Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-06-01T07:46:18.766Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  18 March 2019

Uffe N. Nielsen
Affiliation:
Western Sydney University
Get access
Type
Chapter
Information
Soil Fauna Assemblages
Global to Local Scales
, pp. 291 - 350
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

A'Bear, A. D., Crowther, T. W., Ashfield, R., et al. (2013) Localised invertebrate grazing moderates the effect of warming on competitive fungal interactions. Fungal Ecology, 6, 137140.Google Scholar
A'Bear, A. D., Jones, T. H., Boddy, L. (2014) Potential impacts of climate change on interactions among saprotrophic cord-forming fungal mycelia and grazing soil invertebrates. Fungal Ecology, 10, 3443.Google Scholar
Acosta-Mercado, D., Lynn, D. H. (2002) A preliminary assessment of spatial patterns of soil ciliate diversity in two subtropical forests in Puerto Rico and its implications for designing an appropriate sampling approach. Soil Biology & Biochemistry, 34, 15171520.CrossRefGoogle Scholar
Adams, B. J., Wall, D. H., Gozel, U., et al. (2007) The southernmost worm, Scottnema lindsayae (Nematoda): Diversity, dispersal and ecological stability. Polar Biology, 30, 809815.Google Scholar
Adams, B. J., Wall, D. H., Virginia, R. A., Broos, E., Knox, M. A. (2014) Ecological biogeography of the terrestrial nematodes of Victoria Land, Antarctica. ZooKeys, 419, 2971.CrossRefGoogle Scholar
Addison, J. A. (1977) Population dynamics and biology of Collembola at Truelove lowland. In: Truelove Lowland, Devon Island, Canada: A High Arctic Ecosystem (ed. Bliss, L. C.) pp. 363382. Edmonton, University of Alberta Press.Google Scholar
Adl, S. M., Simpson, A. G., Farmer, M. A., et al. (2005) The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. Journal of Eukaryotic Microbiology, 52, 399451.Google Scholar
Agosti, D., Johnson, N. F. (2010) Antbase. World Wide Web electronic publication. www.antbase.org.Google Scholar
Ainsworth, E. A., Long, D. J. (2005) What have we learned from 15 years of free-air CO2 enrichment (FACE)? A meta-analytic review of the responses of photosynthesis, canopy properties and plant production to rising CO2. New Phytologist, 165, 351372.Google Scholar
Albertengo, J., Bianchini, A., Sylvestre Begnis, A., et al. (2011) Sustainable certified agriculture: The farmer's production alternative. In: World Congress on Conservation Agriculture (WCCA), Australia. www.aciar.gov.au/wccaposters.Google Scholar
Ali, F., Wharton, D. A. (2014) Intracellular freezing in the infective juveniles of Steinernema feltiae: An entomopathogenic nematode. PloS ONE, 9, e94179. doi:10.1371/journal.pone.0094179.Google Scholar
Ali, J. G., Alborn, H. T., Stelinski, L. L. (2011) Constitutive and induced subterranean plant volatiles attract both entomopathogenic and plant parasitic nematodes. Journal of Ecology, 99, 2635.Google Scholar
Allegrucci, G., Carchini, G., Convey, P., Sbordino, V. (2012) Evolutionary geographic relationships among orthocladine chironomid midges from maritime Antarctic and sub-Antarctic islands. Biological Journal of the Linnean Society, 106, 258274.CrossRefGoogle Scholar
Alphey, T. J. W. (1985) Study of spatial distribution and population dynamics of two sympatric species of trichodorid nematodes. Annals of Applied Biology, 107, 497509.Google Scholar
Alroy, J. (2017) Effects of habitat disturbance on tropical forest biodiversity. Proceedings of the National Academy of Sciences of the United States of America, 114, 60616065.Google Scholar
Alvarez, T., Frampton, G. K., Goulson, D. (2001) Epigeic Collembola in winter wheat under organic, integrated and conventional farm management regimes. Agriculture, Ecosystems & Environment, 83, 95110.CrossRefGoogle Scholar
Andersen, A. N., Del Toro, I., Parr, C. L. (2015) Savanna ant species richness is maintained along a bioclimatic gradient of increasing latitude and decreasing rainfall in northern Australia. Journal of Biogeography, 42, 23132322.Google Scholar
Andersen, A. N., Sparling, G. P. (1997) Ants as indicators of restoration success: Relationship with soil microbial biomass in the Australian seasonal tropics. Restoration Ecology, 5, 109114.CrossRefGoogle Scholar
Andersen, C. P. (2003) Source-sink balance and carbon allocation below ground in plants exposed to ozone. New Phytologist, 157, 213228.CrossRefGoogle ScholarPubMed
Anderson, J. M. (1975) The enigma of soil animal species diversity. In: Progress in Soil Zoology (ed. Vanek, J.) pp. 5158. Prague, Academia.CrossRefGoogle Scholar
Anderson, J. M. (1978a) Inter-and intra-habitat relationships between woodland Cryptostigmata species diversity and the diversity of soil and litter microhabitats. Oecologia, 32, 341348.Google Scholar
Anderson, J. M. (1978b) A method to quantify soil-microhabitat complexity and its application to a study of soil animal species diversity. Soil Biology & Biochemistry, 10, 7778.CrossRefGoogle Scholar
Anderson, J. M. (1995) Soil organisms as engineers: Microsite modulation of macroscale processes. In: Linking Species to Ecosystems (eds. Jones, C. G., Lawton, J. H.) pp. 94106. New York, Chapman & Hall.CrossRefGoogle Scholar
Andrássy, I. (1964) Süsswasser-Nematoden aus den grossen Gebirgsgegenden Ostafricas. Acta Zoologica, 10, 159.Google Scholar
André, H. M., Noti, M.-I., Lebrun, P. (1994) The soil fauna: The other last biotic frontier. Biodiversity and Conservation, 3, 4556.Google Scholar
Andriuzzi, W. S., Wall, D. H. (2017) Responses of belowground communities to large aboveground herbivores: Meta-analysis reveals biome-dependent patterns and critical research gaps. Global Change Biology, 23, 38573868.Google Scholar
Andújar, C., Arribas, P., Ruzicka, F., et al. (2015) Phylogenetic community ecology of soil biodiversity using mitochondrial metagenomics. Molecular Ecology, 24, 36033617.Google Scholar
Aoki, Y., Hoshino, M., Matsubara, T. (2007) Silica and testate amoebae in a soil under pine-oak forest. Geoderma, 142, 2935.Google Scholar
Armendáriz, I., Hernández, M. A., Jordana, R. (1996) Temporal evolution of soil nematode communities in Pinus nigra forests of Navarra, Spain. Fundamental and Applied Nematology, 19, 561577.Google Scholar
Arrhenius, O. (1921) Species and area. Journal of Ecology, 9, 9599.Google Scholar
Aslam, T. J., Benton, T. G., Nielsen, U. N., Johnson, S. N. (2015) Impacts of eucalypt plantation management on soil faunal communities and nutrient bioavailability: Trading function for dependence? Biology and Fertility of Soils, 51, 637644.CrossRefGoogle Scholar
Atkin, L., Proctor, J. (1988) Invertebrates in the litter and soil on Volcan Barva, Costa Rica. Journal of Tropical Ecology, 4, 307310.Google Scholar
Aubert, M., Hedde, M., Decaëns, T., et al. (2003) Effects of tree canopy composition on earthworms and other macroinvertebrates in beech forests of Upper Normandy (France). Pedobiologia, 47, 904912.Google Scholar
Ayarbe, J. P., Kieft, T. L. (2000) Mammal mounds stimulate microbial activity in a semiarid shrubland. Ecology, 81, 11501154.Google Scholar
Ayres, E., Dromph, K. M., Cook, R., Ostle, N., Bardgett, R. D. (2007) The influence of below-ground herbivory and defoliation of a legume on nitrogen transfer to neighbouring plants. Functional Ecology, 21, 256263.Google Scholar
Ayres, E., Heath, J., Possell, M., et al. (2004) Tree physiological responses to above-ground herbivory directly modify below-ground processes of soil carbon and nitrogen cycling. Ecology Letters, 7, 469479.Google Scholar
Ayres, E., Wall, D. H., Simmons, B. L., et al. (2008) Belowground nematode herbivores are resistant to elevated atmospheric CO2 concentrations in grassland ecosystems. Soil Biology & Biochemistry, 40, 978985.Google Scholar
Ayuke, F. O., Pulleman, M. M., Vanlauwe, B., et al. (2011) Agricultural management affects earthworm and termite diversity across humid to semi-arid tropical zones. Agriculture, Ecosystems & Environment, 140, 148154.Google Scholar
Baas Becking, L. G. M. (1934) Geobiologie of inleiding tot de milieukunde, The Hague, Van Stockum and Zoon.Google Scholar
Baermann, G. (1917) Eine Einfache Methode zur Auffindung vor Ankylostomum (Nematoden). Larven in Erdproben, pp. 4147. Batavia, Genesk Lab Feestbundel.Google Scholar
Baggen, L. R., Gurr, G. M. (1998) The influence of food on Copidosoma koehleri (Hymenoptera: Encyrtidae), and the use of flowering plants as a habitat management tool to enhance biological control of potato moth, Phthorimaea operculella (Lepidoptera: Gelechiidae). Biological Control, 11, 917.Google Scholar
Bamforth, S. S, Wall, D. H., Virginia, R. A. (2005) Distribution and diversity of soil protozoa in the McMurdo Dry Valleys of Antarctica. Polar Biology, 28, 756762.CrossRefGoogle Scholar
Bardgett, R. D. (2002) Causes and consequences of biological diversity in soil. Zoology, 105, 367374.Google Scholar
Bardgett, R. D. (2005) The Biology of Soils: A Community and Ecosystem Approach, Oxford, Oxford University Press.Google Scholar
Bardgett, R. D., Bowman, W. D., Kaufman, R., Schmidt, S. K. (2005a) A temporal approach to linking aboveground and belowground ecology. Trends in Ecology and Evolution, 20, 634641.Google Scholar
Bardgett, R. D., Chan, K. F. (1999) Experimental evidence that soil fauna enhance nutrient mineralization and plant nutrient uptake in montane grassland ecosystems. Soil Biology & Biochemistry, 31, 10071014.Google Scholar
Bardgett, R. D., Cook, R. G., Yeates, W., Denton, C. S. (1999a) The influence of nematodes on below-ground processes in grassland ecosystems. Plant and Soil, 212, 2333.Google Scholar
Bardgett, R. D., Denton, C. S., Cook, R. (1999b) Belowground herbivory promotes soil nutrient transfer and root growth in grassland. Ecology Letters, 2, 357360.Google Scholar
Bardgett, R. D., van der Putten, W. H. (2014) Belowground biodiversity and ecosystem functioning. Nature, 515, 505511.Google Scholar
Bardgett, R. D., Wardle, D. A. (2003) Herbivore-mediated linkages between aboveground and belowground communities. Ecology, 84, 22582268.Google Scholar
Bardgett, R. D., Wardle, D. A. (2010) Aboveground–Belowground Linkages: Biotic Interactions, Ecosystem Processes, and Global Change, New York, Oxford University Press.Google Scholar
Bardgett, R. D., Yeates, G. W., Anderson, J. M. (2005b) Patterns and determinants of soil biological diversity. In: Biological Diversity and Function in Soils (eds. Bardgett, R. D., Usher, M. B., Hopkins, D. W.) pp. 100118. Cambridge, Cambridge University Press.CrossRefGoogle Scholar
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, 215238.Google Scholar
Barral, M. P., Benayas, J. M. R., Meli, P., Maceira, N. O. (2015) Quantifying the impacts of ecological restoration on biodiversity and ecosystem services in agroecosystems: A global meta-analysis. Agriculture, Ecosystems and Environment, 202, 223231.Google Scholar
