Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-27T04:36:16.685Z Has data issue: false hasContentIssue false

13 - Climate, Diversification and Refugia in the Common Shrew: Evidence from the Fossil Record

Published online by Cambridge University Press:  01 March 2019

Jeremy B. Searle
Affiliation:
Cornell University, New York
P. David Polly
Affiliation:
Indiana University
Jan Zima
Affiliation:
Academy of Sciences of the Czech Republic, Prague
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Agadjanian, A. K. and Serdyuk, N. V. (2005). The history of mammalian communities and paleogeography of the Altai Mountains in the Paleolithic. Paleontological Journal, 39 (Suppl. 6), S645S821.Google Scholar
Altuna, J. (1972). Fauna de Mamíferos de los Yacimientos Prehistóricos de Guipúzcoa, con Catalogo de los Mamíferos Cuaternarios del Cantábrico y del Pirineo Occidental. PhD dissertation. San Sebastián: Sociedad de Ciencias Naturales Aranzadi.Google Scholar
Barron, E. and Pollard, D. (2002). High-resolution climate simulations of Oxygen Isotope Stage 3 in Europe. Quaternary Research, 58, 296309.CrossRefGoogle Scholar
Barron, E., van Andel, T. H., and Pollard, D. (2003). Glacial environments II: reconstructing the climate of Europe in the last glaciation. In Neanderthals and Modern Humans in the European Landscape During the Last Glaciation, ed. van Andel, T. H. and Davies, W.. Cambridge, UK: McDonald Institute for Archaeological Research, pp. 5778.Google Scholar
Bartolomei, G. (1966). Diagramma microfaunistico con Sicista della Grotta della Ferrovia nella Gola della Rossa del fiume Esino presso Iesi (Ancona). Annali dell’Università di Ferrara (Nuova Serie), Sezion 9, 4, 6975.Google Scholar
Bartolomei, G. (1975). Indicazioni paleoecologiche. Il Gravettiano della Grotta Paglicci nel Gargano, A. Palma di Cesnola. Rivista di Scienze Preistoriche, 30, 159‒65.Google Scholar
Basset, P., Yannic, G., and Hausser, J. (2008). Chromosomal rearrangements and genetic structure at different evolutionary levels of the Sorex araneus group. Journal of Evolutionary Biology, 21, 842‒52.Google Scholar
Baum, B. R. (1992). Combining trees as a way of combining data sets for phylogenetic inference, and the desirability of combining gene trees. Taxon, 41, 310.Google Scholar
Bell, C. J., Gauthier, J. A., and Bever, G. S. (2010). Covert biases, circularity, and apomorphies: a critical look at the North American Quaternary herpetofaunal stability hypothesis. Quaternary International, 217, 30‒6.CrossRefGoogle Scholar
Bengtsson, B. O. and Frykman, I. (1990). Karyotype evolution: evidence from the common shrew (Sorex araneus L.). Journal of Evolutionary Biology, 3, 85101.Google Scholar
Bennett, K. D., Tzedakis, P. C., and Willis, K. J. (1991) Quaternary refugia of north European trees. Journal of Biogeography, 18, 103‒15.Google Scholar
Bilton, D. T., Mirol, P. M., Mascheretti, S., et al. (1998). Mediterranean Europe as an area of endemism for small mammals rather than a source for northwards postglacial colonization. Proceedings of the Royal Society of London B, 265, 1219‒26.Google Scholar
Bininda-Emonds, O. R. P., Gittleman, J. L., and Steel, M. A. (2002). The (super) tree of life: procedures, problems, and prospects. Annual Reviews of Systematics and Ecology, 33, 265‒89.Google Scholar
Binney, H.A., Willis, K. J., Edwards, M. E., et al. (2009). The distribution of late-Quaternary woody taxa in northern Eurasia: evidence from a new macrofossil database. Quaternary Science Reviews, 28, 2445‒64.Google Scholar
Borodin, A. V., Strukova, T. V., and Stefanovsky, B. B. (2003). Fossil remains of small mammals from alluvial and lake deposits of the Trans-Urals. In Quaternary Paleozoology of the Urals. Ekaterinburg: Ural University Press, pp. 7385. (In Russian).Google Scholar
Braconnot, P., Otto-Bliesner, B., Harrison, S., et al. (2007). Results of PMIP2 coupled simulations of the mid-Holocene and Last Glacial Maximum. Part 1: experiments and large-scale features. Climate of the Past, 3, 261–77.Google Scholar
Bronk Ramsey, C. (2008). Radiocarbon dating: revolutions in understanding. Archaeometry, 50, 249‒75.Google Scholar
Bronk Ramsey, C. (2009). Bayesian analysis of radiocarbon dates. Radiocarbon, 51, 337‒60.CrossRefGoogle Scholar
Brünner, H., Lugon-Moulin, N., Balloux, F., Fumagalli, L., and Hausser, J. (2002). A taxonomical re-evaluation of the Valais chromosome race of the common shrew Sorex araneus (Insectivora, Soricidae). Acta Theriologica, 47, 245‒75.Google Scholar
Capuzzi, P. and Sala, B. (1980). Il Riparo Tagliente. Analisi delle faune, biostratigrafia e cronologia dei livelli tardiglaciali. In Il Territorio Veronese dalle Origini all’Età Romana, ed. Fasani, L.. Verona: Fiorini, pp. 130‒6.Google Scholar
Cârciumaru, M., Moncel, M.-H., Anghelinu, M., et al. (2002). The Cioarei-Borosteni Cave (Carpathian Mountains, Romania): Middle Palaeolithic finds and technological analysis of the lithic assemblages. Antiquity, 76, 681‒90.Google Scholar
Chaline, J. (1972). Les Rongeurs du Pléistocene Moyen et Supérieur de France. Cahiers de Paléontologie. Paris: Centre National de la Recherche Scientifique.Google Scholar
Chlachula, J., Drozdov, N. I., and Ovodov, N. D. (2003). Last Interglacial peopling of Siberia: the Middle Palaeolithic site Ust’-Izhul’, the upper Yenisei area. Boreas, 32, 506–20.Google Scholar
Colbert, E. H., Hooijer, D. A., and Granger, W. (1953). Pleistocene mammals from the limestone fissures of Szechwan, China. Bulletin of the American Museum of Natural History, 102, 1134.Google Scholar
Coope, G. R., Gibbard, P. L., Hall, A. R., et al. (1997). Climate and environmental reconstructions based on fossil assemblages from the Middle Devensian (Weichselian) deposits of the River Thames at South Kensington, central London. Quaternary Science Reviews, 16, 1163‒96.Google Scholar
