrient uptake to exponential or logistic population growth to the normal, skewed or bimodal distribution of species along environmental gradients. Feedbacks are prevalent, as demonstrated by effects of keystone species on ecosystem function or fuel buildup on fire return frequencies. Ecological communities require the continual transformation of energy, material and information to maintain their highly organized, far-from-equilibrium thermodynamic states. Many ecological processes are historically contingent, with contemporaneous patterns reflecting legacies of past events in both shallow (e.g. the sequence of habitat colonization) and deep (e.g. plate tectonics and evolution) time. And finally, ecological communities are composed of nested hierarchies of complex components, ranging from organic Figure 2 Species–area relationship and species–time relationship analogues for non-ecological systems, including the accumulation of unique mineral varieties by sample area, unique words by total document words, and unique vegetable varieties and Cowboy Junkies songs by time. All data are graphed on log-transformed axes. Best-fit linear regression lines have been plotted for each relationship. Idea and Perspective The wealth of species 191 Ó 2007 Blackwell Publishing Ltd/CNRS 179 molecules and cells through populations, species, guilds and trophic levels. Because of these shared structural and dynamical properties, ecological communities and other kinds of complex systems tend to develop similar patterns of whole-system structure and dynamics. An example of this is the occurrence of many small and few large magnitude events, which can be observed in the cumulative frequency distributions of word use, scientific paper citations, Internet web hits, copies of books sold, telephone calls received on a single day, earthquake magnitudes, lunar crater diameters, solar flare intensity, deaths in wars, wealth of rich people, surname frequencies, and city populations (Newman 2005). These patterns can often be nearly equally well fit by power law or log-normal probability distribution functions (West & Shlesinger 1989). The finding of DOW analogues (rarityenriched SADs, roughly power law SAR and STR, and nonlinear DD) across multiple non-ecological systems thus suggests that these patterns represent shared statistical properties of a large class of complex systems. To discover the ÔlawÕ sought by Preston to explain these striking convergences may therefore require an understanding of the general principles that govern the structure and dynamics of all these systems. The uncanny similarity of ecological DOWs to those exhibited by other complex systems has been largely ignored by community ecologists, who for the most part have sought explanation strictly from biological processes (but see Limpert et al. 2001; McGill 2003b; Halloy & Whigham 2004; Ugland et al. 2005). A few authors have pointed out phenomenological similarities between ecological and other complex systems and suggested that these may hold clues to common causes (e.g. Preston 1950, 1981; May 1975, 1981; Bak et al. 1987; Brown 1994b; Brown et al. 2000, 2002; Allen & Holling 2002; Sole´ & Bascompte 2006), but others have suggested that such similarities are spurious (Root 1989). Until such convergences are taken seriously and efforts are made to identify their root causes, community ecology risks becoming a myopic enterprise. There are, in fact, a number of general processes that might influence wealth accumulation across a wide variety of systems. First, and perhaps most importantly, log-normal and mathematically related power law probability distributions are easily generated by the multiplicative rather than additive interaction of variables. Such processes have been shown to commonly apply to many physical (Meijer et al. 1981), biological (May 1975; McGill 2003b) and human systems (Montroll & Shlesinger 1982). Even the log-normal distribution of publication rates for researchers within scientific institutions is governed by such multiplicative processes (Shockley 1957). Similar distributions may also be generated by the interaction of multiple agents governed by nonlinear processes (Bak et al. 1987; West & Shlesinger 1989). Because of these processes, rarity-enriched SAD analogues may be a universal expectation for many complex systems (McGill 2003b). Second, many physical, biological and human systems exist within environments that exhibit power law or fractal-like spatial and temporal variation Figure 3 Distance decay analogues for non-ecological systems, including the change of compositional similarity of global cuisine ingredient lists over space, and garden seed offerings and Cowboy Junkies song performances over time. Best-fit exponential and power law decay lines have been plotted for each relationship. 192 J. C. Nekola and J. H. Brown Idea and Perspective Ó 2007 Blackwell Publishing Ltd/CNRS 180 (Mandelbrot 1982; Milne 1991; Ritchie & Olff 1999; Brown et al. 2002). Not only can these environmental templates be generated via simple hierarchical random models (Sizling & Storch in press), but can also in turn give rise to typical DOW patterns such as the power law SAR and nonlinear DD (Sizling & Storch 2004; Harte et al. 2005). Third, proximity effects, which occur when the influence of an event on an agent is dependent upon the distance in space and/or time between agent and event, occur in many physical, biological and human systems. Such relationships may be found across phenomena as diverse as electromagnetic radiation and gravitation to natural disturbances, dispersal and diffusion to migration, trade and wars. Proximity effects not only can directly give rise to DD relationships (Nekola & White 1999), but also can be an important source of multiplicative relationships over space or time. It is also possible to identify direct analogues in the processes governing different systems. For instance, both ecological and human systems exhibit hierarchical arrangements of agents, including individuals within populations and soldiers within armies and specialized classes of workers within insect societies and manufacturing firms. The structures and dynamics of both ecological and human systems also reflect a complicated dynamic balance among diverse cohesive and divisive forces, such as mutualism, reciprocity, and cooperation on the one hand and selfinterest, competition, and predatory or parasitic exploitation on the other. The joint effects of such factors conspire to ensure that ÔwealthÕ is distributed highly unequally among agents and nonlinearly over space or time. What lessons can be learned from viewing ecological communities as complex systems and ecological DOWs as shared statistical properties of such systems? First, the mechanistic explanation for many of ecology’s most venerable statistical patterns lie at a level of abstraction extending far beyond the realm of ecological process. The paradox that rarity-enriched SADs, power law SAR/STR, and nonlinear DD are generated by models assuming either complete neutrality or resource competition is therefore explained by realizing both models generate dynamical complexity. In competition models complexity arises as multiple agents, each with individualistic resource utilization and dispersal functions, interact in an elaborate network with multiple other agents often over multiple temporal and spatial scales. Such interactions can be made even more complicated and realistic by inclusion of some degree of stochasticity in resource requirements and dispersal processes (Tilman 2004). In neutral models complexity arises as multiple agents, each with a unique community and metacommunity frequency, undergo a lottery for recruitment of vacated spaces. These interactions are further complicated as new species are supplied to the system across multiple scales via dispersal and speciation. As a result it may prove impossible to infer underlying mechanisms from DOW mathematical forms, both in general and for specific cases. However, this should not suggest that efforts to adduce ecological community assembly mechanisms must be abandoned. For instance, while a power law relationship describes the relationship between mean and variance for heartbeat frequency or interstride interval, the slope of this function supplies information about the relative importance of deterministic vs. stochastic drivers (West & Latka 2005; West 2006b). Similar mechanistic inferences can be made in ecological systems. In the case of the SAR, it has long been recognized that differences in slope over local-to-global scales or between islands and mainlands convey information about the mechanisms that generate and maintain diversity (Rosenzweig 1995). Similarly, DD rates provide information about niche characteristics and dispersal capacities in relation to the environmental template of spatial and temporal variations. So, for example, isolated spruce-fir forests of the Appalachians demonstrate an almost threefold greater DD rate when compared with continuous northern Taiga, and large-fruited, more dispersal-limited plant species have almost twice the DD rate when compared with smaller-seeded taxa (Nekola & White 1999). Second, we question the utility of investing great effort to determine whether competition or neutrality might be responsible for observed DOWs when there is no logical reason to expect these mechanisms to be mutually exclusive – or, for that matter, to represent the only possible alternatives. Such patterns are almost certainly generated by combinations of stochastic and deterministic processes, local and regional processes, current and historical events, biotic interactions and abiotic factors, direct and indirect interactions, and cohesive and divisive forces. Similarly, diverse combinations of factors likely generate analogous properties in some physical and many human social and economic systems. Third, it is long overdue that community ecologists keep abreast of developments in the field of complexity science. We are not suggesting that complexity science has all the solutions to the big important questions of community ecology. Indeed, complexity science is struggling to address similar problems and can currently offer few definitive answers. However, we are suggesting that ecological, physical and social scientists have much to learn from each other. Ideas, models and data from community ecology have the potential to make important contributions to complexity science, just as theoretical and empirical advances in complexity science have the potential to enlighten community ecology. This focus on interdisciplinarity is not a revolutionary concept. Community ecology has a long tradition of borrowing ideas and techniques from many other fields including economics (e.g. law of supply and demand, MacArthur & Wilson 1967; game theory, Smith 1982; Gini coefficients, Weiner & Solbrig 1984; supply side Idea and Perspective The wealth of species 193 Ó 2007 Blackwell Publishing Ltd/CNRS 181 ecology, Roughgarden et al. 1987), geography (DD, Nekola & White 1999; gravity models, Bossenbroek et al. 2001); the behavioural sciences (e.g. multidimensional scaling ordination, Minchin 1987), and physics (e.g. diffusion models, Skellam 1951; chaos, May 1976; percolation theory, O’Neill et al. 1992). Finally, complexity science suggests that certain aspects of community ecology may never be highly predictable. While qualitative forecasts may be possible, the precise quantitative prediction of individual events may pose insurmountable challenges. For example, while it might be possible to predict qualitatively how community-wide SADs and SARs will be altered by invasions of multiple exotic species, it will be much more difficult to exactly predict resultant species abundances and spatial distributions. Indeed, such an effort may be as quixotic as attempting to predict the exact location of an electron in a shell or date and magnitude of the next earthquake in the San Francisco Bay area or price of General Motors stock in 2050. As in weather forecasting, some level of shortterm predictability will often be possible from understanding current drivers, recent trajectories, and spatial/temporal autocorrelations. However, predictions based on these dynamics will necessarily become increasingly imprecise as the forecast period increases. Rather than emphasizing prediction, community ecologists perhaps should spend more effort on understanding the mechanisms and events that have conspired to generate current and past patterns. Even when it may be practically impossible to predict the exact future trajectory of a system, by looking backward it may be possible to deduce quite accurately when and how specific mechanisms came into play. In complex human systems mechanisms are deduced by historians who analyse post hoc the particular combination of initial conditions and drivers that generated pattern. Similarly, community ecologists could enhance their understanding of the spatial and temporal patterns of biodiversity by becoming better natural historians, using post hoc analysis to decipher how past events have left lasting influences. By retrospectively studying systems, community ecologists may thus be able to establish what factors lead to the dominance of a particular taxon or to the occurrence of a particular hantavirus outbreak or simply whether Schro¨dinger’s Cat survived. For community ecology to continue making exciting advances, it must recognize that the distribution of abundance among species has much in common with distributions of ÔwealthÕ in many other non-ecological systems. The fact that the distribution of species abundance is very similar to the distribution of drinking glass longevities or Cowboy Junkies song performances is thus both empowering and humbling in equal parts. While such similarities suggest a possible unification between ecological communities and other complex systems, they also suggest that many of ecology’s cherished fundamental patterns may reflect general phenomena that have more in common with statistical physics than species biology. The apparent universality of the rarity-enriched SAD, power law SAR/STR, and nonlinear DD suggests that community ecologists should pay attention to these more fundamental levels of inquiry and explanation. Complexity researchers of all types, including community ecologists, would be well advised to follow Frank Preston’s path and to focus on the common factors that underlie these seemingly ubiquitous distributions of wealth. ACKN O WLEDG EMEN TS We thank Matt Barthel of the Seed Savers Exchange for granting us access to the digital version of their Garden Seed Inventory. We also wish to thank Nicholas Gotelli, three anonymous referees, Linda Fey and current and former members of Brown Lab for helpful feedback regarding earlier versions. This work was supported in part by a National Science Foundation Biocomplexity grant (DEB 0083422) and a grant from the Los Alamos National Laboratory (W-7405-ENG-36). REFEREN CES Allen, C.R. & Holling, C.S. (2002). Cross-scale structure and scale breaks in ecosystems and other complex systems. Ecosystems, 5, 315–318. Bak, P., Tang, C. & Wiesenfeld, K. (1987). Self-organized criticality: an explanation of 1/f noise. Phys. Rev. Lett., 59, 381–384. Bell, G. (2000). The distribution of abundance in neutral communities. Am. Nat., 155, 606–617. Bossenbroek, J.M., Kraft, C.E. & Nekola, J.C. (2001). Prediction of long-distance dispersal using gravity models: Zebra Mussel invasion of inland lakes. Ecol. Appl., 11, 1178–1788. Brown, J.H. (1994a). Organisms and species as complex adaptive systems: linking the biology of populations with the physics of ecosystems. In: Linking Species and Ecosystems (eds Jones, C.G. & Lawton, J.H.). Chapman and Hall, London, pp. 16–24. Brown, J.H. (1994b). Complex ecological systems. In: Complexity: Metaphors, Models, and Reality (eds Cowan, G.A., Pines, D. & Melzer, D.). Santa Fe Institute Studies in the Science of Complexity, Proceedings Volume XVIII. Addison-Wesley, Reading, pp. 419–449. Brown, J.H., West, G.B. & Enquist, B.J. (2000). Scaling in Biology: Patterns and processes, causes and consequences. In: Scaling in Biology (eds Brown, J.H. & West, G.B.). Oxford University Press, Oxford, pp. 1–24. Brown, J.H., Gupta, V.K., Li, B.L., Milne, B.T., Restrepo, C. & West, G.B. (2002). The fractal nature of nature: power laws, ecological complexity, and biodiversity. Proc. R. Soc. Lond. B, Biol. Sci., 357, 619–626. Chave, J. (2004). Neutral theory and community ecology. Ecol. Lett., 7, 241–253. 194 J. C. Nekola and J. H. Brown Idea and Perspective Ó 2007 Blackwell Publishing Ltd/CNRS 182 Chave, J., Muller-Landau, H.C. & Levin, S.A. (2002). Comparing classical community models: theoretical consequences for patterns of diversity. Am. Nat., 159, 1–23. Connor, E.F. & McCoy, E.D. (1979). The statistics and biology of the species–area relationship. Am. Nat., 113, 791–833. Etienne, R.S. & Olff, H. (2005). Confronting different models of community structure to species–abundance data: a Bayesian model comparison. Ecol. Lett., 8, 493–504. Gaston, K.J. & Blackburn, T.M. (2000). Pattern and Process in Macroecology. Blackwell Science, Oxford. Gaston, K.J. & Chown, S.L. (2005). Neutrality and the niche. Funct. Ecol., 19, 1–6. Gaston, K.J., Blackburn, T.M. & Lawton, J.H. (1993). Comparing animals and automobiles: a vehicle for understanding body size and abundance relationships in species assemblages? Oikos, 66, 172–178. Halloy, S. R. P. & Whigham, P.A. (2004). The lognormal as universal descriptor of unconstrained complex systems: a unifying theory for complexity. In: Proceedings of the 7th Asia-Pacific Complex Systems Conference. Cairns Convention Centre, Cairns, Qld, Australia, pp. 309–320. Harte, J., Conlisk, E., Ostling, A., Green, J.L., Smith, A.B. (2005). A theory of spatial structure in ecological communities at multiple spatial scales. Ecol. Monogr., 75, 179–197. Hubbell, S.P. (2001). The unified neutral theory of biodiversity and biogeography. Monographs in Population Biology #32. Princeton University Press, Princeton, NJ. Limpert, E., Stahel, W.A. & Abbt, M. (2001). Log-normal distributions across the sciences: keys and clues. Bioscience, 51, 341– 352. Loehle, C. (1990). Proper statistical treatment of species–area data. Oikos, 57, 143–146. MacArthur, R.H. (1972). Geographical Ecology. Harper and Row, New York, NY. MacArthur, R.H. & Wilson, E.O. (1967). Theory of island biogeography. Monographs in Population Biology #1. Princeton University Press, Princeton, NJ. Mandelbrot, B.B. (1982). The Fractal Geometry of Nature. W.H. Freeman & Co., New York, NY. May, R.M. (1975). Patterns of species abundance and diversity. In: Ecology and Evolution of Communities (eds Cody, M.L. & Diamond, J.). Belknap Press of Harvard University Press, Cambridge, MA, pp. 81–120. May, R.M. (1976). Simple mathematical models with very complicated dynamics. Nature, 261, 459–467. May, R.M. (1981). Patterns in multi-species communities. In: Theoretical Ecology (ed. May, R.M.). Blackwell Scientific, Oxford, pp. 197–227. McGill, B.J. (2003a). A test of the unified neutral theory of biodiversity. Nature, 422, 881–885. McGill, B.J. (2003b). Strong and weak tests of macroecological theory. Oikos, 102, 679–685. Meijer, P.H.E., Mountain, R.D. & Soulen, R.J. Jr (1981). Sixth International Conference on Noise in Physical Systems. National Bureau of Standards Special Publication #614, Washington, DC. Milne, B.T. (1991). Lessons from applying fractal models to landscape patterns. In: Quantitative Methods in Landscape Ecology (eds Turner, M.G. & Gardner, R.H.). Springer-Verlag, New York, NY, pp. 199–235. Minchin, P.R. (1987). An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio, 69, 89–108. Montroll, E.W. & Shlesinger, M.F. (1982). On 1/f noise and other distributions with long tails. Proc. Natl Acad. Sci. USA, 79, 3380– 3383. Mouquet, N. & Loreau, M. (2003). Community patterns in sourcesink metacommunities. Am. Nat., 162, 544–557. Nee, S., Harvey, P.H. & May, R.M. (1991). Lifting the veil on abundance patterns. Proc. R. Soc. Lond. B, Biol. Sci., 243, 161–163. Nekola, J.C. (2005). Latitudinal richness, evenness, and shell size gradients in eastern North American land snail communities. Rec. West. Aust. Mus. Suppl., 68, 39–51. Nekola, J.C. & White, P.S. (1999). Distance decay of similarity in biogeography and ecology. J. Biogeogr., 26, 867–878. Newman, M.E.J. (2005). Power laws, Pareto distributions, and Zipf’s law. Contemp. Phys., 46, 323–351. O’Neill, R.V., Gardner, R.H. & Turner, M.G. (1992). A hierarchical neutral model for landscape analysis. Landscape Ecol., 7, 55–61. Preston, F.W. (1948). The commonness, and rarity, of species. Ecology, 29, 254–283. Preston, F.W. (1950). Gas laws and wealth laws. Sci. Mon., 71, 309– 311. Preston, F.W. (1962a). The canonical distribution of commonness and rarity – I. Ecology, 43, 185–215. Preston, F.W. (1962b). The canonical distribution of commonness and rarity – II. Ecology, 43, 410–432. Preston, F.W. (1980). Noncanonical distributions of commonness and rarity. Ecology, 6, 88–97. Preston, F.W. (1981). Pseudo-lognormal distributions. Ecology, 62, 355–364. Ritchie, M.E. & Olff, H. (1999). Spatial scaling laws yield a synthetic theory of biodiversity. Nature, 400, 557–560. Root, R.B. (1989). Resolution of respect Frank W. Preston 1896– 1989. Bull. Ecol. Soc. Am. 70, 244–247. Rosenzweig, M.L. (1995). Species Diversity in Space and Time. Cambridge University Press, Cambridge, UK. Roughgarden, J., Gaines, S.D. & Pacala, S.W. (1987). Supply-side ecology: the role of physical transport processes. Symp. Br. Ecol. Soc., 27, 491–518. Shockley, W. (1957). On the statistics of individual variations of productivity in research laboratories. Proc. Inst. Radio Eng., 45, 279–290. Sizling, A.L. & Storch, D. (2004). Power-law species-area relationships and self-similar species distributions within finite areas. Ecol. Lett., 7, 60–68. Sizling, A.L. & Storch, D. (in press). Geometry of species distributions: random clustering and scale invariance. In: Scaling Biodiversity (eds Storch, D., Marquet, P.A. & Brown, J.H.), Cambridge University Press, Cambridge. Skellam, J.G. (1951). Random dispersal in theoretical populations. Biometrika, 38, 196–218. Smith, J.M. (1982). Evolution and the Theory of Games. Cambridge University Press, Cambridge. Smith, J. (1990). The Frugal Gourmet on our Immigrant Ancestors. Harper-Collins, New York, NY. Sole´, R.V. & Bascompte, J. (2006). Self-organization in complex ecosystems. Monographs in Population Biology #42. Princeton University Press, Princeton, NJ. Idea and Perspective The wealth of species 195 Ó 2007 Blackwell Publishing Ltd/CNRS 183 Tilman, D. (2004). Niche tradeoffs, neutrality, and community structure: a stochastic theory of resource competition, invasion, and community assembly. Proc. Natl Acad. Sci. USA, 101, 10854– 10861. Ugland, K.I., Gray, J.S. & Lambshead, J.D. (2005). Species accumulation curves analysed by a class of null models discovered by Arrhenius. Oikos, 108, 263–274. Volkov, I., Banavar, J.R., Hubbell, S.P. & Maritan, A. (2003). Neutral theory and relative species abundance in ecology. Nature, 424, 1035–1037. Weiner, J. & Solbrig, O.T. (1984). The meaning and measurement of size hierarchies in plant populations. Oecologia, 61, 334–336. West, B.J. (2006a). Thoughts on modeling complexity. Complexity, 11, 33–43. West, B.J. (2006b). Complexity, scaling and fractals in biological signals. In: Wiley Encyclopedia of Biomedical Engineering (ed. Akay M.). Wiley-Interscience, New York, NY. West, B.J. & Latka, M. (2005). Fractional Langevin model of gait variability. J. Neuroeng. Rehabil., 2, 24. West, B.J. & Shlesinger, M.F. (1989). On the ubiquity of 1/f noise. Int. J. Mod. Phys. B, 3, 795–819. White, E.P., Adler, P.B., Lauenroth, W.K., Gill, R.A., Greenberg, D., Kaufman, D.M. et al. (2006). A comparison of the species– time relationship across ecosystems and taxonomic groups. Oikos, 112, 185–195. Williams, C.B. (1964). Patterns in the Balance of Nature and Related Problems in Quantitative Biology. Academic Press, London. Editor, Nicholas Gotelli Manuscript received 29 August 2006 First decision made 4 October 2006 Manuscript accepted 30 November 2006 196 J. C. Nekola and J. H. Brown Idea and Perspective Ó 2007 Blackwell Publishing Ltd/CNRS 184 185 186 187 188 189 190 191 192 193 194 195 Section VI: Conservation Biology Because of the precarious status of biodiversity across the globe, it is important for ecologists to apply their pure research findings to the development of optimal conservation strategies. As suggested by Susan Bratton, the current state of conservation biology is not unlike that of the medical profession in 1700’s Vienna, when patients faced a lower chance of survival after treatment than if they had let the disease run its course. Because of the complex nature of ecological systems, we should not search for universally applicable and beneficial reserve design or management strategies or seek exact quantitative predictions regarding particular environmental perturbations. Rather, each reserve will face unique series of threats and will require artist-like managers who develop unique management protocols for each site. Much of my conservation biology research has been focused on documenting the unintended negative impacts posed by fire management policies on biodiversity, especially in fragmented landscapes with low potential for recolonization. Total biodiversity maintenance in these situations requires reduced use of fire across both space and time, with sensitive application of large vertebrate grazing and use of more labor intensive techniques such as hand-removal of woody vegetation. I have also published on the importance of distance and colonization history in reserve design, have documented the negative impact of recreational rock climbing and soil disturbance on land snail diversity, and have investigated the impact of powerline corridors on biodiversity. Representative Publications [number of citations as of October 27, 2017] Nekola, J.C. 2012. The impact of utility corridors on terrestrial gastropod biodiversity. Biodiversity and Conservation. 21:781–795. [9] McMillan, M., J.C. Nekola & D.W. Larson. 2003. Impact of recreational rock climbing on land snail communities of the Niagara Escarpment, southern Ontario, Canada. Conservation Biology. 17:616-621. [41] Nekola, J.C. 2002. Effects of fire management on the richness and abundance of central North American grassland land snail faunas. Animal Biodiversity and Conservation. 25:53-66. [46] Nekola, J.C. & P.S. White. 2002. Conservation, the two pillars of ecological explanation, and the paradigm of distance. Natural Areas Journal. 22:305-310. [23] White, P.S. & J.C. Nekola. 1992. Biological diversity in an ecological context. Pages 10-29 in: Barker, J.R. and D.T. Tingey (eds), Air pollution and biodiversity. Van Nostrand Reinhold, New York. [6] 196 ORIGINAL PAPER The impact of a utility corridor on terrestrial gastropod biodiversity Jeffrey C. Nekola Received: 9 May 2011 / Accepted: 21 December 2011 / Published online: 11 January 2012 Ó Springer Science+Business Media B.V. 2012 Abstract Utility corridors are often thought to be disruptive to biodiversity because they cause habitat fragmentation that may lead to increases in predation, parasitism, disease transmittance and vagrant species while decreasing migration rates, gene flow and genetic diversity for interior species. Species with poor dispersal abilities, sedentary lifestyles, and specialized habitats have been thought to be potentially the most vulnerable to these effects. Terrestrial gastropods thus serve as a valuable system in which to investigate these impacts because they are among the poorest active dispersers in the animal kingdom. To document the impact of corridor formation on land snail biodiversity, a 75-year old powerline right-of-way in the eastern Upper Peninsula of Michigan was chosen for analysis. While terrestrial gastropod richness and abundance was significantly reduced for corridor as compared to adjacent control subsamples, with a 2/3 turnover in species composition, the corridor fauna is similar to nearby nativegrasslandsitesintermsof speciescomposition,abundance distribution, andnumbersand abundance of species of conservation concern. The fauna of control subsamples immediately adjacent to the corridor remained similar to other undisturbed sites in the region, with multiple speciesofconservationconcernpersistingatdistancesofonly30 mfromthecorridor.Thus,the netimpactofcorridorgenerationhas beenarguably positive:while the surroundingforest fauna has not been degraded, within the corridorthe reduction of forest species has beencompensated for by establishment of even rarer grassland species. Keywords Land snail Á Right-of-way Á Landscape Á Community ecology Á Community structure Á Soil architecture Introduction Utility corridors have often been thought to be disruptive to biodiversity (Forman 1986; Andrews 1990) because they fragment habitats, leading to increased predation, parasitism and disease transmittance rates (Simberloff and Cox 1987; Yahner et al. 1989; Bennett J. C. Nekola (&) Department of Biology, University of New Mexico, Albuquerque, NM 87131, USA e-mail: jnekola@unm.edu 123 Biodivers Conserv (2012) 21:781–795 DOI 10.1007/s10531-011-0216-8 Author's personal copy 197 1991), decreased migration abilities and lowering of gene flow and within-population genetic diversity (Fahig and Mirriam 1985; Goosem and Marsh 1997; Gerlach and Mustov 2000), and expansion of vagrant species at the expense of interior species (Arnold et al. 1987). Species with poor dispersal abilities, sedentary habits, and specialized habitat preferences have been thought to be the most vulnerable to these effects (Shaffer and Samson 1985; Andrews 1990). Terrestrial gastropods would therefore appear particularly sensitive to utility corridor formation. The North American fauna, particularly at mid-to-high latitudes, is dominated by species with shells \5 mm in maximum dimension that represent approximately 50–80% of regional species richness and 80–95% of individuals (Nekola 2005). Such species have very poor active dispersal abilities: individuals move perhaps only 1–10 m over a lifetime (Schilthuizen and Lombaerts 1994; Hausdorf and Hennig 2003). Even active movement of larger species (maximum shell dimension [20 mm) has been effectively inhibited by barriers of no more than 3–8 m (Baur 1988; Baur and Baur 1990; Schilthuizen and Lombaerts 1994). Also, terrestrial gastropods exhibit a high degree of habitat specialization, with order-of-magnitude changes in richness and abundance over as little as 1 m (Nekola and Smith 1999). Even though utility corridors may have serious negative impacts on terrestrial gastropod biodiversity and community structure, no prior research has specifically addressed this issue. This study therefore investigates changes in community structure and composition between paired corridor and matrix habitats along a 75-year old utility corridor passing through the Hiawatha National Forest in the eastern Upper Peninsula of Michigan. Comparisons were also made with other undisturbed habitats in the surrounding landscape. Materials and methods Study system ESE-6904/6905 is a 42 km powerline operated by the American Transmission Company in Chippewa and Mackinac counties in the eastern Upper Peninsula of Michigan. The line connects St. Ignace in the south to the Pine River Substation near Rudyard in the north. The corridor was initially constructed in the 1930s and is 30–60 m in width. The line currently runs at 69,000 volts with approximately 395 double-circuit wood pole structures. In Mackinac County a considerable proportion of the corridor passes though the St. Ignace Ranger District of the Hiawatha National Forest, where it has been maintained by mowing and mechanical woody plant removal. In total, 58 land snail species have been previously identified from the eastern Upper Peninsula (Hubricht 1985; Nekola 2004). Of these, 13 are listed as of conservation concern by the Michigan Natural Features Inventory (2011): Catinella exile—Threatened; Euconulus alderi—Threatened; Planogyra asteriscus—Special Concern; Pupilla muscorum— Special Concern; Vallonia gracilicosta—Endangered; Vertigo bollesiana—Threatened; Vertigo cristata—Special Concern; Vertigo elatior—Special Concern; Vertigo hubrichti— Endangered; Vertigo morsei—Endangered; Vertigo nylanderi—Endangered; Vertigo paradoxa—Special Concern and Vertigo pygmaea—Special Concern. Six species (Cochlicopa lubrica, Oxychilus draparnaudi, Pupilla muscorum, Vallonia costata, Vallonia pulchella, Vertigo pygmaea) appear to be Eurasian exotics based on their preference for anthropogenic habitats, known North American history and/or population genetics (Pilsbry 1948; Nekola and Coles 2010). 