Barrett, J. E., Virginia, R. A., Wall, D. H., et al. (2008) Persistent effects of a discrete climate event on a polar desert ecosystem. Global Change Biology, 14, 22492261.Google Scholar
Barrios, E. (2007) Soil biota, ecosystem services and land productivity. Ecological Economics, 64, 269285.Google Scholar
Barrios, E., Sileshi, G. W., Shephard, K., Sinclair, F. (2012) Agroforestry and soil health: Linking trees, soil biota, and ecosystem services. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 315330. Oxford, Oxford University Press.Google Scholar
Bass, D., Thomas, T. A., Matthai, L., Marsh, V., Cavalier-Smith, T. (2007) DNA evidence for global dispersal and probable endemicity of protozoa. BMC Evolutionary Biology, 7, 162.Google Scholar
Bassus, W. (1968) Über Einflüsse von Industrieexhalaten auf den Nematodenbesatz im Boden von Kiefernwäldern. Pedobiologia, 8, 289295.Google Scholar
Bastow, J. (2012) Succession, resource processing, and diversity in detrital food webs. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 117135. Oxford, Oxford University Press.Google Scholar
Bates, S. T., Clemente, J. C., Flores, G. E., et al. (2013) Global biogeography of highly diverse protistan communities in soil. The ISME Journal, 7, 652659.Google Scholar
Battigelli, J. P., Spence, J. R., Langor, D. W., Berch, S. M. (2004) Short-term impact of forest soil compaction and organic matter removal on soil mesofauna density and oribatid mite diversity. Canadian Journal of Forest Research, 34, 11361149.Google Scholar
Beare, M. H., Coleman, D. S., Crossley, D. A. Jr, Hendrix, P. F., Odum, E. P. (1995) A hierarchical approach to evaluating the significance of soil biodiversity to biogeochemical cycling. Plant and Soil, 170, 522.Google Scholar
Beare, M. H., Hu, S., Coleman, D. C., Hendrix, P. F. (1997) Influences of mycelia fungi on soil aggregation and organic matter storage in conventional and no-tillage soils. Applied Soil Ecology, 5, 211219.Google Scholar
Beare, M. H., Parmelee, R. W., Hendrix, P. F., et al. (1992) Microbial and faunal interactions and effects on litter nitrogen and decomposition in agroecosystems. Ecological Monographs, 62, 569591.Google Scholar
Beaulieu, F., Weeks, A. R. (2007) Free-living mesostigmatic mites in Australia: Their roles in biological control and bioindication. Australian Journal of Experimental Agriculture, 47, 460478.Google Scholar
Bedano, J. C., Domínguez, A., Arolfo, R., Wall, L. G. (2016) Effect of good agricultural practices under no-till on litter and soil invertebrates in areas with different soil types. Soil & Tillage Research, 158, 100109.Google Scholar
Beddard, F. E. (1912) Earthworms and Their Allies, Cambridge, Cambridge University Press.Google Scholar
Behan-Pelletier, V. M. (1999) Oribatid mite biodiversity in agroecosystems: Role for bioindication. Agriculture, Ecosystems & Environment, 74, 411423.Google Scholar
Bell, N. L., Adam, K. H., Jones, R. J., et al. (2016) Detection of invertebrate suppressive soils, and identification of a possible biological control agent for Meloidogyne nematodes using high resolution rhizosphere microbial community analysis. Frontiers in Plant Science, 7, 1946.Google Scholar
Belnap, J., Phillips, S. L. (2001) Soil biota in an ungrazed grassland: Response to annual grass (Bromus tectorum) invasion. Ecological Applications, 11, 12611275.Google Scholar
Bender, S. F., van der Heijden, M. G. A. (2015) Soil biota enhance agricultural sustainability by improving crop yield, nutrient uptake and reducing nitrogen leaching losses. Journal of Applied Ecology, 52, 228239.Google Scholar
Bender, S. F., Wagg, C., van der Heijden, M. A. (2016) An underground revolution: Biodiversity and soil ecological engineering for agricultural sustainability. Trends in Ecology and Evolution, 31, 440452.Google Scholar
Bengtsson, G., Rundgren, S. (1983) Respiration and growth of a fungus, Mortierella isabellina, in response to grazing by Onychiurus armatus (Collembola). Soil Biology & Biochemistry, 15, 469473.Google Scholar
Bengtsson, J. (2002) Disturbance and resilience in soil animal communities. European Journal of Soil Biology, 387, 119125.Google Scholar
Bengtsson, J., Ahnström, J., Weibull, A. C. (2005) The effects of organic agriculture on biodiversity and abundance: A meta-analysis. Journal of Applied Ecology, 42, 261269.Google Scholar
Benoit, J. B., Elnitsky, M. A., Schulte, G. G., Lee, R. E. Jr., Denlinger, D. L. (2009) Antarctic collembolans use chemical signals to promote aggregation and egg laying. Journal of Insect Behaviour, 22, 121133.Google Scholar
Berg, M. P. (2012) Patterns of biodiversity at fine and small spatial scales. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 136152. Oxford, Oxford University Press.Google Scholar
Berg, M. P., Bengtsson, J. (2007) Spatial and temporal variation in food web composition. Oikos, 116, 17891804.Google Scholar
Berg, M. P., Kneise, J. P., Bedaux, J. J. M., Verhoef, H. A. (1998) Dynamics and stratification of functional groups of micro- and mesoarthropods in the organic layer of a Scots pine forest. Biology and Fertility of Soils, 26, 268284.Google Scholar
Berg, M. P., Stoffer, M., van den Heuvel, H. H. (2004) Feeding guilds in Collembola based on digestive enzymes. Pedobiologia, 48, 589601.Google Scholar
Bergkvist, G., Stenberg, M., Wetterlind, J., Bãth, B., Elfstrand, S. (2010) Clover cover crops under-sown in winter wheat increase yield of subsequent spring barley – Effect of N dose and companion grass. Field Crop Research, 120, 292298.Google Scholar
Beyens, L., Ledeganck, P., Graae, B. J., Nijs, I. (2009) Are soil biota buffered against climatic extremes? An experimental test on testate amoebae in arctic tundra (Qeqertarsuaq, West Greenland). Polar Biology, 32, 453462.Google Scholar
Bezemer, T. M., De Deyn, G. B., Bossinga, T. M., et al. (2005) Soil community composition drives aboveground plant–herbivore–parasitoid interactions. Ecology Letters, 8, 652661.CrossRefGoogle Scholar
Bezemer, T. M., Fountain, M. T., Barea, J. M., et al. (2010) Divergent composition but similar function of soil food webs of individual plants: Plant species and community effects. Ecology, 91, 30273036.Google Scholar
Bezemer, T. M., Wagenaar, R., van Dam, N. M., Wäckers, F. L. (2003) Interactions between above- and belowground insect herbivores as mediated by the plant defense system. Oikos, 101, 555562.Google Scholar
Bignell, D. E. (2000) Introduction to symbiosis. In: Termites: Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 189208. Dordrecht, Kluwer Academic.Google Scholar
Bignell, D. E. (2006) Termites as soil engineers and soil processors. In: Soil Biology (eds. König, H., Varma, A.) pp. 183220. Berlin, Springer-Verlag.Google Scholar
Bignell, D. E., Eggleton, P. (2000) Termites in ecosystems. In: Termites, Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 363387. Dordrecht, Kluwer Academic Press.Google Scholar
Bihn, J. H., Verhaagh, M., Brändle, M., Brandl, R. (2008) Do secondary forests act as refuges for old growth forest animals? Recovery of ant diversity in the Atlantic forest of Brazil. Biological Conservation, 141, 733743.CrossRefGoogle Scholar
Bik, H. M., Porazinska, D. L., Creer, S., et al. (2012) Sequencing our way towards understanding global eukaryotic biodiversity. Trends in Ecology and Evolution, 27, 233243.Google Scholar
Binet, F., Trehen, P. (1992) Experimental microcosm study of the role of Lumbricus terrestris (Oligochaeta: Lumbricidae) on nitrogen dynamics in cultivated soils. Soil Biology & Biochemistry, 24, 15011506.Google Scholar
Birkhofer, K., Bezemer, T. M., Bloem, J., et al. (2008) Long-term organic farming fosters below and aboveground biota: Implications for soil quality, biological control and productivity. Soil Biology & Biochemistry, 40, 22972308.Google Scholar
Birkhofer, K., Diekötter, T., Boch, S., et al. (2011) Soil fauna feeding activity in temperate grassland soils increases with legume and grass species richness. Soil Biology & Biochemistry, 43, 22002207.Google Scholar
Bishop, T. R., Robertson, M. P., Van Rensburg, B. J., Parr, C. L. (2014) Exploring variation in ant diversity through space and time: An elevational study in the Maloti-Drakensberg Mountains of South Africa. Journal of Biogeography, 41, 22562268.Google Scholar
Blankinship, J. C., Niklaus, P. A., Hungate, B. A. (2011) A meta-analysis of responses of soil biota to global change. Oecologia, 165, 553565.Google Scholar
Block, W. (1983) Low temperature tolerance of soil arthropods – Some recent advances. In: New Trends in Soil Biology (eds. Lebrun, P., André, H. M., De Medts, A., Wauthy, G.) pp. 427431. Ottignies, Diey-Brichart.Google Scholar
Block, W., Webb, N. R., Coulson, S. J., Hodkinson, I. D. (1994) Thermal adaptation in a high arctic collembolan Onychiurus arcticus. Journal of Insect Physiology, 40, 715722.Google Scholar
Bloemers, G. F., Hodda, M., Lambshead, P. J. D., Lawton, J. H., Wanless, F. R. (1997) The effects of forest disturbance on diversity of tropical soil nematodes. Oecologia, 111, 575582.CrossRefGoogle ScholarPubMed
Blomqvist, M. M., Olff, H., Blaauw, M. B., Bongers, T., van der Putten, W. (2000) Interactions between above- and belowground biota: Importance for small-scale vegetation mosaics in a grassland ecosystem. Oikos, 90, 582598.Google Scholar
Blouin, M., Hodson, E., Delgado, E. A., et al. (2013) A review of earthworm impact on soil function and ecosystem services. European Journal of Soil Science, 64, 161182.Google Scholar
Boag, B., Yeates, G. W. (1998) Soil nematode biodiversity in terrestrial ecosystems. Biodiversity and Conservation, 7, 617630.Google Scholar
Boag, B., Yeates, G. W. (2001) The potential impact of the New Zealand flatworm, a predator of earthworms, in western Europe. Ecological Applications, 11, 12761286.CrossRefGoogle Scholar
Bobbink, R., Hicks, K., Galloway, J. N., et al. (2010) Global assessment of nitrogen deposition effects on terrestrial plant diversity: A synthesis. Ecological Applications, 20, 3059.CrossRefGoogle ScholarPubMed
Bohlen, P. J., Scheu, S., Hale, C. M., et al. (2004) Non-native invasive earthworms as agents of change in northern temperate forests. Frontiers in Ecology and the Environment, 2, 427435.Google Scholar
Bokhorst, S., Berg, M. P., Wardle, D. A. (2017a) Micro-arthropod community responses to ecosystem retrogression in boreal forest. Soil Biology & Biochemistry, 110, 7986.Google Scholar
Bokhorst, S., Huiskes, A., Convey, P., et al. (2011) Microclimate impacts of passive warming methods in Antarctica: Implication for climate change studies. Polar Biology, 34, 14211435.Google Scholar