Currant, A. and Jacobi, R. (2001). A formal mammalian biostratigraphy for the Late Pleistocene of Britain. Quaternary Science Reviews, 20, 1707‒16.CrossRefGoogle Scholar
Dahlmann, T. (2001). Die Kleinsäuger der unter-pliozänen Fundstelle Wölfersheim in der Wetterau (Mammalia: Lipotyphla, Chiroptera, Rodentia). Courier Forschungsinstitut Senckenberg, 227, 1129.Google Scholar
Danukalova, G. and Yakovlev, A. (2006). A review of biostratigraphical investigations of Palaeolithic localities in the southern Urals region. Quaternary International, 149, 3743.Google Scholar
Davis, E. B., McGuire, J. L., and Orcutt, J. D. (2014). Ecological niche models of mammalian glacial refugia show consistent bias. Ecography, 37, 1133‒8.Google Scholar
Derevyanko, A. P., Agadjanian, A. K., Kylik, N. A., et al. (2001). Basic results of the study of the multilayer site Ust’-Kanskaya Cave. Problemi Arkheologii, Etnografii, Antropologii Sibirie i Sopredel’nykh Territorii, 7, 109–14. (In Russian).Google Scholar
Dobosi, V. T. and Vörös, I. (1986). Chronological revision of the Pilisszántó Rock Shelter II. Folia Archaeologica, 37, 2545.Google Scholar
Dobosi, V. T. and Vörös, I. (1987). The Pilisszántó Rock Shelter revision. Folia Archaeologica, 38, 764.Google Scholar
Dodonov, A. E., Tesakov, A. S., Titov, V. V., et al. (2007). New data on the stratigraphy of the Plio-Pleistocene of the lower Don River banks on the margins of the Tsimlyansk Reservoir. In Geological Events of the Neogene and Quaternary of Russia, ed. Gladenkov, Y. B.. Moscow: Geological Institute of the Russian Academy of Sciences, pp. 4353. (In Russian).Google Scholar
Dolgov, V. A. (1985). Red-Toothed Shrews of the Old World. Moscow: Moscow University Press. (In Russian).Google Scholar
Duckworth, W. L. H. (1912). Cave exploration at Gibraltar in 1911. Journal of the Royal Anthropological Institute, 42, 515‒28.Google Scholar
Ehlers, J. and Gibbard, P. L. (eds) (2004). Quaternary Glaciations: Extent and Chronology. Part 1: Europe. Amsterdam: Elsevier.Google Scholar
Engelmann, G. F. and Wiley, E. O. (1977). The place of ancestor-descendant relationships in phylogeny reconstruction. Systematic Biology, 26, 111.Google Scholar
Fadeeva, T. (2016). Insectivorous mammals (Lipotyphla, Soricidae) of the Perm Pre-Ural in the Late Pleistocene and Holocene time. Quaternary International, 420, 156‒70.Google Scholar
Field, M. H. and Peglar, S. M. (2010). A palaeobotanical investigation of the sediments from the West Runton Mammoth site. Quaternary International, 228, 3845.Google Scholar
Fisher, D. C. (1994). Stratocladistics: morphological and temporal patterns and their relation to phylogenetic process. In Interpreting the Hierarchy of Nature, ed. Grande, L. and Rieppel, O.. Cambridge, MA: Academic Press, pp. 133‒71.Google Scholar
Fisher, D. C. (2008). Stratocladistics: integrating temporal data and character data in phylogenetic inference. Annual Review of Ecology, Evolution, and Systematics, 39, 365‒85.Google Scholar
Flickert, T., Friend, D., Grünneger, F., et al. (2007). Did debris-covered glaciers serve as Pleistocene refugia for plants? A new hypothesis derived from observations of recent plant growth on glacier surfaces. Arctic, Antarctic, and Alpine Research, 29, 245‒57.Google Scholar
Folland, C. K., Karl, T. R., Vinnikov, K. Y. (1990). Observed climate variations and change. In Climate Change, the IPCC Scientific Assessment. Cambridge, UK: Cambridge University Press, pp. 201‒38.Google Scholar
Fumagalli, L., Taberlet, P., Stewart, D. T., et al. (1999). Molecular phylogeny and evolution of Sorex shrews (Soricidae: Insectivora) inferred from mitochondrial DNA sequence data. Molecular Phylogenetics and Evolution, 11, 222‒35.Google Scholar
Gaffney, V., Thomson, K., and Fitch, S. (2007). Mapping Doggerland: the Mesolithic Landscapes of the Southern North Sea. Oxford: Archaeopress.Google Scholar
Gao, X., Huang, W., Xu, Z., Ma, Z., and Olsen, J. W. (2004). 120–150 ka human tooth and ivory engravings from Xinglongdong Cave, Three Gorges Region, South China. Chinese Science Bulletin, 49, 175‒80.Google Scholar
Gibbons, A. (2011). Who were the Denisovans? Science, 333, 1084‒7.Google Scholar
Gingerich, P. D. (1979). The stratophenetic approach to phylogeny reconstruction in vertebrate paleontology. In Phylogenetic Analysis and Paleontology, ed. Cracraft, J. and Eldredge, N.. New York: Columbia University Press, pp. 4177.Google Scholar
Gómez-Robles, A., Bermúdez de Castro, J. M., Arsuaga, J.-L., Carbonell, E., and Polly, P. D. (2013). No known hominin species matches the expected dental morphology of the last common ancestor of Neanderthals and modern humans. Proceedings of the National Academy of Sciences USA, 110, 18196‒201.Google Scholar
Graham, A. (2011). A Natural History of the New World: the Ecology of Plants in the Americas. Chicago: University of Chicago Press.Google Scholar
Graham, R. W., Lundelius, E. L. Jr, Graham, M. A., et al. (1996). Spatial response of mammals to late Quaternary environmental fluctuations. Science, 272, 1601‒6.Google Scholar
Guiot, J., de Beaulieu, J. L., Cheddadi, R., et al. (1993). The climate in western Europe during the last glacial/interglacial cycle derived from pollen and insect remains. Palaeogeography, Palaeoclimatology, Palaeoecology, 103, 7394.Google Scholar
Guthrie, R. D. (1980). Bison and man in North America. Canadian Journal of Anthropology, 1, 5573.Google Scholar
Guthrie, R. D. (2001). Origin and causes of the mammoth steppe: a story of cloud cover, woolly mammal tooth pits, buckles, and inside-out Beringia. Quaternary Science Reviews, 20, 549‒74.Google Scholar