782 Biodivers Conserv (2012) 21:781–795 123 personal copy 198 Sampling design Eighteen sample sites were identified along the ESE-6904/6905 route as it passes through the Hiawatha National Forest in eastern Mackinac County, Michigan. Sample sites were spread across the entire corridor route, were separated by at least 0.5 km, and represented all natural habitats crossed by the corridor. These included marl ponds, fens, sedge meadows, upland and lowland Northern White Cedar stands, sandy pine forest, mixed deciduous forest, and rocky slopes. Sample sites were located in conjunction with pole numbers 505, 520, 567, 600, 605, 623, 652, 680, 696, 705, 717, 726, 728, 755, 772, 790, 811 and 813 (Fig. 1). At most sites three subsamples were collected: one within the existing corridor, and two controls located within intact vegetation 30 m to both the west and east of the corridor. No private lands were surveyed. For this reason only the corridor and west margin subsamples were collected at Pole 811 and only the corridor subsample was collected at Pole 600. Also, because of abundance of Poison Ivy and lack of leaf litter accumulation within the corridor, only the western margin at Pole 813 was sampled. Comparisons were also made with 32 additional sites in the surrounding landscape, including 11 wooded bedrock outcrops, 5 upland forests, 7 lowland forests, 8 lowland grasslands and a single upland grassland. These capture the full compositional range of regional land snail assemblages (Nekola 2003, 2004, 2005; Fig. 1). Field methods The land snail assemblage of each 100 m2 subsample was documented from September 25–30, 2009 through hand collection of larger shells and litter sampling for smaller taxa following protocols of Cameron and Pokryszko (2005) and Oggier et al. (1998). Fig. 1 Map of the eastern Upper Peninsula of Michigan showing location of the 18 sample sites analyzed along the ESE-6904/6905 corridor along with locations of the 32 additional comparison sites Biodivers Conserv (2012) 21:781–795 783 123 Author's personal copy 199 Approximately 30 min were spent collecting each subsample. Soil litter collections followed protocols of Nekola and Coles (2010), and emphasized microsites known to support high micro-snail densities (e.g., Emberton et al. 1996). Approximately 200 ml of litter were collected per subsample. This generally captured ten times as many individuals as species, which is advocated by Cameron and Pokryszko (2005) for accurate land snail community documentation. The latitude-longitude position of each pole location was determined using a Garmin 12XL hand-held GPS unit. Sampling of comparison sites elsewhere in the eastern Upper Peninsula was conducted from July 1997–October 2009 in a similar fashion (see Nekola 2003, 2004, 2005) although at an order of magnitude larger spatial sampling grain. Any systematic bias would thus result in underestimating the relative richness of ESE-6904/6905 subsamples through random undersampling of the rarest species. However, this will not obscure community composition gradients (Jobe 2007). Laboratory procedures As outlined by Nekola and Coles (2010), dried litter samples were passed through a standard sieve series and handpicked against a neutral-brown background. All shells and shell fragments were removed. All shells were assigned to species (or subspecies) using the author’s reference collection, with the total numbers of shells per species per site being recorded. The number of unassignable immature individuals and fragmentary shells was recorded. Nomenclature is based principally on Hubricht (1985) with updates by Turgeon et al. (1998), Nekola (2004) and Nekola and Coles (2010). Data summary and statistical analysis Species richness and abundance For each species, the total number of subsamples in which it was recorded and the total number of encountered individuals was calculated. Using the non-parametric Kruskal– Wallis Rank Sum Test, these values were compared among the west of corridor, corridor, and east of corridor subsamples, and were graphically illustrated using boxplots. To better visualize the upper and lower bounds, boxplots were constructed using a natural-log scale. However, raw values were used for statistical testing. Variation in community structure among the three subsample positions was illustrated by dominance–diversity curves generated from median proportional abundance values calculated within each of the three subsample locations. Median values were used because they are more robust descriptors of central tendency, being less influenced by outliers as compared to mean values. Variation in richness and abundance of three soil architecture preference classes (duff specialist, generalist, turf specialist) among the three subsample positions was also statistically tested using the Kruskal–Wallis Rank Sum Test and illustrated with boxplots. As before, abundance distributions were visualized using natural-log scaled box-plots, but tested using raw data. The soil architecture nomenclature and preferences of Nekola (2003) were used, with ‘duff specialists’ being species that favor soils with loose organic horizons and ‘turf specialists’ being those that favor organic horizons that are tightly bound with living plant roots. The only exception was Striatura exigua, which is here considered a ‘generalist’ rather than a ‘turf’ species (Nekola 2003) based on additional eastern North American data (Nekola 2010). Differences in richness and abundance for species of conservation concern 784 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 200 and alien species among the three subsample positions was also quantified using the Kruskal–Wallis Rank Sum Test. Community composition The impact of the corridor on community composition was analyzed through global nonmetric multidimensional scaling (NMDS) using DECODA (Minchin 1990). NMDS makes no assumptions regarding the underlying nature of species distributions along compositional gradients, and is thus the most robust form of ordination for detection of ecological patterns (Minchin 1987). All 49 ESE-6904/6905 subsamples in addition to the 32 additional regional comparison sites were included in this analysis. To ordinate sites a matrix of dissimilarity coefficients was calculated based on species abundance data using the Czekanowski index (Faith et al. 1987). Data were untransformed to avoid mathematical artifacts (Nekola et al. 2008). NMDS in one through four dimensions was then performed with 200 iterations, a stress ratio stopping value of 0.9999, and a small stress stopping value of 0.01. Output was scaled in half-change units, so that points in the resulting diagram separated by a distance of 1.0 will correspond, on average, to a 50% turnover in species composition. Because a given NMDS run may locate a local, rather than the global, stress minimum, solution stability was assessed via 50 NMDS runs using different random initial starting points (Minchin 1987). Solutions in each of the four dimensions were compared using a Procrustes transformation to identify those that were statistically similar. The number of unique solutions and number of individual runs that fell into each was then calculated. The global optimum solution was identified as the smallest stress solution that was achieved in a plurality of starts. To assess the impact of corridor construction on community composition, a vector was generated using a given control subsample as the origin and the adjacent corridor subsample as the endpoint. Each of these vectors was plotted on the ordination diagram and its length and angle from vertical calculated. This allowed a median vector of compositional change to be determined across the entire dataset. The significance of these changes in ordination space was estimated using the non-parametric paired Wilcoxon Rank Sum Test. Fisher Exact contingency table tests were used to identify those species that were statistically more abundant within the corridor or control subsamples as compared to a null hypothesis of uniform distribution. Analysis was limited to paired subsamples (i.e. 600C and 813W were eliminated). Species with fewer than ten total occurrences were not analyzed due to the likelihood of Type-II errors. Because this test was repeated for each species, a Bonferroni correction was used to modify the significance threshold. The results of this analysis were then compared to the results of the duff vs. turf soil preference tests of Nekola (2003) by use of a 3 9 3 contingency table. This table was compared to uniform expectations using a Fishers exact test. Results In total, 44 species were observed from the 49 subsamples collected along the ESE-6904/ 6905 corridor (Table 1). Of these, five (Gastrocopta pentodon, Oxychylus draparnaudi, Vertigo bollesiana, Vertigo ovata and Vertigo pygmaea) are reported here for the first time from Mackinac county (Hubricht 1985; Nekola 2004), raising the total county fauna to 54 species. Nine of the thirteen species of conservation concern known from the region Biodivers Conserv (2012) 21:781–795 785 123 Author's personal copy 201 Table 1 Number of occupied subsamples and individuals for all encountered terrestrial gastropod species Species name Soil architecture preference Number of subsamples/individuals West of corridor Powerline corridor East of corridor Anguispira alternata D 2/3 1/1 1/2 Carychium exiguum T 9/293 12/543 7/292 Carychium exile D 6/99 3/19 7/88 *Cochlicopa lubrica D 4/93 3/18 1/1 Columella simplex D 14/127 10/49 16/167 Deroceras spp. G 4/5 7/12 6/9 Discus catskillensis D 9/167 4/5 12/109 Euchemotrema fraternum D 0/0 1/2 1/1 Euconulus alderi T 6/26 8/57 5/19 Euconulus fulvus D 8/31 2/10 9/75 Euconulus polygyratus D 3/16 2/9 2/4 Gastrocopta pentodon D 1/1 1/33 0/0 Gastrocopta tappaniana T 2/25 7/101 3/24 Glyphyalinia indentata D 1/1 1/1 1/1 Helicodiscus parallelus G 0/0 1/2 0/0 Helicodiscus shimeki D 3/11 1/2 9/54 Nesovitrea binneyana D 8/102 0/0 9/144 Nesovitrea electrina T 4/28 10/158 4/43 *Oxychylus draparnaudi G 0/0 1/3 0/0 Oxyloma retusa T 1/1 4/6 0/0 Paravitrea multidentata D 1/8 0/0 0/0 Planogyra asteriscus G 8/814 2/7 8/603 Punctum minutissimum D 16/1,653 7/141 16/1,371 Punctum n.sp. T 6/84 8/410 6/70 *Pupilla muscorum G 1/9 1/8 0/0 Striatura exigua G 13/248 9/195 15/351 Striatura ferrea G 9/176 3/52 12/102 Striatura milium G 12/332 6/65 14/334 Strobilops labyrinthica D 15/1,279 14/275 16/993 Succinea ovalis D 0/0 3/4 0/0 *Vallonia costata G 2/55 1/5 0/0 *Vallonia pulchella T 0/0 2/4 0/0 Vertigo bollesiana D 6/15 0/0 6/14 Vertigo cristata D 7/33 2/20 5/19 Vertigo elatior T 5/42 11/118 7/17 Vertigo gouldii D 2/3 0/0 2/7 Vertigo morsei T 1/9 4/48 0/0 Vertigo nylanderi T 2/24 0/0 2/8 Vertigo ovata T 0/0 3/3 1/1 *Vertigo pygmaea G 1/4 2/2 0/0 Vitrina limpida G 2/8 2/7 0/0 Zonitoides arboreus D 6/14 3/5 5/9 786 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 202 (Euconulus alderi, Planogyra asteriscus, Pupilla muscorum, Vertigo bollesiana, Vertigo cristata, Vertigo elatior, Vertigo morsei, Vertigo nylanderi, and Vertigo pygmaea) were observed. In total, 14, 242 individuals were collected, of which 13,185 were identifiable to species. Thirty-one of the subsamples included at least 200 individuals. The 32 eastern Upper Peninsula comparison sites ranged in richness from 3 to 25 species and in abundance from 19 to 1,480, with a total of 12,831 individuals being identified to species. Impact on species richness and abundance Based on paired observations, both species richness and abundance were lower within the corridor than in adjacent control subsamples (Fig. 2): median richness fell by approximately  (from 14 to 10; P = 0.015) and abundance by approximately ’ (from 400 to 100; P \ 0.00003). However, dominance-diversity curves demonstrated that median proportional abundance was very similar for up to approximately the 7th most abundant species (Fig. 3), after which median abundance fell off more rapidly for the corridor subsamples. Comparison of richness and abundance across species with differing soil architecture preferences (Fig. 4) demonstrates that control subsamples were dominated by duff Fig. 2 Boxplots showing variation in terrestrial gastropod richness and abundance among the three subsample positions along the ESE-6904/6905 corridor. Abundance data are presented along a log-scaled axis. The central line represents the median of the sample, the margins of each box represent the interquartile distances, and the fences represent 1.5 times the interquartile distances. Outliers falling outside of the fences are shown with circles Table 1 continued Species name Soil architecture preference Number of subsamples/individuals West of corridor Powerline corridor East of corridor Zonitoides nitidus T 1/4 0/0 0/0 Zoogenetes harpa D 2/4 1/4 1/2 Total terrestrial gastropods 16/5,847 17/2,404 16/4,934 Species underlined are species of conservation concern in Michigan (Michigan Natural Features Inventory 2011). Species preceded by an asterisk are Eurasian aliens. Soil architecture preference is based on Nekola (2003) with minor revisions as per text; D Duff specialist, T turf specialist, G generalist Biodivers Conserv (2012) 21:781–795 787 123 Author's personal copy 203 Fig. 3 Dominance-diversity curves demonstrating proportional abundance versus rank order of terrestrial gastropod species for the three subample positions along the ESE-6904/6905 corridor Fig. 4 Boxplots showing variation in duff specialist, generalist, and turf specialist terrestrial gastropod species richness and abundance among the three subsample categories along the ESE-6904/6905 corridor. Abundance data are presented along a log-scaled axis. The central line represents the median of the sample, the margins of each box represent the interquartile distances, and the fences represent 1.5 times the interquartile distances. Outliers falling outside of the fences are shown with circles 788 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 204 specialists (*57% of taxa and *80% of individuals), followed by generalists (*27% and *18%) and turf specialists (*16% and *2%; Fig. 4). However, corridor subsamples were dominated by turf specialists (*50% and *63%) followed by duff specialists (*30% and *23%), and generalists (*20% and *14%). While these differences were statistically significant for duff and turf specialists (P = 0.016 and P = 0.0020, respectively, for richness, and P = 0.0095 and P = 0.0037 for individuals), generalist species did not differ significantly (P = 0.1054 for richness and P = 0.0948 for individuals). No significant differences were noted in richness (P = 0.7842) or abundance (P = 0.7350) of species of conservation concern or in exotic species richness (P = 0.2095) or abundance (P = 0.2124) among corridor and adjacent control subsamples. Impact on community composition NMDS along two dimensions was chosen as the most robust solution, given that all five minimum stress configurations fell into a single group. The 1-, 3-, and 4-dimensional solutions were not considered as they either possessed larger stress values or had no single modal solution. The chosen ordination documents a dramatic alteration of terrestrial gastropod community composition within the corridor (Fig. 5). While control subsample assemblages remain similar to other regional lowland conifer forest, upland forest, and lowland grassland assemblages, corridor faunas tend to be displaced down and/or left in the diagram, being similar to the faunas of native lowland grassland sites (e.g., sedge meadows, fens). Displacement vectors from a given control subsample to its paired corridor subsample ranged from 0.18 to 3.4, or 9% to 170% change in composition. Across all pairs, the median distance was 1.27 (indicating roughly a 2/3 turnover in composition) at -110° from the top of the diagram. This displacement was highly significant (P \ 0.00002) along both axes. Fig. 5 NMDS ordination of terrestrial gastropod faunas from the 49 ESE-6904/6905 subamples and 32 comparison sites. Axes are scaled in half-change units. a Optimum two-dimensional NMDS solution. Powerline corridor subamples are represented by hatched diamonds; control subamples are indicated by a circle surrounding the appropriate habitat icon; uncircled habitat icons represent comparison sites. b Vectors demonstrating the change in composition between control (origin) and adjacent corridor (endpoint) samples Biodivers Conserv (2012) 21:781–795 789 123 Author's personal copy 205 Following removal of unpaired subsamples, fifteen species were eliminated from further analysis because they were represented by ten or fewer individuals. Fisher exact tests of the remaining 29 species (Table 2) demonstrated that eight either significantly favored (Bonferroni corrected P-threshold = 0.0017) or tended to favor (0.05 [ P [0.0017) corridor subsamples, 14 significantly favored the adjacent control subsamples, while the remaining six were generalists (0.05 \ P). While three species of conservation concern favored control subsamples (Planogyra asteriscus, Vertigo bollesiana, Vertigo nylanderi) an additional three favored corridor subsamples (Euconulus alderi, Vertigo elatior, Vertigo morsei) with two more being generalists (Pupilla muscorum, Vertigo cristata). Although Vertigo pygmaea was not statistically analyzed because not enough individuals were encountered, 2/3 of its occurrences were within the corridor. Comparison of these responses to known soil architecture preferences (Nekola 2003) indicates that the two factors are significantly linked (P = 0.0154; Table 3): species that favored corridor subsamples tended to favor turf soils, while species favoring control subsamples tended to favor duff soils. Discussion The sampling protocols used in this study were found to be more than adequate to accurately characterize the land snail fauna and its response to powerline corridor formation. First, roughly 80% of the known county and 75% of the eastern Upper Peninsula fauna was encountered despite the fact that the corridor does not intersect the carbonate cliff sites that serve as one of the most important regional reservoirs for land snail biodiversity (Nekola 1999). This sampling design also resulted in documenting well the regional species of conservation concern. Only four such species were not encountered: Vallonia gracilicosta, Vertigo hubrichti and Vertigo paradoxa were not observed because their stable carbonate cliff habitat was not intersected by the ESE-6904/6905 corridor; and Catinella exile was also not recorded even though it has been observed in low numbers on the marl fen at Summerby Swamp to the northwest of the corridor. While seemingly appropriate habitat for this species exists within the ESE-6904/6905 corridor and adjacent habitats, its absence was not surprising as it is also absent from the immediately adjacent Martineau Creek fen (Nekola 1998). Additionally, both the entire dataset as well as ’ of the individual subsamples meet the criteria of Cameron and Pokryszko (2005) for accurate portrayal of land snail community patterns. These data show that creation of a treeless powerline corridor has led to profound changes in the terrestrial gastropod community. Both the richness and abundance of land snails within the corridor have fallen in comparison to adjacent habitats, with composition turning over by roughly 2/3 as turf (i.e., grassland) specialists replace duff (i.e., forest) specialists. This change, however, has not led to a reduction in total land snail biodiversity. First, faunas of control subsamples remain compositionally similar to those of other undisturbed sites in the region, with multiple species of conservation concern persisting only 30 m from the corridor edge. Second, species abundance distributions were generally unaffected, with proportional abundance being similar for corridor and control subsamples up to the seventh-most abundant species. Although abundance levels trailed off more rapidly for corridor subsamples from the eighth-most-abundant species on, this general pattern is characteristic for native grassland faunas throughout the region (Nekola 2002, 2010). As a result, the more rapid fall off of abundance is probably not due to the corridor, per se, but 790 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 206 rather indirectly to the conversion to grassland habitats/faunas. Third, ordination demonstrates that within 75 years the corridor fauna has evolved into assemblages very similar to other native lowland grassland sites in the region, in particular sedge meadows and fens. Table 2 Impact of powerline corridor on species abundance Name Corridor samples Control samples P value Obs. Exp. Obs. Exp. A. Species with too few observations (10 or less) to characterize preference Anguispira alternata; Euchemotrema fraternum; Glyphyalinia indentata; Helicodiscus parallelus; Oxychilus draparnaudi; Oxyloma retusa; Paravitrea multidentata; Succinea ovalis; Vallonia costata; Vallonia pulchella; Vertigo gouldii; Vertigo ovata; Vertigo pygmaea; Vitrina limpida; Zoogenetes harpa B. Generalist species Name Corridor Subamples Control Subamples P value Obs. Exp. Obs. Exp. Deroceras sp. 12 9 14 17 0.572488900 Euconulus polygyratus 9 10 20 19 1.000000000 Punctum n.sp. 85 81 154 158 0.773239300 Pupilla muscorum 8 6 9 11 0.728282100 Vertigo cristata 20 25 52 47 0.472278900 Zonitoides arboreus 5 10 23 18 0.226957500 C. Species that favor (P \ 0.0017) or tend to favor (0.05 [ P [ 0.0017) corridor subsamples Carychium exiguum 465 357 585 693 0.000001676 Cochlicopa lubrica 18 8 6 16 0.008414732 Euconulus alderi 48 32 45 61 0.026030660 Gastrocopta pentodon 33 12 1 22 0.000000047 Gastrocopta tappaniana 95 49 49 95 0.000000090 Nesovitrea electrina 151 76 71 146 0.000000000 Vertigo elatior 118 60 59 117 0.000000001 Vertigo morsei 48 19 9 38 0.000000047 D. Species that favor (P \ 0.0017) or tend to favor (0.05 [ P [ 0.0017) control subsamples Carychium exile 19 70 187 136 0.000000001 Columella simplex 49 116 293 226 0.000000002 Discus catskillensis 5 96 276 185 0.000000000 Euconulus fulvus 10 39 106 77 0.000004013 Helicodiscus shimeki 2 23 65 44 0.000002788 Nesovitrea binneyana 0 84 246 162 0.000000000 Planogyra asteriscus 7 485 1,417 939 0.000000000 Punctum minutissumum 141 1,028 2,879 1,992 0.000000000 Striatura exigua 195 270 599 524 0.000043710 Striatura ferrea 52 112 278 218 0.000000083 Striatura milium 65 249 666 482 0.000000000 Strobilops labyrinthica 275 867 2,271 1,679 0.000000000 Vertigo bollesiana 0 10 29 19 0.000767735 Vertigo nylanderi 0 11 32 21 0.000347028 Species considered of conservation concern by the Michigan Natural Features Inventory are underlined Biodivers Conserv (2012) 21:781–795 791 123 Author's personal copy 207 The ultimate impact of corridor formation has been for the expansion of lowland grassland land snail faunas into the previously largely forested matrix. Fourteen species (*30% of total) were found to favor control subsamples, while eight (*20%) favored the corridor itself. These changes are similar to those previously noted with forest clearing (Nekola 2003). Because species of conservation concern in this region are roughly equally partitioned between forests and open habitats, the decrease in population occurrence/size of forest-favoring rare species (Planogyra asteriscus, Vertigo bollesiana, Vertigo nylanderi) in the corridor has been compensated by the increase of grassland-favoring rare species (Euconulus alderi, Vertigo elatior, Vertigo morsei), with Vertigo morsei representing one of the rarest land snail species in the Great Lakes region (Nekola 2004). As a result, no significant change in either the total richness or abundance of species of conservation concern was noted between corridor and adjacent forest habitats. However, the increase in coverage and population sizes of grassland species that would otherwise be rare or absent from the corridor has enriched total biodiversity. This result holds even when the two alien species listed by the Michigan Natural Features Inventory (Pupilla muscorum, Vertigo pygmaea) are removed from analysis. Even though terrestrial gastropods appear to be ideal candidates for documenting detrimental biodiversity impacts from corridor formation, these data demonstrate that systemwide diversity has not been lowered by corridor formation. Rather, the net impact appears positive, as the corridor has allowed for an expansion of grassland/turf specialists, including one of the rarest species in the landscape, while at the same time not seriously impacting the faunas of adjacent habitats. How have small grassland snails with poor active dispersal abilities been able to occupy the new grassland habitats created by corridor formation within the relatively short ecological time frame that the ESE-6904/6905 corridor has been in existence? The explanation is probably that small land snails, such as those that dominate the Upper Peninsula fauna, possess great passive dispersal capabilities. For example, DNA sequence data demonstrate that members of the genus Balea have been repeatedly carried across 9,000 km of open eastern Atlantic Ocean by migrating birds (Gittenberger et al. 2006). Similarly, in North America many members of the Vertigo gouldii group have ranges exceeding 5,000 km in extent even though they inhabit areas covered by continental ice as recently as 10 ka (Nekola et al. 2009). Passive dispersal of some small land snail species is also facilitated by their ability to reproduce uniparentally (Pokryszko 1987), allowing only single individuals to found populations. As a result, small land snails tend to possess much larger ranges and to more completely saturate their available habitats than larger taxa (McClain and Nekola 2008; Nekola 2009). Table 3 Species response to subsamples placement versus duff/turf soil architecture preference, with numbers in each cell representing number of species Soil architecture preference Duff species Generalist species Turf species Corridor preference Control subsamples 9 4 1 Generalist 3 1 1 Corridor subsamples 2 0 6 Fisher’s exact test: P = 0.0154 792 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 208 The most likely vectors for passive land snail dispersal in the ESE-6904/6905 corridor are birds and small mammals. For both of these groups, utility corridors represent travel paths for grassland species through forested landscapes (Kroodsma 1982; Gates 1991). This will allow for movement of grassland/turf specialist land snails through corridors as well. Additionally, narrow corridors such as that for ESE-6904/6905 have been found to not restrict movements of small mammals in eastern North American forests (Schreiber and Graves 1977; Gates 1991), and gliding mammals (Asari et al. 2010), small mammals, reptiles, and amphibians (Carthew et al. 2009) in Australia. Consequently, the ESE-6904/ 6905 corridor probably does not serve as a barrier to passive movements of minute forest snails. It should thus not be surprising that the richness, abundance, and community structure of land snail assemblages on either side of the corridor were found to be very similar. These results mirror numerous studies that have shown that utility corridors serve as refuges for grassland species within forested landscapes. This effect has been repeatedly shown for birds (Kroodsma 1982; King et al. 2009; Meehan and Haas 1997) and mammals (Gates 1991) in eastern North America and for a variety of small vertebrates in Australia (Goosem and Marsh 1997; Clarke et al. 2006; Clarke and White 2008a, b). As was found in the current study, these grassland species may be quite uncommon elsewhere within the landscape (Clarke and White 2008b). Similar patterns exist for vascular plants in the southeastern USA, where powerline corridors serve as refuges in an otherwise forested matrix for a number of critically endangered grassland/savanna species (Sorrie and Weakley 2001). Thus, as long as a given utility corridor (1) does not substantially impact the overall population of a given forest-specialist species, (2) exists within a forest-dominated landscape that also includes native grassland habitats and (3) is maintained through manual vegetation cutting, it may actually benefit terrestrial gastropod biodiversity by maintaining refuges for both common and rare grassland species while not leading to the extirpation of sensitive forest species. For the current alignment of ESE-6904/6905 this equates to roughly 3–6 ha of grassland habitat per kilometer of corridor. However, it should not be assumed that these positive benefits will also accrue in landscapes lacking native grassland habitats, or for corridors maintained through herbicide application (Nowak et al. 1993). Although quantitative sampling for land snails was not conducted in areas subjected to this type of management, it appeared that terrestrial gastropod richness and abundance was much lower in herbicide-managed sections of the corridor just outside of the National Forest boundary. Acknowledgments Thanks are extended to Todd Miller and the American Transmission Company for facilitating access to the study site and for underwriting field work and analysis expenses. Joe Pagliara of Natural Resources Consulting, Inc./Stantec helped interface with ATC and also assisted with field work. Melinda Fey, Robert Cowie, and two anonymous reviewers provided invaluable comments on earlier manuscript drafts. References Andrews A (1990) Fragmentation of habitat by roads and utility corridors: a review. Aust Zool 26:130–141 Arnold GW, Algar D, Hobbs RJ, Atkins L (1987) A survey of vegetation and its relationship to vertebrate fauna present in winter on road verges in the Kellerberrin District, Western Australia. Technical Report 18, Department of Conservation and Land Management, Western Australia Asari Y, Johnson CN, Parsons M, Larson J (2010) Gap-crossing in fragmented habitats by mahogany gliders (Petaurus gracilis). Do they cross roads and powerline corridors? Aust Mammol 32:10–15 Biodivers Conserv (2012) 21:781–795 793 123 Author's personal copy 209 Baur B (1988) Microgeographical variation in shell size of the land snail Chondrina clienta. Biol J Linn Soc 35:247–259 Baur A, Baur B (1990) Are roads barriers to dispersal in the land snail Arianta arbustorum? Can J Zool 68:613–617 Bennett AF (1991) Roads, roadsides, and wildlife conservation: a review. In: Saunders DA, Hobbs RJ (eds) Conservation 2: the role of corridors. Surrey Beatty & Sons, Chipping Norton Cameron RAD, Pokryszko BM (2005) Estimating the species richness and composition of land mollusc communities. J Conchol 38:529–547 Carthew SM, Horner B, Jones KMW (2009) Do utility corridors affect movements of small terrestrial fauna? Wildl Res 36:488–495 Clarke DJ, White JG (2008a) Towards ecological management of Australian powerline corridor vegetation. Landsc Urban Plan 86:257–266 Clarke DJ, White JG (2008b) Recolonization of powerline corridor vegetation by small mammals: timing and influence of vegetation management. Landsc Urban Plann 87:108–116 Clarke DJ, Pearce KA, White JG (2006) Powerline corridors: degraded ecosystems or wildlife havens? Wildl Res 33:615–626 Emberton KC, Pearce TA, Randalana R (1996) Quantitatively sampling land-snail species richness in Madagascan rainforests. Malacologia 38:203–212 Fahig L, Mirriam G (1985) Habitat patch connectivity and population survival. Ecology 66:1762–1768 Faith DP, Minchin PR, Belbin L (1987) Compositional dissimilarity as a robust measure of ecological distance. Vegetatio 69:57–68 Forman RTT (1986) Landscape ecology. Wiley, New York Gates JE (1991) Powerline corridors, edge effects, and wildlife in forested landscapes of the central Appalachians. In: Rodiek JE, Bolen EG (eds) Wildlife and habitats in managed landscapes: an overview. Island Press, Washington Gerlach G, Mustov K (2000) Fragmentation of landscape as a cause for genetic subdivision in bank voles. Conserv Biol 14:1066–1074 Gittenberger E, Groenenberg DSJ, Kokshoorn B, Preece RC (2006) Molecular trails from hitch-hiking snails. Nature 439:409 Goosem M, Marsh H (1997) Fragmentation of a small mammal community by a powerline corridor through tropical rainforest. Wildl Res 24:613–629 Hausdorf B, Hennig C (2003) Nestedness of north-west European land snail ranges as a consequence of differential immigration from Pleistocene glacial refuges. Oecologia 35:102–109 Hubricht L (1985) The distributions of the native land mollusks of the eastern United States. Fieldiana New Ser 24:1–191 Jobe RT (2007) Estimating landscape-scale species richness: reconciling frequency- and turnover-based approaches. Ecology 89:174–182 King DI, Chandler RB, Collins JM, Petersen WR, Lautzenheiser TE (2009) Effects of width, edge and habitat on the abundance and nesting success of scrub-shrub birds in powerline corridors. Biol Conserv 142:2672–2680 Kroodsma RL (1982) Edge effect on breeding forest birds along a power-line corridor. J Appl Ecol 19:361–370 McClain C, Nekola JC (2008) The role of local-scale on terrestrial and deep-sea Gastropod body size distributions across multiple scales. Evol Ecol Res 10:129–146 Meehan AL, Haas CA (1997) Use of a powerline corridor by breeding and wintering birds in Giles County, Virginia. Va J Sci 48:259–264 Michigan Natural Features Inventory (2011) Michigan’s Special Animals. http://web4.msue.msu.edu/mnfi/ data/specialanimals.cfm. Accessed 7 Mar 2011 Minchin PR (1987) An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio 69:89–108 Minchin PR (1990) DECODA software. Australian National University, Canberra Nekola JC (1998) Terrestrial gastropod inventory of the Niagaran escarpment and Keweenaw volcanic belt in Michigan’s Upper Peninsula. Final Report, Michigan Department of Natural Resources Small Grants Program, Lansin Nekola JC (1999) Terrestrial gastropod richness of carbonate cliff and associated habitats in the Great Lakes region of North America. Malacologia 41:231–252 Nekola JC (2002) Effects of fire management on the richness and abundance of central North American grassland land snail faunas. Animal Biodivers Conserv 25:53–66 Nekola JC (2003) Large-scale terrestrial gastropod community composition patterns in the Great Lakes region of North America. Divers Distrib 9:55–71 794 Biodivers Conserv (2012) 21:781–795 123 Author's personal copy 210 Nekola JC (2004) Terrestrial gastropod fauna of northeastern Wisconsin and the southern Upper Peninsula of Michigan. Am Malacol Bull 18:21–44 Nekola JC (2005) Latitudinal richness, evenness, and shell size gradients in eastern North American land snail communities. Rec West Aust Mus Suppl 68:39–51 Nekola JC (2009) Big ranges from small packages: North American vertiginids more widespread than thought. Tentacle 17:26–27 Nekola JC (2010) Acidophilic terrestrial gastropod communities of North America. J Mollus Stud 76:144–156 Nekola JC, Coles BF (2010) Pupillid land snails of eastern North America. Am Malacol Bull 28:29–57 Nekola JC, Smith TA (1999) Terrestrial gastropod richness patterns in Wisconsin carbonate cliff communities. Malacologia 41:253–269 Nekola JC, Sˇizling AL, Boyer AG, Storch D (2008) Artifactions in the log-transformation of species abundance distributions. Folia Geobot 43:259–468 Nekola JC, Coles BF, Bergthorsson U (2009) Evolutionary pattern and process in the Vertigo gouldii (Mollusca: Pulmonata, Pupillidae) group of minute North American land snails. Mol Phylogenet Evol 53:1010–1024 Nowak CA, Abrahamson LP, Raynal DJ (1993) Powerline corridor vegetation management trends in New York State: has a post-herbicide era begun? J Arboric 19:20–26 Oggier P, Zschokke S, Baur B (1998) A comparison of three methods for assessing the gastropod community in dry grasslands. Pedobiologia 42:348–357 Pilsbry HA (1948) Land Mollusca of North America (North of Mexico). Academy of Natural Sciences of Philadelphia, Monograph #3 Pokryszko BM (1987) On aphally in the Vertiginidae (Gastrocopta: Pulmonata: Orthurethra). J Conchol 32:365–375 Schilthuizen M, Lombaerts M (1994) Population structure and levels of gene flow in the Mediterranean land snail Albinaria corrugata (Pulmonata: Clausiliidae). Evolution 48:577–586 Schreiber RK, Graves JH (1977) Powerline corridors as possible barriers to movements of small mammals. Am Midl Nat 97:504–508 Shaffer ML, Samson FB (1985) Population size and extinction: a note on determining critical sizes. Am Nat 125:144–152 Simberloff D, Cox J (1987) Consequences and costs of conservation corridors. Conserv Biol 1:63–71 Sorrie BA, Weakley AS (2001) Coastal plain vascular plant endemics and phytogeographic patterns. Castanea 66:50–82 Turgeon DD, Quinn JF Jr, Bogan AE, Coan EV, Hochberg FG, Lyons WG, Mikkelsen P, Neves RJ, Roper CFE, Rosenberg G, Roth B, Scheltema A, Thompson FG, Vecchione M, Williams JD (1998) Common and scientific names of aquatic invertebrates from the United States and Canada: Mollusks, 2nd edn. Bethesda: American Fisheries Society Special Publication 26 Yahner RH, Morrell TE, Rachael JS (1989) Effects and edge contrast on depredation of artificial avian nests. J Wildl Manag 53:1135–1138 Biodivers Conserv (2012) 21:781–795 795 123 Author's personal copy 211 Research Note 616 Conservation Biology, Pages 616–621 Volume 17, No. 2, April 2003 Effects of Rock Climbing on the Land Snail Community of the Niagara Escarpment in Southern Ontario, Canada MICHELE A. MCMILLAN,* JEFFREY C. NEKOLA,† AND DOUGLAS W. LARSON*‡ *Department of Botany, University of Guelph, Guelph, Ontario N1G 2W1, Canada †Department of Natural and Applied Sciences, University of Wisconsin, Green Bay, WI 54311, U.S.A. Abstract: The cliffs of the Niagara Escarpment provide habitat for extremely diverse communities of land snails that may be at risk as a result of recreational rock climbing. We examined the effects of rock climbing on the density, richness, diversity, and community composition of snails on the Niagara Escarpment in southern Ontario, Canada. We sampled from randomly selected climbed and unclimbed sections of cliffs on the plateau (cliff edge), cliff face, and talus (cliff base). Snail density, richness, and diversity were lower along climbing routes than in unclimbed areas, and community composition differed between climbed and unclimbed samples. These results suggest that rock climbing has significant negative effects on all aspects of the snail community on cliffs; therefore, we recommend the inclusion of gastropods in conservation plans for protected areas containing cliffs. Efectos del Alpinismo en la Comunidad de Caracoles Terrestres del Acantilado del Niagara en Ontario Meridional, Canadá Resumen: Las barrancas del acantilado del Niagara proveen hábitat para comunidades extremadamente diversas de caracoles terrestres que pueden estar en riesgo debido al alpinismo recreativo. Examinamos los efectos del alpinismo en la densidad, riqueza, diversidad y composición de comunidades de caracoles del acantilado del Niagara en Ontario Meridional, Canadá. Tomamos muestras de las mesetas (bordes de acantilados), la cara del acantilado y el talud (base del acantilado) de secciones de barrancas usadas y no usadas para el alpinismo y seleccionadas al azar. La densidad, la riqueza y la diversidad de caracoles fueron más bajas en las rutas escaladas que en aquellas áreas no escaladas y la composición de la comunidad difirió entre muestras escaladas y no escaladas. Estos resultados sugieren que el alpinismo tiene impactos negativos significativos en todos los aspectos de la comunidad de caracoles en acantilados; por lo tanto, recomendamos la inclusión de gasterópodos en los planes de conservación para áreas protegidas que contengan acantilados. ‡Address correspondence to D.W. Larson, email dwlarson@uoguelph.ca Paper submitted July 25, 2001; revised manuscript accepted May 31, 2002. Introduction Since its introduction to eastern North America 50 years ago, recreational rock climbing has continually increased in popularity, with the most dramatic increases occurring over the past 20 years (Valis 1991). Cliff ecosystems contribute greatly to the regional biodiversity of plants and animals (Larson et al. 2000). Research conducted thus far has demonstrated that rock climbing can lead to decreased abundance and richness of vascular and nonvascular plants and lichens (Nuzzo 1995; Herter 1996; Nuzzo 1996; Kelly & Larson 1997; Camp & Knight 1998; Farris 1998; McMillan & Larson 2002), suggesting that the entire biotic community might be affected. 212 Conservation Biology Volume 17, No. 2, April 2003 McMillan et al. Effects of Rock Climbing on Land Snails 617 The potential exists for rock climbing to have a particularly strong influence on land snails because land snail diversity is generally higher on limestone outcrops than in other habitats (Burch 1962; Baur et al. 1995). In fact, maximum global richness levels for terrestrial gastropods are often associated with wooded carbonate cliffs and talus slopes. This has been demonstrated in New South Wales (Stanisic 1997), Scotland (Cameron & Greenwood 1991), Germany (Schmid 1966), Sweden (Waldén 1981), the Appalachian mountains (Getz & Uetz 1994), and the Great Lakes region of North America (Nekola 1999). The Niagara Escarpment is a series of dolomitic limestone outcrops that extends from the Bruce Peninsula to the Niagara Region in Ontario and through Michigan, Wisconsin, Illinois, and Iowa. These cliffs support the most ancient forest ecosystem east of the Rocky Mountains (Larson & Kelly 1991), with eastern white cedar trees (Thuja occidentalis) living in excess of 900 years (Larson et al. 2000). These cliffs also support extremely high levels of snail diversity. Up to 34 species have been observed in soil collected from a surface area of 1000 m2 , up to 23 species in 1 m2 , and up to 21 species in 0.04 m2 (Nekola 1999; Nekola & Smith 1999). Unfortunately, these cliffs are also exposed to the highest levels of recreational rock climbing in Ontario (Bracken et al. 1991). The importance of the Niagara Escarpment as a reservoir for biotic biodiversity necessitates the creation of formal conservation policies that protect this community. Here, we document the effects of rock climbing on the density, richness, diversity, and community composition of land snails on dolomitic limestone cliffs in Ontario. Methods Study Site We conducted field work from late August to early October 1998 along a 35-km section of the Niagara Escarpment near Milton, Ontario (Universal Transverse Mercator Zone 17, 4806800–4818000 mN, 587500–591500 mE). The public and privately owned cliffs we sampled are composed of Silurian dolomitic limestone (Niagara Escarpment Commission 1979) surrounded by a green belt of secondgrowth deciduous forest that runs through larger areas of urban development and agricultural land. Climbed cliffs were sampled at Buffalo Crag (90 graded routes), Rattlesnake Point (142 graded routes), Mt. Nemo (236 graded routes), and Kelso (54 graded routes) conservation areas. Unclimbed cliffs were sampled at Mt. Nemo and Crawford Lake conservation areas and from an adjacent cliff on the property of Milton Limestone Incorporated. Sections of climbed and unclimbed cliffs were within 15 km of one another. Sampling Design We determined the impact of climbing through the comparison of “climbed” and “unclimbed” areas. A region was classified as climbed when climbing manuals described an established climbing route for that area (Bracken et al. 1991; Oates & Bracken 1997). A section of the cliff was classified as unclimbed when no routes were described for the area and when no other obvious signs of climbing were present (e.g., implanted climbing hardware, chalk marks on hand holds). We sampled from 50 vertical belt transects, each 1 m in width, that extended the height of the cliff. Twentyfive climbed transects were chosen randomly from a pool of 101 climbing routes, and 25 unclimbed transects were chosen randomly from a pool of 106 transects. The pool of climbed transects included all documented climbing routes for the area with a difficulty rating between 5.7 and 5.9 and a star rating in the climbing manual (Bracken et al. 1991). We chose these routes because they represented the best routes for climbers with an intermediate skill level. As such, they likely attract higher and more regular amounts of climbing activity than other routes. We restricted unclimbed transects to sections of the cliff that were Ն7 m wide, that appeared physically suitable for climbing, and that two experienced climbers and the first author agreed would be rated between 5.7 and 5.9 if graded. Climbed and unclimbed transects were rejected if they were Ͻ10 m in height or contained a roof or overhang of more than 1 m, continuous water seeps, or loose rocks. Land snails were sampled from three positions within each transect for a total of 150 observations: cliff face (vertical surface), plateau (or cliff edge), and talus (cliff base). These adjacent habitats were sampled in addition to the cliff face because they are also subjected to climbing-related disturbances such as being used as a place from which to belay (hold the ropes for the climber), to anchor climbing ropes, to store packs and other equipment, and for climbers to rest between ascents. Transects extended from 2 m beyond the cliff face on the plateau to 2 m beyond the cliff base in the subtending talus. We collected 250 mL of soil from each of the three positions within each transect. In plateau and talus habitats, we collected soil from the upper 5 cm of the organic soil horizon within a 1 ϫ 2 m area. Because of the paucity of soil on the cliff face, we collected soil samples from ledges and cracks extending the entire height of the cliff within the 1-m transect (ranging from 11.5 to 30 m). Although the sample extent for the cliff-face samples was greater, the total surface area over which the samples were taken did not exceed 2 m2 . To minimize the effects of sampling on the cliff community, no ledge was deprived of more than half its soil. Each 1 ϫ 2 m quadrat sampled from the plateau and talus was divided into 50 20 ϫ 20 cm subquadrats, and 213 618 Effects of Rock Climbing on Land Snails McMillan et al. Conservation Biology Volume 17, No. 2, April 2003 Table1.Frequency(%ofquadrats)inwhicheachspeciesoflandsnailwaspresentforunclimbed(U)andclimbed(C)quadratssampledin1998fromtheNiagaraEscarpmentinsouthern Ontario,Canada. Freq.(%)Freq.(%) SpeciesGranka Srankb UC␹2c,d SpeciesGranka Srankb UC␹2c,d AnguispiraalternataG5S549.324.09.30*NesovitreabinneyanaG?S4-S528.02.716.64** CarychiumexileG4?S3-S428.09.37.42ParavitreamultidentataG4-G5S2-S36.72.70.60 CarychiumnannodesG4-G5S1-S26.71.31.56Punctumminutissimum//85.32.750.03**** CochlicopalubricaG4-G5S514.70.09.81*Pupillamuscorum//1.332.023.23**** CochlicopalubricellaG4-G5S42.71.30.00Stenotremafraternumfraternum//5.30.02.31 Columellasimplex//28.00.022.15***StriaturaexiguaG4S417.01.39.53* DiscuscatskillensisG3-G5S572.032.022.46****StriaturaferreaG4-G5S51.30.00.00 EuconulusfulvusG4-G5S4-S528.00.022.15***StriaturamiliumG4S4-S540.02.728.96**** EuconuluspolygyratusG?S46.70.03.31StrobilopslabyrinthicaG5S518.710.71.33 GastrocoptacontractaG5S514.718.70.19SuccineaovalisG5S3-S420.00.014.52** GastrocoptacorticariaG4-G5S232.038.70.47Triodopsisalbolabris//2.70.00.51 GastrocoptaholzingeriG4-G5S2-S353.349.30.11Triodopsisdenotata//1.30.00.00 GastrocoptapentodonG5S561.317.328.61****TriodopsistridentataG5S3-S46.74.00.13 Glyphyaliniaindentata//4.01.30.26ValloniacostataG4-G5S50.01.30.00 GlyphyalinarhoadsiG5S3-S416.04.04.74ValloniagracilicostaG?S1-S265.372.00.50 GuppyasterkiiG4-G5SH5.31.30.83ValloniapulchellaG4-G5S50.01.30.00 HawaiiaminisculaG5S30.08.04.34VertigobollesianaG3S326.75.311.16* HelicodiscusparallelusG5S512.06.70.71VertigogouldiG4-G5S378.733.329.46**** HelicodiscusshimekiG4S4-S52.76.70.60VertigoparadoxaG2-G4S2-S376.042.715.91** MesomphixcupreusG5S31.30.00.00ZonitoidesarboreusG5S550.126.78.12 a Globalconservationrank:G1,extremelyrare;G2,veryrare;G3,raretouncommon;G4,common;G5,verycommon;G?,unrankedornotlisted. b Provincialconservationranking:S1,extremelyrare;S2,veryrare;S3,raretouncommon;S4,common;S5,verycommon;SH,historicallypresentinOntariobutnotobservedrecently(typ- icallynotrecordedfor20years)ornotlisted. c AdjustedusingYates’correctionforcontinuityassuggestedforcasesinwhichthereisasingledegreeoffreedom.TheBonferronicorrectionformultiplecomparisonswasusedtocalculate significancelevels(MinistryofNaturalResources2002). d Probability:*pϽ0.1,**pϽ0.01,***pϽ0.001,****pϽ0.0001. 214 Conservation Biology Volume 17, No. 2, April 2003 McMillan et al. Effects of Rock Climbing on Land Snails 619 the number of subquadrats containing exposed rock was recorded. A 1 ϫ 2 m quadrat was similarly sampled from the vertical center of each cliff-face transect. We recorded this data prior to soil collection. We collected soil samples in muslin bags. We placed fresh samples in a drying oven at 80Њ C for 4–7 days and then stored them at room temperature for several months. We soaked each bag in water from 12–24 hours, after which we removed the soil and washed it through a graduated sieve series (4, 2, 1, and 0.42 mm). We discarded particles that passed through the smallest sieve. We returned the soil to muslin bags and redried them. We then dry-sieved the soil through the graduated sieve series and examined the factions of soil under an 8ϫ dissecting microscope against a neutral brown background. We removed intact shells from the soil with a moist paintbrush and identified them to species (or subspecies) following the nomenclature of Hubricht (1985). We checked our identifications against the Hubricht Collection at the Field Museum of Natural History and the second author’s reference collection at the University of Wisconsin, Green Bay. We counted and recorded the number of shells for each species. We also calculated the species richness and heterogeneity (Shannon-Wiener index) of each sample. The reference collection from this study is housed in the Department of Botany at the University of Guelph. Statistical Analysis We used a split-plot experimental design (Steel & Torrie 1980; Kuehl 1994) to compare the density, richness, and heterogeneity of the sampled snails. We used climbing intensity (climbed vs. unclimbed) as the whole-plot factor and habitat (plateau, cliff face, and talus) as the splitplot factor. To minimize the impact of factors other than climbing and habitat on the outcome of the statistical analysis, we included cliff height, percent canopy cover, and compass direction of the cliff face in the model as covariates. The compass direction of the cliff was broken down into a north component and an east component, with the former being equal to the cosine (compass reading * ␲) and the latter being equal to sine (compass reading * ␲). We used these conversions because the use of an untransformed compass reading would determine 1Њ and 359Њ to be very different from one another, when they are both almost due north. We analyzed the effects of rock climbing and habitat type on snail density, richness, and heterogeneity through analysis of variance (ANOVA) tables generated by SAS computer software (Anonymous 1995). We used the MIXED procedure due to the combination of covariates within the split-plot design. We analyzed the results for both main and simple effects, the main effects with F statistics and the simple effects with Tukey’s procedure (Steel & Torrie 1980). We employed chi-square tests to determine the impact of climbing and habitat type on the abundance of individual species. We used the Bonferroni correction for multiple comparisons (Kuehl 1994) and Yates’s correction for continuity when there was only a single degree of freedom (Steel & Torrie 1980). Results We retrieved 14,023 intact shells from the 37.5 L of soil collected. These individuals belonged to 40 different species, which represents almost half the diversity of land snails in Ontario (88 taxa). The species varied in global conservation ranking from G3 (Vertigo bollesiana) to G5 and in provincial ranking from S1/S2 (Carychium nannodes, Vallonia gracilicosta) to S5 (Table 1). Most of the species were Ͻ5 mm at their widest point and are thus referred to as “minute” snails (Emberton 1995). We found a large amount of broken shell material but did not include this in our analysis. Shell density was over five times greater in unclimbed samples than in climbed samples (78.26 vs. 14.86 individuals per 250 mL of soil; p Ͻ 0.00001). Shell density was significantly lower along climbing routes in each of the three habitat positions (Fig. 1). In addition, shell density varied significantly between the three habitat positions ( p Ͻ 0.005). The average density of shells found in Figure 1. Mean density of land snails per 250 mL soil sample collected from the Niagara Escarpment, Ontario, Canada. Each bar represents the mean from 25 samples. Bars that share a letter code are not significantly different from one another at ␣ ϭ 0.05. 215 620 Effects of Rock Climbing on Land Snails McMillan et al. Conservation Biology Volume 17, No. 2, April 2003 cliff face and talus samples (46.63 and 45.41 individuals per 250 mL soil) was more than twice the average density of shells found in plateau samples (19.15 individuals per 250 mL soil; p Ͻ 0.005). The species richness of land snails was significantly lower along climbing routes (p Ͻ 0.00001). Unclimbed areas supported almost two times more species than undisturbed areas (9.84 vs. 5.20 species per 250 mL soil). This significant reduction in richness along climbing routes occurred in all habitat positions (Fig. 2). Similarly, climbed areas showed significantly lower levels of Shannon-Wiener diversity than did unclimbed areas (1.11 vs. 1.53; p Ͻ 0.001). Species richness was lowest on the plateau (4.94 species per 250 mL soil) and highest on the talus (10.56 species per 250 mL soil; p Ͻ0.00001). Fourteen of the 40 species had significantly greater frequency in unclimbed samples, whereas only one species had a significantly greater frequency in climbed areas (Table 1). The remaining species showed no statistical preference for either climbed or unclimbed areas. The proportion of subquadrats containing exposed rock was significantly higher in climbed than unclimbed quadrats (79.6% vs. 48.0%; p Ͻ 0.00001). This trend occurred in each of the three habitat types, but was significant for the plateau (10.6% vs. 66.8%; p Ͻ 0.0001) and talus (40.2% vs. 73.9%; p Ͻ 0.0001) habitats only. Unclimbed and climbed quadrats contained similar amounts of exposed rock (93.1% vs. 98.0%; p ϭ 0.98). Discussion Our results suggest that rock climbing has strong negative effects on the extremely diverse and abundant community of land snails that normally occurs on undisturbed cliffs of the Niagara Escarpment. Snail density, richness, and diversity were all significantly lower in climbed areas than in undisturbed areas, in spite of the fact that the amount of soil collected was constant. Because the surface area of soil was also lower in climbed areas, it is likely that rock climbing has an even larger impact on land snails than is suggested by our results. The enormous local abundance of land snails and the unusually high species richness in undisturbed sites suggest that snails may be processing energy and matter at high rates and therefore may be important natural components of cliff-ecosystem food webs. Whatever their role, their recolonization and the subsequent restoration of function are likely to occur very slowly on denuded cliffs in view of the slow rates of dispersal of such snails ( Baur & Baur 1994). Rock climbing also appears to affect the composition of land snail communities. There are several possible mechanisms that might account for this change. First, the ledges and cracks in the vertical rock surface that are the primary habitat for cliff land snails are also the primary means by which climbers ascend the cliff face. The use of these microsites for hand and foot holds causes removal of unconsolidated soil and organic matter, thereby decreasing the amount and quality of available habitat for snails. In some instances, climbers purposefully remove soil to increase the security of the hold and to reduce the chance of slipping. The soil that is left is then subject to direct pressure from the hands and feet of climbers, which probably causes soil compaction and a reduction in organic material, as is found in other disturbed sites. Soil compaction and removal from staging areas in the talus and at tie-off sites on the plateau also decrease the amount of habitat and resources available for these species. Future research is required to determine the effects of climbing on physical soil properties and chemistry. Due to the detrimental effect of rock climbing on the land snail community we observed in this study, we recommend that management plans for the Niagara Escarpment be modified to include specific policies regarding recreational rock climbing. The most effective management practices for the protection of gastropod species on the Niagara Escarpment would limit the aerial extent of climbing activity. We believe that land snail communities on other cliffs may react similarly to climbing pressure, because other cliffs around the world also support extremely diverse communities of terrestrial gastropods (Schmid 1966; Waldén 1981; Cameron & Greenwood 1991; Getz & Uetz 1994; Stanisic 1997; Nekola 1999). We therefore recommend that gastropods also be included in management policies for other areas containFigure 2. Mean number of land snail species present in 250 mL soil samples collected from the Niagara Escarpment, Ontario, Canada. Each bar represents the mean from 25 samples. Bars that share a letter code are not significantly different from one another at ␣ ϭ 0.05. 216 Conservation Biology Volume 17, No. 2, April 2003 McMillan et al. Effects of Rock Climbing on Land Snails 621 ing cliffs. Special attention should be given to areas in which rare species are known to exist. Acknowledgments Funding for this project was provided by the Natural Sciences and Engineering Research Council of Canada, the Mountain Equipment Co-op, the Lattornell Graduate Student Travel Fund, and the Access Fund. We thank X. Benjamin, I. Brodo, J. Buck, P. Clarke, J. Crowe, A. Cullen, M. deGruchy, C. Fietsch, S. Girard, K. Gormley, P. Kelly, C. A. Lacroix, G. Lecanu, J. Lundholm, G. Mackie, U. Matthes, A. Millward, A. Nassuth, S. Newman, R. Pal, R. Reader, J. Stewart, S. Thevenet, and P. Wong for their help in the field and laboratory. We thank Halton Region Conservation Authority and Milton Limestone Ltd. for allowing us to sample on their properties. Literature Cited Anonymous. 1995. SAS system for Windows. Version 6.11. SAS Institute, Cary, North Carolina. Baur, B., and A. Baur. 1994. Dispersal of the land snail Chondrina avenacea on vertical rock walls. Malacological Review 27:53–59. Baur, B., L. Fröberg, and A. Baur. 1995. Species diversity and grazing damage in a calcicolous lichen community on top of stone walls in Öland, Sweden. Annales Botanici Fennici 32:239–250. Bracken, M., J. Barnes, and C. Oates. 1991. The escarpment: a climbers guide. Borealis Press, Toronto. Burch, J. B. 1962. How to know the eastern land snails. WM.C. Brown, Dubuque, Iowa. Cameron, R. A. D., and J. J. D. Greenwood. 1991. Some montane and forest molluscan fauna from eastern Scotland: effects of altitude, disturbance, and isolation. Pages 437–442 in Proceedings of the tenth international malacological congress. Unitas Malacologica, Tübingen, Germany. Camp, R. J., and R. L. Knight. 1998. Effects of rock climbing on cliff plant communities at Joshua Tree National Park, California. Conservation Biology 12:1302–1306. Emberton, K. C. 1995. Land-snail community morphologies of the highest-diversity sites of Madagascar, North America, and New Zealand, with recommended alternatives to height-diameter plots. Malacologia 36:43–66. Farris, M. A. 1998. The effects of rock climbing on the vegetation of three Minnesota cliff systems. Canadian Journal of Botany 76: 1981–1990. Getz, L. L., and G. W. Uetz. 1994. Species diversity of terrestrial snails in the southern Appalachian mountains, U.S.A. Malacological Review 27:61–74. Herter, W. 1996. Die Xerothermvegetation des Oberen Donautals— Gefaerhrdung der Vegetation durchMensch und Wild sowie Schutz—und Erhaltungsvorschlaege. Veroeffentlichungen des projekts AAngewandte Oekologie@ PAoe 10. Veroef, Karlsruhe, Ger- many. Hubricht, L. 1985. The distributions of the native land mollusks of the eastern United States. Fieldiana 24:1–191. Kelly, P. E., and D. W. Larson. 1997. Effects of rock climbing on populations of presettlement eastern white cedar (Thuja occidentalis) on cliffs of the Niagara Escarpment, Canada. Conservation Biology 11:1125–1132. Kuehl, R. O. 1994. Statistical principles of research design and analysis. Wadsworth Publishing, Belmont, California. Larson, D. W., and P. E. Kelly. 1991. The extent of old-growth Thuja occidentalis on cliffs of the Niagara Escarpment. Canadian Journal of Botany 69:1628–1636. Larson, D. W., U. Matthes, and P. E. Kelly. 2000. Cliff ecology: pattern and process in cliff ecosystems. Cambridge University Press, Cambridge, United Kingdom. McMillan, M. A., and D. W. Larson. 2002. The effects of rock climbing on the vegetation of the Niagara Escarpment in Southern Ontario, Canada. Conservation Biology 16:389–398. Ministry of Natural Resources. 2002. Natural heritage information centre list of Ontario mollusks. Ministry of Natural Resources, Peterborough, Ontario, Canada. Available from http://www.mnr.gov.on. ca/MNR/nhic/nhic.html (accessed March 2002). Nekola, J. C. 1999. Terrestrial gastropod richness of carbonate cliff and associated habitats in the great lakes region of North America. Malacologia 41:231–252. Nekola, J. C., and T. M. Smith. 1999. Terrestrial gastropod richness patterns in Wisconsin and carbonate cliff communities. Malacologia 41:253–269. Niagara Escarpment Commission. 1979. Geology: the Niagara Escarpment. Niagara Escarpment Commission, Georgetown, Ontario, Canada. Nuzzo, V. A. 1995. Effects of rock climbing on cliff goldenrod (Solidago sciaphila Steele) in Northwest Illinois. American Midland Naturalist 133:229–241. Nuzzo, V. A. 1996. Structure of cliff vegetation on exposed cliffs and the effect of rock climbing. Canadian Journal of Botany 74:607–617. Oates, C., and M. Bracken. 1997. A sport climbers guide to Ontario limestone. Borealis Press, Toronto. Schmid, G. 1966. Die mollusken des Spitzbergs. Der Spitzberg bei Tübingen, Natur und Landschaftsschutzgebiete. Baden-Würtemburg 3:596–701. Stanisic, J. 1997. An area of exceptional land snail diversity: the Macleay Valley, north-eastern New South Wales. Memoirs of the Museum of Victoria 56:441–448. Steel, R. G. D., and J. H. Torrie. 1980. Principles and procedures of statistics: a biometrical approach. 