Bokhorst, S., Huiskes, A., Convey, P., Van Bodegom, P. M., Aerts, R. (2008) Climate change effects on soil arthropod communities from the Falkland Islands and the Maritime Antarctic. Soil Biology & Biochemistry, 40, 15471556.Google Scholar
Bokhorst, S., Kardol, P., Bellingham, P. J., et al. (2017b) Response of communities of soil organisms and plants to soil aging at two contrasting long-term chronosequences. Soil Biology & Biochemistry, 106, 6979.Google Scholar
Bokhorst, S., Phoenix, G. K., Bjerke, J. W., et al. (2012) Extreme winter warming events more negatively impact small rather than large soil fauna: Shift in community composition explained by traits not taxa. Global Change Biology, 18, 11521162.Google Scholar
Bommarco, R., Kleijn, D., Potts, S. J. (2016) Ecological intensification: Harnessing ecosystem services for food security. Trends in Ecology and Evolution, 28, 230238.Google Scholar
Bongers, T. (1990) The maturity index: An ecological measure of environmental disturbance based on nematode species composition. Oecologia, 83, 1419.Google Scholar
Bongers, T. (1999) The maturity index, the evolution of nematode life history traits, adaptive radiation and cp-scaling. Plant and Soil, 212, 1322.Google Scholar
Bonkowski, M. (2004) Soil protozoa and plant growth: The microbial loop in soil revisited. New Phytologist, 162, 617631.Google Scholar
Bonkowski, M., Brandt, F. (2002) Do soil protozoa enhance plant growth by hormonal effects? Soil Biology & Biochemistry, 34, 17091715.Google Scholar
Bonkowski, M., Geoghegan, I. E., Birch, A. N. E., Griffiths, B. S. (2001) Effects of soil decomposer invertebrates (protozoa and earthworms) on an above-ground phytophagous insect (cereal aphid), mediated through changes in the host plant. Oikos, 95, 441450.Google Scholar
Bonkowski, M., Griffiths, B. S., Scrimgeour, C. (2000) Substrate heterogeneity and microfauna in soil organic ‘hotspots’ as determinants of nitrogen capture and growth of rye-grass. Applied Soil Ecology, 14, 3753.Google Scholar
Bonkowski, M., Roy, J. (2012) Decomposer community complexity affects plant competition in a model early successional grassland community. Soil Biology & Biochemistry, 46, 4148.Google Scholar
Bonkowski, M., Scheu, S. (2004) Biotic interactions in the rhizosphere: Effects on plant growth and herbivore development. In: Insects and Ecosystem Functioning (eds. Weisser, W., Sieman, E.) pp. 7191. Heidelberg, Germany, Ecological Studies: Springer Verlag.Google Scholar
Bonkowski, M., Villenave, C., Griffiths, B. (2009) Rhizosphere fauna: The functional and structural diversity of intimate interactions of soil fauna with plant roots. Plant and Soil, 321, 213233.Google Scholar
Borcard, D., Legendre, P. (2002) All-scale spatial analysis of ecological data by means of principal coordinates of neighbour matrices. Ecological Modelling, 153, 5168.Google Scholar
Borcard, D., Legendre, P., Drapeau, P. (1992) Partialling out the spatial component of ecological variation. Ecology, 73, 10451055.Google Scholar
Borcard, D., Matthey, M. (1995) Effect of a controlled trampling of sphagnum mosses on their oribatid mite assemblages (Acari, Oribatei). Pedobiologia, 39, 219230.Google Scholar
Borgonie, G., Garcia-Moyano, A., Litthauer, D., et al. (2011) Nematoda from the terrestrial deep subsurface of South Africa. Nature, 474, 7982.Google Scholar
Bornebusch, C. H. (1930) The fauna of forest soil. Forst. ForsVaes. Danm., 11, 1224.Google Scholar
Bouché, M. B. (1983) The establishment of earthworm communities. In: Earthworm Ecology: From Darwin to Vermiculture (ed. Satchell, J. E.) pp. 431448. London, Chapman and Hall.Google Scholar
Bouwman, L. A., Zwart, K. B. (1994) The ecology of bacterivorous protozoans and nematodes in arable soil. Agriculture, Ecosystems & Environment, 51, 145160.Google Scholar
Bowman, W. D., Cleveland, C. C., Halada, Ĺ., Hresko, J., Baron, J. S. (2008) Negative impact of nitrogen deposition on soil buffering capacity. Nature Geoscience, 1, 767770.Google Scholar
Boyd, J. N. (1958) The ecology of earthworms in cattle-grazed machair in Tiree, Argyll. Journal of Animal Ecology, 27, 147157.Google Scholar
Boyer, S., Blakemore, R. J., Wratten, S. D. (2011) An integrative taxonomic approach to the identification of three new New Zealand endemic earthworm species (Acanthodrilidae, Octochaetidae: Oligochaeta). Zootaxa, 2994, 2132.Google Scholar
Boyer, S., Kim, Y.-N., Bowie, M. H., Lefort, M.-C., Dickinson, N. M. (2016) Response of endemic and exotic earthworm communities to ecological restoration. Restoration Ecology, 24, 717721.Google Scholar
Boyer, S., Wratten, S. D. (2010) The potential of earthworms to restore ecosystem services after opencast mining – A review. Basic and Applied Ecology, 11, 196203.Google Scholar
Brady, S. G. (2003) Evolution of the army ant syndrome: The origin and long-term evolutionary stasis of a complex of behavioural and reproductive adaptations. Proceedings of the National Academy of Sciences of the United States of America, 100, 65756579.Google Scholar
Brady, S. G., Schultz, T. R., Fisher, B. L., Ward, P. S. (2006) Evaluating alternative hypotheses for the early evolution and diversification of ants. Proceedings of the National Academy of Sciences of the United States of America, 103, 1817218177.Google Scholar
Brandl, R., Topp, W. (1985) Size structure of Pterostichus spp. (Carabidae): Aspects of competition. Oikos, 44, 234238.Google Scholar
Brandsæter, L. O., Netland, J. (1999) Winter annual legumes for use as cover crops in row crops in northern regions: I. Field experiments. Crop Science, 39, 13691379.Google Scholar
Brennan, A., Fortune, T., Bolger, T. (2006) Collembola abundances and assemblage structures in conventionally tilled and conservation tillage arable systems. Pedobiologia, 50, 135145.Google Scholar
Bretherton, S., Tordoff, G. M., Jones, T. H., Boddy, L. (2006) Compensatory growth of Phanerochaete velutina mycelial systems grazed by Folsomia candida (Collembola). FEMS Microbiology Ecology, 58, 3340.Google Scholar
Breznak, J. A., Brune, A. (1994) Role of microorganisms in the digestion of lignocellulose by termites. Annual Review of Entomology, 39, 453487.Google Scholar
Bridges, E. M. (1997) World Soils, Cambridge, Cambridge University Press.Google Scholar
Bridges, E. M., Oldeman, L. R. (1999) Global assessment of human-induced soil degradation. Arid Soil Research and Rehabilitation, 13, 319325.Google Scholar
Briones, M. I., Garnett, M. H., Piearce, T. G. (2005) Earthworm ecological groupings based on 14C analysis. Soil Biology & Biochemistry, 37, 21452149.Google Scholar
Briones, M. J. I. (2009) Uncertainties related to the temperature sensitivity of soil carbon decomposition. In: Uncertainties in Environmental Modelling and Consequences for Policy Making (eds. Baveye, P., Mysiak, J., Laba, M.) pp. 317335. New York, Springer.Google Scholar
Briones, M. J. I., Ineson, P. (2002) Use of 14C carbon dating to determine feeding behaviour of enchytraeids. Soil Biology & Biochemistry, 34, 881881.Google Scholar
Briones, M. J. I., Ineson, P., Heinemeyer, A. (2007a) Predicting potential impacts of climate change on the geographical distribution of enchytraeids: A meta-analysis approach. Global Change Biology, 13, 22522269.Google Scholar
Briones, M. J. I., Ineson, P., Poskitt, J. (1998) Climate change and Cognettia sphagnetorum: Effects on carbon dynamics in organic soil. Functional Ecology, 12, 528535.Google Scholar
Briones, M. J. I., Ostle, N., Garnett, M. H. (2007b) Invertebrates increase the sensitivity of non-labile soil carbon to climate change. Soil Biology & Biochemistry, 39, 81698818.Google Scholar
Briones, M. J. I., Poskitt, J., Ostle, N. (2004) Influence of warming and enchytraeid activities on soil CO2 and CH4 fluxes. Soil Biology & Biochemistry, 36, 18511859.Google Scholar
Brown, J. (1984) On the relationship between abundance and distribution of species. American Naturalist, 124, 255279.Google Scholar
Brown, V. K., Gange, A. C. (1990) Insect herbivory below ground. Advances in Ecological Research, 20, 158.Google Scholar
Brown, V. K., Gange, A. C. (1992) Secondary plant succession – how is it modified by insect herbivory? Vegetatio, 101, 313.Google Scholar
Brown, W. L. (1973) A comparison of the Hylean and Congo-West African rain forest ant faunas. In: Tropical Forest Ecosystems in Africa and South America: A Comparative Review (eds. Meggers, B. J., Ayensu, E. S., Duckworth, W. D.) pp. 161185. Washington, DC, Smithsonian Institute Press.Google Scholar
Brückner, A., Hilpert, A., Heethoff, M. (2017) Biomarker function and nutritional stoichiometry of neutral lipid fatty acids and amino acids in oribatid mites. Soil Biology & Biochemistry, 115, 3543.Google Scholar
Brühl, C. A., Mohamed, M., Linsenmair, K. E. (1999) Altitudinal distribution of leaf litter ants along a transect in primary forests on Mount Kinabalu, Malaysia. Journal of Tropical Ecology, 15, 265277.Google Scholar
Brussaard, L., Aanen, D. K., Briones, M. J. I., et al. (2012) Biogeography and phylogenetic community structure of soil invertebrate ecosystem engineers: Global to local patterns, implications for ecosystem functioning and services and global environmental change impacts. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 201232. Oxford, Oxford University Press.Google Scholar
Brussaard, L., Behan-Pelletier, V. M., Bignell, D. E., et al. (1997) Biodiversity and ecosystem functioning in soil. Ambio, 26, 563570.Google Scholar
Buchowski, R. W., Bradford, M. A., Grandy, A. S., Schmitz, O. J., Wieder, W. R. (2017) Applying population and community ecology theory to advance understanding of belowground biogeochemistry. Ecology Letters, 20, 231245.Google Scholar
Butt, K. R. (2008) Earthworms in soil restoration: Lessons learned from United Kingdom case studies of land reclamation. Restoration Ecology, 16, 637641.Google Scholar
Buyer, J. S., Teasdale, J. R., Roberts, D. P., Zasada, I. A., Maul, J. E. (2010) Factors affecting soil microbial community structure in tomato cropping systems. Soil Biology & Biochemistry, 42, 831841.Google Scholar
Callaghan, T. V., Tweedie, C. E., Akerman, J., et al. (2011) Multi-decadal changes in tundra environments and ecosystems: Synthesis of the International Polar Year-Back to the Future Project (IPY-BTF). Ambio, 40, 705716.Google Scholar
Callaham, M. A., Rhoades, C. C., Heneghan, L. (2008) A striking profile: Soil ecological knowledge in restoration management and science. Restoration Ecology, 16, 604607.Google Scholar
Cameron, E. K., Bayne, E. M., Clapperton, M. J. (2007) Human-facilitated invasion of exotic earthworms into northern boreal forests. Ecoscience, 14, 482490.Google Scholar