Harrison, D. L. (1996). Systematic status of Kennard’s shrew (Sorex kennardi Hinton, 1911, Insectivora: Soricidae): a study based on British and Polish material. Acta Zoologica Cracoviensia, 39, 201‒12.Google Scholar
Hastings, D. and Dunbar, P. (1998). Development and assessment of the Global Land One-km Base Elevation digital elevation model (GLOBE). In IAPRS Commission IV Symposium on GIS – Between Visions and Applications, ed. Fritsch, D., Englich, M., and Sester, M.. IAPRS, 32, 218‒21.Google Scholar
Hatfield, T., Barton, N., and Searle, J. B. (1992). A model of a hybrid zone between two chromosomal races of the common shrew (Sorex araneus). Evolution, 46, 1129‒45.Google Scholar
Hausser, J. (1990). Sorex coronatus Millet, 1882 ‒ Schabrackenspitzmaus. In Handbuch der Säugetiere Europas, Band 3/1, ed. Niethammer, J. and Krapp, F.. Wiesbaden: Aula-Verlag, pp. 279‒86.Google Scholar
Hausser, J. (1994). The Sorex of the araneus-arcticus group (Mammalia: Soricidae): do they actually speciate? In Advances in the Biology of Shrews, ed. Merritt, J. F., Kirkland, G. L. Jr, and Rose, R. K.. Pittsburgh: Carnegie Museum of Natural History, Special Publication No. 18, pp. 295306.Google Scholar
Hausser, J. and Jammot, D. (1974). Etude biométrique des mâchoires chez les Sorex du groupe araneus en Europe continentale (Mammalia, Insectivora). Mammalia, 38, 324‒43.Google Scholar
Hays, J., Imbrie, J., and Shackleton, N. (1976). Variations in the Earth’s orbit: pacemaker of the ice ages. Science, 194, 1121‒32.Google ScholarPubMed
Hearty, P. J., Hollin, J. T., Neumann, A. C., et al. (2007). Global sea-level fluctuations during the Last Interglaciation (MIS 5e). Quaternary Science Reviews, 26, 2090‒112.Google Scholar
Heller, F. (1930). Eine Forest-Bed Fauna aus der Sackdillinger Hohle (Oberpfalz). Neues Jahrbuch für Mineralogie, Geologie und Palaontologie. Abteilung B, 63, 247‒98.Google Scholar
Herman, J. S. and Searle, J. B. (2011). Post-glacial partitioning of mitochondrial genetic variation in the field vole. Proceedings of the Royal Society B, 278, 3601–7.Google Scholar
Hewitt, G. M. (1996). Some genetic consequences of ice ages, and their role in divergence and speciation. Biological Journal of the Linnean Society, 58, 247‒76.Google Scholar
Hewitt, G. M. (1999). Post‐glacial re‐colonization of European biota. Biological Journal of the Linnean Society, 68, 87112.Google Scholar
Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G., and Jarvis, A. (2005). Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology, 25, 1965‒78.Google Scholar
Hillson, S. (2005). Teeth. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Hoek Ostende, L. W., van den, Reumer, J. W. F., and Doukas, C. S. (2010). The nature of the fossil record of Neogene insectivores. Hellenic Journal of Geosciences, 44, 117‒24.Google Scholar
Horáček, I. (1985). Quaternary morphoclines and changes in community structure in European shrews (Mammalia, Soricidae). In Evolution and Morphogenesis, ed. Mlíkovský, J. and Novák, V. J. A.. Prague: Academia, pp. 799810.Google Scholar
Horáček, I. (2005). Small vertebrates of the Weichselian series in Dzeravá Skala Cave: list of the samples and a brief summary. In Pleistocene Environments and Archaeology of the Dzeravá Skala Cave, Lesser Carpathians, Slovakia, ed. Kominská, L., Kozłowski, J. K., and Svoboda, J.. Kraków: Polska Akademia Umiejętności, pp. 157‒67.Google Scholar
Horáček, I., Ložek, V., Knitlová, M., et al. (2015). Darkness under candlestick: glacial refugia on mountain glaciers. In Forgotten Times and Spaces, ed. Sázelová, S., Novák, M., and Mizerová, A.. Brno: Institute of Archaeology of the Czech Academy of Sciences and Masaryk University, pp. 363‒77.Google Scholar
Horáček, I. and Sánchez-Marco, A. (1984). Comments on the Weichselian small mammal assemblages in Czechoslovakia and their stratigraphical interpretation. Neues Jahrbuch für Geologie und Paläontologie Monatshefte, 9, 560‒76.Google Scholar
Huertas, A. D., Iacumin, P., and Longinelli, A. (1997). A stable isotope study of fossil mammal remains from the Paglicci cave, southern Italy, 13 to 33 ka BP: palaeoclimatological considerations. Chemical Geology, 141, 211‒23.Google Scholar
Huntley, B. and Allen, J. R. M. (2003). Glacial environments III: Palaeo-vegetation patterns in Late Glacial Europe. In Neanderthals and Modern Humans in the European Landscape During the Last Glaciation, ed. van Andel, T. H. and Davies, W.. Cambridge, UK: McDonald Institute for Archaeological Research, pp. 79102.Google Scholar
Jammot, D. (1977). Les Musaraignes (Soricidae, Insectivora) du Plio-Pléstocène d’Europe. Dissertation, University of Bourgogne, Dijon.Google Scholar
Janis, C. M., Gunnell, G. F., and Uhen, M. D. (eds) (2008). Evolution of Tertiary Mammals of North America. Vol. 2. Cambridge, UK: Cambridge University Press.Google Scholar
Janis, C. M., Scott, K. M., and Jacobs, L. L. (eds) (1998). Evolution of Tertiary Mammals of North America. Vol. 1. Cambridge, UK: Cambridge University Press.Google Scholar
Jánossy, D. (1969). Stratigraphische Auswertung der europäischen mittelpleistozänen Wirbeltierfauna. Berichte der Deutschen Gesellschaft für Geologische Wissenschaften, Reihe A, 14, 367438.Google Scholar
Jánossy, D. (1986). Pleistocene vertebrate faunas of Hungary. Developments in Paleontology and Stratigraphy, 8, 2208.Google Scholar
Jensen, J. B., Bennike, O., Witkowski, A., et al. (1999). Early Holocene history of the southwestern Baltic Sea: the Ancylus Lake stage. Boreas, 28, 437‒53.Google Scholar
Kahlke, R.-D. (2006). Untermassfeld: a Late Early Pleistocene (Epivillafranchian) Fossil Site near Meiningen (Thuringia, Germany) and Its Position in the Development of the European Mammal Fauna, BAR International Series 1578. Oxford: Archaeopress.Google Scholar