2nd edition. McGraw-Hill, New York. Valis, T. 1991. Escarpment climbing history. Pages 3–6 in M. Bracken, J. Barnes, and C. Oates, editors. The escarpment: a climbers guide. Borealis Press, Toronto. Waldén, H. W. 1981. Communities and diversity of land molluscs in Scandinavian woodlands. I. High diversity communities in taluses and boulder slopes in SW Sweden. Journal of Conchology 30:351–372. 217 53Animal Biodiversity and Conservation 25.2 (2002) © 2002 Museu de Ciències NaturalsISSN: 1578–665X Nekola, J. C., 2002. Effects of fire management on the richness and abundance of central North American grassland land snail faunas. Animal Biodiversity and Conservation, 25.2: 53–66. AbstractAbstractAbstractAbstractAbstract Effects of fire management on the richness and abundance of central North American grassland land snail faunas.— The land snail faunas from 72 upland and lowland grassland sites from central North America were analyzed. Sixteen of these had been exposed to fire management within the last 15 years, while the remainder had not. A total of 91,074 individuals in 72 different species were observed. Richness was reduced by approximately 30% on burned sites, while abundance was reduced by 50–90%. One–way ANOVA of all sites (using management type as the independent variable), a full 2–way ANOVA (using management and grassland type) of all sites, and a 2–way ANOVA limited to 26 sites paired according to their habitat type and geographic location, demonstrated in all cases a highly significant (up to p < 0.0005) reduction in richness and abundance on fire managed sites. Contingency table analysis of individual species demonstrated that 44% experienced a significant reduction in abundance on firemanaged sites. Only six species positively responded to fire. Comparisons of fire response to the general ecological preferences of these species demonstrated that fully 72% of turf–specialists were negatively impacted by fire, while 67% of duff–specialists demonstrated no significant response. These differences were highly significant (p = 0.0006). Thus, frequent use of fire management represents a significant threat to the health and diversity of North American grassland land snail communities. Protecting this fauna will require the preservation of site organic litter layers, which will require the increase of fire return intervals to 15+ years in conjunction with use of more diversified methods to remove woody and invasive plants. Key words: Land snail, Biodiversity, Conservation, Fire management, Grassland, North America. ResumenResumenResumenResumenResumen Efectos de la gestión con fuego sobre la riqueza y abundancia de la fauna de caracoles terrestres de las praderas de América del Norte.— Se analiza la fauna de caracoles terrestres de 72 praderas en mesetas y llanuras de la región central de América del Norte. En 16 de ellas se habían efectuado intervenciones de incendio controlado durante los últimos 15 años, mientras en el resto no. Se observaron un total de 91.074 individuos de 72 especies diferentes. La riqueza en especies estaba reducida en un 30% en las áreas quemadas, mientras que la abundancia de individuos estaba reducida en un 50–90%. Un ANOVA unidireccional de todas las áreas (usando como variable independiente el tipo de intervención), un ANOVA bidireccional completo (usando el tipo de intervención y el tipo de pradera) en todas las áreas y un ANOVA bidireccional limitado a 26 áreas agrupadas según su tipo de hábitat y localización geográfica, demostró en todos los casos una reducción altamente significativa de la riqueza y de la abundancia (hasta p < 0,0005) en áreas sometidas a incendio. Un análisis individual de las especies mediante tablas de contingencia demostró que el 44% experimentaron una reducción significativa de su abundancia en las áreas quemadas. Sólo seis especies respondieron positivamente al fuego. Comparando la respuesta al fuego con las preferencias ecológicas generales de estas especies se demostró que al menos el 72% de las especialistas que viven en sustrato herbáceo fueron afectadas negativamente por el fuego mientras que el 67% de las que viven en sustrato húmico no demostraron ninguna respuesta significativa. Estas diferencias fueron altamente significativas (p = 0,0006). Así pues, el uso frecuente del fuego representa una amenaza significativa para la salud y diversidad de las comunidades de caracoles terrestres de las praderas de América del Norte. La protección de esta fauna requerirá la preservación de las capas de materia orgánica y la ampliación de los intervalos entre las actuaciones de quema a periodos superiores a 15 años, así como el uso de métodos más diversos para eliminar las plantas leñosas e invasivas. Palabras clave: Caracol terrestre, Biodiversidad, Conservación, Gestión con fuego, Praderas, América del Norte. (Received: 9 IV 02; Final acceptance: 18 VI 02) Jeffrey C. Nekola, Dept. of Natural and Applied Sciences, Univ. of Wisconsin–Green Bay, Green Bay, Wisconsin 54311 USA. E–mail: nekolaj@uwgb.edu Effects of fire management on the richness and abundance of central North American grassland land snail faunas J. C. Nekola 218 54 Nekola Introduction Fire has long been implicated in the maintenance of central North American grassland communities (WEAVER, 1954; CURTIS, 1959). Numerous native plant species respond to fire by increasing their growth and reproductive rates (EHRENREICH & AIKMAN, 1963; KUCERA & KOELLING, 1964; TOWNE & OWENSBY, 1984). One of the most direct effects of prairie fire is the removal of the soil mulch layer, which has been implicated in the ‘stagnation’ of prairie plant communities through the delay of initial spring growth, thinning of grass stem density, and prevention of herbaceous understory development (WEAVER & ROWLAND, 1952; KUCERA & KOELLING, 1964). Fire is also thought to limit invasion of woody and exotic plants into native prairie habitats (e.g., PAULY, 1985; ROOSA, 1984). For these reasons, prescribed fire has become the management tool of choice by prairie conservation groups throughout the midwestern USA (COLLINS & WALLACE, 1990). However, an increasing body of research suggests that fire is not universally beneficial all prairie biota. Fire depresses growth and reproductive rates of native C3 prairie plants (DIX, 1960; HADLEY, 1970; HILL & PLATT, 1975), which make up at least 50% of the native flora north of 44o N (STOWE & TEERI, 1978; SIMS, 1988). Fire has also been implicated in the loss and/or reduction of numerous native prairie invertebrate species including Lepidoptera, Homoptera, Hymenoptera, and Araneae (SWENGEL, 1996, 1998; HARPER et al., 2000). The effects of such practices on prairie soil biodiversity are largely undocumented. Combustion of mulch through repeated fire episodes will remove the detritusphere, one of the most important reservoirs for soil biodiversity (COLEMAN & CROSSLEY, 1996). HARPER et al. (2000) documented significant reductions in Collembola following Illinois prairie fires. As the soil fauna represents one of the largest species pools in terrestrial ecosystems (BEHAN–PELLETIER & NEWTON, 1999), the potential impacts of such processes on total site biodiversity may be large. Although not as hyper–diverse as bacteria, fungi, nematodes, and arthropods, molluscs still represent one of the more important components of soil biodiversity (RUSSELL–HUNTER, 1983). Almost 600 species are known from eastern North America (HUBRICHT, 1985), with up to 21 taxa cooccurring within 400 cm2 microhabitats (NEKOLA & SMITH, 1999). Most of these taxa represent generalist detritivores that live in and on dead organic material (BURCH & PEARCE, 1990) As almost 90% of snails occur within 5 cm of the soil surface (HAWKINS et al., 1998), protection of this fauna will likely be tied to the fate of mulch layers. Disturbances such as logging, recreational or urban development, or bedrock and soil removal cause dramatic changes in woodland snail communities with duff soil surfaces (NEKOLA, in press a). The impact of fire, and associated detritusphere removal, on snail communities is unclear. Fire has been suggested to negatively influence the faunas of Aegean islands (WELTER–SCHULTES & WILLIAMS, 1999), Queensland fens (STANISIC 1996), and Tasmanian woodlands (REGAN et al., 2001). However, FREST & JOHANNES (1995) state that molluscs are able to survive natural fires in northwestern North America, and THELER (1997) argued that xeric prairie faunas in Wisconsin owe their existence to frequent fires that keep grassland areas treeless. Unfortunately, no data was presented by these various authors to validate such conflicting statements. To evaluate this issue, the richness and abundance of land snails was quantitatively compared between unburned and recently (< 15 year) burned sites in the midwestern USA, including 13 pairs of sites which possess similar habitats and are spatiall proximate. From these, the following questions will be considered: 1. Is there a significant difference in land snail community richness between burned and unburned grasslands? 2. Is there a significant difference in land snail abundance between burned and unburned grasslands? 3. What species show positive, negative, or no response to fire? What ecological factors (if any) may help explain these responses? Materials and methods Study Sites Seventy two grassland sites were surveyed between May 1996–November 2001 for terrestrial molluscs across a 850 km extent of central North America (fig. 1, table 1). Sites are generally centered on northwestern Minnesota and northeastern Iowa. Forty–two occur in Minnesota, 25 in Iowa, and 5 in Wisconsin. Thirty–two sites represent upland habitats (including tallgrass prairie, sand prairie, and bedrock glades), while the remaining 40 are lowland sites (including wet prairie, sedge meadow, and fens). Previous use of fire management on sites was assessed by either observing carbonized woody plant stems or other debris on the ground surface, or through interviews with site managers or other knowledgeable individuals. No use of fire management was noted from 56 sites (88% of total), while 16 (22%) had been subjected to some amount of prescribed burning. Eleven of these burned sites occur in Minnesota, while the remaining five occur in Iowa. The latitude– longitude location of each site was determined using either USGS 7.5 minute topographic maps or a hand–held GPS. Field Methods Documentation of terrestrial gastropod faunas from each site was accomplished by hand 219 Animal Biodiversity and Conservation 25.2 (2002) 55 collection of larger shells and litter sampling for smaller taxa within 100–1,000 m2 areas that contained examples of all major microhabitats and were thus representative of the larger site. The actual grain size employed was determined by the minimum size necessary to emcompass all microhabitats. Soil litter sampling was primary used as it provides the most complete assessment of grassland faunas (OGGIER et al., 1998). A single site sample consisted of a composite of individual soil litter subsamples of approximately 200 ml collected from appropriate microhabitats. As suggested by EMBERTON et al. (1996), litter collections were made at places of high micromollusc density, with a constant volume (approximately 4 liters) being gathered from each site. Sampling was generally comprised of: 1. Small blocks (ca. 125 cm3 ) of turf; 2. Loose soil and leaf litter accumulations under or adjacent to shrubs, cobbles, boulders, and/or hummocks; and 3. Other microsites supporting relatively thick mulch layers. LLLLLaboratory procedures Samples were slowly and completely dried in either a low–temperature soil oven (ca. 80–95o C) or in full sun in a greenhouse. Dried samples were then soaked in water for 3–24 hours, and subjected to careful but vigorous water disaggregation through a standard sieve series (ASTME 3/8" (9.5 mm), #10 (2.0 mm), #20 (0.85), and #40 (0.425 mm) mesh screens). Sieved sample fractions were then dried and passed again through the same sieve series. These dry, resorted fractions were hand picked against a neutralbrown background. All shells and shell fragments were removed. All identifiable shells from each site were assigned to species (or subspecies) using the author’s reference collection and the Hubricht Collection at the Field Museum of Natural History (FMNH), with the total number of shells per species per site being recorded. The total number of unassignable, immature individuals was also counted from each site. All specimens have been catalogued and are housed in the author’s collection at the University of Wisconsin–Green Bay. Nomenclature generally follows that of HUBRICHT (1985), with updates and corrections by FREST (1990, 1991) and NEKOLA (in press b). The general ecological preferences (turf specialist, duff–specialist or generalist) of each species is based upon analyses presented in NEKOLA (in press a). Statistical procedures Differences in species richness and total shell abundance between burned and unburned grassland sites were analyzed via ANOVA. Initially, 1–way ANOVAs were preformed on the entire dataset. However, the effect of fire may be obscured in this analysis due to confounding effects of habitat type and geographic location. To help control for this, two additional sets of ANOVAs were conducted. First, full 2–way ANOVAs were calculated for all sites using grassland type (upland vs. lowland) and management history (burned vs. unburned) as the independent variables. Second, 13 pairs of sites representing closely similar habitats within the same geographic region, but differing in their fire management history, were selected. These site pairs are (first site is burned, second is unburned): Malmberg Prairie vs. Sandpiper Prairie; Pankratz Mesic Prairie vs. Radium NE; Pankratz Low Prairie vs. Bjornson WMA; Pankratz Fen vs. Faith South; Marcoux WMA vs. Cyr Creek; East Park WMA vs. Goose Lake; Felton Fen 1 vs. Ogema West; Waubun SE vs. Eastlund Lake; Chicog vs. Tansen; Beemis Creek vs. Hampton East; Fayette vs. Decorah Glade; Baty Glade vs. Canton Glade; Brayton–Horsley vs. Stapleton Church. A 2–way ANOVA without interaction was then calculated for these sites, with site pair identity and management type representing independent variables. Fig. 1. Map of study region, showing location of surveyed grassland sites: F Unburned upland; M Burned upland; H Unburned lowland; F Burned lowland. Fig. 1. Mapa del área de estudio que muestra la localización de las praderas estudiadas: F Meseta no quemada; M Meseta quemada; H Llanura no quemada; F Llanura quemada. 220 56 Nekola Table 1. Location, grassland type, management, species richness and total number of collected individuals from sample sites: GT. Grassland type; M. Management; R. Richness; I. Individuals. Tabla 1. Localización, tipo de pradera, gestión, riqueza de especies y número total de individuos recogidos en cada área de estudio: GT. Tipo de pradera; M. Gestión; R. Riqueza; I. Individuos. State / County / Site Name Location GT M R I Iowa Allamakee County Fish Farm Mounds 91°17'12" W – 43°27'13" N Upland Unburned 21 632 Williams Creek 3 91°29'1" W – 43°8'1" N Upland Unburned 23 2,708 Bremer County Brayton–Horsley Fen 92°6'29" W – 42°48'36" N Lowland Burned 16 627 Buchanan County Rowley Fen 91°51'7" W – 42°22'27" N Lowland Unburned 16 3,217 Rowley North Fen 91°51'3" W – 42°22'35" N Lowland Unburned 17 3,231 Rowley West Fen 91°54'40" W – 42°22'15" N Lowland Unburned 22 2,250 Cerro Gordo County Buffalo Slough 93°11'11" W – 43°10'36" N Lowland Unburned 19 4,770 Chickasaw County Stapelton Church Fen 92°6'14" W – 43°1'35" N Lowland Unburned 18 1,065 Clayton County Postville Fen 91°33'59" W – 43°2'3" N Lowland Unburned 12 252 Turkey River Mounds 91°2'11" W – 42°42'46" N Upland Unburned 22 870 Clinton County Maquoketa South 90°39'5" W – 42°1'12" N Upland Unburned 12 310 Dubuque County Roosevelt Road 90°44'30" W – 42°32'55" N Upland Unburned 18 375 Fayette County Fayette 91°47'28" W – 42°50'11" N Upland Burned 13 254 Turner Creek 1 Fen 91°52'11" W – 42°58'15" N Lowland Unburned 16 1,071 Floyd County Beemis Creek 93°1'18" W – 42°59'39" N Upland Burned 8 192 Juniper Hill 92°59'2" W – 43°3'10" N Upland Unburned 12 206 Franklin County Hampton East 93°8'13" W – 42°43'42" N Upland Unburned 15 381 Howard County Hayden Prairie 92°23'4" W – 43°26'30" N Upland Burned 12 132 Staff Creek Fen 92°30'34" W – 43°26'41" N Lowland Unburned 15 1,599 Jackson County Hamilton Glade 90°34'9" W – 42°4'23" N Upland Unburned 15 340 Jones County Canton Glade 90°59'52" W – 42°10'46" N Upland Unburned 19 446 Linn County Baty Glade 91°39'14" W – 42°11'44" N Upland Burned 16 345 Paris Fen 91°35'42" W – 42°13'40" N Lowland Unburned 12 1,254 221 Animal Biodiversity and Conservation 25.2 (2002) 57 State / County / Site Name Location GT M R I Mitchell County Stone School Fen 92°38'11" W – 43°22'49" N Lowland Unburned 18 2,926 Winneshiek County Decorah Glade 91°46'11" W – 43°18'55" N Upland Unburned 18 605 Minnesota Becker County Audubon South Fen 95°58'47" W – 46°49'58" N Lowland Unburned 15 1,816 Callaway North 95°55'22" W – 47°3'57" N Upland Unburned 19 362 Greenwater Lake Fen 95°29'59" W – 46°59'20" N Lowland Unburned 20 2,132 Ogema West Fen 95°55'59" W – 47°6'32" N Lowland Unburned 16 5,001 Straight Lake 95°18'40" W – 46°58'40" N Upland Unburned 13 281 Beltrami County Fourtown Fen 95°18'21" W – 48°15'56" N Lowland Unburned 14 1,403 Clay County Barnesville WMA 96°17'34" W –46°43'5" N Upland Unburned 11 469 Barnesville WMA Fen 96°17'38" W – 46°43'9" N Lowland Unburned 13 436 Bjornson WMA 96°21'24" W – 46°45'44" N Lowland Unburned 14 436 Bluestem Prairie 96°28'45" W – 46°51'18" N Upland Burned 15 371 Felton Prairie 1 Fen 96°26'21" W – 47°3'51" N Lowland Burned 15 2,370 Felton Prairie 2 Fen 96°26'20" W – 47°4'0" N Lowland Unburned 14 3,131 Felton Prairie 96°26'1" W – 47°3'34" N Upland Unburned 5 63 Tansen 96°11'17" W – 46°42'14" N Upland Unburned 10 146 Clearwater County Bagley Lake Fen 95°14'35" W – 47°45'41" N Lowland Unburned 9 126 Filmore County Vesta Creek 91°45'0" W – 43°40'5" N Upland Unburned 21 1,151 Houston County Twin Pines Farm 91°22'45" W – 43°44'48" N Upland Unburned 24 591 Yucatan Twp. 91°38'28" W – 43°43'23" N Upland Unburned 20 765 Mahnomen County Eastlund Lake 95°47'5" W – 47°26'41" N Upland Unburned 13 490 Mahnomen North 95°58'8" W – 47°21'27" N Upland Unburned 18 806 Waubun SE 95°54'55" W – 47°9'57" N Lowland Unburned 18 2,915 Waubun SE 95°55'4" W – 47°10'5" N Upland Burned 8 220 Marshall County East Park WMA 96°16'44" W – 48°31'57" N Lowland Burned 14 735 Florian WMA 96°33'21" W – 48°26'33" N Lowland Unburned 17 3,923 Radium NE 96°32'38" W – 48°16'49" N Upland Unburned 12 493 Norman County Faith South 96°5'12" W – 47°15'42" N Lowland Unburned 16 3,047 Prairie Smoke Dunes 96°18'22" W – 47°27'44" N Upland Unburned 3 19 Sandpiper Prairie 96°24'22" W – 47°14'43" N Lowland Unburned 12 1,261 Table 1. (Cont.) 222 58 Nekola The central tendencies in these various relationships were graphically represented via box plots. In box plots, the central line represents the median of the sample, the margins of the box represent the interquartile distances, and the fences represent 1.5 times the interquartile distances. For data having a Gaussian distribution, approximately 99.3% of the data will fall inside of the fences (VELLEMAN & HOAGLIN, 1981). Outliers falling outside of the fences are shown with asterisks. The average number of individuals per species per site was determined for burned uplands, unburned uplands, burned lowlands, and unburned lowlands. The average proportion of each species in the total community for each site was calculated for each management/habitat type. These proportions were placed in rank order, and plotted vs. log–transformed frequency to create dominance–diversity curves (WHITTAKER, 1975). The response of individual species to fire was analyzed through log–linear modelling, as predicted values in the associated contingency table were sparse (< 5) in more than one–fifth of cells (ZAR, 1984). The total number of individuals within all burned or unburned sites was compared to a null expectation of equal occurrence Pennington County Goose Lake 96°27'44" W – 48°5'37" N Lowland Unburned 17 996 Higenbotham WMA 96°17'41" W – 48° 0'22" N Lowland Unburned 22 1,114 Sanders Fen 96°21'9" W – 48°3'52" N Lowland Unburned 15 2,218 Polk County Chicog Prairie 96°23'14" W – 47°35'53" N Upland Burned 2 153 Erskine North 96°0'3" W – 47°44'17" N Lowland Unburned 19 741 Gulley Fen 95°37'22" W – 47°48'13" N Lowland Unburned 19 2,032 Malmberg Prairie 96°49'25" W – 47°43'52" N Lowland Burned 7 563 Pankratz Prairie 96°26'37" W – 47°43'23" N Lowland Burned 12 314 Pankratz Prairie 96°26'31" W – 47°43'23" N Upland Burned 11 159 Pankratz Prairie 96°26'48" W – 47°43'9" N Lowland Burned 7 190 Red Lake County Crane WMA 95°42'49" W – 47°53'27" N Lowland Unburned 15 425 Cyr Creek 96°16'12" W – 47°48'10" N Lowland Unburned 22 1,845 Marcoux WMA 96°13'27" W – 47°47'55" N Lowland Burned 12 688 Winona County Great River Bluffs 91°23'28" W – 43°56'53" N Upland Burned 19 788 Wisconsin Green Lake County Berlin Fen 88°54'20" W – 43°57'47" N Lowland Unburned 20 3,454 Manitowoc County Point Beach St. Forest 87°30'40" W – 44°11'52" N Upland Unburned 4 6 Walworth County Bluff Creek Fen 88°40'54" W – 42°48'2" N Lowland Unburned 20 1,106 Washington County Allenton Fen 88°18'25" W – 43°22'42" N Lowland Unburned 20 2,858 Waushara County Bass Lake Fen 89°16'59" W – 44°0'16" N Lowland Unburned 19 1,466 Table 1. (Cont.) State / County / Site Name Location GT M R I 223 Animal Biodiversity and Conservation 25.2 (2002) 59 Table 2. List of encountered species, with their average abundances from burned and unburned sites. P–values are based on log-likelihood ratio tests, with the two–tailed significance threshold being lowered to p = 0.000347 to account for the 72 tested species. General ecological preferences are based on NEKOLA (in press a). Turf–specialists represent those species demonstrating at least a p < 0.05 preference to sites with a friable upper A soil horizon supporting few living plant roots. Turf specialists represent those species demonstrating at least a p < 0.05 preference to sites with an upper A soil horizon that is bound together with living plant roots. Species without preferences were too infrequently encountered by NEKOLA (in press a) to be statistically assigned: AvU. Average unburned; Abb. Abundance burned; Ecp. Ecological preference (T. Turf; D. Duff; G. Generalist.) Tabla 2. Lista de especies detectadas, con sus abundancias medias en áreas quemadas y no quemadas. Los valores de P se basan en tests de cociente de probabilidad logarítmica, con el umbral de significación de doble cola reducido hasta p = 0,000247 para las 72 especies estudiadas. Las preferencias ecológicas generales se basan en NEKOLA (in press a). Las especies que viven en sustrato herbáceo presentan una preferencia de al menos p < 0,05 por las zonas con horizonte de tierra friable superior A provisto de escasas raíces de plantas vivas. Las especies que viven en sustrato herbáceo presentan una preferencia de al menos p < 0,05 por las zonas cuyo horizonte se mantiene unido por raíces de plantas vivas. Las especies sin preferencias resultaron excesivamente infrecuentes según NEKOLA (in press a) para consignarlas estadísticamente: AvU. Medias en áreas no quemadas; Abb. Abundancia en áreas quemadas; Ecp. Preferencia ecológica (T. Sustrato herbáceo; D. Sustrato húmico; G. Generalista.) Species AvU Abb P–value Ecp Negative responses Carychium exiguum (Say, 1822) 273.607 90.250 0.0000000 T Carychium exile H. C. Lea, 1842 5.196 0.000 0.0000000 D Catinella exile (Leonard, 1972) 58.446 1.625 0.0000000 T Catinella "vermeta" 1.482 0.000 0.0000001 T Deroceras laeve (Müller, 1774) 4.036 1.188 0.0000050 G Discus cronkhitei (Newcomb, 1865) 16.143 5.000 0.0000000 G Euconulus alderi (Gray, 1840) 43.054 8.375 0.0000000 T Gastrocopta contracta (Say, 1822) 21.232 5.875 0.0000000 G Gastrocopta holzingeri (Sterki, 1889) 36.196 17.938 0.0000000 D Gastrocopta pentodon (Say, 1821) 9.661 4.875 0.0000072 D Gastrocopta procera Gould, 1840 2.304 0.000 0.0000000 T Gastrocopta rogersensis Nekola & Coles, 2001 5.518 0.062 0.0000000 T Gastrocopta similis (Sterki, 1909) 21.518 4.125 0.0000000 T Gastrocopta tappaniana (C. B. Adams, 1842) 112.929 33.375 0.0000000 T Hawaiia minuscula (A. Binney, 1840) 36.286 23.062 0.0000000 G Helicodiscus n. sp. 1.071 0.062 0.0001509 T Nesovitrea binneyana (Morse, 1864) 1.500 0.000 0.0000001 D Nesovitrea electrina (Gould, 1841) 80.179 22.688 0.0000000 T Oxyloma retusa (I. Lea, 1834) 21.268 8.750 0.0000000 T Pomatiopsis lapidaria (Say, 1817) 1.196 0.000 0.0000021 – Punctum minutissimum (I. Lea, 1841) 26.286 10.500 0.0000000 D Punctum n. sp. 41.982 12.625 0.0000000 T Punctum vitreum H. B. Baker, 1930 16.536 2.812 0.0000000 D Stenotrema leai leai (A. Binney) 2.696 0.375 0.0000004 T Striatura milium (Morse, 1859) 0.714 0.000 0.0002470 G Strobilops affinis Pilsbry, 1893 74.911 4.312 0.0000000 T Triodopsis multilineata (Say, 1821) 1.571 0.062 0.0000016 G Vallonia pulchella (Müller, 1774) 23.321 11.688 0.0000000 T 224 60 Nekola Species AvU Abb P–value Ecp Vertigo elatior Sterki, 1894 41.875 5.875 0.0000000 T Vertigo milium (Gould, 1840) 59.357 36.375 0.0000000 T Vertigo morsei Sterki, 1894 3.589 0.375 0.0000000 T Vitrina limpida Gould, 1850 1.143 0.000 0.0000035 G No response Anguispira alternata (Say, 1817) 0.018 0.000 0.5622176 D Catinella avara (Say, 1824) 7.286 8.000 0.5185276 T Cochlicopa lubrica (Müller, 1774) 0.464 0.000 0.0031253 D Cochlicopa lubricella (Porro, 1838) 1.714 1.938 0.6798546 G Columella simplex (Gould, 1841) 0.071 0.188 0.4068651 D Discus catskillensis (Pilsbry, 1898) 0.357 0.000 0.0095467 D Euconulus fulvus (Müller, 1774) 3.429 1.625 0.0040433 D Euconulus polygyratus (Pilsbry, 1899) 0.018 0.000 0.5622176 D Gastrocopta abbreviata (Sterki, 1909) 0.000 0.062 0.2039785 – Gastrocopta armifera (Say, 1821) 1.232 1.188 0.9197258 G Glyphyalinia indentata (Say, 1823) 2.732 3.250 0.4543545 D Haplotrema concavum (Say, 1821) 0.411 0.000 0.0054455 G Hawaiia n. sp. 2.571 1.750 0.1616094 T Helicodiscus inermis H. B. Baker, 1929 0.679 1.812 0.0089186 – Helicodiscus parallelus (Say, 1817) 5.393 6.438 0.2831711 G Helicodiscus shimeki Hubricht, 1962 0.286 0.000 0.0204383 D Helicodiscus singleyanus (Pilsbry, 1890) 0.375 1.125 0.0247064 G Hendersonia occulta (Say, 1831) 0.036 0.062 0.7606162 D Mesodon clausus clausus (Say, 1821) 0.054 0.000 0.3154693 D Oxyloma peoriensis (Wolf in Walker, 1892) 0.125 0.000 0.1251903 – Pupoides albilabris (C. B. Adams, 1821) 8.393 7.062 0.2324971 T Stenotrema barbatum (Clapp, 1904) 0.107 0.000 0.1557229 D Stenotrema fraternum fraternum (Say, 1824) 0.054 0.062 0.9262276 D Succinea indiana Pilsbry, 1905 0.000 0.188 0.0277912 – Succinea ovalis Say, 1817 0.143 0.188 0.7834904 D Triodopsis alleni (Wetherby in Sampson, 1883) 0.071 0.000 0.2464148 D Vallonia gracilicosta Reinhardt, 1883 11.500 11.062 0.7459095 D Vertigo arthuri (von Martens, 1884) 0.643 0.000 0.0005065 – Vertigo gouldi (A. Binney, 1843) 0.018 0.000 0.5622176 D Vertigo nylanderi Sterki, 1909 0.036 0.000 0.4124393 T Vertigo ovata Say, 1822 5.750 4.000 0.0472312 T Vertigo tridentata Wolf, 1870 0.375 0.062 0.0738709 D Zonitoides arboreus (Say, 1816) 4.143 4.625 0.5650976 D Zonitoides nitidus (Müller, 1774) 0.464 0.000 0.0031253 T Positive response Gastrocopta corticaria (Say, 1816) 0.732 3.500 0.0000003 D Strobilops labyrinthica (Say, 1817) 9.661 22.375 0.0000000 D Vallonia costata (Müller, 1774) 1.804 6.438 0.0000000 G Vallonia parvula Sterki, 1892 8.250 13.250 0.0001267 T Vallonia perspectiva Sterki, 1892 2.143 5.625 0.0000055 D Vertigo pygmaea (Draparnaud, 1801) 0.000 0.562 0.0001385 G Tabla 2. (Cont.) 225 Animal Biodiversity and Conservation 25.2 (2002) 61 Fig. 2. Box–plot diagram of the response of species richness and abundance to management type on all sampled sites. Fig. 2. Diagrama de la respuesta en riqueza y abundancia de especies al tipo de actuación desarrollado en todas las áreas estudiadas. 3030303030 2020202020 1010101010 00000 SpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichness Unburned BurnedUnburned BurnedUnburned BurnedUnburned BurnedUnburned Burned Unburned BurnedUnburned BurnedUnburned BurnedUnburned BurnedUnburned Burned Management typeManagement typeManagement typeManagement typeManagement type Management typeManagement typeManagement typeManagement typeManagement type 6,0006,0006,0006,0006,000 5,0005,0005,0005,0005,000 4,0004,0004,0004,0004,000 3,0003,0003,0003,0003,000 2,0002,0002,0002,0002,000 1,0001,0001,0001,0001,000 00000 TTTTTotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividuals p = 0.001p = 0.001p = 0.001p = 0.001p = 0.001 p = 0.008p = 0.008p = 0.008p = 0.008p = 0.008 frequency. This null expectation was calculated by assigning 88% of all encountered individuals to unburned sites, with the remaining 22% to burned sites. This procedure was necessary as the number of unburned vs. burned sites was not balanced (88% vs. 22%). A two–tailed significance threshold was employed so that species with positive and negative responses to fire could both be identified. As these analyses were repeated for each species, a Bonferroni correction was used to adjust the significance threshold. Differences between fire responses across the three general ecological preference types were documented via a contingency table, with significance being estimated using both log–linear modelling and Fisher’s Exact test. Results These grassland habitats were generally found to support a diverse and abundant land snail fauna. A total of 91,074 individuals in 72 different species were recovered from the 72 surveyed sites (tables 1, 2). The number of species per each 4 l litter sample ranged from two (Chicog gravel prairie) to 24 (Twin Pines Farm sandstone glade). Average richness ranged from roughly 15 in upland sites, to 17 in lowland. Snail abundance per site ranged from 6 (Point Beach State Forest dunes) to 5,001 (Ogema West fen). Average abundance ranged from roughly 500 in upland sites to 2,000 in lowlands. One–way ANOVA, using all sites, demonstrated that both species richness (p = 0.001) and abundance (p = 0.008) were significantly lower on sites that had experienced fire management (fig. 2). Median species richness was approximately 18 on unburned vs. 12 on burned sites. Likewise, median shell abundance was 1,000 on unburned vs. 300 on burned sites. Full 2–way ANOVA, using all sites and considering both management type and habitat type (upland vs. lowland) as independent variables, demonstrated a highly significant (p = 0.002) reduction (approximately 30%) in species richness in both upland and lowland sites (fig. 3). Habitat type and the interaction between habitat and fire history were not significant predictors (p = 0.209 and p = 0.628, respectively). Likewise, a significant (p = 0.010) reduction in shell abundance (50–70%) was noted on burned sites (fig. 3). In this case, however, habitat type was a more significant (p < 0.0005) predictor, with lowlands having 4–10 times the number of shells as uplands. Additionally, a marginally significant (p = 0.088) interaction between management and habitat was observed, with the reduction appearing to be roughly 50% greater in lowlands. Two–way ANOVA restricted to the 26 paired sites (fig. 4) demonstrated that even after blocking of variation due to site pair identity, a significant reduction in richness (p < 0.0005) and abundance (p = 0.015) still occurred on fire– managed sites. 