Cameron, E. K., Vilà, M., Cabeza, M. (2016) Global meta-analysis of the impacts of terrestrial invertebrate invaders on species, communities and ecosystems. Global Ecology and Biogeography, 25, 596606.Google Scholar
Carbajo, V., den Braber, B., van der Putten, W. H., De Deyn, G. B. (2011) Enhancement of late successional plants on ex-arable land by soil inoculations. PloS ONE, 6, e21943. doi:10.1371/journal.pone.0021943.Google Scholar
Cardosa, E. J. B. N., Vasconcellas, R. L. F., Bini, D., et al. (2013) Soil health: Looking for suitable indicators. What should be considered to assess the effects of use and management on soil health? Acientia Agricola, 70, 274289.Google Scholar
Caro, G., Decaëns, T., Lecarpentier, C., Mathieu, J. (2013) Are dispersal behaviours of earthworms related to their functional group? Soil Biology & Biochemistry, 58, 181187.Google Scholar
Carpenter, D., Hodson, M. E., Eggleton, P., Kirk, C. (2007) Earthworm induced mineral weathering: Preliminary results. European Journal of Soil Biology, 43, S176S183.Google Scholar
Carrillo, C., Ball, B. A., Bradford, M. A., Jordan, C. F., Molina, M. (2011) Soil fauna alter the effects of litter composition on nitrogen cycling in a mineral soil. Soil Biology & Biochemistry, 43, 14401449.Google Scholar
Caruso, T., Hogg, I. D., Carapelli, A., Frati, F., Bargagli, R. (2009) Large-scale spatial patterns in the distribution of Collembola (Hexapoda) species in Antarctic terrestrial ecosystems. Journal of Biogeography, 36, 879886.Google Scholar
Caruso, T., Trokhymets, V., Bargagli, R., Convey, P. (2013) Biotic interactions as a structuring force in soil communities: Evidence from the micro-arthropods of an Antarctic moss model system. Oecologia, 172, 495503.Google Scholar
Caruso, T., Taormina, M., Migliorini, M. (2012) Relative role of deterministic and stochastic determinants of soil animal community: A spatially explicit analysis of oribatid mites. Journal of Animal Ecology, 81, 214221.Google Scholar
Chahartaghi, M., Langel, R., Scheu, S., Ruess, L. (2005) Feeding guilds in Collembola based on nitrogen stable isotope ratios. Soil Biology & Biochemistry, 37, 17181725.Google Scholar
Chaladze, G. (2012) Climate-based model of spatial pattern of the species richness of ants in Georgia. Journal of Insect Conservation, 16, 791800.Google Scholar
Chamberlain, P., Mcnamara, N., Chaplow, J., Stott, A., Black, H. (2006) Translocation of surface litter carbon into soil by Collembola. Soil Biology & Biochemistry, 38, 26552664.Google Scholar
Chao, A., Li, P. C., Agatha, S., Foissner, W. (2006) A statistical approach to estimate soil ciliate diversity and distribution based on data from five continents. Oikos, 114, 479493.Google Scholar
Chapin, F. S. (1980) The mineral nutrition of wild plants. Annual Review of Ecology and Systematics, 11, 233260.Google Scholar
Chauvat, M., Trap, J., Perez, G., Delporte, P., Aubert, M. (2011) Assemblages of Collembola across a 130-year chronosequence of beech forest. Soil Organisms, 83, 405418.Google Scholar
Chauvat, M., Wolters, V., Dauber, J. (2007) Response of collembolan communities to land-use change and grassland succession. Ecography, 30, 183192.Google Scholar
Chauvel, A., Grimaldi, M., Barros, E., et al. (1999) Pasture damage by an Amazonian earthworm. Nature, 398, 3233.Google Scholar
Chazdon, R. L. (2008) Beyond deforestation: Restoring forests and ecosystem services on degraded lands. Science, 320, 14581460.Google Scholar
Chelinho, S., Sautter, K. D., Cachada, A., et al. (2011) Carbofuran effects in soil nematode communities: Using trait and taxonomic based approaches. Ecotoxicology and Environmental Safety, 74, 20022012.Google Scholar
Chen, D., Cheng, J., Chu, P., et al. (2016) Effect of diversity on biomass across grasslands on the Mongolian Plateau: Contrasting effects between plants and soil nematodes. Journal of Biogeography, 43, 955966.Google Scholar
Chen, H. L., Li, B., Fang, C. M., Chen, J. K., Wu, J. H. (2007) Exotic plant influences soil nematode communities through litter input. Soil Biology & Biochemistry, 39, 17821793.Google Scholar
Chen, X., Adams, B., Bergeron, C., Sabo, A., Hooper-Bùi, L. (2015) Ant community structure and response to disturbances on coastal dunes of Gulf of Mexico. Journal of Insect Conservation, 19, 113.Google Scholar
Chen, Z., Wang, X. K., Feng, Z. Z., Xiao, Q., Duan, X. (2009) Impact of elevated O3 on soil microbial community function under wheat crop. Water, Air, & Soil Pollution, 198, 189198.Google Scholar
Chernova, N. M., Potapov, M. B., Savenkova, Y. Y., Bokova, A. I. (2009) Ecological significance of parthenogenesis in collembola. Zoologichesky Zhurnal, 88, 14551470.Google Scholar
Chown, S. L., Huiskes, A. H. L., Gremmen, N. J. M., et al. (2012) Continent-wide risk assessment for the establishment of nonindigenous species in Antarctica. Proceedings of the National Academy of Sciences of the United States of America, 109, 49384943.Google Scholar
Christensen, B. (1956) Studies on Enchytraeidae 6. Technique for culturing Enchytraeidae, with notes on cocoon types. Oikos, 7, 303307.Google Scholar
Christensen, B., Dózka-Farkas, K. (2006) Invasion of terrestrial enchytraeids into two postglacial tundras: North-eastern Greenland and the Arctic archipelago of Canada (Enchytraeidae, Oligochaeta). Polar Biology, 29, 454466.Google Scholar
Christensen, B., Glenner, H. (2010) Molecular phylogeny of Enchytraeidae (Oligochaeta) indicates separate invasions of the terrestrial environment. Journal of Zoological Systematics and Evolutionary Research, 48, 208212.Google Scholar
Christensen, S., Griffiths, B. S., Ekelund, F., Rønn, R. (1992) Huge increase in bacterivores on freshly killed barley roots. FEMS Microbiology Ecology, 86, 303310.Google Scholar
Chust, G., Pretus, J. L., Ducrot, D., Bedòs, A., Deharveng, L. (2003) Response of soil fauna to landscape heterogeneity: Determining optimal scales for biodiversity. Conservation Biology, 17, 17121723.Google Scholar
Clarholm, M. (1981) Protozoan grazing of bacteria in soil – Impact and importance. Microbial Ecology, 7, 343350.Google Scholar
Clarholm, M. (1985) Interactions of bacteria, protozoa and plants leading to mineralization of soil nitrogen. Soil Biology & Biochemistry, 17, 181187.Google Scholar
Clarholm, M. (1989) Effects of plant–bacterial–amoebal interactions on plant uptake of nitrogen under field conditions. Biology and Fertility of Soils, 8, 373378.Google Scholar
Clark, C. M., Cleland, E. E., Collins, S. L., et al. (2007) Environmental and plant community determinants of species loss following nitrogen enrichment. Ecology Letters, 10, 596607.Google Scholar
Cobb, N. A. (1915) Nematodes and Their Relationships, Washington, DC, U.S. G.P.O.Google Scholar
Cock, M. J. W., Biesmeijer, J. C., Cannon, R. J. C., et al. (2012) The positive contribution of invertebrates to sustainable agriculture and food security. CAB Reviews, 7, 127.Google Scholar
Cohn, E., Koltai, H., Sharon, E., Spiegel, Y. (2002) Root-nematode interactions: Recognition and pathogenicity. In: Plant Roots – The Hidden Half (eds. Waisel, J., Eshel, A., Kalkaf, U.) pp. 783796. New York, Marcel Dekker.Google Scholar
Čoja, T., Bruckner, A. (2003) Soil microhabitat diversity of a temperate Norway spruce (Picea abies) forest does not influence the community composition of gamasid mites (Gamasida, Acari). European Journal of Soil Biology, 39, 7984.Google Scholar
Cole, A. C. J. (1940) A guide to the ants of the Great Smoky Mountains National Park, Tennessee. The American Midland Naturalist, 24, 188.Google Scholar
Cole, L., Bardgett, R. D., Ineson, P. (2000) Enchytraeid worms (Oligochaeta) enhances mineralization of carbon in organic upland soils. European Journal of Soil Science, 51, 185192.Google Scholar
Cole, R. J., Holl, K. D., Zahawi, R. A., Wickey, P., Townsend, A. R. (2016) Leaf litter arthropod responses to tropical forest restoration. Ecology and Evolution, 6, 51585168.Google Scholar
Coleman, D. C. (1994) The microbial loop concept as used in terrestrial soil ecology studies. Microbial Ecology, 28, 245250.Google Scholar
Coleman, D. C. (2008) From peds to paradoxes: Linkages between soil biota and their influences on ecological processes. Soil Biology & Biochemistry, 40, 271289.Google Scholar
Coleman, D. C., Callaham, M. A. Jr, Crossley, D. A. Jr (2018) Fundamentals of Soil Ecology, 3rd edn, London, Academic Press.Google Scholar
Coleman, D. C., Crossley, D. A. Jr, Hendrix, P. F. (2004) Fundamentals of Soil Ecology, 2nd edn, San Diego, CA, Academic Press.Google Scholar
Coleman, D. C., Ingham, R. E., Mcclellan, J. F., Trofymow, J. A. (1984) Soil nutrient transformations in the rhizosphere via animal–microbial interactions. In: Invertebrates–Microbial Interactions (eds. Anderson, J. M., Rayner, A. D. M., Walton, D. W. H.) pp. 3558. Cambridge, Cambridge University Press.Google Scholar
Coleman, D. C., Macfadyen, A. (1966) The recolonization of gamma-irradiated soil by small arthropods. Oikos, 17, 6270.Google Scholar
Coleman, D. C., McGinnis, J. T. (1970) Quantification of fungus – Small arthropod food chains in the soil. Oikos, 21, 134137.Google Scholar
Coleman, D. C., Whitman, W. B. (2005) Linking species richness, biodiversity and ecosystem function in soil systems. Pedobiologia, 49, 479497.Google Scholar
Collins, N. M. (1980) The distribution of soil macrofauna on the West Ridge of Gunung (Mount) Mulu, Sarawak. Oecologia, 44, 263275.Google Scholar
Collison, E. J., Ruitta, T., Slade, E. M. (2013) Macrofauna assemblage composition and soil moisture interact to affect soil ecosystem functions. Acta Oecologica, 47, 3036.Google Scholar
Colwell, R. K., Coddington, J. A. (1994) Estimating terrestrial biodiversity through extrapolation. Philosophical Transactions of the Royal Society B: Biological Sciences, 345, 101118.Google Scholar
Comiso, J. C. (2006) Arctic warming signals from satellite observations. Weather, 61, 7076.Google Scholar
Convey, P., Chown, S. L., Clarke, A., et al. (2014) The spatial structure of Antarctic biodiversity. Ecological Monographs, 84, 203244.Google Scholar
Convey, P. A., McInnis, S. J. (2005) Exceptional tardigrade-dominated ecosystems in Ellsworth Land, Antarctica. Ecology, 86, 519527.Google Scholar
Costa, S. R., Kerry, B. R., Bardgett, R. D., Davies, K. G. (2012) Interactions between nematodes and their microbial enemies in coastal sand dunes. Oecologia, 170, 10531066.Google Scholar
Cotrufo, M. F., Soong, J., Vandegehuchte, M. L., et al. (2014) Naphthalene addition to soil surfaces: A feasible method to reduce soil micro-arthropods with negligible direct effects on soil C dynamics. Applied Soil Ecology, 74, 2129.Google Scholar
Coulson, J. C., Whittaker, J. B. (1978) The ecology of moorland animals. In: Production Ecology of British Moors and Montane Grasslands (eds. Heal, O. W., Perkins, D. F.) pp. 5293. Berlin, Springer.Google Scholar