Kalthoff, D. C. (1998). Die Kleinsäuger (Mammalia) der Fundstelle Kettig (Rheinland-Pfalz, Deutschland) im Rahmen der allerödzeitlichen Säugetierfauna Mittel-und Süddeutschlands. Paläontologische Zeitschrift, 72, 407‒24.Google Scholar
Kalthoff, D. C., Mörs, T., and Tesakov, A. (2007). Late Pleistocene small mammals from the Wannenköpfe volcanoes (Neuweid Basin, western Germany) with remarks on the stratigraphic range of Arvicola terrestris. Geobios, 40, 609‒23.Google Scholar
Khenzykhenova, F. I. (2008). Paleoenvironments of Palaeolithic humans in the Baikal region. Quaternary International, 179, 53‒7.Google Scholar
Kidwell, S. M. and Holland, S. M. (2002). The quality of the fossil record: implications for evolutionary analyses. Annual Reviews of Ecology and Systematics, 33, 561‒88.Google Scholar
Kleiven, H., Jansen, E., Fronval, T., and Smith, T. (2002). Intensification of northern hemisphere glaciations in the circum-Atlantic region (3.5–2.4 ma) ‒ ice-rafted detritus evidence. Palaeogeography, Palaeoclimatology, Palaeoecology, 184, 213–23.Google Scholar
Knitlová, M. and Horáček, I. (2017). Late Pleistocene-Holocene paleobiogeography of the genus Apodemus in central Europe. PLoS ONE, 12, e0173668.Google Scholar
Kotsakis, T., Abbazzi, L., Angelone, C., et al. (2003). Plio-Pleistocene biogeography of Italian mainland micromammals. Deinsea, 10, 313‒42.Google Scholar
Kotsakis, T., Marcolini, F., De Rita, D., Conti, M., and Esu, D. (2011). Three Late Pleistocene small mammal faunas from the Baccano maar (Rome, central Italy). Bollettino della Società Paleontologica Italiana, 50, 103‒10.Google Scholar
Kowalski, K. (2001). Pleistocene rodents of Europe. Folia Quaternaria, 72, 1389.Google Scholar
Krokhmal, A. I. and Rekovets, L. I. (2007). Pleistocene Small Mammal Localities from the Ukraine and Adjacent Territories. Kiev: LAT & K Publishers. (In Ukrainian).Google Scholar
Kurtén, B. (1968). Pleistocene Mammals of Europe. Chicago: Aldine Publishing.Google Scholar
Lambeck, K., Purcell, A., Zhao, J., et al. (2010). The Scandinavian ice sheet: from MIS 4 to the end of the Last Glacial Maximum. Boreas, 39, 410‒35.Google Scholar
Lawing, A. M. and Polly, P. D. (2011). Pleistocene climate, phylogeny, and climate envelope models: an integrative approach to better understand species’ response to climate change. PLoS ONE, 16, e28554.Google Scholar
Leroi-Gourhan, A. and Leroi-Gourhan, A. (1964). Chronologie des grottes d’Arcy-sur-Cure (Yonne). Gallia Préhistoire, 7, 164.Google Scholar
Li, C. L. and Xue, X. X. (1996). Biogeography and the age of the fossil rodent fauna from Xishuidong, Lantian, Shaanxi. Vertebrata Palasiatica, 34, 156‒62.Google Scholar
Li, Y. X., Zhang, Y. X., and Ao, H. (2013). Sorex fossils (Soricidae, Insectivora) from the Middle Pleistocene cave site of Shanyangzhai, Hebei Province, China. Quaternary International, 298, 187‒95.Google Scholar
Lindsay, E. H., Opdyke, N. D., and Fejfar, O. (1997). Correlation of selected late Cenozoic European mammal faunas with the magnetic polarity time scale. Palaeogeography, Palaeoclimatology, Palaeoecology, 133, 205‒26.Google Scholar
Lisiecki, L. E. and Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleooceanography, 20, PA1003.Google Scholar
López Martínez, N. (1972). Los Micromamíferos del Cuaternario de Rincón de la Victoria. Boletín de la Real Societad Espaniola de Historia Natural (Geologia), 70, 223‒33.Google Scholar
Ložek, V., Tyráček, J., and Fejfar, O. (1958). Die Quärtaren Sedimenten der Felschnische auf der Velká Kobylanka bei Hranice (Weisskirchen). Anthropozoikum, 8, 177203.Google Scholar
McKenna, M. C. and Bell, S. K. (1997). Classification of Mammals Above the Species Level. New York: Columbia University Press.Google Scholar
Mackiewicz, P., Moska, M., Wierzbicki, H., et al. (2017). Evolutionary history and phylogeographic relationships of shrews from the Sorex araneus group. PLoS ONE, 12, e0179760.Google Scholar
Markova, A. K. (2007). Pleistocene mammal faunas of Eastern Europe. Quaternary International, 160, 100‒11.Google Scholar
Masini, F., Giannini, T., Abbazzi, L., et al. (2005). A latest Biharian small vertebrate fauna from the lacustrine succession of San Lorenzo (Sant’Arcangelo Basin, Basilicata, Italy). Quaternary International, 131, 7993.Google Scholar
Maul, L. C. (1990). Überblick über die unterpleistozänen Kleinsäugerfaunen Europas. Quartärpaläeontologie, 8, 153‒91.Google Scholar
Maul, L. C. and Parfitt, S. A. (2010). Micromammals from the 1995 Mammoth Excavation at West Runton, Norfolk, UK: morphometric data, biostratigraphy, and taxonomic reappraisal. Quaternary International, 228, 91115.Google Scholar
Mészáros, L. G. (2004). Taxonomical revision of the Late Würm Sorex (Mammalia, Insectivora) remains of Hungary, for proving the presence of an alpine ecotype in the Pilisszántó Horizon. Annales Universitatis Scientiarum Budapestinensis, Sectio Geologica, 34, 925.Google Scholar
Mezhzherin, V. A. (1972). Shrews (Sorex, Insectivora, Mammalia) from Pleistocene deposits of the USSR. In Theriologica 1, ed. Kolosov, L. D. and Lukyanov, I. V.. Novosibirsk: Nauka, pp. 117‒30. (In Russian).Google Scholar
Mitchell-Jones, A. J., Amori, G., Bogdanowicz, W., et al. (1999). The Atlas of European Mammals. London: Poyser.Google Scholar
Movius, H. L. Jr. (1977). Excavation of the Abri Pataud, Les Eyzies (Dordogne). Cambridge, MA: Peabody Museum of Archaeology and Ethnology.Google Scholar
Myers, C. E., Stigall, A. L., and Lieberman, B. S. (2015). PaleoENM: applying ecological niche modeling to the fossil record. Paleobiology, 41, 226‒44.Google Scholar
Nadachowski, A., Stefaniak, K., Szynkiewicz, A., et al. (2011). Biostratigraphic importance of the Early Pleistocene fauna of Żabia Cave (Poland) in Central Europe. Quaternary International, 243, 204‒18.Google Scholar