226 62 Nekola Fig. 3. Box–plot diagram of the response of species richness and abundance to management and habitat type on all sampled sites. Fig. 3. Diagrama de la respuesta en riqueza y abundancia de especies al tipo de actuación desarrollado y el hábitat en todas las áreas estudiadas. Comparison of dominance–diversity diagrams for these sites (fig. 5) demonstrates that both burned upland and lowland sites have truncated curves, with the rarest 40–50% of species being much less common as compared to unburned sites. However, the more common species appear to have largely similar dominance–diversity diagrams. Contingency table analysis of individual species responses to fire (table 2) indicate that 32 (44%) experience a significant reduction in abundance on fire–managed sites, even following use of a Bonferroni–corrected two–tailed significance threshold (p = 0.000347). Only six species (8%) demonstrated positive responses to fire, while the remaining 34 (47%) demonstrated no ANOVANOVANOVANOVANOVA results: fire p = 0.002A results: fire p = 0.002A results: fire p = 0.002A results: fire p = 0.002A results: fire p = 0.002 habitat p = 0.209habitat p = 0.209habitat p = 0.209habitat p = 0.209habitat p = 0.209 interaction p = 0.628interaction p = 0.628interaction p = 0.628interaction p = 0.628interaction p = 0.628 ANOVANOVANOVANOVANOVA results:A results:A results:A results:A results: fire p = 0.010fire p = 0.010fire p = 0.010fire p = 0.010fire p = 0.010 habitat p < 0.0006habitat p < 0.0006habitat p < 0.0006habitat p < 0.0006habitat p < 0.0006 interaction p = 0.088interaction p = 0.088interaction p = 0.088interaction p = 0.088interaction p = 0.088 Lowland sitesLowland sitesLowland sitesLowland sitesLowland sites Upland sitesUpland sitesUpland sitesUpland sitesUpland sites 6,0006,0006,0006,0006,000 5,0005,0005,0005,0005,000 4,0004,0004,0004,0004,000 3,0003,0003,0003,0003,000 2,0002,0002,0002,0002,000 1,0001,0001,0001,0001,000 00000 3333300000 2020202020 1010101010 00000 SpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichness TTTTTotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividuals Unburned BurnedUnburned BurnedUnburned BurnedUnburned BurnedUnburned Burned Unburned BurnedUnburned BurnedUnburned BurnedUnburned BurnedUnburned Burned Management typeManagement typeManagement typeManagement typeManagement type Management typeManagement typeManagement typeManagement typeManagement type 227 Animal Biodiversity and Conservation 25.2 (2002) 63 Fig. 4. Box–plot diagram of the response of species richness and abundance to management on 26 sites paired by habitat type and geographic location. Fig. 4. Diagrama de la respuesta en riqueza y abundancia de especies a la actuación desarrollada en 26 áreas emparejadas según el tipo de hábitat y localización geográfica. Fig. 5. Dominance–diversity curve for upland/lowland sites which have been burned/unburned. Fig. 5. Curva de la dominancia–diversidad para las mesetas/llanuras que hayan sido quemadas/ no quemadas. 3030303030 2020202020 1010101010 00000 SpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichnessSpeciesrichness 6,0006,0006,0006,0006,000 5,0005,0005,0005,0005,000 4,0004,0004,0004,0004,000 2,0002,0002,0002,0002,000 3,0003,0003,0003,0003,000 1,0001,0001,0001,0001,000 00000 TTTTTotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividualsotal#ofindividuals Unburned BurnedUnburned BurnedUnburned BurnedUnburned BurnedUnburned Burned UnburnedUnburnedUnburnedUnburnedUnburned BurnedBurnedBurnedBurnedBurned Management typeManagement typeManagement typeManagement typeManagement type Management typeManagement typeManagement typeManagement typeManagement type ANOVANOVANOVANOVANOVA results:A results:A results:A results:A results: ANOVANOVANOVANOVANOVA results:A results:A results:A results:A results: Pairs p = 0.001Pairs p = 0.001Pairs p = 0.001Pairs p = 0.001Pairs p = 0.001 Pairs p = 0.009Pairs p = 0.009Pairs p = 0.009Pairs p = 0.009Pairs p = 0.009 Fire p < 0.0005Fire p < 0.0005Fire p < 0.0005Fire p < 0.0005Fire p < 0.0005 Fire p = 0.015Fire p = 0.015Fire p = 0.015Fire p = 0.015Fire p = 0.015 Lowland sitesLowland sitesLowland sitesLowland sitesLowland sites Upland sitesUpland sitesUpland sitesUpland sitesUpland sites –1–1–1–1–1 –4–4–4–4–4 –7–7–7–7–7 –10–10–10–10–10 00000 1010101010 2020202020 3030303030 4040404040 5050505050 6060606060 00000 10 20 3010 20 3010 20 3010 20 3010 20 30 4040404040 5050505050 6060606060 Rank orderRank orderRank orderRank orderRank order Rank orderRank orderRank orderRank orderRank order BurnedBurnedBurnedBurnedBurned UnburnedUnburnedUnburnedUnburnedUnburned ln(%frecuency)ln(%frecuency)ln(%frecuency)ln(%frecuency)ln(%frecuency) 228 64 Nekola significant changes in population size. Contingency table analysis of ecological preference vs. fire response indicated that fully 72% of turf– specialists were negatively impacted by fire (table 3). However, only 22% of duff–specialists exhibited negative responses. While 67% of duff– specialists demonstrated no significant response to fire only 24% of turf–specialists were unaffected. Generalist species demonstrated little discernable trend to fire, with seven decreasing, two increasing, and five with no response. Log– likelihood ratio and Fisher’s Exact tests both indicated these differences as being highly significant (p = 0.0006 and p = 0.004, respectively). Discussion These data clearly indicate that fire management causes significant reductions in land snail community richness and abundance in both upland and lowland grasslands throughout a significant section of the tallgrass prairie biome in central North America. At a species–level, fire most strongly impacts the rarest species, and causes significant population reductions in 44% of the 72 encountered taxa. These negative impacts were most strongly felt in turf–specialists, where almost 75% experienced significant reductions. Thus, statements regarding the benign nature of fire on snail populations (FREST & JOHANNES, 1995), and the beneficial impact of fire on North American grassland faunas (THELER, 1997) can be proven false. Rather, frequent use of fire management appears to represent a significant threat to the health and diversity of North American grassland land snails. It is not possible through these analyses to definitively identify the factors that directly lead to these impacts. However, at least part of the answer must lay in grassland detritusphere removal. This will lead to direct mortality, as the great majority of land snails are limited to this layer (HAWKINS et al., 1998). As land snail abundance (BERRY, 1973), diversity (CAIN, 1983; LOCASCIULLI & BOAG, 1987), and composition (CAMERON & MORGAN–HUWS, 1975; BAUER et al., 1996; BARKER & MAYHILL, 1999) is often positively correlated with litter depth, detritusphere removal would be expected to have a strong negative impact on land snail community structure. Redevelopment of an equilibrium thickness of organic detritus takes at least five years in southern Plains grasslands (KUCERA & KOELLING, 1964), with even longer intervals being required in more northern locations (HILL & PLATT, 1975). The optimal interval between fires for land snails might be even longer, depending upon the time required for more refractory plant debris (such as lignified grass stems) to break down, allowing a complete suite of decompositional microhabitats to develop. Litter architecture is known to effect snail community composition in forests of Virginia (BURCH, 1956), British Columbia (CAMERON, 1986), and Puerto Rico (ÁLVAREZ & WILLIG, 1993) and grasslands of England (YOUNG & EVANS, 1991). It should thus not be surprising that in the current data set, sites burned up to 15 years ago have maintained lowered land snail richness and abundance as compared to unburned sites. As grassland land snails presumably evolved in conjunction with natural fire regimes, it is also intriguing to note that turf–specialists experienced the most severe negative impacts to fire. If fire was a common process structuring central North American grasslands, evolution should have selected for individuals that were more tolerant of, or favored by, this disturbance. Like other native grassland invertebrate groups (SWENGEL, 1996; HARPER et al., 2000), land snails in the presettlement landscape may have been able to tolerate fires by being able to easily recolonize from source pools in adjacent unburned areas. Even when such adjacent source pools are present, recolonization may take over a dozen years (MÄND et al., 2001). In modern landscapes, where grasslands are highly fragmented and surrounded by agricultural, urban, or forest habitats, such recolonization has become much more difficult. Thus, turf–specialist taxa may continue to decrease in burned grasslands due to a lack of recolonization sources, while generalist and duff-specialist woodland taxa, which are more common in the surrounding landscape, may be able to maintain their populations through mass effect (SHMIDA & ELLNER, 1984). The depression of land snail richness and abundance following fire episodes, the length of time required to redevelop a mature detritusphere, and the greater sensitivity of turf– specialist taxa to fire casts doubt on the wide– Table 3. Contingency table analysis of fire response vs. general ecological preferences. Log–likelihood ratio p = 0.000634; Fisher’s Exact Test p = 0.004 (Ecological preferences: T. Turf; D. Duff; G. Generalist.) Tabla 3. Análisis de la tabla de contingencia de la respuesta al fuego frente a las preferencias ecológicas generales. Logaritmo de la razón de verosimilitudes p = 0,000634; Test exacto de Fisher p = 0,004 (Preferencias ecológicas: T. Sustrato herbáceo; D. Sustrato húmico; G. Generalista.) Ecological preferences Fire response T D G Negative 18 6 7 None 6 18 5 Positive 1 3 2 229 Animal Biodiversity and Conservation 25.2 (2002) 65 held belief (e.g., PAULY, 1985) that North American grasslands should be burned at 2–6 year intervals. Rather, these data support the contention that presettlement return intervals ranged between 20–30 years (SIMS, 1988). These data also strongly suggest that other factors, such as large herbivore grazing (COLLINS et al., 1998) and periodic drought (BORCHERT, 1950), may have also played essential roles in keeping prairies treeless, as these processes do not lead to the wholesale detritusphere removal. Protecting the health of North American grassland land snail populations will require the preservation of mulch layers on sites. Such efforts will also help protect a large percentage of the entire grassland soil biota. The detritusphere can only be protected if more realistic fire return intervals (20–30 years) are adopted by conservation agencies, and used in conjunction with more diversified approaches towards woody and invasive plant removal. Activities like grazing, haying, and hand cutting/pulling will not cause widespread removal of the detritusphere, and should thus be more compatible with land snail (and soil biodiversity) conservation. Acknowledgements Alyssa Barnes, Tracy Kuklinski, J. J. Schiefelbein and Angela Sette helped processed many soil litter samples, and assisted in shell counting. Additional assistance in litter processing was also provided by students of the Land Snail Ecology Practicum at the University of Wisconsin – Green Bay. Funding was provided by the Minnesota Nongame Wildlife Tax Checkoff and Minnesota State Park Nature Store Sales through the Minnesota Department of Natural Resources Natural Heritage and Nongame Research Program. References ÁLVAREZ, J. & WILLIG, M. R., 1993. Effects of treefall gaps on the density of land snails in the Luquillo Experimental Forest of Puerto Rico. Biotropica, 25: 100–110. BARKER, G. M. & MAYHILL, P. C., 1999. Patterns of diversity and habitat relationships in terrestrial mollusc communities of the Pukeamaru Ecological District, northeastern New Zealand. Journal of Biogeography, 26: 215–238. BAUER, B., JOSHI, J., SCHMID, B., HÄNGGI, A., BORCARD, D., STARY, J., PEDROLI–CHRISTEN, A., THOMMEN, G. H., LUKA, H., RUSTERHOLZ, H. P., OGGIER, P., LEDERGERBER, S. & ERHARDT, A., 1996. Variation in species richness of plants and diverse groups of invertebrates in three calcareous grasslands of the Swiss Jura mountains. Revue Suisse de Zooligie, 103: 801–833. BEHAN–PELLETIER, V. & NEWTON, G., 1999. Linking soil biodiversity and ecosystem function–the taxonomic dilemma. Bioscience, 49: 149–153. BERRY, F. G., 1973. Patterns of snail distribution at Ham Street Woods National Nature Reserve, East Kent. Journal of Conchology, 28: 23–35. BOUCHERT, J. R., 1950. The climate of the central North American grassland. Annals of the Association of American Geographers, 40: 1–39. BURCH, J. B., 1956. Distribution of land snails in plant associations in eastern Virginia. The Nautilus, 70: 60–64. BURCH, J. B. & PEARCE, T. A., 1990. Terrestrial Gastropoda. In: Soil biology guide: 201–309 (D. L. Dindal, Ed.). John Wiley & Sons, New York. CAIN, A. J., 1983. Ecology and ecogenetics of terrestrial molluscan populations. In: The Mollusca, Volume 6, Ecology: 597-647 (W. D. Russell–Hunter, Ed.). Academic Press, New York. CAMERON, R. A. D., 1986. Environment and diversities of forest snail faunas from coastal British Columbia. Malacologia, 27: 341–355. CAMERON, R. A. D. & MORGAN–HUWS, D. I., 1975. Snail faunas in the early stages of a chalk grassland succession. Biological Journal of the Linnean Society, 7: 215–229. COLEMAN, D. C. & CROSSLEY, D. A. JR., 1996. Fundamentals of soil ecology. Academic Press, New York. COLLINS, S. L., KNAPP, A. K., BLAIR, J. M. & STEINAUER, E. M., 1998. Modulation of diversity by grazing and mowing on native tallgrass prairie. Science, 280: 745–747. COLLINS, S. L. & WALLACE, L. L., 1990. Fire in North American tallgrass prairie. University of Oklahoma Press, Norman, Oklahoma. CURTIS, J. T., 1959. The vegetation of Wisconsin. U. Wisconsin Press, Madison. DIX, R. L., 1960. The effects of burning on the mulch structure and species composition of grasslands in western North Dakota. Ecology, 41: 49–56. EHRENREICH, J. H. & AIKMAN, J. M., 1963. An ecological study of the effect of certain management practices on native prairie in Iowa. Ecological Monographs, 33: 113–134. EMBERTON, K. C., PEARCE, T. A. & RANDALANA, R., 1996. Quantitatively sampling land–snail species richness in Madagascan rainforests. Malacologia, 38: 203–212. FREST, T. J., 1990. Final report, field survey of Iowa spring fens, contract #65–2454. Iowa Department of Natural Resources, Des Moines. – 1991. Summary status reports on eight species of candidate land snails from the Driftless Area (Paleozoic Plateau), Upper Midwest. Final Report, Contract #301–01366, USFWS Region 3, Ft. Snelling, Minnesota. FREST, T. J. & JOHANNES, E. J., 1995. Interior Columbia Basin mollusc species of special concern. Final Report, Contract #43–0E00–4–9112, Interior Columbia Basin Ecosystem Management Project, Walla Walla, Washington. HADLEY, E. B., 1970. Net productivity and burning responses of native eastern North Dakota 230 66 Nekola prairie communities. American Midland Naturalist, 84: 121–135. HARPER, M. G., DIETRICH, C. H., LARIMORE, R. L. & TESSENE, P. A., 2000. Effects of prescribed fire on prairie Arthropods: an enclosure study. Natural Areas Journal, 20: 325–335. HAWKINS, J. W., LANKESTER, M. W. & NELSON, R. R. A., 1998. Sampling terrestrial gastropods using cardboard sheets. Malacologia, 39: 1–9. HILL, G. R. & PLATT, W. J., 1975. Some effects of fire upon a tallgrass prairie plant community in northwestern Iowa. In: Prairie: a multiple view: 103–113 (M. K. Wali, Ed.). University of North Dakota Press, Grand Forks. HUBRICHT, L., 1985. The distributions of the native land molluscs of the eastern United States. Fieldiana, n.s., 24: 1–191. KUCERA, C. L. & KOELLING, M., 1964. The influence of fire on composition of central Missouri prairie. American Midland Naturalist, 72: 142–147. LOCASCIULLI, O. & BOAG, D. A., 1987. Microdistribution of terrestrial snails (Stylommatophora) in forest litter. Canadian Field Naturalist, 101: 76–81. MÄND, R., EHLVEST, A. & KIRISTAJA, P., 2001. Land snail in an afforested oil–shale mining area. Proceedings of the Estonian Academy of Sciences, Biology and Ecology, 50: 37–41. NEKOLA, J. C. (in press a). Large–scale Terrestrial Gastropod Community Composition Gradients in the Great Lakes region of North America. Journal of Biogeography. – (in press b). Terrestrial gastropod fauna of northeastern Wisconsin and the southern Upper Peninsula of Michigan. American Malacological Bulletin. NEKOLA, J. C. & SMITH, T. A., 1999. Terrestrial gastropod richness patterns in Wisconsin carbonate cliff communities. Malacologia, 41: 253–269. OGGIER, P., ZSCHOKKE, S. & BAUR, B., 1998. A comparison of three methods for assessing the gastropod community in dry grasslands. Pedobiologia, 42: 348–357. PAULY, W. R., 1985. How to manage small prairie fires. Dane County Environmental Council, Madison, Wisconsin. REGAN, T. J., REGAN, H. M., BONHAM, K., TAYLOR, R. J. & BURGMAN, M. A., 2001. Modelling the impact of timber harvesting on a rare carnivorous land snail (Tasmaphena lamproides) in northwest Tasmania, Australia. Ecological Modelling, 139: 253–264. ROOSA, D. M., 1984. Fury on the prairie. Iowa Conservationist, 43: 13. RUSSELL–HUNTER, W. D., 1983. Overview: planetary distribution of and the ecological constraints upon the mollusca. In: The Mollusca, Volume 6, Ecology: 1–27 (W. D. Russell–Hunter, Ed.). Academic Press, New York. SHMIDA, A. & ELLNER, S., 1984. Coexistence of plant species with similar niches. Vegetatio, 58: 29–55. SIMS, P. L., 1988. Grasslands. In: North American terrestrial vegetation: 266–286 (M. G. Barbour & W. D. Billings, Eds.). Cambridge University Press, New York. STANISIC, J., 1996. New land snails from boggomoss environments in the Dawson Valley, southeastern Queensland (Eupulmonata: Charopidae and Camaenidae). Memoirs of the Queensland Museum, 39: 343–354. STOWE, L. G. & TEERI, J. A., 1978. The geographic distribution of C4 species of the Dicotyledonae in relation to climate. American Naturalist, 112: 609–621. SWENGEL, A. B., 1996. Effects of fire and hay management on the abundance of prairie butterflies. Biological Conservation, 76: 73–85. – 1998. Comparisons of butterfly richness and abundance measures in prairie and barrens. Biodiversity and Conservation, 7: 1,639–1,659. THELER, J. L., 1997. The modern terrestrial gastropod (land snail) fauna of western Wisconsin’s hill prairies. The Nautilus, 110: 111–121. TOWNE, G. & OWENSBY, C., 1984. Long–term effects of annual burning at different dates in ungrazed Kansas tallgrass prairie. Journal of Range Management, 37: 392–397. VELLEMAN, P. F. & HOAGLIN, D. C., 1981. Applications, basics, and computing of exploratory data analysis. Addison–Wesley Press, Reading, Massachusetts. WEAVER, J. E., 1954. North American prairie. Johnson Publishing Company, Lincoln, Nebraska. WEAVER, J. E. & ROWLAND, N. W., 1952. Effects of excessive natural mulch on development, yield, and structure of native grassland. Botanical Gazette, 114: 1–19. WELTER–SCHULTES, F. W. & WILLIAMS, M. R., 1999. History, island area and habitat availability determine land snail species richness of Aegean islands. Journal of Biogeography, 26: 239–249. WHITTAKER, R. H., 1975. Communities and ecosystems. MacMillan Publishing, New York. YOUNG, M. S. & EVANS, J. G., 1991. Modern land mollusc communities from Flat Holm, South Glamorgan. Journal of Conchology, 34: 63–70. ZAR, J. H., 1984. Biostatistical analysis. Prentice– Hall, Inc., Englewood Cliffs, New Jersey. 231 Section VII: Human macroecology and sustainability The goal of creating a human society which exists in a sustainable fashion with the finite Earth requires not only that our actions be assessed for individual farms, villages, and cities, but also collectively across humanity. However, such a macroecological perspective is largely missing from sustainability studies. Because human civilization is a complex adaptive system which is maintained far from thermodynamic equilibrium, it requires energy inputs via vast quantities of increasingly exhausted fossil fuel stocks. It also requires other essential and non-substitutable commodities such as metal ores, radio-nucleotides, rare earths, phosphate fertilizer, arable land, and fresh water that are becoming evermore scarce. The dynamics of such systems are highly unpredictable. Small perturbations can cause wholesale changes, including total collapse. Ensuring that global civilization does not suffer this fate will require that valuable insights from the natural – and not just social – sciences be considered. Our most recent paper is a ecological consideration of the history food storage and transportation, published in Bioscience in the summer of 2015 (see right). The next project considers the ecological dynamics of human innovation, in particular whether high population density cities are needed to generate the technological advancements needed to support the ever growing human population. Representative Publications [number of citations as of October 27, 2017] Hammond, S.T., J.H. Brown, J.R. Burger, T.P. Flanagan, T.S. Fristoe, N. Mercado-Silva, J.C. Nekola & J.G. Okie. 2015. Food spoilage, storage and transport: implications for a sustainable future. Bioscience. 65:758–768. [12] Brown, J.H., J.R. Burger, W.R. Burnside, M. Chang, A.D. Davidson, T.S. Fristoe, M.J. Hamilton, S.T. Hammond, A. Kodric-Brown, N. Mercado-Silva, J.C. Nekola & J.G. Okie. 2014. Macroecology meets macroeconomics: resource scarcity and global sustainability. Ecological Engineering. 65:24-32. [32] Nekola, J.C., C.D. Allen, J.H. Brown, J.R. Burger, A.D. Davidson, T.S. Fristoe, M.J. Hamilton, S.T. Hammond, A. Kodric-Brown, N. Mercado-Silva & J.G. Okie. 2013. The Malthusian-Darwinian dynamic and the trajectory of civilization. Trends in Ecology and Evolution. 28:127-130. [41] Nekola, J.C., J.H. Brown, A. Kodric-Brown & J.G. Okie. 2013. Global sustainability vs. the MDD: a response to Rull. Trends in Ecology and Evolution. 28: 444. [3] Relationship between energetic costs and speed for goods transportation fueled by animal metabolism (black dots) and extrametabolic processes in marine (blue), terrestrial (blue), and aerial (red) domains. The trajectory of the relationship over time is shown within each of the latter three groups by best-fit vectors. 232 Burger, J.R., J.H. Brown, C.D. Allen,, W.R. Burnside, A.D. Davidson, T.S. Fristoe, M.J. Hamilton, N. Mercado-Silva, J.C. Nekola, J.G. Okie & W. Zuo. 2012. The macroecology of sustainability. PLoS Biology. 10:e1001345. [84] Brown, J.H., W.R. Burnside, A.D. Davidson, J.P. DeLong, W.C. Dunn, M.J. Hamilton, J.C. Nekola, J.G. Okie, N. Mercado-Silva, W.H. Woodruff & W. Zuo. 2011. Energetic Limits to Economic Growth. Bioscience. 61:19-26. [179] Smith, V.H., G.D. Tilman & J.C. Nekola. 1999. Eutrophication impacts of excess nutrient inputs on fresh water, marine, and terrestrial ecosystems. Environmental Pollution. 100: 179-196. [2017] 233 The Malthusian–Darwinian dynamic and the trajectory of civilization Jeffrey C. Nekola1 , Craig D. Allen2 , James H. Brown1,3 , Joseph R. Burger1 , Ana D. Davidson1,4 , Trevor S. Fristoe1 , Marcus J. Hamilton1,3,5 , Sean T. Hammond1 , Astrid Kodric-Brown1 , Norman Mercado-Silva1,6 , and Jordan G. Okie1,7 1 Department of Biology, University of New Mexico, Albuquerque, NM, USA 2 US Geological Survey, Fort Collins Science Center, Jemez Mountains Field Station, Los Alamos, NM, USA 3 Santa Fe Institute, Santa Fe, NM, USA 4 Department of Ecology and Evolution, Stony Brook University, Stony Brook, NY, USA 5 Department of Anthropology, University of New Mexico, Albuquerque, NM, USA 6 School of Natural Resources and the Environment, University of Arizona, Tucson, AZ, USA 7 School of Earth and Space Exploration, Arizona State University, Tempe, AZ, USA Two interacting forces influence all populations: the Malthusian dynamic of exponential growth until resource limits are reached, and the Darwinian dynamic of innovation and adaptation to circumvent these limits through biological and/or cultural evolution. The specific manifestations of these forces in modern human society provide an important context for determining how humans can establish a sustainable relationship with the finite Earth. Malthus, Darwin, and population dynamics In 1798 Thomas Malthus laid out the concept of exponential population growth that became the foundation of demography and population biology. He noted that the ‘increase of population is necessarily limited by the means of subsistence.’ Population growth can thus continue only as long as environmental conditions remain favorable. As numbers increase, sooner or later environmental limits cause birth rates to decrease and/or death rates to increase, ultimately leading to an end to population growth. These concepts profoundly influenced Charles Darwin half a century later: because more offspring are born than can survive, only the fittest individuals reproduce and pass their superior traits on to their offspring. The result is adaption or innovation in the form of either genetic or cultural evolution. The Malthusian dynamic pushes a population to increase until it reaches its environmental limits. The Darwinian dynamic pushes against these limits by incorporating new traits and technologies that enhance survival and reproduction. There are restrictions to this Malthusian–Darwinian Dynamic (MDD), however: it is logically, physically, and biologically impossible for exponential growth to continue indefinitely within a finite world. Ecological and historical perspective on the rise of human civilization Humans are an exceptional species. Our population has increased almost continuously from less than a million individuals in sub-Saharan Africa 50 000 years ago to a current population of 7 billion spanning the entire globe. The human population is projected to reach between 9 and 10 billion by 2050 [1]. In the process, humans have created complex social, technological, and economic systems. We have transformed the atmosphere, water, land, and biodiversity of the planet. We have become the most dominant single species the earth has ever seen. Our success to date is a consequence of our ability to continually develop new innovations that push back environmental limits and to transfer this information across generations (Figure 1). Operating in this way, the MDD has led to rapid cultural evolution and made possible the transition from hunter–gatherer to agricultural to industrial– technological–informational economies. A central feature of human ecology has been the positive feedback between growth and innovation. As populations grew and aggregated into larger and more complex social groups, more information was acquired and processed. This led to new technologies that further pushed back ecological limits, allowing for continued population growth. The result has been an ascending spiral of exponential processes feeding back on each other: population growth and aggregation begot technological innovation, which in turn allowed for more resource extraction and a greater ability to overcome ecological constraints, begetting still more population growth and socioeconomic development [2]. Our ability to evade local resource shortages and population crashes through trade, migration, and innovation has allowed for continued growth not just of the global population but also of its resource use and economic activity [3]. For more than 200 years, so-called Malthusians [4] have argued, however, that this cannot indefinitely continue because essential resources supplied by the finite Earth will ultimately become limiting. This Malthusian perspective has historically been countered by so-called Cornucopians, who have argued that there is no hard limit to human population size and economic activity because human ingenuity and technological innovation provide an effectively infinite capacity to increase resource supply [5]. Forum: Science & Society Corresponding author: Nekola, J.C. (jnekola@unm.edu). Keywords: resource limitation; behavioral constraints; sustainability science; macroecology; evolutionary biology. 127 Author's Personal Copypy234 Malthus warned of population and economic collapse in the British Isles because the exponentially increasing population would eventually outstrip linearly increasing food supplies. Although millions did die from widespread famines in Ireland and the Scottish Highlands in the mid-19th century, more widespread starvation was averted because of newly acquired agricultural technologies and emigration of people to overseas colonies [2]. As a result, the British economy did not collapse and by 1900 had become the dominant global empire. In the 1960s and 1970s Malthusians again warned that the global human population was nearing its environmental limits. Outbreaks of famine and disease did occur, but advances in agriculture, medicine, technology, and commerce – most notably the green revolution and economic globalization – pushed back the limits again and our population has more than doubled from approximately three billion in 1960 to more than seven billion today [1]. Our ability to evade previous environmental limits through the MDD should not imply, however, that such outcomes are inevitable. First, although the MDD has served as the engine for our extraordinary success, it also contains within it the potential seeds for our ultimate downfall. Motives driven by the MDD, such as selfishness and cheating, benefit individuals at the expense of society as a whole (Box 1). Second, because natural selection can only operate on current conditions, there is little tendency to recognize approaching environmental limits and curb growth before resources are overexploited [6]. Third, the assertion that adaptation and innovation will always prevent collapse – because they have in the past – is logically untenable and akin to the statistical fallacy of extrapolating beyond a data set. Whether we wish to acknowledge it or not, there are hard limits to innovation: according to the first law of thermodynamics, we cannot create energy or matter; according to the second law of thermodynamics, large energy inputs are required to maintain highly organized systems. The ruins of Mohenjo Daro, Mesopotamia, Egypt, Greece, Rome, the Maya, Hohokam, Angkor Wat, and Easter Island are enduring evidence that many earlier societies were unable to innovate their way out of local limits and therefore collapsed despite attaining dense populations and advanced cultures. Although the proximate causes of these declines are debated and undoubtedly differ, the ultimate causes lie in an inability to sustain resource supplies and protect against parasites, diseases, and other human groups [7]. Until now, both Malthusians and Cornucopians have been correct: some populations have crashed and cultures have vanished, but our species has endured because these events have been localized. However, behavioral changes and technological innovations over the last century now intricately interconnect us in a single global society. As a result, local perturbations currently have the ability to reverberate across all of humanity. For instance, the 2008 meltdown of the USA real estate and mortgage markets and the 2011 To¯hoku earthquake and tsunami in Japan both led to global interruptions in economic activity, impacting most individuals in some way. Within the context of our now highly globalized society, the essential question is how much potential exists for the Bronze: IrrigaƟon: CiƟes: Wheel: Bow & arrow: Animal domesƟcaƟon: 5000 5000 5700 6000 10000 10000 0 10000 8000 6000 4000 Years before present GlobalpopulaƟon(Billions) 2000 0 1 2 3 4 5 6 7 Steam engine: Windmill: Gunpowder: Water mills: Roads: Metal plows: 300 1200 1200 2300 2500 2500 Green revoluƟon: Computer: AnƟbioƟcs: PlasƟcs: Airplane: Internal combusƟon engine: Vaccine: 40 60 70 90 100 150 200 years ago TRENDS in Ecology & Evolution Bronze: IrrigaƟon: CiƟes: Wheel: Bow & arrow: Animal domesƟcaƟon: 5000 5000 5700 6000 10000 10000 Steam engine: Windmill: Gunpowder: Water mills: Roads: Metal plows: 300 1200 1200 2300 2500 2500 Green revoluƟon: Computer: AnƟbioƟcs: PlasƟcs: Airplane: Internal combusƟon engine: Vaccine: 40 60 70 90 100 150 200 years ago Figure 1. Global human population over the last 10 000 years. Data are based on estimates from [11], the US census bureau, and http://en.wikipedia.org/wiki/ World_population_estimates. Average estimates of population size across these sources are plotted. Examples of major innovations that have helped to expand human carrying capacity (based in part on [12]) are listed for three time periods in years before present (ybp): 10 000–3000 ybp (green), 3000–300 ybp (orange), and 300 ybp–present (red). Box 1. The MDD and challenges to global sustainability Although the MDD has spurred innovations and adaptations that have allowed us to colonize the entire planet, it has also shaped our biological makeup, selecting for traits and behaviors that favor selfishness and may ultimately make us ill-equipped to act in concert to address global issues. Natural selection generates the tendency for individuals to preferentially promote themselves and their families. Social groups often establish acceptable behaviors to minimize the inherent conflict between such individual self-interest and group welfare. Although disobeying group rules can result in increased individual fitness, this is counteracted by the detection and punishment of cheaters by social groups. Whether through luck or drive, some will still manage to possess a disproportionate share of resources. Although the degree of inequality can be moderated, highly asymmetric wealth distributions have proven robust to the best efforts of social philosophies and political movements to impose more egalitarian patterns [13]. All of these factors contribute to the tragedy of the commons, which occurs when selfish individuals reap the benefits from shared resources while spreading the costs across the group [14]. The increase in global atmospheric CO2 concentrations and the reduction in global fisheries stocks (Figure 2) both represent pressing global environmental issues related to this mechanism. In the case of CO2 emissions, individuals benefit from using carbon-based energy while global society is burdened by the costs related to increased atmospheric CO2 concentrations. Similarly, a single fishing vessel has no incentive to curb harvests because the costs of reduced stocks are paid for by humanity as a whole. Efforts to implement and enforce international regulations to moderate these two activities have failed because of the selfishness of countries and individuals in combination with the difficulty of detecting cheating at the global scale [15]. Forum: Science & Society Trends in Ecology & Evolution March 2013, Vol. 28, No. 3 128 Author's Personal Copypy235 Darwinian side of the MDD to allow for continued adaptation and innovation to push back against global scale constraints. Many are now asking questions such as ‘What are the limits to growth?’ and ‘What will happen when these limits are met?’ [5,8]. The road forward We cannot provide definitive answers to these questions. Contemporary human civilization is a complex adaptive system, maintained far from thermodynamic equilibrium largely via the throughput of vast quantities of increasingly exhausted fossil fuel stocks [9]. This system also requires other essential and non-substitutable commodities such as metal ores, radionucleotides, rare earth elements, phosphate fertilizer, arable land, and fresh water that are becoming ever more scarce [10]. The dynamics of such systems are highly unpredictable. Small perturbations can cause wholesale changes, including total collapse. The bad news is that the MDD has left humans ill prepared to make the necessary ecological and behavioral changes required to avoid civilization collapse. The universal underlying biological imperatives of the MDD lie at the root of many of the most challenging impediments to longterm sustainability: exponential population growth, exploitation of all available resources, and the expression of behaviors that promote the competitive abilities of individuals, their families, and social groups over the species as a whole. However,thegoodnewsisthattheMDDmayalsoprovide valuable insights into potential solutions from both natural (in particular evolutionary biology and ecology) and social (in particular economics and sociology) science perspectives. We must recognize that a sustainable future will ultimately require: (i) negative population growth for a number of generations, followed by zero growth; (ii) a steady-state economy based on sustainable use of renewable energy and material resources; and (iii) new social norms that favor the welfare of the entire global population over that of specific individuals and groups. It is also essential that we recognize that humanity has not yet evolved the genetic or cultural adaptations needed to accomplish these tasks. Our exceptional brains give us the ability to appreciate the present situation and envision alternate futures before catastrophe occurs. The challenge will be to facilitate a rapid cultural evolution that, for the good of the entire species, rewards individual sacrifices in fitness and quality of life. Genuine collaboration between natural and social scientists will be essential to inform society as a whole – and policy makers specifically – of this difficult but necessary adaptation required to accommodate our species in a finite and now full world. References 1 Bongaarts, J. (2009) Human population growth and the demographic transition. Philos. Trans. R. Soc. B: Biol. Sci. 364, 2985–2990 2 Boserup, E. (1981) Population and Technological Change, University of Chicago Press 3 Hall, C.A.S. and Klitgaard, K.A. (2011) Energy and the Wealth of Nations: Understanding the Biophysical Economy, Springer 4 Ehrlich, P.R. (1968) The Population Bomb, Sierra Club/Ballantine Books 5 Simon, J.L. (1981) The Ultimate Resource, Princeton University Press 6 Levin, S.A. (2000) Fragile Dominion: Complexity and the Commons, Perseus Publishing 7 Tainter, J.A. (1988) The Collapse of Complex Societies, Cambridge University Press 8 Heinberg, R. (2011) The End of Growth: Adapting to Our New Economic Reality, New Society Publishers 1960 1970 1980 Year Wildfisheriesharvest(tonnes) AtmosphericconcentraƟonofCO2(ppm) GlobalhumanpopulaƟon 1990 2000 2010 2e+6 3e+6 4e+6 5e+6 6e+6 7e+6 8e+6 1e+7 2e+7 3e+7 4e+7 5e+7 6e+7 7e+7 8e+7 300 400 380 360 340 320 Human populaƟon Key: Fisheries CO2 TRENDS in Ecology & Evolution Figure 2. The trajectories of atmospheric CO2 and wild fisheries harvest in relation to global population since 1955. Note that the increase in CO2 concentration has accelerated, whereas fisheries harvest reached a peak in the 1990s and has since declined. The world population size is from the World Resources Institute (http:// earthtrends.wri.org). The wild fisheries harvest data are from the FAO Fishery Statistical Collection Global Capture Production Database (http://www.fao.org/fishery/ statistics/global-capture-production/en) and are limited to diadromous and marine species. Yearly mean CO2 concentrations as measured at the Mauna Loa Observatory were obtained from ftp://ftp.cmdl.noaa.gov/ccg/co2/trends/co2_annmean_mlo.txt. Forum: Science & Society Trends in Ecology & Evolution March 2013, Vol. 28, No. 3 129 Author's Personal Copypy236 9 Smil, V. (2008) Energy in Nature and Society: General Energetics of Complex Systems, MIT Press 10 Burger, J.R. et al. (2012) The macroecology of sustainability. PLoS Biol. 10, e1001345 11 Durand, J.D. (1977) Historical estimates of world population: an evaluation. Popul. Dev. Rev. 3, 253–296 12 Vermeij, G.J. and Leigh, E.G., Jr (2011) Natural and human economies compared. Ecosphere 2, art39 13 Chatterjee, A. et al., eds (2005) Econophysics of Wealth Distributions, Springer Verlag 14 Hardin, G. (1968) The tragedy of the commons. Science 162, 1243–1248 15 Ostrom, E. et al. (1999) Revisiting the commons: local lessons, global challenges. Science 284, 278–282 0169-5347/$ – see front matter ß 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.tree.2012.12.001 Trends in Ecology & Evolution, March 2013, Vol. 28, No. 3 Forum: Science & Society Trends in Ecology & Evolution March 2013, Vol. 28, No. 3 130 Author's Personal Copypy237 Essay The Macroecology of Sustainability Joseph R. Burger1 *, Craig D. Allen2 , James H. Brown1,3 , William R. Burnside1 , Ana D. Davidson1,4 , Trevor S. Fristoe1 , Marcus J. Hamilton1,3,5 , Norman Mercado-Silva1,6 , Jeffrey C. Nekola1 , Jordan G. Okie1¤ , Wenyun Zuo1 1 Department of Biology, University of New Mexico, Albuquerque, New Mexico, United States of America, 2 Geological Survey, Fort Collins Science Center, Jemez Mountains Field Station, Los Alamos, New Mexico, United States of America, 3 Santa Fe Institute, Santa Fe, New Mexico, United States of America, 4 Instituto de Ecologı´a, Universidad Nacional Auto´noma de Me´xico, Me´xico Federal, Me´xico, 5 Department of Anthropology, University of New Mexico, Albuquerque, New Mexico, United States of America, 6 School of Natural Resources and the Environment, University of Arizona, Tucson, Arizona, United States of America abstract: The discipline of sustainability science has emerged in response to concerns of natural and social scientists, policymakers, and lay people about whether the Earth can continue to support human population growth and economic prosperity. Yet, sustainability science has developed largely independently from and with little reference to key ecological principles that govern life on Earth. A macroecological perspective highlights three principles that should be integral to sustainability science: 1) physical conservation laws govern the flows of energy and materials between human systems and the environment, 2) smaller systems are connected by these flows to larger systems in which they are embedded, and 3) global constraints ultimately limit flows at smaller scales. Over the past few decades, decreasing per capita rates of consumption of petroleum, phosphate, agricultural land, fresh water, fish, and wood indicate that the growing human population has surpassed the capacity of the Earth to supply enough of these essential resources to sustain even the current population and level of socioeconomic development. ‘‘Sustainability’’ has become a key concern of scientists, politicians, and lay people—and for good reason. There is increasing evidence that we have approached, or perhaps even surpassed, the capacity of the planet to support continued human population growth and socioeconomic development [1–3]. Currently, humans are appropriating 20%–40% of the Earth’s terrestrial primary production [4–6], depleting finite supplies of fossil fuels and minerals, and overharvesting ‘‘renewable’’ natural resources such as fresh water and marine fisheries [7–10]. In the process, we are producing greenhouse gases and other wastes faster than the environment can assimilate them, altering global climate and landscapes, and drastically reducing biodiversity [2]. Concern about whether current trajectories of human demography and socioeconomic activity can continue in the face of such environmental impacts has led to calls for ‘‘sustainability.’’ A seminal event was the Brundtland commission report [11], which defined ‘‘sustainable development (as) development that meets the needs of the present without compromising the ability of future generations to meet their own needs.’’ One result has been the emergence of the discipline of sustainability science. ‘‘Sustainability science (is) an emerging field of research dealing with the interactions between natural and social systems, and with how those interactions affect the challenge of sustainability: meeting the needs of present and future generations while substantially reducing poverty and conserving the planet’s life support systems’’ (Proceedings of the National Academy of Sciences of the USA [PNAS], http://www. pnas.org/site/misc/sustainability.shtml). It is the subject of numerous books, at least three journals (Sustainability Science [Springer]; Sustainability: Science, Practice, & Policy [ProQuest-CSA]; International Journal of Sustainability Science and Studies [Polo Publishing]), and a special section of the PNAS. In ‘‘A Survey of University-Based Sustainability Science Programs’’, conducted in 2007, (http://sustainabilityscience.org/ content.html?contentid =1484), the American Association for the Advancement of Science listed 103 academic programs, including 64 in the United States and Canada, and many more have been established subsequently. Interestingly, despite the above definition, the majority of sustainability science appears to emphasize social science while largely neglecting natural science. A survey of the published literature from 1980 through November 2010 using the Web of Science reveals striking results. Of the 23,535 published papers that include ‘‘sustainability’’ in the title, abstract, or key Essays articulate a specific perspective on a topic of broad interest to scientists. Citation: Burger JR, Allen CD, Brown JH, Burnside WR, Davidson AD, et al. (2012) The Macroecology of Sustainability. PLoS Biol 10(6): e1001345. doi:10.1371/journal.pbio.1001345 Academic Editor: Georgina M. Mace, Imperial College London, United Kingdom Published June 19, 2012 This is an open-access article, free of all copyright, and may be freely reproduced, distributed, transmitted, modified, built upon, or otherwise used by anyone for any lawful purpose. The work is made available under the Creative Commons CC0 public domain dedication. Funding: Financial support was provided by the Howard Hughes Medical Institute (Interfaces grant to JHB, JGO, and WZ), National Institutes of Health (grant T32EB009414 to JHB and JRB), National Science Foundation (grant OISE-0653296 to ADD and grant DEB-0541625 to MJH), and Rockefeller Foundation (grant to MJH). The content is solely the responsibility of the authors and does not necessarily reflect the views of the organizations listed above. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript. Competing Interests: The authors have declared that no competing interests exist. Abbreviations: GDP, gross domestic product * E-mail: jrburger@unm.edu ¤ Current address: School of Earth and Space Exploration, Arizona State University, Tempe, Arizona, United States of America PLoS Biology | www.plosbiology.org 1 June 2012 | Volume 10 | Issue 6 | e1001345 238 words, 48% include ‘‘development’’ or ‘‘economics’’. In contrast, only 17% include any mention of ‘‘ecology’’ or ‘‘ecological’’, 12% ‘‘energy’’, 2% ‘‘limits’’, and fewer than 1% ‘‘thermodynamic’’ or ‘‘steady state’’. Any assessment of sustainability is necessarily incomplete without incorporating these concepts from the natural sciences. Human Macroecology A macroecological approach to sustainability aims to understand how humans are integrated into and constrained by the Earth’s systems [12]. In just the last 50,000 years, Homo sapiens has expanded out of Africa to become the most dominant species the Earth has ever experienced. Near-exponential population growth, global colonization, and socioeconomic development have been fueled by extracting resources from the environment and transforming them into people, goods, and services. Hunter-gatherers had subsistence economies based on harvesting local biological resources for food and fiber and on burning wood and dung to supplement energy from human metabolism. With the transition to agricultural societies after the last ice age [13] and then to industrial societies within the last two centuries, per capita energy use has increased from approximately 120 watts of human biological metabolism to over 10,000 watts, mostly from fossil fuels [3,14]. Modern economies rely on global networks of extraction, trade, and communication to rapidly distribute vast quantities of energy, materials, and information. The capacity of the environment to support the requirements of contemporary human societies is not just a matter of political and economic concern. It is also a central aspect of ecology— the study of the interactions between organisms, including humans, and their environments. These relationships always involve exchanges of energy, matter, or information. The scientific principles that govern the flows and transformations of these commodities are fundamental to ecology and directly relevant to sustainability and to the maintenance of ecosystem services, especially in times of energy scarcity [15]. A macroecological perspective highlights three principles that should be combined with perspectives from the social sciences to achieve an integrated science of sustainability. Principle 1: Thermodynamics and the Zero-Sum Game The laws of thermodynamics and conservation of energy, mass, and chemical stoichiometry are universal and without exception. These principles are fundamental to biology and ecology [16–18]. They also apply equally to humans and their activities at all spatial and temporal scales. The laws of thermodynamics mean that continual flows and transformations of energy are required to maintain highly organized, far-from-equilibrium states of complex systems, including human societies. For example, increased rates of energy use are required to fuel economic growth and development, raising formidable challenges in a time of growing energy scarcity and insecurity [3,15,19]. Conservation of mass and stoichiometry means that the planetary quantities of chemical elements are effectively finite [15,18]. Human use of material resources, such as nitrogen and phosphorus, alters flows and affects the distribution and local concentrations in the environment [18]. This is illustrated by the Bristol Bay salmon fishery, which is frequently cited as a success story in sustainable fisheries management [20,21]. In three years for which good data are available (2007– 2009), about 70% of the annual wild salmon run was harvested commercially, with one species, sockeye, accounting for about 95% of the catch [22]. From a management perspective, the Bristol Bay sockeye fishery has been sustainable, because annual runs have not declined. Additional implications for sustainability, however, come from considering the effect of human harvest on the flows of energy and materials in the upstream ecosystem (Figure 1). When humans take about 70% of Bristol Bay sockeye runs as commercial catch, this means a 70% reduction in the number of mature salmon returning to their native waters to spawn and complete their life cycles. It also means a concomitant reduction in the supply of salmon to support populations of predators, such as grizzly bears, bald eagles, and indigenous people, all of which historically relied on salmon for a large proportion of their diet [23,24]. Additionally, a 70% harvest means annual removal of more than 83,000 metric tonnes of salmon biomass, consisting of approximately 12,000, 2,500, and 330 tonnes of carbon, nitrogen, and phosphorus, respectively (see Text S1 for sources and calculations). These marine-derived materials are no longer deposited inland in the Bristol Bay watershed, where they once provided important nutrient subsidies to stream, lake, riparian, and terrestrial ecosystems [24–27]. So, for example, one apparent consequence is that net primary production in one oligotrophic lake in the Bristol Bay watershed has decreased ‘‘to about 1/3 of its level before commercial fishing’’ [28]. Seventy percent of Bristol Bay salmon biomass and nutrients are now exported to eastern Asia, western Europe, and the continental US, which are the primary markets for commercially harvested wild Alaskan salmon. Our macroecological assessment of the Bristol Bay fishery suggests that ‘‘sustainable harvest’’ of the focal salmon species does not consider the indirect impacts of human take on critical resource flows in the ecosystem (Figure 1). So the Bristol Bay salmon fishery is probably not entirely sustainable even at the ‘‘local’’ scale. Principle 2: Scale and Embeddedness Most published examples of sustainability focus on maintaining or improving environmental conditions or quality of life in a localized human system, such as a farm, village, city, industry, or country ([29,30] and articles following [31]). These socioeconomic systems are not closed or isolated, but instead are open, interconnected, and embedded in larger environmental systems. Human economies extract energy and material resources from the environment and transform them into goods and services. In the process, they create waste products that are released back into the environment. The laws of conservation and thermodynamics mean that the embedded human systems are absolutely dependent on these flows: population growth and economic development require increased rates of consumption of energy and materials and increased production of wastes. The degree of dependence is a function of the size of the economy and its level of socioeconomic development [3]. Most organic farms import fuel, tools, machinery, social services, and even fertilizer, and export their products to markets. A small village in a developing country harvests food, water, and fuel from the surrounding landscape. Large, complex human systems, such as corporations, cities, and countries, are even more dependent on exchanges with the broader environment and consequently pose formidable challenges for sustainability. Modern cities and nation states are embedded in the global economy, and supported by trade and communication networks that transport people, other organisms, energy, materials, and information. High densities of people and concentrations of socioeconomic activities require massive inputs of energy and materials and produce proportionately large amounts of wastes. Claims that such PLoS Biology | www.plosbiology.org 2 June 2012 | Volume 10 | Issue 6 | e1001345 239 systems are ‘‘sustainable’’ usually only mean that they are comparatively ‘‘green’’—that they aim to minimize environmental impacts while offering their inhabitants happy, healthy lifestyles. A macroecological perspective on the sustainability of local systems emphasizes their interrelations with the larger systems in which they are embedded, rather than viewing these systems in isolation. Portland, Oregon offers an illuminating example. The city of Portland and surrounding Multnomah County, with a population of 715,000 and a median per capita income of US$51,000, bills itself and is often hailed by the media as ‘‘the most sustainable city in America’’ (e.g., SustainLane.com, 2008). On the one hand, there can be little question that Portland is relatively green and offers its citizens a pleasant, healthy lifestyle, with exemplary bike paths, parks, gardens, farmers’ markets, and recycling programs. About 8% of its electricity comes from renewable non-hydroelectric sources (http://apps3.eere.energy.gov/ greenpower/resources/tables/topten.shtml). On the other hand, there also can be no question that Portland is embedded in and completely dependent on environments and economies at regional, national, and global scales (Figure 2). A compilation and quantitative analysis of the flows into and out of the city are informative (see Text S1 for sources and calculations). Each year the Portland metropolitan area consumes at least 1.25 billion liters of gasoline, 28.8 billion megajoules of natural gas, 31.1 billion megajoules of electricity, 136 billion liters of water, and 0.5 million tonnes of food, and the city releases 8.5 million tonnes of carbon as CO2, 99 billion liters of liquid sewage, and 1 million tonnes of solid waste into the environment. Total domestic and international trade amounts to 24 million tonnes of materials annually. With respect to these flows, Portland is not conspicuously ‘‘green’’; the above figures are about average for a US city of comparable size (e.g., [32]). A good way to see the embedding problem is to imagine the consequences of cutting off all flows in and out, as military sieges of European castles and cities attempted to do in the Middle Ages. From this point of view and in the short term of days to months, some farms and ranches would be reasonably sustainable, but the residents of a large city or an apartment building would rapidly succumb to thirst, starvation, or disease. Viewed from this perspective, even though Portland may be the greenest and by some definitions ‘‘the most sustainable city in America’’, it is definitely not self-sustaining. Massive flows of energy and materials across the city’s boundaries are required just to keep its residents alive, let alone provide them with the lifestyles to which they have become accustomed. Any complete ecological assessment of the sustainability of a local system should consider its connectedness with and dependence on the larger systems in which it is embedded. Figure 1. Pictorial illustration of important flows of salmon and contained biomass, energy, and nutrients within and out of the Bristol Bay ecosystem. Brown arrows depict the flows within the ecosystem, green arrows depict inputs due to growth in fresh water or the sea, and red arrows represent human harvest. Seventy percent of salmon are extracted by humans and are no longer available to the Bristol Bay ecosystem. doi:10.1371/journal.pbio.1001345.g001 PLoS Biology | www.plosbiology.org 3 June 2012 | Volume 10 | Issue 6 | e1001345 240 Principle 3: Global Constraints For thousands of years, humans have harvested fish, other animals, and plants with varying degrees of ‘‘sustainability’’ and lived in settlements that depend on imports and exports of energy and materials. Throughout history, humans have relied on the environment for goods and services and used trade to compensate for imbalances between extraction, production, and consumption at local to regional scales. What is different now are the enormous magnitudes and global scales of the fluxes of energy and materials into and out of human systems. Every year fisheries export thousands of tonnes of salmon biomass and the contained energy and nutrients from the Bristol Bay ecosystem to consumers in Asia, Europe, and the US. Every year Portland imports ever larger quantities of energy and materials to support its lifestyle and economy. Collectively, such activities, replicated thousands of times across the globe, are transforming the biosphere. Can the Earth support even current levels of human resource use and waste production, let alone provide for projected population growth and economic development? From our perspective, this should be the critical issue for sustainability science. The emphasis on local and regional scales—as seen in the majority of the sustainability literature and the above two examples—is largely irrelevant if the human demand for essential energy and materials exceeds the capacity of the Earth to supply these resources and if the release of wastes exceeds the capacity of the biosphere to absorb or detoxify these substances. Human-caused climate change is an obvious and timely case in point. Carbon dioxide has always been a waste product of human metabolism—not only the biological metabolism that consumes oxygen and produces carbon dioxide as it converts food into usable energy for biological activities, but also the extra-biological metabolism that also produces CO2 as it burns biofuels and fossil fuels to power the maintenance and development of huntergatherer, agricultural, and industrial-technological societies. Only in the last century or so, however, has the increasing production of CO2 by humans overwhelmed the Earth’s capacity to absorb it, increasing atmospheric concentrations and warming the planet more each decade. So, for example, efforts to achieve a ‘‘sustainable’’ local economy for a coastal fishing village in a developing country will be overwhelmed if, in only a few decades, a rising sea level caused by global climate change inundates the community. This shows the importance of analyzing sustainability on a global as well as a local and regional scale. A macroecological approach to sustainability science emphasizes how human socioeconomic systems at any scale depend on the flows of essential energy and material resources at the scale of the biosphere as a whole. The finite Earth system imposes absolute limits on the ecological processes and human activities embedded within it. The impossibility of continued exponential growth of population and resource use in a finite world has long been recognized [33–35]. But repeated failures to reach the limits in the predicted time frames have caused much of the economic establishment and general public to discredit or at least discount Malthusian dynamics. Now, however, there is increasing evidence that humans are pushing if not exceeding global limits [2,3,36,37]. For example, the Global Footprint Network estimates that the ecological footprint, the amount of land required to maintain the human population at a steady state [9], had exceeded the available land area by more than 50% by 2007, and the imbalance is increasing (http:// www.footprintnetwork.org/en/index.php/ GFN). Here we present additional evidence that humans have approached or surpassed the capacity of the biosphere to provide essential and often non-substitutable natural resources. Figure 3 plots trends in the total and per capita use of agricultural land, fresh water, fisheries, wood, phosphate, petroleum, copper, and coal, as well as gross domestic product (GDP), from 1961 to 2008. Note that only oil, copper, coal, and perhaps fresh water show consistent increases in total consumption. Consumption of the other resources peaked in the 1980s or 1990s and has since declined. Dividing the total use of each resource by the human population gives the per capita rate of resource use, which has decreased conspicuously for all commodities except copper and coal. This means that production of these commodities has not kept pace with population growth. Consumption by the present generation is already ‘‘compromising the ability of future generations to meet their own needs.’’ And this does not account for continued population growth, which is projected to increase the global population to 9–10 billion by 2050 and would result in substantial further decreases in per capita consumption. Figure 3 shows results consistent with other analyses reporting ‘‘peak’’ oil, fresh water, and phosphate, meaning that global stocks of these important resources have been depleted to the point that global consumption will soon decrease if it has not already done so [10,37]. Decreased per capita consumption of essential resources might be taken as an encouraging sign of increased efficiency. But the increase in efficiency is also a response to higher prices as a result of decreasing supply and increasing demand. We have included plots for copper and coal to show that overall production of some more abundant commodities has kept pace with population growth, even though the richest stocks have already been exploited. This is typical in ecology: not all essential resources are equally limiting at any given time. Diminishing supplies of some critical Figure 2. Pictorial illustration of important flows of resources into and wastes out of Portland, Oregon. This ‘‘most sustainable city in America’’ depends on exchanges with the local, regional, and global environments and economies in which it is embedded. doi:10.1371/journal.pbio.1001345.g002 PLoS Biology | www.plosbiology.org 4 June 2012 | Volume 10 | Issue 6 | e1001345 241 Figure 3. Global trends in total and per capita consumption of resources and GDP from 1961 to 2008. Total global use/production is represented by the grey line using the axis scale on the left side of each diagram. Per capita use/production is represented by the black line using the axis scale on the right side of each diagram. Per capita values represent the total values divided by global population size as reported by the World Resources Institute (http://earthtrends.wri.org/). The y-axes are untransformed and scaled to allow for maximum dispersion of variance. Individual sources for global use/production values are as follows: Agricultural land in square-km is from the World Development Indicators Database of the World Bank (http://data. worldbank.org/data-catalog/world-development-indicators) and represents the sum of arable, permanent crop, and permanent pasture lands (see also [46]). Freshwater withdrawal in cubic-km from 1960, 1970, 1980, and 1990 is from UNESCO (http://webworld.unesco.org/water/ihp/db/shiklomanov/ part%273/HTML/Tb_14.html) and for 2000 from The Pacific Institute (http://www.worldwater.org/data.html). Wild fisheries harvest in tonnes is from the FAO Fishery Statistical Collection Global Capture Production Database (http://www.fao.org/fishery/statistics/global-capture-production/en) and is limited to diadromous and marine species. Wood building material production in tonnes is based on the FAO ForeSTAT database (http://faostat.fao.org/site/626/ default.aspx), and represents the sum of compressed fiberboard, pulpwood+particles (conifer and non-conifer [C & NC]), chips and particles, hardboard, insulating board, medium density fiberboard, other industrial roundwood (C & NC), particle board, plywood, sawlogs+veneer logs (C & NC), sawnwood (C & NC), veneer sheets, and wood residues. Phosphate, copper, and combustible coal production in tonnes is based on World Production values reported in the USGS Historical Statistics for Mineral and Material Commodities (http://minerals.usgs.gov/ds/2005/140/). Global coal production data is limited to 1966– 2008. Petroleum production in barrels from 1965 to 2008 is based on The Statistical Review of World Energy (http://www.bp.com/sectiongenericarticle800. do?categoryId=9037130&contentId=7068669) and represents all crude oil, shale oil, and oil sands plus the liquid content of natural gas where this is separately recovered. These data are reported in 1,000 barrels/day units, and were transformed to total barrels produced per year. GDP in 1990 US dollars are from the World Resources Institute (http://earthtrends.wri.org/). All data were accessed May 15–June 15, 2011. doi:10.1371/journal.pbio.1001345.g003 PLoS Biology | www.plosbiology.org 5 June 2012 | Volume 10 | Issue 6 | e1001345 242 resources, such as oil, phosphorus, arable land, and fresh water, jeopardize the capacity to maintain even the current human population and standard of living. What are the consequences of these trends? Many economists and sustainability scientists suggest that there is little cause for concern, at least in the short term of years to decades. They give several reasons: i) the finite stocks have not been totally exhausted, just depleted; there are still fish in the sea, and oil, water, phosphate, copper, and coal in the ground; they are just getting harder to find and extract; ii) conservation and substitution can compensate for depletion, allowing economies to grow and provide for increases in population and standard of living; iii) production depends more on the relationship between supply and demand as reflected in price than on absolute availability; and iv) the socioeconomic status of contemporary humans depends not so much on raw materials and conventional goods as on electronic information, service industries, and the traditional economic variables of money, capital, labor, wages, prices, and debt. There are several reasons to question this optimistic scenario. First, the fact that GDP has so far kept pace with population does not imply that resource production will do likewise. Indeed, we have shown that production of some critical resources is not keeping pace. Second, there is limited or zero scope to substitute for some resources. For most of them, all known substitutes are inferior, scarcer, and more costly. For example, there is no substitute for phosphate, which is an essential requirement of all living things and a major constituent of fertilizer. No other element has the special properties of copper, which is used extensively in electronics. Despite extensive recycling of copper, iron, aluminum, and other metals, there is increasing concern about maintaining supplies as the rich natural ores have been depleted (e.g., [38], but see [39]). Third, several of the critical resources have interacting limiting effects. For example, the roughly constant area of land in cultivation since 1990 indicates that modern agriculture has fed the increasing human population by achieving higher yields per unit area. But such increased yields have required increased inputs of oil for powering machinery, fresh water for irrigation, and phosphate for fertilizer. Similarly, increased use of finite fossil fuels has been required to synthesize nitrogen fertilizers and to maintain supplies of mineral resources, such as copper, nickel, and iron, as the richest ores have been depleted and increased energy is required to extract the remaining stocks. An optimistic scenario would suggest that increased use of coal and renewable energy sources such as solar and wind can substitute for depleted reserves of petroleum, but Figure 3 shows a similar pattern of per capita consumption for coal as for other limiting resources, and the capacity of renewables to substitute for fossil fuels is limited by thermodynamic constraints due to low energy density and economic constraints of low energy and monetary return on investment [40–43]. Fourth, these and similar results (e.g., [3]) are starting to illuminate the necessary interdependencies between the energetic and material currencies of ecology and the monetary currencies of economics. The relationship between decreasing supply and increasing demand is causing prices of natural resources to increase as they are depleted, and also causing prices of food to increase as fisheries are overharvested and agriculture requires increasing energy and material subsidies [2,8,43]. The bottom line is that the growing human population and economy are being fed by unsustainable use of finite resources of fossil fuel energy, fertilizers, and arable land and by unsustainable harvests of ‘‘renewable resources’’ such as fish, wood, and fresh water. Furthermore, attaining sustainability is additionally complicated by inevitable yet unpredictable changes in both human socioeconomic conditions and the extrinsic global environment [44]. Sustainability will always be a moving target and there cannot be a single long-term stable solution. Most sustainability science focuses on efforts to improve standards of living and reduce environmental impacts at local to regional scales. These efforts will ultimately and inevitably fail unless the global system is sustainable. There is increasing evidence that modern humans have already exceeded global limits on population and socioeconomic development, because essential resources are being consumed at unsustainable rates. Attaining sustainability at the global scale will require some combination of two things: a decrease in population and/or a decrease in per capita resource consumption (see also [45]). Neither will be easy to achieve. Whether population and resource use can be reduced sufficiently and in time to avoid socioeconomic collapse and attendant human suffering is an open question. Critics will point out that our examination of sustainability from a macroecological and natural science perspective conveys a message of ‘‘doom and gloom’’ and does not offer ‘‘a way forward’’. It is true that humanity is faced with difficult choices, and there are no easy solutions. But the role of science is to understand how the world works, not to tell us what we want to hear. The advances of modern medicine have cured some diseases and improved health, but they have not given us immortality, because fundamental limits on human biology constrain us to a finite lifespan. Similarly, fundamental limits on the flows of energy and materials must ultimately limit the human population and level of socioeconomic development. If civilization in anything like its present form is to persist, it must take account of the finite nature of the biosphere. Conclusion If sustainability science is to achieve its stated goals of ‘‘dealing with the interactions between natural and social systems’’ so as to ‘‘[meet] the needs of present and future generations while substantially reducing poverty and conserving the planet’s life support systems’’, it must take account of the ecological limits on human systems and the inherently ecological nature of the human enterprise. The human economy depends on flows of energy and materials extracted from the environment and transformed by technology to create goods and services. These flows are governed by physical conservation laws. These flows rarely balance at local or regional scales. More importantly, however, because these systems are all embedded in the global system, the flows of critical resources that currently sustain socioeconomic systems at these scales are jeopardized by unsustainable consumption at the scale of the biosphere. These ecological relationships will determine whether ‘‘sustainability’’ means anything more than ‘‘green’’, and whether ‘‘future generations [will be able] to meet their own needs’’. Supporting Information Text S1 Supplementary data. Calculations for salmon nutrient inputs to terrestrial and riparian ecosystems. (DOCX) Acknowledgments This paper is a product of the Human Macroecology Group, an informal collaboration of scientists associated with the University of New Mexico and the Santa Fe Institute. It benefits from discussions with many colleagues on the important but controversial topic of sustainability. We thank Woody Woodruff, Matthew Moerschbaecher, and John Day for providing helpful comments on the manuscript. PLoS Biology | www.plosbiology.org 6 June 2012 | Volume 10 | Issue 6 | e1001345 243 References 1. Goodland R (1995) The concept of environmental sustainability. Ann Rev Ecol System 26: 1–24. 2. Rockstro¨m J, Steffen W, Noone K, Persson A, Chapin FS 3rd, et al. (2009) A safe operating space for humanity. Nature 461: 472–475. 3. Brown JH, Burnside WR, Davidson AD, DeLong JP, Dunn WC, et al. (2011) Energetic limits to economic growth. BioScience 61: 19–26. 4. Vitousek PM, Mooney HA, Lubchenco J, Melillo JM (1997) Human domination of Earth’s ecosystems. Science 277: 494–499. 5. Imhoff ML, Bounoua L, Ricketts T, Loucks C, Harriss R, et al. (2004) Global patterns in human consumption of net primary production. Nature 429: 870–873. 6. Haberl H, Erb KH, Krausmann F, Gaube V, Bondeau A, et al. (2007) Quantifying and mapping the human appropriation of net primary production in earth’s terrestrial ecosystems. Proc Natl Acad Sci U S A 104: 12942–12947. 7. Pauly D, Watson R, Alder J (2005) Global trends in world fisheries: impacts on marine ecosystems and food security. Phil Trans Roy Soc B 360: 5– 12. 8. Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, et al. (2006) Impacts of biodiversity loss on ocean ecosystem services. Science 314: 787– 790. 9. Wackernagel M, Rees WE (1996) Our ecological footprint: reducing human impact on the earth. New Society Publications. 10. Gleick PH, Palaniappan M (2010) Peak water limits to freshwater withdrawal and use. Proc Natl Acad Sci U S A 107: 11155–11162. 11. Brundtland GH (1987) Our common future. World Commission on Environment and Devel- opment. 12. Burnside WR, Brown JH, Burger O, Hamilton MJ, Moses M, et al. (2012) Human macroecology: linking pattern and process in big picture human ecology. Biol Rev 87: 194–208. 13. Day JW Jr, Boesch DF, Clairain EJ, Kemp GP, Laska SB, et al. (2007) Restoration of the Mississippi Delta: lessons from hurricanes Katrina and Rita. Science 315: 1679–1684. 14. Moses ME, Brown JH (2003) Allometry of human fertility and energy use. Ecol Let 6: 295–300. 15. Day JW Jr, et al. (2009) Ecology in times of scarcity. BioScience 59: 321–331. 16. Odum HT (1971) Environment, power and society. New York: Wiley-Interscience. 17. Odum HT (2007) Environment, power, and society for the twenty-first century: the hierarchy of energy. Columbia University Press. 18. Sterner RW, Elser JJ (2002) Ecological stoichiometry: the biology of elements from molecules to the biosphere. Princeton: Princeton University Press. 19. Czu´cz B, Gathman JP, Mcpherson GUY (2010) The impending peak and decline of petroleum production: an underestimated challenge for conservation of ecological integrity. Con Biol 24: 948–956. 20. Hilborn R, Quinn TP, Schindler DE, Rogers DE (2003) Biocomplexity and fisheries sustainability. Proc Natl Acad Sci U S A 100: 6564–6568. 21. Hilborn R (2006) Salmon-farming impacts on wild salmon. Proc Natl Acad Sci U S A 103: 15277. 22. ADG&F (2010) Alaska historical commercial salmon catches, 1878–2010. Alaska Department of Game and Fish, Division of Commercial Fisheries. 23. Coupland G, Stewart K, Patton K (2010) Do you never get tired of salmon? Evidence for extreme salmon specialization at Prince Rupert harbour, British Columbia. J Anthro Arch 29: 189–207. 24. Cederholm CJ, Kunze MD, Murota T, Sibatani A (1999) A Pacific salmon carcasses: essential contributions of nutrients and energy for aquatic and terrestrial ecosystems. Fisheries 24: 6–15. 25. Gende SM, Edwards RT, Willson MF, Wipfli MS (2002) Pacific salmon in aquatic and terrestrial ecosystems. BioScience 52: 917–928. 26. Naiman RJ, Bilby RE, Schindler DE, Helfield JM (2002) Pacific salmon, nutrients, and the dynamics of freshwater and riparian ecosystems. Ecosystems 5: 399–417. 27. Schindler DE, Scheuerell MD, Moore JW, Gende SM, Francis TB, et al. (2003) Pacific salmon and the ecology of coastal ecosystems. Front Ecol Enviro 1: 31–37. 28. Schindler DE, Leavitt PR, Brock CS, Johnson SP, Quay PD (2005) Marine-derived nutrients, commercial fisheries, and production of salmon and lake algae in Alaska. Ecology 86: 3225–3231. 29. International Council for Science (2002) Science and technology for sustainable development. Paris: International Council for Science. 30. Millenium Ecosystem Assessment (2005) Ecosystems and human well-being: synthesis. Island Press. 31. Turner BL, Lambin EF, Reenberg A (2007) The emergence of land change science for global environmental change and sustainability. Proc Natl Acad Sci U S A 104: 20666–20671. 32. Hillman T, Ramaswami A (2010) Greenhouse gas emission footprints and energy use benchmarks for eight US cities. Environ Science & Tech 44: 1902–1910. 33. Malthus TR (1798) An essay on the principle of population. Prometheus. 34. Ehrlich PR (1968) The population bomb. New York. 35. Meadows DH (1972) The limits of growth. A report for The Club of Rome. 36. Holdren JP (2008) Science and technology for sustainable well-being. Science 319: 424–434. 37. Nel WP, Van Zyl G (2010) Defining limits: energy constrained economic growth. Applied Energy 87: 168–177. 38. Gordon RB, Bertram M, Graedel TE (2006) Metal stocks and sustainability. Proc Natl Acad Sci U S A 103: 1209–1214. 39. Kesler SE, Wilkinson BH (2008) Earth’s copper resources estimated from tectonic diffusion of porphyry copper deposits. Geology 36: 255. 40. Fargione J, Hill J, Tilman D, Polasky S, Hawthorne P (2008) Land clearing and the biofuel carbon debt. Science 319: 1235–1238. 41. Hall CAS, Day JW (2009) Revisiting the limits to growth after peak oil. American Scientist 97: 230– 237. 42. Smil V (2008) Energy in nature and society: general energetics of complex systems. MIT Press. 43. The Royal Society (2009) Reaping the benefits: science and the sustainable intensification of global agriculture. London: The Royal Society. 44. Milly PC, Betancourt J, Falkenmark M, Hirsch RM, Kundzewicz ZW, et al. (2008) Stationarity is dead: whither water management? Science 319: 573–574. 45. Cohen JE (1996) How many people can the earth support? WW Norton & Company. 46. Foley JA, Ramankutty N, Brauman KA, Cassidy ES, Gerber JS, et al. (2011) Solutions for a cultivated planet. Nature 478: 337–342. PLoS Biology | www.plosbiology.org 7 June 2012 | Volume 10 | Issue 6 | e1001345 244 Articles www.biosciencemag.org January 2011 / Vol. 61 No. 1 • BioScience 19 Energetic Limits to Economic Growth James H. Brown, William R. Burnside, Ana D. Davidson, John P. DeLong, William C. Dunn, Marcus J. Hamilton, Jeffrey C. Nekola, Jordan G. Okie, Norman Mercado-Silva, William H. Woodruff, and Wenyun Zuo The human population and economy have grown exponentially and now have impacts on climate, ecosystem processes, and biodiversity far exceeding those of any other species. Like all organisms, humans are subject to natural laws and are limited by energy and other resources. In this article, we use a macroecological approach to integrate perspectives of physics, ecology, and economics with an analysis of extensive global data to show how energy imposes fundamental constraints on economic growth and development. We demonstrate a positive scaling relationship between per capita energy use and per capita gross domestic product (GDP) both across nations and within nations over time. Other indicators of socioeconomic status and ecological impact are correlated with energy use and GDP. We estimate global energy consumption for alternative future scenarios of population growth and standards of living. Large amounts of energy will be required to fuel economic growth, increase standards of living, and lift developing nations out of poverty. Keywords: energy, economic growth, economy, human macroecology, scaling metabolic ecology (McMahon and Bonner 1983, Schneider and Kay 1995, Brown et al. 2004) and empirical approaches from macroecology (Brown 1995) to document energetic constraints on human ecology that have important implications for modern humans. The central role of energy Economic growth and development require that energy and other resources be extracted from the environment to manufacture goods, provide services, and create capital. The central role of energy is substantiated by both theory and data. Key theoretical underpinnings come from the laws of thermodynamics: first, that energy can be neither created nor destroyed, and second, that some capacity to perform useful work is lost as heat when energy is converted from one form to another. Complex, highly organized systems, including human economies, are maintained in states far from thermodynamic equilibrium by the continual intake and transformation of energy (Soddy 1926, Odum 1971, Georgescu-Roegen 1977, Ruth 1993, Schneider and Kay 1995, Hall et al. 2001, Chen 2005, Smil 2008). Empirically, the central role of energy in modern human economies is demonstrated by the positive relationship between energy use and economic growth (Shafiee and Topal 2008, Smil 2008, Payne 2010). Here, we take a macroecological perspective and quantify statistical relationships between energy use and economic activity for 220 nations over 24 years, using data from the International Energy Agency (IEA; www.iea.org/stats/index.asp) and World Resources Institute (WRI; http://earthtrends.wri.org/index.php).Per capita energy consumption for each country is calculated as the sum of human biological metabolism plus the energy obtained from The human species has an interesting duality. On the one  hand, Homo sapiens is just another species, subject to the same scientific laws as the millions of other animals, plants, and microbes. On the other hand, humans are unique. No other species in the history of Earth has achieved such ecological dominance and created such complex socioeconomic systems. Because of this duality, humans have been studied by both natural and social scientists, but often from very different perspectives (Arrow et al. 2004). In just a few thousand years the human population has colonized the entire world and grown to almost 7 billion. Humans now appropriate 20% to 40% of terrestrial annual net primary production, and have transformed the atmosphere, water, land, and biodiversity of the planet (Vitousek et al. 1997, Haberl et al. 2007). For centuries some have questioned how long a finite planet can continue to support near-exponential population and economic growth (e.g., Malthus 1798, Ehrlich 1968, Meadows et al. 1972). Recent issues such as climate change, the global decline in population growth rate, the depletion of petroleum reserves and resulting increase in oil prices, and the recent economic downturn have prompted renewed concerns about whether longstanding trajectories of population and economic growth can continue (e.g., Arrow et al. 2004). These serious issues fall within the purview of both the natural and social sciences—especially ecology and economics. This article integrates perspectives from physics, ecology, and economics with an analysis of extensive global data to show how scientific laws governing the flows of energy in the biosphere affect socioeconomic activity. Our purpose is neither to pit ecology against economics nor to predict future population and economic trends; rather, we use theoretical perspectives from thermodynamics, allometry, and BioScience 61: 19–26. ISSN 0006-3568, electronic ISSN 1525-3244. © 2011 by American Institute of Biological Sciences. All rights reserved. Request permission to photocopy or reproduce article content at the University of California Press’s Rights and Permissions Web site at www.ucpressjournals.com/reprintinfo.asp. doi:10.1525/bio.2011.61.1.7 245 20 BioScience • January 2011 / Vol. 61 No. 1 www.biosciencemag.org Articles all other sources. Individual biological metabolism was estimated from data on daily caloric intake for each country and converted to watts per individual. Energy use from all other sources, including fossil fuels and renewable energy supplies, was obtained from the IEA. For our measure of economic activity, we used the WRI’s data for gross domestic product (GDP), the market value of all goods and services produced within a country per year (Mankiw 2006). Both energy use and GDP are expressed on a per capita basis. So, per capita GDP can be thought of as an index of an average individual’s share of his or her country’s economy, and per capita energy use as the power required to sustain that level of economic activity.  Figure 1 shows the relationship between energy use and GDP plotted on logarithmic axes, with each colored line indicating the trajectory for a single country over the period 1980–2003. A regression through the mean GDP, G for mean energy consumption and E for each country over the 24-year period, accounts for 76% of the variation. The fitted regression describes the scaling of per capita energy use with per capita GDP as a power law: E = 4.13G 0.76 (figure 1; see also figure S1c in supplemental online materials at http://caliber. ucpress.net/doi/suppl/10.1525/ bio.2011.61.1.7). The sublinear slope, 0.76, indicates that the rate of per capita energy consumption associated with greater economic activity increases less rapidly than GDP itself. Countries with larger economies take advantage of economies of scale and new technologies to use energy more efficiently on a per capita basis (Hoffert et al. 1998). For example, there is both a positive relationship and an economy of scale between economic growth and the amount of infrastructure, such as roads, pipelines, and power lines, that distributes energy resources (Easterly and Rebelo 1993). The relationship between energy use and GDP holds across countries spanning the entire range of economic development from poorest to richest, encompassing two orders of magnitude in both energy use (100 to 10,000 watts) and wealth ($500 to $50,000).  A similar trend occurs within countries over time. The vast majority of nations we analyzed (74%) increased both energy use and GDP from 1980 to 2003 and exhibited positive correlations across the 24 years (mean slope = 0.59; 95% confidence interval = 0.45–0.72; figure 2; see also figure S1 at http://caliber.ucpress.net/doi/suppl/10.1525/bio.2011.61.1.7). Countries notable for sustained recent development, such as Figure 1. The relationship between per capita energy use and per capita gross domestic product (GDP; in US dollars) of countries, plotted on logarithmic axes, from 1980 to 2003. Note that the slope or exponent, 0.76 (95% confidence interval: 0.69–0.82), is close to three-quarters, which is the canonical value of the exponent for the scaling of metabolic rate with body mass in animals. If per capita GDP is taken as the size of an average individual’s economy and per capita energy use as the rate of energy consumption required to support that economy, this relationship may not be coincidental. Total per capita energy consumption is calculated as the caloric intake of humans (about 130 watts) plus the energy derived from all other sources, including fossil fuels and renewables. The thin colored lines show trends for individual countries from 1980 to 2003. The thick black line is a regression model fit to the mean values for each nation during this period. GDP data are from the World Resources Institute (http://earthtrends.wri.org/index.php). Total energy consumption data are calculated from the sum of energy consumption from eating (data from the World Resources Institute) plus all other sources of energy consumed for other purposes such as utilities, manufacturing, and transportation. Source: Data are from the International Energy Agency at www.iea.org/stats/index.asp. 246 www.biosciencemag.org January 2011 / Vol. 61 No. 1 • BioScience 21 Articles China and India, show trajectories of continually increasing energy use. The patterns in figures 1 and 2 are consistent with the increasing energy use that fueled socioeconomic development throughout history (Tainter 1988, Smil 2008). For example, from 1850 to 2000, while the global human population grew fivefold, world energy use increased 20-fold and fossil fuel use rose more than 150-fold (Holdren 2008).  The relationship between energy use and GDP depicted in figure 1 raises several important issues. One is the considerable variation around the regression line: Countries with similar per capita GDPs differ by more than an order of magnitude in per capita energy consumption.We did not analyze this residual variation quantitatively, although it would be illuminating to do so. For example, several oil-exporting nations (e.g., the United Arab Emirates, Bahrain, Tajikistan, and Azerbaijan) are consistent outliers, with high values of energy use relative to GDP. We hypothesize that much of this energy is used to extract oil that is then exported to and consumed by industrialized nations. Although the United States is among the countries with highest per capita energy use, several other developed countries have comparably high values. We hypothesize that in some of them, such as Iceland, Sweden, and Norway, large quantities of energy are used to heat houses and workplaces during cold winters at high latitudes.  Additionally, some of the variation in figure 1 may reflect error or a lack of standardization in the ways different countries have calculated energy use and GDP. The data sources that we used (IEA and WRI), although probably the best available, have their limitations. The original data are selfreported by the countries, and some of the abrupt, seemingly inexplicable changes in the trajectories in figure 1 may simply reflect errors or changing methods of estimating energy use, GDP, or both. One strength of our macroecological approach (Brown 1995), however, is that the hundreds of data points distributed over orders of magnitude variation mean that modest errors and other sources of uncontrolled noise in the data do not obscure the strong signals that are manifest in the robust patterns within and across countries over time.  Another question is, what are the independent and dependent variables? Does energy use support economic development or does economic development drive energy consumption? Financial and energy economists have used econometric techniques to analyze time series of energy consumption and economic growth within countries in an effort to assess causal relationships, but they have reached no clear consensus about whether energy use causes economic growth, or vice versa (see Mahadevan and Asafu-Adjaye 2007, Payne 2010). By analogy to biological allometry, we plotted per capita energy use as the dependent variable and per capita GDP as the independent variable; this is analogous to plotting the rate of energy use of an animal as a function of its body size. The exponent for the scaling of energy use as a function of GDP, 0.76, is reminiscent of the three-quarter-power scaling of metabolic rate with body mass in animals (Kleiber 1961, McMahon and Bonner 1983). This may not be coincidental. In a very real sense both animals and economies have “metabolisms.” Both consume, transform, and allocate energy to maintain complex adaptive systems far from thermodynamic equilibrium. The energy and other resources that sustain these systems are supplied by hierarchically branching networks, such as the blood vessels and lungs of mammals and the oil pipelines, power grids, and transportation networks of nations. Models of these networks suggest that three-quarter-power scaling optimizes distribution of resources (West et al. 1997, Banavar et al. 2010).  Some may be concerned that the relationships in figures 1 and 2 are “just correlations” that do not necessarily imply any underlying mechanism or causality. We disagree. All science is ultimately based on correlations—between dependent and independent variables, model predictions and empirical measurements, or experimental treatments and controls. Any mechanism or causation comes from logical inference. We infer that energy limits economic activity through direct causal mechanisms. The evidence for this inference is presented above and comes from three sources: (1) theory, the application of the second law of thermodynamics to complex adaptive systems; (2) data, the robust relationship between per capita energy use and per capita GDP across both space (the 220 nations of the world) and time (24 years); and (3) analogy, the similarity Figure 2. The frequency distribution of exponents (slopes from log-log plots) of economic trajectories of per capita energy use as a function of per capita gross domestic product (GDP) within countries over the time series from 1980 to 2003. The average slope is approximately 0.6 and nearly all countries showed positive growth over the 24-year period. Source: Data are from the World Resources Institute (http:// earthtrends.wri.org/index.php) and the International Energy Agency (www.iea.org/stats/index.asp). 