Coulson, S. J., Hodkinson, I. D., Webb, N. R. (2003) Microscale distribution patterns in high arctic soil arthropod communities: The influence of plant species within the vegetation mosaic. Ecography, 26, 801809.Google Scholar
Coulson, S. J., Hodkinson, I. D., Webb, N. R. et al. (1996) Effects of experimental temperature elevation on high arctic soil microarthropod populations. Polar Biology, 16, 147153.Google Scholar
Coulson, S. J., Leinaas, H. P., Ims, R. A., Sovik, G. (2000) Experimental manipulation of the winter surface ice layer: The effects on a high Arctic soil microarthropod community. Ecography, 23, 299306.Google Scholar
Cox, B., Moore, P. D., Ladle, R. (2016) Biogeography: An Ecological and Evolutionary Approach, Chichester, UK, Wiley-Blackwell.Google Scholar
Cox, G. W., Mills, J. N., Ellis, B. A. (1992) Fire ants (Hymenoptera: Formicidae) as major agents of landscape development. Environmental Entomology, 21, 281286.Google Scholar
Cragg, J. B. (1961) Some aspects of the ecology of moorland animals. Journal of Ecology, 49, 477506.Google Scholar
Creamer, R. E., Hannula, S. E., Van Leeuwen, J. P., et al. (2016) Ecological network analysis reveals the inter-connection between soil biodiversity and ecosystem function as affected by land use across Europe. Applied Soil Ecology, 97, 112124.Google Scholar
Creevy, A. L., Fisher, J., Puppe, D., Wilkinson, D. A. (2016) Protist diversity on a nature reserve in NW England – With particular reference to their role in soil biogenic silicon pools. Pedobiologia, 59, 5159.Google Scholar
Crist, T. O. (1998) The spatial distribution of termites in shortgrass steppe – A geostatistical approach. Oecologia, 114, 410416.Google Scholar
Crotty, F. V., Adl, S. M., Blackshaw, R. P., Murray, P. J. (2012a) Protozoan pulses unveil their pivotal position within the soil food web. Microbial Ecology, 63, 905918.Google Scholar
Crotty, F. V., Adl, S. M., Blackshaw, R. P., Murray, P. J. (2012b) Using stable isotopes to differentiate trophic feeding channels within soil food webs. Journal of Eukaryotic Microbiology, 59, 520526.Google Scholar
Crotty, F. V., Fychan, R., Scullion, J., Sanderson, R., Marley, C. L. (2015) Assessing the impact of agricultural forage crops on soil biodiversity and abundance. Soil Biology & Biochemistry, 91, 119126.CrossRefGoogle Scholar
Crowther, T. W., A'Bear, A. D. (2012) Impacts of grazing soil fauna on decomposer fungi are species-specific and density-dependent. Fungal Ecology, 5, 277281.Google Scholar
Crowther, T. W., Boddy, L., Jones, T. H. (2011a) Outcomes of fungal interactions are determined by soil invertebrate grazers. Ecology Letters, 14, 11341142.Google Scholar
Crowther, T. W., Hefin Jones, H., Boddy, L., Baldrian, P. (2011b) Invertebrate grazing determines enzyme production by basidiomycete fungi. Soil Biology & Biochemistry, 43, 20602068.Google Scholar
Crowther, T. W., Jones, T. H., Boddy, L. (2011c) Species-specific effects of grazing invertebrates on mycelial emergence and growth from woody resources into soil. Fungal Ecology, 5, 333341.Google Scholar
Crowther, T. W., Stanton, D. W. G., Thomas, S. M., et al. (2013) Top-down control of soil fungal community composition by a globally distributed keystone consumer. Ecology, 94, 25182528.Google Scholar
Crowther, T. W., Thomas, S. M., Maynard, D. S., et al. (2015) Biotic interactions mediate soil microbial feedbacks to climate change. Proceedings of the National Academy of Sciences of the United States of America, 112, 70337038.Google Scholar
Culman, S. W., Young-Mathews, A., Hollander, A. D., et al. (2010) Biodiversity is associated with indicators of soil ecosystem functions over a landscape gradient of agricultural intensification. Landscape Ecology, 25, 13331348.Google Scholar
Cunha, A., Azevedo, R. B. R., Emmons, S. W., Leroi, A. M. (1999) Variable cell numbers in nematodes. Nature, 402, 253.Google Scholar
Curtis, P. S., Wang, X. (1998) A meta-analysis of elevated CO2 effects on woody plant mass, form, and physiology. Oecologia, 113, 299313.Google Scholar
Cushman, J. H., Lawton, J. H., Manly, B. F. J. (1993) Latitudinal patterns in European ant assemblages: Variation in species richness and body size. Oecologia, 95, 3037.Google Scholar
Czechowski, P., Clarke, L. J., Breen, J., Cooper, A., Stevens, M. I. (2016) Antarctic eukaryotic soil diversity of the Prince Charles Mountains revealed by high-throughput sequencing. Soil Biology & Biochemistry, 95, 112121.Google Scholar
Dam, M., Vestergård, M., Christensen, S. (2012) Freezing eliminates efficient colonizers from nematode communities in frost-free temperate soils. Soil Biology & Biochemistry, 48, 167174.Google Scholar
Darwin, C. (1859) On the Origin of Species by Means of Natural Selection, Or the Preservation of Favoured Races in the Struggle for Life, London, Murray.Google Scholar
Darwin, C. (1881) The Formation of Vegetable Mould, through the Action of Worms, with Observations of Their Habits, London, Murray.Google Scholar
Dash, M. C. (1990) Enchytraeidae. In: Soil Biology Guide (ed. Dindal, D. L.) pp. 311340. New York, John Wiley.Google Scholar
Dauber, J., Purtauf, T., Allspach, A., et al. (2005) Local vs. landscape controls on diversity: A test using surface-dwelling soil macroinvertebrates of differing mobility. Global Ecology and Biogeography, 14, 213221.CrossRefGoogle Scholar
Davidson, D. A., Bruneau, P. M. C., Grieve, I. C., Young, I. M. (2002) Impacts of fauna on an upland grassland soil as determined by micromorphological analysis. Applied Soil Ecology, 20, 133143.Google Scholar
Davis, E. L., Hussey, R. S., Baum, T. J. (2004) Getting to the root of parasitism by nematodes. Trends in Parasitology, 20, 134141.Google Scholar
Day, T. A., Ruhland, C. T., Strauss, S. L., et al. (2009) Response of plants and the dominant microarthropod, Cryptopygus antarcticus, to warming and contrasting precipitation regimes in Antarctic tundra. Global Change Biology, 15, 16401651.Google Scholar
De Deyn, G. B., Raaijmakers, C. E., van Ruijven, J., Berendse, F., van der Putten, W. H. (2004) Plant species identity and diversity effects on different trophic levels of nematodes in the soil food web. Oikos, 106, 576586.Google Scholar
De Deyn, G. B., Raaijmakers, C. E., Zoomer, H. R., et al. (2003) Soil invertebrate fauna enhances grassland succession and diversity. Nature, 422, 711713.Google Scholar
De Deyn, G. B., van der Putten, W. H. (2005) Linking aboveground and belowground diversity. Trends in Ecology and Evolution, 20, 625633.Google Scholar
de Graaff, M.-A., Adkins, J., Kardon, P., Throop, H. L. (2015) A meta-analysis of soil biodiversity impacts on the carbon cycle. Soil, 1, 257271.Google Scholar
de Groot, G. A., Jagers Op Akkerhuis, G. A. J. M., Dimmers, W. J., Charrier, X., Faber, J. H. (2016) Biomass and diversity of soil mite functional groups respond to extensification of land management, potentially affecting soil ecosystem services. Frontiers in Environmental Science, 4, 15.Google Scholar
De Ruiter, P. C., Moore, J. C., Zwart, K. B., et al. (1993) Simulation of nitrogen mineralization in the below-ground food webs of two winter wheat fields. Journal of Applied Ecology, 30, 95106.Google Scholar
de Vries, F. T., Bloem, J., Quirk, H., et al. (2012a) Extensive management promotes plant and microbial nitrogen retention in temperate grassland. PloS ONE, 7, e51201.Google Scholar
de Vries, F. T., Liiri, M., Bjørnlund, L., et al. (2012b) Land use alters the resistance and resilience of soil food webs to drought. Nature Climate Change, 2, 276280.Google Scholar
Decaëns, T. (2010) Macroecological patterns in soil communities. Global Ecology and Biogeography, 19, 287302.Google Scholar
Decaëns, T., Bureau, F., Margerie, P. (2003) Earthworm communities in a wet agricultural landscape of the Seine Valley (Upper Normandy, France). Pedobiologia, 47, 479489.Google Scholar
Decaëns, T., Jiménez, J. J., Barros, A. E., et al. (2004) Soil macrofaunal communities in permanent pastures derived from tropical forest or savanna. Agriculture, Ecosystems and Environment, 103, 301312.Google Scholar
Decaëns, T., Jiménez, J. J., Gioia, C., Measey, G. J., Lavelle, P. (2006) The values of soil animals for conservation biology. European Journal of Soil Biology, 42, S23S38.Google Scholar
Decaëns, T., Jiménez, J. J., Rossi, J. P. (2009) A null-model analysis of the spatio-temporal distribution of earthworm communities in Colombian grasslands. Journal of Tropical Ecology, 25, 415427.Google Scholar
Decaëns, T., Lavelle, P., Jiménez, J. J., Escobar, G., Rippstein, G. (1994) Impact of land management on soil macrofauna in the Oriental Llanos of Colombia. European Journal of Soil Biology, 30, 157168.Google Scholar
Decaëns, T., Margerie, P., Aubert, M., Hedde, M., Bureau, F. (2008) Assembly rules within earthworm communities in north-western France: A regional analysis. Applied Soil Ecology, 39, 321335.Google Scholar
Decaëns, T., Mariani, L., Lavelle, P. (1999) Soil surface macrofaunal communities associated with earthworm casts in grasslands of the Eastern Plains of Colombia. Applied Soil Ecology, 13, 87100.Google Scholar
Decaëns, T., Rossi, J. P. (2001) Spatio-temporal structure of earthworm community and soil heterogeneity in a tropical pasture. Ecography, 24, 671682.Google Scholar
Degenkolb, T., Vilcinskas, A. (2016) Metabolites from nematophagous fungi and nematicidal natural products from fungi as an alternative for biological control. Part I: Metabolites from nematophagous ascomycetes. Applied Microbiology and Biotechnology, 100, 37993812.Google Scholar
Delabie, J. H. C., Fowler, G. W. (1995) Soil and litter cryptic ant assemblages of Bahian cocoa plantations. Pedobiologia, 39, 423433.Google Scholar
Delaville, L., Rossi, J.-P., Quénéhervé, P. (1996) Plant row and soil factors influencing the microspatial patterns of plant parasitic nematodes on sugarcane in Martinique. Fundamental and Applied Nematology, 19, 321328.Google Scholar
Delgado-Baquerizo, M., Powell, J. R., Hamonts, K., et al. (2017) Circular linkages between soil biodiversity, fertility and plant productivity are limited to topsoil at the continental scale. New Phytologist, 215, 11861196.Google Scholar
Denno, R. F., Gruner, D. S., Kaplan, I. (2008) Potential for entomopathogenic nematodes in biological control: A meta-analytical synthesis and insights from trophic cascade theory. Journal of Nematology, 40, 6172.Google Scholar
Devetter, M., Háněl, L., Řeháková, K., Doležal, J. (2017) Diversity and feeding strategies of soil microfauna along elevation gradients in Himalayan cold deserts. PloS ONE, 12, e0187646.Google Scholar