Nagel, D., Rabeder, G., and Reiner, G. (1995). Die Insektivoren und Rodentier (Mammalia) aus dem Spätglazial der Gamssulzenhöhle im Toten Gebirge (Oberösterreich). Mittellungen der Kommission für Quartärforschung der Österreichischen Akademie der Wien, 9, 61‒8.Google Scholar
Neet, C. R. and Hausser, J. (1990). Habitat selection in zones of parapatric contact between the common shrew Sorex araneus and Millet’s shrew S. coronatus. Journal of Animal Ecology, 59, 235–50.CrossRefGoogle Scholar
NGRIP members (2004). High-resolution record of Northern Hemisphere climate extending into the last interglacial period. Nature, 431, 147‒51.Google Scholar
Nikolova, I., Yin, Q., Berger, A., et al. (2013). The last interglacial (Eemian) climate simulated by LOVECLIM and CCSM3. Climate of the Past, 9, 1789‒806.Google Scholar
Nogués-Bravo, D. (2009). Predicting the past distribution of species climatic niches. Global Ecology and Biogeography, 18, 521‒31.Google Scholar
Norell, M. A. (1993). Tree-based approaches to understanding history: comments on ranks, rules, and the quality of the fossil record. American Journal of Science, 293A, 407‒17.Google Scholar
Ohdachi, S. D., Hasegawa, H., Iwasa, M. A., et al. (2006). Molecular phylogenetics of soricid shrews (Mammalia) based on mitochondrial cytochrome b gene sequences: with special reference to the Soricinae. Journal of Zoology, 270, 177‒91.Google Scholar
Ohdachi, S. D., Masuda, R., Abe, H., et al. (1997). Phylogeny of Eurasian soricine shrews (Insectivora, Mammalia) inferred from the mitochondrial cytochrome b gene sequences. Zoological Science, 14, 527‒32.Google Scholar
Opdyke, N., Mein, P., Lindsay, E., et al. (1997). Continental deposits, magnetostratigraphy and vertebrate paleontology, late Neogene of eastern Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 133, 129‒48.Google Scholar
Oppo, D. W., McManus, J. F., and Cullen, J. L. (2006). Evolution and demise of the Last Interglacial warmth in the subpolar North Atlantic. Quaternary Science Reviews, 25, 3268‒77.Google Scholar
Osipova, V. A., Rzebik-Kowalska, B., and Zaitsev, M. V. (2006). Intraspecific variability and phylogenetic relationship of the Pleistocene shrew Sorex runtonensis (Soricidae). Acta Theriologica, 51, 129‒38.Google Scholar
Pääbo, S., Poinar, H., Serre, D., et al. (2004). Genetic analyses from ancient DNA. Annual Reviews of Genetics, 38, 645‒79.Google Scholar
Pagani, M., Liu, Z., LaRiviere, J., and Ravelo, A. C. (2010). High Earth-system climate sensitivity determined from Pliocene carbon dioxide concentrations. Nature Geoscience, 3, 2730.Google Scholar
Parducci, L., Jørgensen, T., Tollefsrud, M. M., et al. (2012). Glacial survival of boreal trees in northern Scandinavia. Science, 335, 1083–6.Google Scholar
Parfitt, S. A. (1998). The interglacial mammalian fauna from Barnham. In Excavations at the Lower Palaeolithic site at East Farm, Barnham, Suffolk 1989–94, ed. Ashton, N. A., Lewis, S. G., and Parfitt, S. A.. London: British Museum Press, pp. 111‒47.Google Scholar
Parham, J. F., Donoghue, P. C., Bell, C. J., et al. (2011). Best practices for justifying fossil calibrations. Systematic Biology, 61, 346‒59.Google Scholar
Pavlova, S. V., Borisov, S. A., Timoshenko, A. F., and Sheftel, B. I. (2017). ‘European’ race-specific metacentrics in East Siberian common shrews (Sorex araneus): a description of two new chromosomal races, Irkutsk and Zima. Comparative Cytogenetics, 11, 797806.Google Scholar
Pei, W. C. (1936). The mammalian remains from Locality 3 at Choukoutien. Palaeontologica Sinica C, 7, 1108.Google Scholar
Peterson, A. T. (2003). Predicting the geography of species’ invasions via ecological niche modeling. Quarterly Review of Biology, 78, 419‒33.Google Scholar
Phillips, S.J. and Dudik, M. (2008). Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography, 31, 161–75.Google Scholar
Polly, P. D. (1997). Ancestry and species definition in paleontology: a stratocladistic analysis of Viverravidae (Carnivora, Mammalia) from Wyoming. Contributions from the Museum of Paleontology, University of Michigan, 30, 153Google Scholar
Polly, P. D. (2001). On morphological clocks and paleophylogeography: towards a timescale for Sorex hybrid zones. Genetica, 112/113, 339‒57.Google Scholar
Polly, P. D. (2003). Paleophylogeography of Sorex araneus: molar shape as a morphological marker for fossil shrews. Mammalia, 68, 233‒43.Google Scholar
Polly, P. D. (2007). Phylogeographic differentiation in Sorex araneus: morphology in relation to geography and karyotype. Russian Journal of Theriology, 6, 7384.Google Scholar
Polly, P. D. (2010). Tiptoeing through the trophics: geographic variation in carnivoran locomotor ecomorphology in relation to environment. In Carnivoran Evolution: New Views on Phylogeny, Form, and Function, ed. Goswami, A. and Friscia, A.. Cambridge, UK: Cambridge University Press, pp. 374410.Google Scholar
Polly, P. D. and Eronen, J. T. (2011). Mammal associations in the Pleistocene of Britain: implications of ecological niche modelling and a method for reconstructing palaeoclimate. In The Ancient Human Occupation of Britain, ed. Ashton, N., Lewis, S. G., and Stringer, C.. Amsterdam: Elsevier, pp. 279304.Google Scholar
Polly, P. D., Le Comber, S. C., and Burland, T. M. (2005). On the occlusal fit of tribosphenic molars: are we underestimating species diversity in the Mesozoic? Journal of Mammalian Evolution, 12, 283‒99.Google Scholar
Polly, P. D., Polyakov, A. V., Ilyashenko, V. B., et al. (2013). Phenotypic variation across chromosomal hybrid zones of the common shrew (Sorex araneus) indicates reduced gene flow. PLoS ONE, 8, e67455.Google Scholar