247 22 BioScience • January 2011 / Vol. 61 No. 1 www.biosciencemag.org Articles between biological and socioeconomic metabolism. We find the last to be especially compelling. Just as a body has a metabolism that burns food energy to survive and grow, a city or national economy has a metabolism that must burn fuel in order to sustain itself and grow. Just as higher metabolic rates are required to sustain and grow larger, more complex bodies (Kleiber 1961, McMahon and Bonner 1983), so higher rates of energy consumption are required to sustain and grow larger, more developed economies that provide greater levels of technological development and higher standards of living. Quantitative relationships among energy use, GDP, and other socioeconomic indicators Some may be concerned that the relationships in figures 1 and 2 do not reflect what is “really important,” which might be some aspect of quality of life rather than GDP. However, nearly all measures of economic activity and standard of living are closely correlated with both GDP and energy use (figure 3; for additional variables, see figure S2 in the supplemental online materials at http://caliber.ucpress.net/ doi/suppl/10.1525/bio.2011.61.1.7). These include measures of nutrition, education, health care, resource use, technology, and innovation. These relationships are not surprising and reflect mechanistic underpinnings. It takes money and energy to train engineers, MDs, and PhDs; to produce vaccines, drugs, and medical equipment; and to construct and maintain road, rail, airplane, cell phone, and Internet networks, hospitals and research centers, parks and conservation areas, and modern buildings and cities. The ecological footprint, an aggregate measure of per capita resource consumption and waste production, also increases with energy use and GDP (figure 3; Dietz et al. 2007). Figure 3 shows that it has not been possible to increase socially desirable goods and services substantially without concomitantly increasing the consumption of energy and other natural resources and without increasing environmental impacts that now include climate change, pollution, altered biogeochemical cycles, and reduced biodiversity. Energetic implications for future economic growth These empirical patterns, together with their theoretical underpinnings, raise the question of whether economic growth and associated increases in human population, resource use, technological development, and standard of living can continue their present trajectories (Grossman and Krueger 1995, Ausubel 1996). In figure 4 we develop some quantitative scenarios. We caution that these are not intended to be predictions of the future; rather, they are extrapolations of the power-law relationship shown in figure 1 to estimate the quantity of energy that would be required to support different global populations and levels of economic development. So, for example, raising the current global population to the standard of living in the United States would require a nearly fivefold increase in the rate of energy consumption, from 17 to 77 terawatts (1 terawatt = 1012 watts). Population growth must also be considered in any future scenario. To support a projected global population of 9.5 billion in 2050 with an average standard of living equivalent to the current US lifestyle would require about 268 terawatts, 16 times the current global energy use. Even maintaining this increased population at the more modest Chinese standard of living would require 2.5 times more energy than is used today (figure 4).  There are good reasons, however, such simple scenarios based on extrapolations of current population and economic trends may be imprecise. Our calculations incorporate the economy of scale implicit in the nonlinear scaling of energy use with GDP, but do not take into account many potentially important factors, such as greater efficiency that may be triggered by energy shortages; technological innovations that may increase energy supplies; and socioeconomic, demographic, and behavioral changes. Indeed, the global human socioeconomic system is complex, poised far from thermodynamic equilibrium by high rates of energy input and transformation. Such systems have unpredictable nonlinear dynamics, making it nearly impossible to predict very far into the future (Schneider and Kay 1995).  One thing is clear: If the relationships depicted in figures 1–3 characterize fundamental causal relationships among the rate of energy use, level of economic activity, and standard of living, then additional economic growth and development will require some combination of (a) increased energy supply, (b) decreased per capita energy use, and (c) decreased human population. We consider each in turn. Increased energy supply.  The sources of energy that may be used to support future economic growth include finite stocks of fossil fuels as well as nuclear, renewable, and other proposed but unproven technologies. Fossil fuels currently provide 85% of humankind’s energy needs (figure 5), but they are effectively fixed stores that are being depleted rapidly (Heinberg 2003, IEA 2008, Hall and Day 2009). Conventional nuclear energy currently supplies only about 6% of global energy; fuel supplies are also finite, and future developments are plagued by concerns about safety, waste storage,anddisposal(NelandCooper2009).Abreakthrough in nuclear fusion, which has remained elusive for the last 50 years, could potentially generate enormous quantities of energy, but would likely produce large and unpredictable socioeconomic and environmental consequences. Solar, hydro, wind, and tidal renewable energy sources are abundant, but environmental impacts and the time, resources, and expenses required to capture their energy limit their potential (Hall and Day 2009). Biofuels may be renewable, but ecological constraints and environmental impacts constrain their contribution (Fargione et al. 2008). More generally, most efforts to develop new sources of energy face economic problems of diminishing returns on energy and monetary investment (Hall et al. 1986, Tainter 1988, Allen et al. 2001, Tainter et al. 2003). 248 www.biosciencemag.org January 2011 / Vol. 61 No. 1 • BioScience 23 Articles Decreased per capita energy use.  The Malthusian-Darwinian dynamic that has shaped the evolution of human behavior and demography has created powerful tendencies for individuals and societies to exploit all available resources and use all available technologies to enhance personal status, biological fitness, and societal wealth (Lotka 1922). Poor people migrate to cities and to other countries to improve their prospects. Citizens of developing countries such as China and India are not usually satisfied with the status quo, and understandably want to live like those in the deFigure 3. Variables reflecting socioeconomic status and standard of living are strongly correlated with per capita energy use (upper panels) and per capita gross domestic product (GDP; lower panels). The variables include measures of health and wellness (population growth rate, doctors per 100,000 people, life expectancy, infant mortality, caloric intake, national poverty), energy use (electricity [kilowatt hours], residential energy [kilograms of oil equivalent]), resource consumption (meat, televisions, aluminum [kilotons], waste [kilograms]), intellectual and technological contributions (Nobel Prizes, patents), and ecological impacts (ecological footprint [in hectares]). All correlations (r) are significant (P < 0.05). Data sources are provided in supplemental online materials. 249 24 BioScience • January 2011 / Vol. 61 No. 1 www.biosciencemag.org Articles veloped world. People in the richest nations are reluctant to sacrifice economic growth—much less give up their automobiles, electronics, and organ transplants—so that people in poorer countries can have bicycles, personal computers, and flu shots. Decreased human population.  With growing standards of living and rates of energy use, parents tend to invest more resources in fewer children (Moses and Brown 2003). This trade-off between the number and quality of offspring contributes to demographic transitions, where family size and the rate of population growth decrease with increasing economic development (Thompson 1929). The global population growth rate has declined in the last decade, but only a few developed countries currently have zero or negative population growth (WRI, http://earthtrends.wri. org/index.php). The relationship between family size and per capita energy use suggests that five times the current rates of energy supply will be required to achieve a global level of socioeconomic development capable of stabilizing the human population without infringing on the freedom of individuals to have as many children as they choose (Moses and Brown 2003, DeLong et al. 2010).  The bottom line is that an enormous increase in energy supply will be required to meet the demands of projected population growth and lift the developing world out of poverty without jeopardizing current standards of living in the most developed countries. And the possibilities for substantially increasing energy supplies are highly uncertain. Moreover, the nonlinear, complex nature of the global economy raises the possibility that energy shortages might trigger massive socioeconomic disruption. Again, consider the analogy to biological metabolism: Gradually reducing an individual’s food supply leads initially to physiological adjustments, but then to death from starvation, well before all food supplies have been exhausted.  Mainstream economists historically have dismissed warnings that resource shortages might permanently limit economic growth. Many believe that the capacity for technological innovation to meet the demand for resources is as much a law of human nature as the MalthusianDarwinian dynamic that creates the demand (Barro and Sala-i-Martin 2003, Durlauf et al. 2005, Mankiw 2006). However, there is no scientific support for this proposition; it is either an article of faith or based on statistically flawed extrapolations of historical trends. The ruins of Mohenjo Daro, Mesopotamia, Egypt, Rome, the Maya, Angkor, Easter Island, and many other complex civilizations provide incontrovertible evidence that innovation does not always prevent socioeconomic collapse (Tainter 1988, Diamond 2004). Conclusions We are by no means the first to write about the limits to economic growth and the fundamental energetic constraints that stem directly from the laws of thermodynamics and the principles of ecology. Beginning with Malthus (1798), both ecologists and economists have called attention to the essential dependence of economies on natural resources and have pointed out that near-exponential growth of the human population and economy cannot be sustained indefinitely in a world of finite resources (e.g., Soddy 1922, Odum 1971, Daly 1977, Georgescu-Roegen 1977, Cleveland et al. 1984, Costanza and Daly 1992, Hall et al. 2001, Arrow et al. 2004, Stern 2004, Nel and van Zyl 2010). Some ecological economists and systems ecologists have made Figure 4. Current and projected global energy consumption based on alternative scenarios of population growth (2006, 2025, and 2050) and standard of living (equivalent to contemporary Uganda, China, and United States). Dashed line is total global terrestrial net primary productivity (NPP), 75 terawatts (Haberl et al. 2007). Data sources and calculation methods can be found in supplemental online materials. Figure 5. Sources of energy currently consumed by the global human economy. Total annual consumption is approximately 15.9 terawatts (TW, 1 terawatt = 1012 watts), of which about 85% comes from fossil fuels, 6% from nuclear energy, and the remaining 9% from solar, hydro, wind, and other renewable sources (BP 2009, REN21 2009). 250 www.biosciencemag.org January 2011 / Vol. 61 No. 1 • BioScience 25 Articles similar theoretical arguments for energetic constraints on economic systems (e.g., Odum 1971, Hall et al. 1986). However, these perspectives have not been incorporated into mainstream economic theory, practice, or pedagogy (e.g., Barro and Sala-i-Martin 2003, Mankiw 2006), and they have been downplayed in consensus statements by influential ecologists (e.g., Lubchenco et al. 1991, Palmer et al. 2004, ESA 2009) and sustainability scientists (e.g., NRC 1999, Kates et al. 2001, ICS 2002, Kates and Parris 2003, Parris and Kates 2003, Clark 2007).  Our explicitly macroecological and metabolic approach uses new data and analyses to provide quantitative, mechanistic, and practically relevant insights into energetic limits on economic growth. We hope the evidence and interpretations presented here will call the attention of scientists, policymakers, world leaders, and the public to the central but largely underappreciated role of energetic limits to economic growth. Acknowledgments For support we thank the Howard Hughes Medical Institute and National Institute of Biomedical Imaging and Bioengineering Interfaces grant to JHB, JGO, and WZ; National Science Foundation (NSF) Grant DEB-0541625 and the Rockefeller Foundation to MJH; NSF Grant OISE-0653296 to ADD; and National Institutes of Health Grant DK36263 to WHW. We thank the many colleagues who have discussed these ideas with us and encouraged us to write this article. Charles A. S. Hall, Charles Fowler, Joseph A. Tainter, and several anonymous reviewers provided helpful comments on earlier drafts of the manuscript. References cited Allen TFH, Tainter JA, Pires JC, Hoekstra TW. 2001. Dragnet ecology— “Just the facts, ma’am”: The privilege of science in a postmodern world. BioScience 51: 475–485. Arrow K, Dasgupta P, Goulder L, Daily G, Ehrlich P, Heal G, Levin S, Schneider S, Starrett D, Walker B. 2004. Are we consuming too much? Journal of Economic Perspectives 18: 147–172. Ausubel JH. 1996. Can technology spare the Earth? American Scientist 84: 166–178. Banavar JR, Moses ME, Brown JH, Damuth J, Rinaldo A, Sibly RM, Maritan A. 2010. A general basis for quarter power scaling in animals. Proceedings of the National Academy of Sciences 107: 15816–15820. Barro RJ, Sala-i-Martin X. 2003. Economic Growth. 2nd ed. MIT Press. [BP] British Petroleum.2009.BP Statistical Review of World Energy June 2006. BP. (11October2010;www.bp.com/liveassets/bp_internet/globalbp/globalbp_ uk_english/reports_and_publications/statistical_energy_review_2006/ STAGING/local_assets/downloads/spreadsheets/statistical_review_full_ report_workbook_2006.xls) Brown JH. 1995. Macroecology. University of Chicago Press. Brown JH, Gillooly JF, Allen AP, Savage VM, West GB. 2004. Toward a metabolic theory of ecology. Ecology 85: 1771–1789. Chen J. 2005. The Physical Foundation of Economics: An Analytical Thermodynamic Theory. World Scientific. Clark WC. 2007. Sustainability science: A room of its own. Proceedings of the National Academy of Sciences 104: 1737–1738. Cleveland CJ, Costanza R, Hall CAS, Kaufmann RK. 1984. Energy and the U.S. economy: A biophysical perspective. Science 225: 890–897. Costanza R, Daly HE. 1992. Natural capital and sustainable development. Conservation Biology 6: 37–46. Daly HE. 1977. Steady-state Economics. Freeman. DeLong JP, Burger O, Hamilton MJ. 2010. Current demographics suggest that future energy supplies will be inadequate to slow human population growth. PLoS ONE 5: e13206. doi:10.1371/journal.pone.0013206 Diamond J. 2004. Collapse: How Societies Choose to Fail or Succeed.Viking Adult. Dietz T, Rosa EA, York R. 2007. Driving the human ecological footprint. Frontiers in Ecology and the Environment 5: 13–18. Durlauf SN, Johnson PA, Temple JRW. 2005. Growth econometrics. Pages 555–677 in Aghion P, Durlauf SN, eds. Handbook of Economic Growth. Elsevier. Easterly W, Rebelo S. 1993. Fiscal policy and economic growth: An empirical investigation. Journal of Monetary Economics 32: 417–458. Ehrlich P. 1968. The Population Bomb. Ballantine. [ESA] Ecological Society of America. 2009. Ecological Impacts of Economic Activities. ESA. (14 October 2010; www.esa.org/pao/economic_activities. php) Fargione J, Hill J, Tilman D, Polasky S, Hawthorne P. 2008. Land clearing and the biofuel carbon debt. Science 319: 1235–1238. Georgescu-Roegen N. 1977. The steady state and ecological salvation: A thermodynamic analysis. BioScience 27: 266–270. Grossman GM, Krueger AB. 1995. Economic growth and the environment. Quarterly Journal of Economics 110: 353–377. Haberl H, Erb K-H, Krausmann F, Gaube V, Bondeau A, Plutzar C, Gingrich S, Lucht W, Fischer-Kowalski M. 2007. Quantifying and mapping the human appropriation of net primary production in Earth’s terrestrial ecosystems. Proceedings of the National Academy of Sciences 104: 12942–12947. Hall CAS, Day JW Jr. 2009. Revisiting the limits to growth after peak oil. American Scientist 97: 230–237. Hall CAS, Cleveland CJ, Kaufmann RK. 1986. Energy and Resource Quality: The Ecology of the Economic Process. Wiley. Hall C, Lindenberger D, Kümmel R, Kroeger T, Eichhorn W. 2001. The need to reintegrate the natural sciences with economics. BioScience 51: 663–673. Heinberg R. 2003. The end of the oil age. Earth Island Journal 18: 3. Hoffert MI, et al. 1998. Energy implications of future stabilization of atmospheric CO2 content. Nature 395: 881–884. Holdren JP. 2008. Science and technology for sustainable well-being. Science 319: 424–434. [ICS] International Council for Science. 2002. Report of the Scientific and Technological Community to the World Summit on Sustainable Development (WSSD). ICS. Report no. 1. [IEA] International Energy Agency. 2008. World Energy Outlook 2008. Organisation for Economic Co-operation and Development and IEA. Kates RW, Parris TM. 2003. Long-term trends and a sustainability transition. Proceedings of the National Academy of Sciences 100: 8062–8067. Kates RW, et al. 2001. Environment and development: Sustainability science. Science 292: 641–642. Kleiber M. 1961. The Fire of Life: An Introduction to Animal Energetics. Wiley. Lotka AJ. 1922. Contribution to the energetics of evolution. Proceedings of the National Academy of Sciences 8: 147–151. Lubchenco J, et al. 1991. The sustainable biosphere initiative: An ecological research agenda: A report from the Ecological Society of America. Ecology 72: 371–412. Mahadevan R, Asafu-Adjaye J. 2007. Energy consumption, economic growth and prices: A reassessment using panel VECM for developed and developing countries. Energy Policy 35: 2481–2490. Malthus TR. 1798. An Essay on the Principle of Population. Prometheus. Mankiw NG. 2006. Macroeconomics. 6th ed. Worth. McMahon TA, Bonner JT. 1983. On Size and Life. Scientific American Library. Meadows DH, Meadows DL, Randers J, Behrens WW. 1972. The Limits to Growth: A Report for the Club of Rome’s Project on the Predicament of Mankind. Earth Island. 251 26 BioScience • January 2011 / Vol. 61 No. 1 www.biosciencemag.org Articles ———. 1926. Wealth, Virtual Wealth, and Debt: The Solution of the Economic Paradox. Omni. Stern DI. 2004. Economic growth and energy. Pages 35–78 in Encyclopedia of Energy, vol 2. Academic Press. Tainter JA. 1988. The Collapse of Complex Societies. Cambridge University Press. Tainter JA, Allen THF, Little A, Hoekstra TW. 2003. Resource transitions and energy gain: Contexts of organization. Conservation Ecology 7: 4. Thompson WS. 1929. Population. American Journal of Sociology 34: 959–975. Vitousek PM, Mooney HA, Lubchenco J, Melillo JM. 1997. Human domination of Earth’s ecosystems. Science 277: 494–499. West GB, Brown JH, Enquist BJ. 1997. A general model for the origin of allometric scaling laws in biology. Science 276: 122–126. James H. Brown (jhbrown@unm.edu) is a distinguished professor at the University of New Mexico and external faculty of the Santa Fe Institute. William R. Burnside, William C. Dunn, Jordan G. Okie, and Wenyun Zuo are PhD candidates in the Department of Biology at the University of New Mexico. Ana D. Davidson is a postdoctoral researcher at the National University of Mexico and adjunct professor of biology at the University of New Mexico. John P. DeLong is a postdoctoral associate at Yale University in the Department of Ecology and Evolutionary Biology. Marcus J. Hamilton is an archaeological anthropologist at the University of New Mexico and the Santa Fe Institute. Jeffrey C. Nekola is an ecologist at the University of New Mexico. Norman Mercado-Silva is a research specialist with the School of Natural Resources and the Environment, Arizona Cooperative Fish and Wildlife Research Unit, at the University of Arizona, in Tucson. William H. Woodruff is a scientist at Los Alamos National Laboratory and external faculty at the Santa Fe Institute. Moses ME, Brown JH. 2003. Allometry of human fertility and energy use. Ecology Letters 6: 295–300. [NRC] National Research Council. 1999. Our Common Journey. The National Academies Press. Nel WP, Cooper CJ. 2009. Implications of fossil fuel constraints on economic growth and global warming. Energy Policy 37: 166–180. Nel WP, van Zyl G. 2010. Defining limits: Energy constrained economic growth. Applied Energy 87: 168–177. Odum HT. 1971. Environment, Power and Society. Wiley. Palmer M, et al. 2004. Ecology: Ecology for a crowded planet. Science 304: 1251–1252. Parris TM, Kates RW. 2003. Characterizing a sustainability transition: Goals, targets, trends, and driving forces. Proceedings of the National Academy of Sciences 100: 8068–8073. Payne JE. 2010. Survey of the international evidence on the causal relationship between energy consumption and growth. Journal of Economic Studies 37: 53–95. [REN21] Renewable Energy Policy Network for the 21st Century.2006.Renewables, Global Status Report: 2006 Update. REN21 Steering Committee. Ruth M. 1993. Integrating Economics, Ecology, and Thermodynamics. Springer. Schneider ED, Kay JJ. 1995. Order from Disorder: The Thermodynamics of Complexity in Biology. Pages 161–172 in Murphy MP, O’Neill LAJ, eds. What Is Life: The Next Fifty Years. Reflections on the Future of Biology. Cambridge University Press. Shafiee S, Topal E. 2008. An econometrics view of worldwide fossil fuel consumption and the role of US. Energy Policy 36: 775–786. Smil V. 2008. Energy in Nature and Society: General Energetics of Complex Systems. MIT Press. Soddy F. 1922. Cartesian Economics: The Bearing of Physical Science upon State Stewardship. Hendersons. 252 CONCLUDING REMARKS What does ‘mechanism’ mean within an ecological context? It is interesting to note the strongly negative reactions of some ecologists to even the idea that stochastic process may underlie a number of classic ecological patterns. There appear to be two main reasons ecologists have tended to reject stochastic mechanism when attempting to explain ecological pattern: they are ‘too’ random and they ‘have nothing to do with ecology’. With regards to the first critique, why should pattern-producing random process not be considered a type of mechanism? This bias against stochastic models may be partly driven by the field of ecology having historically adopted a definition of ‘mechanism’ used by biological, molecular and cellular biologists, in which mechanism is quite explicit, concrete and deterministic – this protein bumps into that protein causing a conformation change and exposing an enzymatic site. But physics has seen a progression from deterministic laws such as those of Newton to the stochastic techniques and laws of Heisenberg and Schrödinger. Now as molecular and cellular biologists have begun scaling up into systems biology, stochastic models are of increasing importance. Why should ecology be any different? Perhaps the increasing incorporation of stochasticity is a sign of disciplinary maturity. Stochastic models should be seen as just another class of possible explanatory models that should be judged on their ability to elucidate our understanding of and make novel predictions about ecological systems. Although almost never verbalized, it also seems likely that considerable resistance to stochastic thinking is also based on the “it has nothing to do with ecology” argument. Ecology is relatively unique as a field because many scientists choose their discipline not because they are enamored with the process of science but also because they are enthralled with the organisms they study and the landscapes those organisms live in. Few physicists felt a loss at reducing atoms to randomly colliding abstractions, but many ecologists feel a loss at similarly reducing organisms to grist for an abstract, general principle like the central limit theorem or maximum entropy. But therein lies an irreconcilable problem: many of the patterns ecologists like to hold up as their own, such as the species abundance distribution, species area relationship or distance decay of similarity are in fact frequently repeated across entirely non-ecological systems such as geology, meteorology, economics, computer science, sociology, and the arts. Are we willing to throw these essential ecological descriptors away as not being truly ‘ecological’? Or, should we be willing to accept mechanisms that are general enough to be common between all these fields? Through my eclectic research program, I have rounded up my ‘blind men’ (the various fields that I conduct active research in) and asked them to report back on the nature of this thing called ‘ecology’. When viewed across multiple fields and types of observation, some general perspectives come into view – perspectives which are generally not articulated by practitioners in the field. Yet they appear to help define ‘ecology’: (1) The more general the pattern the more likely its ultimate cause lies in statistical mechanics. In ecological subdisciplines like community ecology and macroecology – which by their vary nature consider hundreds to billions of interacting individuals and dozens to thousands of species – it is highly probable that almost all emergent pattern is 253 generated from principles of statistical mechanics. This realization underlies the fact that the distribution of species abundance is very similar to the distribution of drinking glass longevities and Cowboy Junkies song performances. (2) Conversely, the smaller the system under investigation, and the more specific / idiosyncratic its patterns, the more likely it is governed by system-restricted process. Process restricted to purely ecological process (i.e. the interactions between individuals and their environment) is most apt to operate at the small scale including few interacting individuals and species and may not be generalizable. While typically such research has been dismissed (especially by theorists) as ‘natural history’ and ‘just-so stories’ yet it remains the most likely path to identifying ecological mechanism. (3) The presence of deterministic mechanistic process at the small scale does not mean that the sum across all such processes at larger scales will not be stochastic. While the evolutionary trajectory for a given species or clade and the explicit makeup of a particular individual ecological community may well be explained via deterministic, mechanistic ecological process, the emergent patterns created across the larger system including these components may well take on patterns indicative of stochastic process. This will happen when individual mechanistic processes – and their individual/species-specific parameters – are themselves randomly distributed. (4) The existence of universal patterns cannot be used to test for underlying mechanism beyond that of stochastic logic. The paradox that the same statistical distributions can be generated in model communities from complete resource competition or neutral models is explained by realizing both generate dynamical complexities, and it is these which generate expected patterns. In competition models complexity arises as multiple agents, each with individualistic resource utilization and dispersal functions, interact in an elaborate network with multiple other agents often over multiple temporal and spatial scales. Such interactions can be made even more complicated, but also more realistic, by inclusion of some degree of stochasticity in resource requirements and dispersal processes. In neutral models complexity arises as multiple agents, each with a unique community and metacommunity frequency, undergo a lottery process of replacement in space. These interactions are further complicated as new species are supplied to the system across multiple scales via dispersal and speciation. (5) Important mechanistic insights may be gained by looking for differences in the parameterized values and functional form of universal distributions, and deviations of pattern from universal expectations. For instance, in the case of the species area relationship, it has long been recognized that differences in slope on continents vs. across archipelagos conveys information about the mechanisms that generate and maintain diversity. Similarly, distance decay rates provide vital information about niche characteristics and dispersal capacities in relation to the environmental template of spatial and temporal variation. So, for example, isolated spruce-fir forests of the Appalachians demonstrate an almost 3-fold greater DD rate as compared to continuous northern taiga, and large-fruited, more dispersal-limited plant species have almost twice the DD rate as compared to smaller-seeded taxa. And the shape of the distance decay relationship 254 informs about whether samples were collected within the same (power-law) or across different (exponential) communities. Important insights may also be gained by identifying domains where universal expectations are not generated, such as the lower than expected frequency of rare and abundant organisms in ecological communities vs. other complex systems. (6) Efforts to accurately predict ecological outcomes, even at small deterministic scales, must be held with great suspect. The mathematics for >2 interacting individuals is highly complicated and falls within the framework of the Three Body Problem in astrophysics, which shows there is no single general outcome emerging from the gravitational interaction of 3 or more bodies (see right panel). Accurate prediction in such systems is entirely dependent upon exact knowledge of starting parameters, and as shown by the Heisenberg Uncertainty Principle and Observer Effects, such knowledge is impossible to obtain. While qualitative forecasts may be possible, precise, quantitative predictions of individual events pose great challenges. For example, it might be possible to predict qualitatively how community-wide species abundance distributions will be altered by invasions of multiple exotic species. It will be much more difficult, however, to predict the relative abundance and extent of spatial distribution of each exotic and native species. Indeed, such an effort may be as Quixotic as attempting to predict the exact date and magnitude of the next earthquake in the San Francisco Bay area or the price of a given stock in 2050. As in weather forecasting, while some level of short-term predictability will often be possible due to understanding current drivers, recent trajectories, and spatial/temporal autocorrelations, predictions will necessarily become increasingly imprecise as the forecast period increases. Rather than emphasizing prediction, ecologists perhaps should spend more effort on understanding the mechanisms and events that have conspired to generate current and past patterns. Even when it may be practically impossible to predict the future trajectory of a system, by looking backward it may be possible to deduce quite accurately when and how specific mechanisms came into play. In complex, human systems mechanisms are deduced by historians who analyze post hoc the particular combination of initial conditions and drivers that generated pattern. Similarly, community ecologists could enhance their understanding of the spatial and tempral patterns of biodiversity by becoming better natural historians, and using post hoc analysis to decipher how past events that have left a lasting influence on current conditions. While it may not be possible to predict future events with precision, by studying systems over time community ecologists may be able to establish what factors caused a particular outbreak of hantavirus or bird flu or simply whether Schrödinger's Cat survived. 255 (7) Statistical mechanics is not physics even though physicists got there first. Statistical mechanics simply represents the fundamental rules of logic for systems made of a large number of quasi-independent and variable parts. While this exists for physical systems (for instance the speed, trajectory, and atomic weights of gas molecules) it is even more applicable for ecological systems where agents possess many additional degrees of freedom. In some ways statistical mechanics may become even more fully expressed in ecological systems as compared to the simpler systems expressed by electrons, gasses, and fluids. (8) When searching for mechanism for a given ecological pattern, we must first consider processes larger than ecology. These include: mathematical logic; statistical mechanics (mathematical logic of large stochastic systems); physical laws; and chemical laws. What is left unexplained after this may be due to explicit ecological process (the interaction of organisms with themselves and the physical environment). 256