Diamond, J. M. (1975) Assembly of species communities. In: Ecology and Evolution of Communities (eds. Cody, M. L., Diamond, J. M.) pp. 342444. Cambridge, MA, Harvard University Press.Google Scholar
Didden, W., Römbke, J. (2001) Enchytraeids as indicator organisms for chemical stress in terrestrial ecosystems. Ecotoxicology and Environmental Safety, 50, 2543.Google Scholar
Didden, W. A. M. (1990) Involvement of Enchytraeidae (Oligochaeta) in soil structure evolution in agricultural fields. Biology and Fertility of Soils, 9, 152158.Google Scholar
Didden, W. A. M. (1993) Ecology of terrestrial Enchytraeidae. Pedobiologia, 37, 229.Google Scholar
Dijkstra, P., Ishizu, A., Doucett, R., et al. (2006) 13C and 12N natural abundance of soil microbial biomass. Soil Biology & Biochemistry, 38, 32573266.Google Scholar
Dindal, D. L. (1980) Soil Biology Guide, New York, John Wiley & Sons.Google Scholar
Dirilgen, T., Jucevia, E., Melecis, V., Querner, P., Bolger, T. (2018) Analysis of spatial patterns informs community assembly and sampling requirements for Collembola in forest soils. Acta Oecologica, 86, 2330.Google Scholar
Doblas-Miranda, E., Wardle, D. A., Peltzer, D. A., Yeates, G. W. (2008) Changes in the community structure and diversity of soil invertebrates across the Franz Josef Glacier chronosequence. Soil Biology & Biochemistry, 40, 10691081.Google Scholar
Domínguez, A., Bedano, J. C., Becker, A. R. (2014) Organic farming fosters agroecosystem functioning in Argentinian temperate soils: Evidence from litter decomposition and soil fauna. Applied Soil Ecology, 83, 170176.Google Scholar
Donner, J. (1966) Rotifers. London, Warne.Google Scholar
Dósza-Farkas, K. (1996) Reproduction strategies in some enchytraeid species. In: Newsletter on Enchytraeidae No. 5 (ed. Dósza-Farkas, K.) pp. 2533. Budapest, Hungary, Eötvös Loránd University.Google Scholar
Dray, S., Legendre, P., Peres-Neto, P. R. (2006) Spatial modelling: A comprehensive framework for principal coordinate analysis of neighbor matrices (PCNM). Ecological Modelling, 196, 483493.Google Scholar
Drigo, B., Kowalchuk, G. A., Yergeau, E., et al. (2007) Impact of elevated carbon dioxide on the rhizosphere communities of Carex arenaria and Festuca rubra. Global Change Biology, 13, 23962410.Google Scholar
Dromph, K. M., Cook, R., Ostle, N. J., Bardgett, R. D. (2006) Root parasite induced nitrogen transfer between plants is density dependent. Soil Biology & Biochemistry, 38, 24952498.Google Scholar
Ducarme, X. D., André, H. M., Wauthy, G., Lebrun, P. (2004) Are there real endogeic species in temperate forest mites? Pedobiologia, 48, 139147.Google Scholar
Dunn, R. R. (2004a) Managing the tropical landscape: A comparison of the effects of logging and forest conversion to agriculture on ants, birds, and lepidoptera. Forest Ecology and Management, 191, 215224.Google Scholar
Dunn, R. R. (2004b) Recovery of faunal communities during tropical forest regeneration. Conservation Biology, 18, 302309.Google Scholar
Dunn, R. R., Agosti, D., Andersen, A. N., et al. (2009) Climatic drivers of hemispheric asymmetry in global patterns of ant species richness. Ecology Letters, 12, 324333.Google Scholar
Edgecombe, G. D., Giribet, G. (2007) Evolutionary biology of centipedes (Myriapoda: Chilopoda). Annual Review of Entomology, 52, 151170.Google Scholar
Edwards, C. A., Lofty, J. R. (1969) The influence of agricultural practices on soil micro-arthropod populations. In: The Soil Ecosystem (ed. Sheals, J. G.) pp. 237248. London, Publs. Syst. Assoc.Google Scholar
Edwards, C. A., Lofty, J. R. (1977) Biology of Earthworms, 2nd edn, London, Chapman and Hall.Google Scholar
Edwards, C. A., Reichle, D. E., Crossley, D. A. Jr. (1970) The role of soil invertebrates in turnover of organic matter and nutrients. Ecological Studies – Analysis and Synthesis, 1, 147172.Google Scholar
Eggleton, P. (1994) Termites live in a pear-shaped world: A response to Platnik. Journal of Natural History, 28, 12091212.Google Scholar
Eggleton, P. (2000) Global patterns of termite diversity. In: Termites: Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 2551. Dordrecht, Kluwer Academic Press.Google Scholar
Eggleton, P., Bignell, D. E. (1995) Monitoring the response of tropical insects to changes in the environment: Troubles with termites. In: Insects in a Changing Environment (eds. Harrington, R., Stork, N. E.) pp. 473497. London, Academic Press.Google Scholar
Eggleton, P., Bignell, D. E., Sands, W. A., et al. (1996) The diversity, abundance and biomass of termites under differing levels of forest disturbance in the Mbalmayo Forest Reserve, southern Cameroon. Philosophical Transactions of the Royal Society B: Biological Sciences, 351, 5168.Google Scholar
Eggleton, P., Vanbergen, A. J., Jones, D. T., et al. (2005) Assemblages of soil macrofauna across a Scottish land-use intensification gradient: Influences of habitat quality, heterogeneity and area. Journal of Applied Ecology, 42, 11531164.Google Scholar
Ehnes, R. B., Rall, B. C., Brose, U. (2011) Phylogenetic grouping, curvature and metabolic scaling in terrestrial invertebrates. Ecology Letters, 14, 9931000.Google Scholar
Eisen, G. (1900) Researches in American Oligochaeta, with especial reference to those of the Pacific Coast and adjacent islands. Proceedings of the California Academy of Sciences 3rd Series: Zoology, 2, 85276.Google Scholar
Eisenhauer, N. (2010) The action of an animal ecosystem engineer: Identification of the main mechanisms of earthworm impacts on soil microarthropods. Pedobiologia, 53, 343352.Google Scholar
Eisenhauer, N., Antunes, P. M., Bennett, A. E., et al. (2017) Priorities for research in soil ecology. Pedobiologia – Journal of Soil Ecology, 63, 17.Google Scholar
Eisenhauer, N., Bowker, M. A., Grace, J. B., Powell, J. R. (2015) From patterns to causal understanding: Structural equation modeling (SEM) in soil ecology. Pedobiologia, 58, 6572.Google Scholar
Eisenhauer, N., Dobies, T., Cesarz, S., et al. (2013) Plant diversity effects on soil food webs are stronger than those of elevated CO2 and N deposition in a long-term grassland experiment. Proceedings of the National Academy of Sciences of the United States of America, 110, 68896894.Google Scholar
Eisenhauer, N., Klier, M., Partsch, S., et al. (2009) No interactive effects of pesticides and plant diversity on soil microbial biomass and respiration. Applied Soil Ecology, 42, 3136.Google Scholar
Eisenhauer, N., Milcu, A., Sabais, A. C. W., et al. (2011a) Plant diversity surpasses plant functional groups and plant productivity as drivers of soil biota in the long term. PloS ONE, 6 (1), e16055. doi:10.1371/journal.pone.0016055.Google Scholar
Eisenhauer, N., Reich, P. B. (2012) Above- and below-ground plant inputs both fuel soil food webs. Soil Biology & Biochemistry, 45, 156160.Google Scholar
Eisenhauer, N., Sabais, A. C. W., Scheu, S. (2011b) Collembola species composition and diversity effects on ecosystem functioning vary with plant functional group identity. Soil Biology & Biochemistry, 43, 16971704.Google Scholar
Eisenhauer, N., Schädler, M. (2011) Inconsistent impacts of decomposer diversity on the stability of aboveground and belowground ecosystem functions. Oecologia, 165, 403415.Google Scholar
Eisenhauer, N., Scheu, S., Reich, P. B. (2012) Increasing plant diversity effects on productivity with time due to delayed soil biota effects on plants. Basic and Applied Ecology, 13, 571578.Google Scholar
Ekelund, F., Rønn, R. (1994) Notes on protozoa in agricultural soil with emphasis on heterotrophic flagellates and naked amoebae and their ecology. FEMS Microbiology Reviews, 15, 321353.Google Scholar
Ekelund, F., Rønn, R., Christensen, S. (2001) Distribution with depth of protozoa, bacteria and fungi in soil profiles from three Danish forest sites. Soil Biology & Biochemistry, 33, 475481.Google Scholar
Eldridge, D. J., Koen, T. B. (2008) Formation of nutrient-poor soil patches in a semi-arid woodland by the European rabbit (Oryctolagus cuniculus L.). Austral Ecology, 33, 8898.Google Scholar
Emerson, B. C., Cicconardi, F., Fanciulli, P. P., Shaw, P. J. A. (2011) Phylogeny, phylogeography, phylobetadiversity and the molecular analysis of biological communities. Philosophical Transactions of the Royal Society B: Biological Sciences, 366, 23912402.Google Scholar
Endlweber, K., Ruess, L., Scheu, S. (2009) Collembola switch diet in presence of plant roots thereby functioning as herbivores. Soil Biology & Biochemistry, 41, 11511154.Google Scholar
Endlweber, K., Scheu, S. (2006) Effects of Collembola on root properties of two competing ruderal plant species. Soil Biology & Biochemistry, 38, 20252031.Google Scholar
Erséus, C. (2005) Phylogeny of Oligochaetous Clitellata. Hydrobiologia, 535, 357372.Google Scholar
Escobar, F., Lobo, J. M., Halffter, G. (2005) Altitudinal variation of dung beetle (Scarabaeidae: Scarabaeinae) assemblages in the Colombian Andes. Global Ecology and Biogeography, 14, 327337.Google Scholar
Ettema, C. H. (1998) Soil nematode diversity: Species coexistence and ecosystem function. Journal of Nematology, 30, 159169.Google Scholar
Ettema, C. H., Coleman, C. D., Vellidis, G., Lowrance, R., Rathbun, S. L. (1998) Spatiotemporal distributions of bacterivorous nematodes and soil resources in a restored riparian wetland. Ecology, 79, 27212734.Google Scholar
Ettema, C. H., Rathbun, S. L., Coleman, D. C. (2000) On spatiotemporal patchiness and the coexistence of five species of Chronogaster (Nematoda: Chronogasteridae) in a riparian wetland. Oecologia, 125, 444452.Google Scholar
Ettema, C. H., Wardle, D. A. (2002) Spatial soil ecology. Trends in Ecology and Evolution, 17, 177183.Google Scholar
Eviner, V. T., Chapin, F. S. III (2001) Plant species provide vital ecosystem functions for sustainable agriculture, rangeland management and restoration. California Agriculture, 55, 5459.Google Scholar
Faber, J. H. (1991) Functional classification of soil fauna. A new approach. Oikos, 62, 110117.Google Scholar
Fanin, N., Gundale, M. J., Farrell, M., et al. (2018) Consistent effects of biodiversity loss on multifunctionality across contrasting ecosystems. Nature Ecology & Evolution, 2, 269278.Google Scholar
Fattorini, S. (2009) On the general dynamic model of oceanic island biogeography. Journal of Biogeography, 36, 11001110.Google Scholar
Faucon, M. P., Houben, D., Lambers, H. (2017) Plant functional traits: Soil and ecosystem services. Trends in Plant Science, 22, 385394.Google Scholar
Fayle, T. M., Turner, E. C., Snaddon, J. L. (2010) Oil palms expansion into rain forest greatly reduces ant biodiversity in canopy, epiphytes and leaf-litter. Basic and Applied Ecology, 11, 337345.Google Scholar
Feldman, L. J. (1988) The habits of roots. Bioscience, 38, 612618.Google Scholar