Polyakov, A. V., Onischenko, S. S., Ilyashenko, V.B., et al. (2002). Morphometric difference between the Novosibirsk and Tomsk chromosome races of Sorex araneus in a zone of parapatry. Acta Theriologica, 47, 381‒7.Google Scholar
Polyakov, A. V., Panov, V. V., Ladygina, T. Y., et al. (2001). Chromosomal evolution of the common shrew Sorex araneus L. from the southern Urals and Siberia in the Postglacial Period. Russian Journal of Genetics, 37, 351‒7.Google Scholar
Polyakov, A. V., Volobouev, V. T., Borodin, P. M., et al. (1996). Karyotypic races of the common shrew (Sorex araneus) with exceptionally large ranges: the Novosibirsk and Tomsk races of Siberia. Hereditas, 125, 109‒15.Google Scholar
Polyakov, A. V., Zima, J., Banaszek, A., et al. (2000a). New chromosome races of the common shrew Sorex araneus from Eastern Siberia. Acta Theriologica, 45 (Suppl. 1), 1117.Google Scholar
Polyakov, A. V., Zima, J., Searle, J. B., et al. (2000b). Chromosome races of the common shrew Sorex araneus in the Ural Mts: a link between Siberia and Scandinavia? Acta Theriologica, 45 (Suppl. 1), 1926.Google Scholar
Popov, V. V. (1988). Middle Pleistocene small mammals (Mammalia: Insectivora, Lagomorpha, Rodentia) from Varbeshnitsa (Bulgaria). Acta Zoologica Cracoviensia, 31, 193234.Google Scholar
Popov, V. V. (2000). The small mammals (Mammalia: Insectivora, Chiroptera, Lagomorpha, Rodentia) from Cave 16 and the paleoenvironmental changes during the Late Pleistocene. In Temnata Cave: Excavation in Karlukova Karst Area, Bulgaria, ed. Kozłowski, J. K., Laville, H., and Ginters, B.. Kraków: Jagellonian University, pp. 159240.Google Scholar
Popova, L. V. (2014). Small mammal fauna as an evidence of environmental dynamics in the Holocene of Ukrainian area. Quaternary International, 357, 8292.Google Scholar
Price, C. R. (2003). Late Pleistocene and Early Holocene Small Mammals in Southwest Britain: Environmental and Taphonomic Implications and their Role in Archaeological Research. Oxford: Archaeopress.Google Scholar
Prost, S., Klietmann, J., Van Kolfschoten, T., et al. (2013). Effects of Late Quaternary climate change on Palearctic shrews. Global Change Biology, 19, 1865‒74.Google Scholar
Qiu, Z., Zhang, Y., and Hu, S. (1985). Human tooth and paleoliths found at locality 2 of Longtanshan, Chenggong, Kunming. Acta Anthropologica Sinica, 4, 233‒41.Google Scholar
Rabeder, G. (1974). Die Kleinsäugerfauna des Jungpliozäns von Stranzendorf. Mitteilungen den Quartärkommission der Österreichische Akademie Wissenschaften, 1, 137‒9.Google Scholar
Rabeder, G. (1995). Die Gamssulzenhöhle im Toten Gebirge. Mitteilungen den Quartärkommission der Österreichische Akademie Wissenschaften, 9, 1133.Google Scholar
Rasmussen, S. O., Andersen, K. K., Svensson, A. M., et al. (2006). A new Greenland ice core chronology for the last glacial termination. Journal of Geophysical Research, 111, D06102.Google Scholar
Raymo, M. E. and Ruddiman, W. F. (1992). Tectonic forcing of late Cenozoic climate. Nature, 359, 117‒22.Google Scholar
Reumer, J. W. F. (1984). Ruscinian and early Pleistocene Soricidae (Insectivora, Mammalia) from Tegelen (The Netherlands) and Hungary. Scripta Geologica, 73, 1173.Google Scholar
Reumer, J. W. F. and Meylan, A. (1986). New developments in vertebrate cytotaxonomy IX: chromosome numbers in the order Insectivora (Mammalia). Genetica, 70, 119‒51.Google Scholar
Rofes, J. and Cuenca-Bescós, G. (2009). A new genus of red-toothed shrew (Mammalia, Soricidae) from the Early Pleistocene of Gran Dolina (Atapuerca, Burgos, Spain), and a phylogenetic approach to the Eurasiatic Soricinae. Zoological Journal of the Linnean Society, 155, 904‒25.Google Scholar
Rofes, J., Moya-Costa, R., Bennàsar, M., Blain, H.-A., and Cuenca-Bescós, G. (2016). Biostratigraphy, palaeogeography, and palaeoenvironmental significance of Sorex runtonensis Hinton, 1911 (Mammalia, Soricidae): first record from the Iberian Peninsula. Palaeogeography, Palaeoclimatology, Palaeoecology, 459, 508‒17.Google Scholar
Rose, K. D. and Bown, T. M. (1993). Species concepts and species recognition in Eocene primates. In Species, Species Concepts, and Species Evolution, ed. Kimbel, W. H. and Martin, L. B.. New York: Plenum Press, pp. 299330.Google Scholar
Rössner, G. E. and Heissig, K. (1999). The Miocene Land Mammals of Europe. Munich: Verlag Dr. Friedrich Pfeil.Google Scholar
Rudenko, S. I., Wormington, H. M., and Chard, C. S. (1961). The Ust’-Kanskaia Palaeolithic cave site, Siberia. American Antiquity, 27, 203‒15.Google Scholar
Rzebik-Kowalska, B. (1994). Pliocene and Quaternary Insectivora (Mammalia) of Poland. Acta Zoologica Cracoviensia, 37, 77136.Google Scholar
Rzebik-Kowalska, B. (1998). Fossil history of shrews in Europe. In Evolution of Shrews, ed. Wójcik, J. M. and Wolsan, M.. Białowieża: Mammal Research Institute, pp. 2392.Google Scholar
Rzebik-Kowalska, B. (2000). Insectivora (Mammalia) from the Early and Middle Pleistocene of Betfia in Romania. I. Soricidae Fischer von Waldheim, 1817. Acta Zoologica Cracoviensia, 43, 153.Google Scholar
Rzebik-Kowalska, B. (2005). Poland. In The Fossil Record of the Eurasian Neogene Insectivores (Erinaceomorpha, Soricomorpha, Mammalia), Part I, ed. van den Hoek Ostende, L. W., Doukas, C. S., and Reumer, J. W. F.. Scripta Geologica Special Issue, 5, 119‒34.Google Scholar
Rzebik-Kowalska, B. (2006). Erinaceomorpha and Soricomorpha (Mammalia) from the Late Pleistocene and Holocene of Krucza Skała Rock Shelter and Komorowa Cave (Poland). Acta Zoologica Cracoviensia, 49A, 83118.CrossRefGoogle Scholar
Rzebik-Kowalska, B. (2007). New data on Soricomorpha (Lipotyphla, Mammalia) from the Pliocene and Pleistocene of Transbaikalia and Irkutsk region (Russia). Acta Zoologica Cracoviensia, 50A, 1548.Google Scholar