Feller, C., Brown, G. G., Blanchart, E., Deleporte, P., Chernyanskii, S. S. (2003) Charles Darwin, earthworms and the natural sciences: Various lessons from past to future. Agriculture, Ecosystems & Environment, 99, 2949.Google Scholar
Ferris, H., Bongers, T., De Goede, R. G. M. (2001) A framework for soil food web diagnostics: Extension of the nematode faunal analysis concept. Applied Soil Ecology, 18, 1329.Google Scholar
Ferris, H., Venette, R. C., Lau, S. S. (1997) Population energetics of bacterial-feeding nematodes: Carbon and nitrogen budgets. Soil Biology & Biochemistry, 29, 11831194.Google Scholar
Ferris, V. R., Ferris, J. M., Goseco, C. G. (1981) Phylogenetic and biogeographic hypotheses in Leptonchidae (Nematoda: Dorylaimida) and a new classification. In: Proceedings of the Helminthological Society. pp. 163171. Washington.Google Scholar
Ferris, V. R., Goseco, C. G., Ferris, J. M. (1976) Biogeography of free-living soil nematodes from the perspective of plate tectonics. Science, 193, 508510.Google Scholar
Fiedler, A. K., Landis, D. A., Wratten, S. D. (2008) Maximizing ecosystem services from conservation biological control: The role of habitat management. Biological Control, 45, 254271.Google Scholar
Fiera, C. (2014) Application of stable isotopes and lipid analysis to understand trophic interactions in springtails. North-western Journal of Zoology, 10, 227235.Google Scholar
Fierer, N., Strickland, M. S., Litzin, D., Bradford, M. A., Cleveland, C. C. (2009) Global patterns in belowground communities. Ecology Letters, 12, 12381249.Google Scholar
Filser, J. (2003) The role of Collembola in carbon and nitrogen cycling in soil. Pedobiologia, 46, 234245.Google Scholar
Finlay, B. G. (2002) Global dispersal of free-living microbial eukaryote species. Science, 296, 10611063.Google Scholar
Finlay, B. J., Fenchel, T. (1999) Divergent perspectives on protist species richness. Protist, 150, 229233.Google Scholar
Fisher, B. L. (1999a) Ant diversity patterns along an elevational gradient in the reserve Naturelle Integrale d'Andohahela, Madagascar. Fieldiana Zoology, 94, 129147.Google Scholar
Fisher, B. L. (1999b) Improving inventory efficiency: A case study of leaf-litter ant diversity in Madagascar. Ecological Applications, 9, 714731.Google Scholar
Fisher, R., Mcdowell, N., Purvis, D., et al. (2010) Assessing uncertainties in a second-generation dynamic vegetation model caused by ecological scale limitations. New Phytologist, 187, 666681.Google Scholar
Floren, A., Freking, A., Biehl, M., Linsenmair, K. E. (2001) Anthropogenic disturbance changes the structure of arboreal tropical ant communities. Ecography, 24, 547554.Google Scholar
Flury, P., Aellen, N., Ruffner, B., et al. (2016) Insect pathogenicity in plant-beneficial pseudomonads: Phylogenetic distribution and comparative genomics. The ISME Journal, 10, 25272542.Google Scholar
Foissner, W. (1987) Soil protozoa: Fundamental problems, ecological significance, adaptations in ciliates and testaceans, bioindicators, and guide to the literature. Progress in Protistology, 2, 69212.Google Scholar
Foissner, W. (1999a) Protist diversity: Estimates of the near-imponderable. Protist, 150, 363368.Google Scholar
Foissner, W. (1999b) Soil protozoa as bioindicators: Pros and cons, methods, diversity, representative examples. Agriculture, Ecosystems and Environment, 74, 95112.Google Scholar
Foissner, W. (2006) Biogeography and dispersal of micro-organisms: A review emphasizing protists. Acta Protozoologica, 45, 111136.Google Scholar
Foley, J., Ramankutty, N., Brauman, K., et al. (2011) Solutions for a cultivated planet. Nature, 478, 337342.Google Scholar
Fontaneto, D., Ricci, C. (2006) Spatial gradients in species diversity of microscopic animals: The case of bdelloid rotifers at high altitude. Journal of Biogeography, 33, 13051313.Google Scholar
Fontaneto, D., Westberg, M., Hortal, J. (2011) Evidence of weak habitat specialisation in microscopic animals. PloS ONE, 6, e23969. doi:10.1371/journal.pone.0023969.Google Scholar
Forey, E., Barot, S., Decaëns, T., et al. (2011) Importance of earthworm seed interactions for the composition and structure of plant communities: A review. Acta Oecologica, 37, 594603.Google Scholar
Fortin, M.-J., Dale, M. (2005) Spatial Ecology: A Guide for Ecologists. Cambridge, UK, Cambridge University Press.Google Scholar
Fowbert, J. A., Smith, R. I. L. (1994) Rapid population increases in native vascular plants in the Argentine islands, Antarctic Peninsula. Arctic and Alpine Research, 26, 290296.Google Scholar
Fox, C. A., Fonseca, E. J. A., Miller, J. J., Tomlin, A. D. (1999) The influence of row position and selected soil attributes on Acarina and Collembola in no-till and conventional continuous corn on a clay loam soil. Applied Soil Ecology, 13, 18.Google Scholar
Francini, G., Hui, N., Jumpponen, A., et al. (2018) Soil biota in boreal urban greenspace: Responses to plant type and age. Soil Biology & Biochemistry, 118, 145155.Google Scholar
Francl, L. J. (1993) Interactions of Nematodes with Mycorrhizae and Mycorrhizal Fungi. New York, Chapman & Hall.Google Scholar
Fraser, C. I., Terauds, A., Smellie, J., Convey, P., Chown, S. L. (2014) Geothermal activity helps life survive glacial cycles. Proceedings of the National Academy of Sciences of the United States of America, 111, 56345639.Google Scholar
Fraser, P. M., Schon, N. L., Piercy, J. E., Mackay, A. D., Minor, M. A. (2012) Influence of summer irrigation on soil invertebrate populations in a long-term sheep irrigation trial at Winchmore (Canterbury). New Zealand Journal of Agricultural Research, 55, 165180.Google Scholar
Freckman, D. W. (1988) Bacterivorous nematodes and organic matter decomposition. Agriculture, Ecosystems & Environment, 24, 195217.Google Scholar
Freckman, D. W., Ettema, C. H. (1993) Assessing nematode communities in agroecosystems of varying human intervention. Agriculture, Ecosystems & Environment, 45, 239261.Google Scholar
Freckman, D. W., Virginia, R. A. (1989) Plant-feeding nematodes in deep-rooting desert ecosystems. Ecology, 70, 16651678.Google Scholar
Frederiksen, H. B., Rønn, R., Christensen, S. (2001) Effect of elevated atmospheric CO2 and vegetation type on microbiota associated with decomposing straw. Global Change Biology, 7, 313321.Google Scholar
Frew, A., Barnett, K., Riegler, M., Nielsen, U. N., Johnson, S. N. (2016) Belowground ecology of scarabs feeding on grass roots: Current knowledge and future directions for management in Australasia. Frontiers in Plant Science, 7, 321. doi:10.3389/fpls.2016.00321.Google Scholar
Fromm, H., Winter, K., Filser, J., Hantschel, R., Beese, F. (1993) The influence of soil type and cultivation system on the spatial distributions of the soil fauna and microorganisms and their interactions. Geoderma, 60, 109118.Google Scholar
Frostegård, Å., Petersen, S., Bååth, E., Nielsen, T. (1997) Dynamics of a microbial community associated with manure hot spots as revealed by phospholipid fatty acid analyses. Applied and Environmental Microbiology, 63, 22242231.Google Scholar
Frouz, J., Thébault, E., Pižl, V., et al. (2013) Soil food web changes during spontaneous succession at post mining sites: A possible ecosystem engineering effect on food web organization? PloS ONE, 8, e79694.Google Scholar
Fukima, T., Nakajima, M. (2013) Complex plant–soil interactions enhance plant species diversity by delaying community convergence. Journal of Ecology, 101, 316324.Google Scholar
Gabbutt, P. D. (1967) Quantitative sampling of the pseudoscorpion Chthonius ischnocheles from beech litter. Journal of Zoology, London, 151, 469478.Google Scholar
Gabriel, A. G. A., Chown, S. L., Barendse, J., et al. (2001) Biological invasions of Southern Ocean islands: The Collembola of Marion Island as a test of generalities. Ecography, 24, 421430.Google Scholar
Galloway, J. N., Townsend, A. R., Erisman, J. W., et al. (2008) Transformation of the nitrogen cycle: Recent trends, questions and potential solutions. Science, 320, 889892.Google Scholar
Gange, A. C., Stagg, P. G., Ward, L. K. (2002) Arbuscular mycorrhizal fungi affect phytophagous insect specialism. Ecology Letters, 5, 1115.Google Scholar
Gao, M., He, P., Zhang, X., Liu, D., Wu, D. (2014) Relative roles of spatial factors, environmental filtering and biotic interactions in fine-scale structuring of a soil mite community. Soil Biology & Biochemistry, 79, 6877.Google Scholar
Garćia-Palacios, P., Maestre, F. T., Kattge, J., Wall, D. H. (2013) Climate and litter quality differently modulate the effects of soil fauna on litter decomposition across biomes. Ecology Letters, 16, 10451053.Google Scholar
Gaston, K. J. (2000) Global patterns in biodiversity. Nature, 405, 220227.Google Scholar
Gaston, K. J., Blackburn, T., Lawton, J. H. (1997) Interspecific abundance range size relationships: An appraisal of mechanisms. Journal of Animal Ecology, 66, 579601.Google Scholar
Geisen, S. (2016) The bacterial–fungal energy channel concept challenged by enormous functional versatility of soil protists. Soil Biology & Biochemistry, 102, 2225.Google Scholar
Geisen, S., Cornelia, B., Römbke, J., Bonkowski, M. (2014) Soil water availability strongly alters the community composition of soil protists. Pedobiologia, 57, 205213.Google Scholar
Geisen, S., Koller, R., Hünninghaus, M., Dumack, K., Bonkowski, M. (2016) The soil food web revisited: Diverse and widespread mycophagous soil protists. Soil Biology & Biochemistry, 94, 1018.Google Scholar
Geisen, S., Mitchell, E. A. D., Wilkinson, D. M., et al. (2017) Soil protistology rebooted: 30 fundamental questions to start with. Soil Biology & Biochemistry, 111, 94103.Google Scholar
Geisen, S., Rosengarten, J., Koller, R., et al. (2015a) Pack hunting by a common soil amoeba on nematodes. Environmental Microbiology, 17, 45384546.Google Scholar
Geisen, S., Viscaíno, A., Bonkowski, M., De Groot, G. A. (2015b) Not all are free-living: High-throughput DNA metabarcoding reveals a diverse community of protists parasitizing soil metazoa. Molecular Ecology, 24, 45564569.Google Scholar
George, P. B. L., Keith, A. M., Creer, S., et al. (2017) Evaluation of mesofauna communities as soil quality indicators in a national-level monitoring programme. Soil Biology & Biochemistry, 115, 537546.Google Scholar
Georgieva, S. S., Mcgrath, S. P., Hooper, D. J., Chambers, B. S. (2002) Nematode communities under stress: The long-term effects of heavy metals in soil treated with sewage sludge. Applied Soil Ecology, 20, 2742.Google Scholar
Gerard, B. M. (1967) Factors affecting earthworms in pastures. Journal of Animal Ecology, 36, 235252.Google Scholar
Gilbert, O. W. (1956) The natural histories of four species of Calathus (Coleoptera, Carabidae) living in sand dunes in Anglesey, N. Wales. Oikos, 7, 2247.Google Scholar