Rzebik-Kowalska, B. (2008). Insectivores (Soricomorpha, Mammalia) from the Pliocene and Pleistocene of Transbaikalia and Irkutsk region (Russia). Quaternary International, 179, 96100.Google Scholar
Rzebik-Kowalska, B. (2013). Sorex bifidus n. sp. and the rich insectivore mammal fauna (Erinaceomorpha, Soricomorpha, Mammalia) from the Early Pleistocene of Żabia Cave in Poland. Palaeontologia Electronica, 16.2.12A, 135.Google Scholar
Rzebik-Kowalska, B. and Rekovets, L. I. (2015). Recapitulation of data on Ukrainian fossil insectivore mammals (Eulipotyphla, Insectivora, Mammalia). Acta Zoologica Cracoviensia, 58, 137‒71.Google Scholar
Sala, B. (1980). Interpretazione crono-bio-stratigrafica dei depositi pleistocenici della Grotta del Broion (Vicenza). Geografia Fisica e Dinamica Quaternaria, 3, 6671.Google Scholar
Sala, B. (1983). Variations climatiques et séquences chronologiques sur la bas des variations des associations fauniques à grands mammifères. Revista di Scienze Preistoriche, 38, 161‒80.Google Scholar
Sala, B. and Masini, F. (2007). Late Pliocene and Pleistocene small mammal chronology in the Italian peninsula. Quaternary International, 160, 416.Google Scholar
Schaeffer, H. (1975). Die Spitzmäuse der Hohen Tatra seit 30,000 Jahren (Mandibular Studie). Zoologischer Anzeiger, 95, 89111.Google Scholar
Schreve, D. C. (2000). The vertebrate assemblage from Hoxne, Suffolk. In The Quaternary of Norfolk and Suffolk Field Guide, ed. Lewis, S. G., Whiteman, C. A., and Preece, R. C.. London: Quaternary Research Association, pp. 155‒63.Google Scholar
Searle, J. B. (1984). Three new karyotypic races of the common shrew Sorex araneus (Mammalia: Insectivora) and a phylogeny. Systematic Zoology, 33, 184‒94.Google Scholar
Searle, J. B. (1986). Factors responsible for a karyotypic polymorphism in the common shrew, Sorex araneus. Proceedings of the Royal Society of London B, 229, 277–98.Google Scholar
Searle, J. B., Kotlík, P., Rambau, R. V., et al. (2009). The Celtic fringe of Britain: insights from small mammal phylogeography. Proceedings of the Royal Society B, 276, 4287–94.Google Scholar
Searle, J. B. and Wilkinson, P. J. (1987). Karyotypic variation in the common shrew (Sorex araneus) in Britain – a ‘Celtic fringe’. Heredity, 59, 345‒51.Google Scholar
Serduyk, N. and Zenin, A. (2016). Small mammals from the Strashnaya cave (Northwest Altai, West Siberia, Russia). Quaternary International, 406B, 162‒8.Google Scholar
Sesé, C. (1994). Paleoclimatological interpretation of the Quaternary small mammals of Spain. Geobios, 27, 753‒67.Google Scholar
Sher, A. V. (1971). Mammals and stratigraphy of the Pleistocene of the extreme northeast of the USSR and North America. International Geology Review, 16, 1284. (In Russian).Google Scholar
Sher, A. V., Weinstock, J., Baryshnikov, G. F., et al. (2011). The first record of ‘spelaeoid’ bears in Arctic Siberia. Quaternary Science Reviews, 30, 2238‒49.Google Scholar
Shotton, F. W., Keen, D. H., Coope, G. R. et al. (1993). The Middle Pleistocene deposits of Waverly Wood Pit, Warwickshire, England. Journal of Quaternary Science, 8, 293325.Google Scholar
Sommer, R. S. and Nadachowski, A. (2006). Glacial refugia of mammals in Europe: evidence from fossil records. Mammal Review, 36, 251‒65.Google Scholar
Srdoč, D., Sliepčevic, A., Obelic, B., et al. (1979). Rudjer Bošković Institute radiocarbon measurements V. Radiocarbon, 21, 131‒7.Google Scholar
Steininger, F. F., Berggren, W. A., Kent, D. V., et al. (1996). Circum-Mediterranean Neogene (Miocene and Pliocene) marine-continental chronological correlations of European mammal units. In The Evolution of Western Eurasian Neogene Mammal Faunas, ed. Bernor, R. L., Fahlbusch, V., and Mittmann, H.-W.. New York: Columbia University Press, pp. 746.Google Scholar
Stewart, J. R. (2008). The progressive effect of the individualistic response of species to Quaternary climate change: an analysis of British mammalian faunas. Quaternary Science Reviews, 27, 2499‒508.Google Scholar
Stewart, J. R. and Lister, A. M. (2001). Cryptic northern refugia and the origins of the modern biota. Trends in Ecology and Evolution, 16, 608‒13.Google Scholar
Stewart, J. R., Lister, A. M., Barnes, I., and Dalén, L. (2010). Refugia revisited: individualistic responses of species in space and time. Proceedings of the Royal Society B, 277, 661–71.Google Scholar
Stockwell, D. R. B. and Noble, I. R. (1992). Induction of sets of rules from animal distribution data: a robust and informative method of data analysis. Mathematics and Computers in Simulation, 33, 385‒90.Google Scholar
Storch, G. (1995). The Neogene mammalian faunas of Ertemte and Harr Obo in Inner Mongolia (Nei Mongol), China. Senckenbergiana Lethaea, 75, 221‒51.Google Scholar
Storch, G. (1998). Fossil history of shrews in Asia. In Evolution of Shrews, ed. Wójcik, J. M. and Wolsan, M.. Białowieża: Mammal Research Institute, pp. 93120.Google Scholar
Storch, G. and Qiu, Z. (1991). Insectivores (Mammalia: Erinaceidae, Soricidae, Talpidae) from the Lufeng hominoid locality, Late Miocene of China. Geobios, 24, 601‒21.Google Scholar
Strauss, D. and Sadler, P. M. (1989). Classical confidence intervals and Bayesian probability estimates for ends of local taxon ranges. Mathematical Geology, 21, 411‒27.Google Scholar
Strukova, T. V., Bachura, O. P., Borodin, A. V., and Stefanovskii, V. V. (2006). First report of a mammalian fauna from cave deposits from the Late Pleistocene and Holocene of the northern Urals (Cheremukhova-1). Stratigraphy, Geology, Correlation, 14, 98108. (In Russian).Google Scholar
Stuart, A. J. (1976). The history of the mammal fauna during the Ipswichian/Last Interglacial in England. Philosophical Transactions of the Royal Society of London B, 276, 221‒50.Google Scholar
Sulimski, A. (1962). Supplementary studies on the insectivores from Węże 1 (Poland). Acta Palaeontologica Polonica, 7, 441498.Google Scholar
Sun, Y. F., Jin, C. Z., and Xu, Q. Q. (1992). Dalian Haimao Fauna. Dalian: Dalian University of Technology Press.Google Scholar
Svendsen, J. I., Alexanderson, H., Astakhov, V. I., et al. (2004). Late Quaternary ice sheet history of northern Eurasia. Quaternary Science Reviews, 23, 1229‒71.Google Scholar
Svenning, J. C., Fløjgaard, C., Marske, K. A., Nógues-Bravo, D., and Normand, S. (2011). Applications of species distribution modeling to paleobiology. Quaternary Science Reviews, 30, 2930‒47.Google Scholar
Taberlet, P., Fumagalli, L., and Hausser, J. (1994). Chromosomal versus mitochondrial DNA evolution: tracking the evolutionary history of the southwestern European populations of the Sorex araneus group (Mammalia, Insectivora). Evolution, 48, 623‒36.Google Scholar
Tarasov, P. E., Volkova, V. S., Webb, T. I., et al. (2000). Last glacial maximum biomes reconstructed from pollen and plant macrofossil data from northern EurasiaJournal of Biogeography, 27, 609‒20.Google Scholar
Terzea, E. (1987). La faune du Pléistocène Supérieur de la Grotte ‘Pestera Cioarei’ de Borosteni (Départ. de Gorj). Traveux de l’Institut de Spéologie ‘Émile Racovitza’, 26, 5566.Google Scholar
Tesakov, A. S. (2004). Biostratigraphy of Middle Pliocene-Eopleistocene of eastern Europe based on small mammals. Transactions of the Geological Institute, 554, 1248. (In Russian).Google Scholar
Thomassen, H. (1996). De Midden Paleolitische kleine zoogdierfauna uit de Sesselfels-grot (Zuid-Duitsland), met nadruk op de spitsmuizen (Mammalia, Insectivora, Soricidae). Cranium, 13, 4752.Google Scholar
Ungar, P. S. (2010). Mammal Teeth: Origin, Evolution, and Diversity. Baltimore: Johns Hopkins University Press.Google Scholar
Van Andel, T. H. (2002). The climate and landscape of the middle part of the Weichselian glaciation in Europe: the stage 3 project. Quaternary Research, 57, 28.Google Scholar
Van Vliet-Lanoë, B. and Hallégouët, B. (2001). European permafrost at the LGM ‒ and at its maximal extentNATO Science, Series 2, Environmental Security, 76, 195214.Google Scholar
Vangengeim, E. A., Pevzner, M. A., and Tesakov, A. S. (2001). Zonal subdivisions of the Quaternary in eastern Europe based on small mammals. Stratigraphy and Geological Correlations, 9, 280‒92.Google Scholar
Velichko, A. A., Wright, H. E., and Barnosky, C. W. (1980). Late Quaternary Environments of the Soviet Union. Minneapolis: University of Minnesota Press.Google Scholar
Wagner, P. J. (1998). A likelihood approach for evaluating estimates of phylogenetic relationships among fossil taxa. Paleobiology, 24, 430‒49.Google Scholar
Warnock, R. C., Yang, Z., and Donoghue, P. C. (2012). Exploring uncertainty in the calibration of the molecular clock. Biology Letters, 8, 156–9.Google Scholar
White, T. A., Bordewich, M., and Searle, J. B. (2010). A network approach to study karyotypic evolution: the chromosomal races of the common shrew (Sorex araneus) and house mouse (Mus musculus) as model systems. Systematic Biology, 59, 262‒76.Google Scholar
Whittaker, R. H. (1975). Communities and Ecosystems. New York: Macmillan Publishing.Google Scholar
Willis, K. J. (1994). The vegetational history of the Balkans. Quaternary Science Reviews, 13, 769‒88.Google Scholar
Willis, K. J., Rudner, E., and Sümegi, P. (2000). The full-glacial forests of central and southeastern Europe. Quaternary Research, 53, 203‒13.Google Scholar
Willis, K. J. and van Andel, T. H. (2004). Trees or no trees? The environments of central and eastern Europe during the last glaciation. Quaternary Science Reviews, 23, 2369‒87.Google Scholar
Willis, K. J. and Whittaker, R. J. (2000). The refugial debate. Science, 287, 1406‒7.Google Scholar
Young, C. C. (1935). Note on a mammalian microfauna from Yenchingkou near Wanhsien, Szechuan. Bulletin of the Geological Society of China, 6, 247‒8.Google Scholar
Zagwijn, W. H. (1996). An analysis of Eemian climate in western and central Europe. Quaternary Science Reviews, 15, 451‒69.Google Scholar
Zaitsev, M. V. (1998). Late Anthropogene Insectivora from the South Urals with a special reference to diagnostics of the red-toothed shrews of the genus Sorex. Illinois State Museum Scientific Papers, 27, 145‒58.Google Scholar
Zaitsev, M. V. and Baryshnikov, G. F. (2002). Pleistocene Soricidae (Lipotyphla, Insectivora, Mammalia) from Teugolnaya Cave, Northern Caucasus, Russia. Acta Zoologica Cracoviensia, 45, 283305.Google Scholar
Zaitsev, M. V. and Rzebik-Kowalska, B. (2003). Variation and taxonomic value of some mandibular characters in red-tooth shrews of the genus Sorex L. (Insectivora: Soricidae). Russian Journal of Theriology, 2, 97104.Google Scholar
Zaitsev, M. V., Voyta, L. L., and Sheftel, B. I. (2014). The Mammals of Russia and Adjacent Territories: Lipotyphlans. St Petersburg: Nauka. (In Russian).Google Scholar
Zhang, S. S. (1993). Comprehensive study on the Jinniushan Paleolithic site. Memoirs of Institute of Vertebrate Palaeontology and Palaeoanthropology, Chinese Academy of Sciences, 19, 1164. (In Chinese).Google Scholar
Zheng, S. H. and Cai, B. Q. (1991). Micromammalian fossils from Danangou of Yuxian, Hebei. In Institute of Vertebrate Paleontology and Paleoanthropology, Contribution to the XIII INQUA Congress, ed. Sinica, Academia. Beijing: Beijing Science and Technology Publishing House, pp. 100‒31Google Scholar
Zheng, S. H. (2004). Jianshi Hominid Site. Beijing: Science Press. (In Chinese).Google Scholar
Zima, J., Lukáčová, L., and Macholán, M. (1998). Chromosomal evolution in shrews. In Evolution of Shrews, ed. Wójcik, J. M. and Wolsan, M.. Białowieża: Mammal Research Institute, pp. 175218.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×