Gill, R. W. (1969) Soil microarthropod abundance following old-field litter manipulation. Ecology, 50, 805816.Google Scholar
Giller, K. E., Beare, M. H., Lavelle, P., Izac, A. M. N, Swift, M. J. (1997) Agricultural intensification, soil biodiversity and agroecosystem function. Applied Soil Ecology, 6, 316.Google Scholar
Giller, P. S. (1996) The diversity of soil communities, the ‘poor man's tropical rainforest’. Biodiversity and Conservation, 5, 135168.Google Scholar
Gillison, A. N., Jones, D. T., Susilo, F.-X., Bignell, D. E. (2003) Vegetation indicates diversity of soil macroinvertebrates: A case study with termites along a land-use intensification gradient in lowland Sumatra. Organisms, Diversity and Evolution, 3, 111126.Google Scholar
Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. B., Charnov, E. L. (2001) Effects of size and temperature on metabolic rate. Science, 293, 22482251.Google Scholar
Godfray, H. C. J., Beddington, J. R., Crute, I. R., et al. (2010) Food security: The challenge of feeding 9 billion people. Science, 327, 812818.Google Scholar
González, G., García, E., Cruz, V., et al. (2007) Earthworm communities along an elevation gradient in northeastern Puerto Rico. European Journal of Soil Biology, 43, S24S32.Google Scholar
Gooseff, M. N., Barrett, J. E., Doran, P. T., et al. (2003) Snowpack influence on soil biogeochemical processes and invertebrate distribution in the McMurdo Dry Valleys, Antarctica. Arctic, Antarctica and Alpine Research, 35, 9199.Google Scholar
Goralczyk, K. (1998) Nematodes in a coastal dune succession: Indicators of soil properties? Applied Soil Ecology, 9, 465469.Google Scholar
Görres, J. H., Dichiaro, M. J., Lyons, J. B., Amador, J. A. (1998) Spatial and temporal patterns of soil biological activity in a forest and an old field. Soil Biology & Biochemistry, 30, 219230.Google Scholar
Gossner, M. M., Lewinsohn, T. M., Kahl, T., et al. (2016) Land-use intensification causes multitrophic homogenization of grassland communities. Nature, 540, 266269.Google Scholar
Gotelli, N. J., Ellison, A. M. (2002) Biogeography at a regional scale: Determinants of ant species density in New England bogs and forests. Ecology, 83, 16041609.Google Scholar
Gotelli, N. J., Graves, G. R. (1996) Null Models in Ecology, Washington, DC, Smithsonian Institution Press.Google Scholar
Grace, J. B. (2006) Structural Equation Modeling and Natural Systems, Cambridge, Cambridge University Press.Google Scholar
Graham, J. H. (2001) What do root pathogens see in mycorrhizas? New Phytologist, 149, 357359.Google Scholar
Grandy, A. S., Wieder, W. R., Wickings, K., Kyker-Snowman, E. (2016) Beyond microbes: Are fauna the next frontier in soil biogeochemical models? Soil Biology & Biochemistry, 102, 4044.Google Scholar
Greenslade, P. (2008) Has survey effort of Australia's islands reflected conservation and biogeographical significance? An assessment using Collembola. European Journal of Soil Biology, 44, 458462.Google Scholar
Greenslade, P., Convey, P. (2012) Exotic Collembola on subantarctic islands: Pathways, origins and biology. Biological Invasions, 14, 405417.Google Scholar
Greenstone, M. H., Weber, D. C., Coudron, T. A., Payton, M. E., Hu, J. S. (2012) Removing external DNA contamination from arthropod predators destined for molecular gut-content analysis. Molecular Ecology Resources, 12, 464469.Google Scholar
Grgič, T., Kos, I. (2005) Influence of forest development phase on centipede diversity in managed beech forests in Slovenia. Biodiversity and Conservation, 14, 18411862.Google Scholar
Griffiths, B. S. (1989) Enhanced nitrification in the presence of bacteriophagous protozoa. Soil Biology & Biochemistry, 21, 10451051.Google Scholar
Griffiths, B. S. (1990) A comparison of microbial-feeding nematodes and protozoa in the rhizosphere of different plants. Biology and Fertility of Soils, 9, 8388.Google Scholar
Griffiths, B. S. (1994) Soil nutrient flow. In: Soil Protozoa (ed. Darbyshire, J. F.) pp. 6591. Wallingford, CAB International.Google Scholar
Griffiths, B. S., Caul, S. (1993) Migration of bacterial-feeding nematodes, but not protozoa, to decomposing grass residues. Biology and Fertility of Soils, 15, 201207.Google Scholar
Grime, J. P. (1973) Control of species diversity in herbaceous vegetation. Journal of Environmental Management, 1, 151167.Google Scholar
Gunnarson, T., Tunlid, A. (1986) Recycling of fecal pellets in isopods: Microorganisms and nitrogen compounds as potential food for Oniscus asellus L. Soil Biology & Biochemistry, 18, 595600.Google Scholar
Gutt, J., Isla, E., Bertler, N., et al. (2018) Cross-disciplinarity in the advance of Antarctic ecosystem research. Marine Genomics, 37, 117.Google Scholar
Haase, S., Ruess, L., Neumann, G., Marhan, S., Kandeler, E. (2007) Low-level herbivory by root-knot nematodes (Meloidogyne incognita) modifies root hair morphology and rhizodeposition in host plants (Hordeum vulgare). Plant and Soil, 301, 151164.Google Scholar
Hågvar, S. (1998) The relevance of the Rio-Convention on biodiversity to conserving the biodiversity of soils. Applied Soil Ecology, 9, 17.Google Scholar
Hågvar, S., Klanderud, K. (2009) Effect of simulated environmental change on alpine soil arthropods. Global Change Biology, 15, 29722980.Google Scholar
Haimi, J., Huhta, V. (1990) Effects of earthworms on decomposition processes in a raw humus soil: A microcosm study. Biology and Fertility of Soils, 10, 178183.Google Scholar
Hale, C. M., Frelich, L. E., Reich, P. B., Pastor, J. (2005) Effects of European earthworm invasion on soil characteristics in northern hardwood forests of Minnesota, USA. Ecosystems, 8, 911927.Google Scholar
Hamilton, W. E., Dindal, D. L. (1983) The vermisphere concept: Earthworm activity and sewage sludge. Biocycle, 24, 5455.Google Scholar
Handa, I. T., Aerts, R., Berendse, F., et al. (2007) Consequences of biodiversity loss for litter decomposition across biomes. Nature, 509, 218221.Google Scholar
Hanlon, R. D. G., Anderson, J. M. (1979) The effects of Collembola grazing on microbial activity in decomposing leaf litter. Oecologia, 38, 9399.Google Scholar
Hanlon, R. D. G., Anderson, J. M. (1980) Influence of macroarthropod feeding activities on microflora in decomposing oak leaves. Soil Biology & Biochemistry, 12, 255261.Google Scholar
Hansen, R. (2000) Effects of habitat complexity and composition on a diverse litter microarthropod assemblage. Ecology, 81, 11201132.Google Scholar
Hansen, R. A., Coleman, D. C. (1998) Litter complexity and composition are determinants of the diversity and species composition of oribatid mites (Acari: Oribatida) in litterbags. Applied Soil Ecology, 9, 1723.Google Scholar
Harris, J. (2009) Soil microbial communities and restoration ecology: Facilitators or followers? Science, 325, 573575.Google Scholar
Hart, S. P., Usinowicz, J., Levine, J. M. (2017) The spatial scales of species coexistence. Nature Ecology & Evolution, 1, 10661073.Google Scholar
Hassall, M., Turner, J. G., Rands, M. R. W. (1987) Effects of terrestrial isopods on the decomposition of woodland leaf litter. Oecologia, 72, 597604.Google Scholar
Hassink, J., Bouwman, L. A., Zwart, K. B., Brussard, L. (1993) Relationship between habitable pore space, soil biota and mineralization rates in grassland soil. Soil Biology & Biochemistry, 25, 4755.Google Scholar
Hättenschwiler, S., Tiunov, A. V., Scheu, S. (2005) Biodiversity and litter decomposition in terrestrial ecosystems. Annual Review of Ecology, Evolution and Systematics, 36, 191218.Google Scholar
Haynes, R. J., Naidu, R. (1998) Influence of lime, fertilizer and manure applications on soil organic matter content and soil physical conditions: A review. Nutrient Cycling in Agroecosystems, 51, 123137.Google Scholar
Heckel, D. G. (2012) Insecticide resistance after Silent spring. Science, 337, 16121614.Google Scholar
Hedlund, K., Boddy, L., Preston, C. M. (1991) Mycelial responses of the soil fungus, Mortierella isabellina, to grazing by Onychiurus armatus (Collembola). Soil Biology & Biochemistry, 23, 361366.Google Scholar
Hedlund, K., Öhrn, O. (2000) Tritrophic interactions in a soil community enhance decomposition rates. Oikos, 88, 585591.Google Scholar
Hedlund, K., Santa Regina, I., van der Putten, W. H., et al. (2003) Plant species diversity, plant biomass and responses of the soil community on abandoned land across Europe: Idiosyncracy or above-belowground time lags. Oikos, 103, 4558.Google Scholar
Heemsbergen, D. A., Berg, M. P., Loreau, M., et al. (2004) Biodiversity effects on soil processes explained by interspecific functional dissimilarity. Science, 306, 10191020.Google Scholar
Heger, T. J., Lara, E., Mitchell, E. A. D. (2011) Arcellinida testate amoebae (Amoebozoa: Arcellinida): Model of organisms for assessing microbial biogeography. In: Biogeography of Microscopic Organisms (ed. Fontaneto, D.) pp. 111129. Cambridge, Cambridge University Press.Google Scholar
Heidemann, K., Scheu, S., Ruess, L., Maraun, M. (2011) Molecular detection of nematode predation and scavenging in oribatid mites: Laboratory and field experiments. Soil Biology & Biochemistry, 43, 22292236.Google Scholar
Heisler, C., Kaiser, E.-A. (1995) Influence of agricultural traffic and crop management on collembolan and microbial biomass in arable soil. Biology and Fertility of Soils, 19, 159165.Google Scholar
Helmus, M. R., Savage, K., Diebel, M. W., Maxted, J. T., Ives, A. R. (2007) Separating the determinants of phylogenetic community structure. Ecology Letters, 10, 917925.Google Scholar
Hendrix, P. F., Baker, G. H., Callaham, M. A. Jr, et al. (2006) Invasion of exotic earthworms into ecosystems inhabited by native earthworms. Biological Invasions, 8, 12871300.Google Scholar
Hendrix, P. F., Callaham, M. A. J., Drake, J. M., et al. (2008) Pandora's box contained bait: The global problem of introduced earthworms. Annual Review of Ecology, Evolution and Systematics, 39, 593613.Google Scholar
Heneghan, L., Coleman, C. D., Zou, X., Crossley, D. A. Jr., Haines, B. L. (1998) Soil microarthropod community structure and litter decomposition dynamics: A study of tropical and temperate sites. Applied Soil Ecology, 9, 3338.Google Scholar
Hillebrand, H. (2004) On the generality of the latitudinal diversity gradient. American Naturalist, 163, 192211.Google Scholar
Hillebrand, H., Blenckner, T. (2002) Regional and local impact on species diversity – From pattern to processes. Oecologia, 132, 479491.Google Scholar
Hillebrand, H., Watermann, F., Karez, R., Berninger, U.-G. (2001) Differences in species richness patterns between unicellular and multicellular organisms. Oecologia, 126, 114124.Google Scholar
HilleRisLambers, J., Adler, P. B., Harpole, W. S., Levine, J. M., Mayfield, M. M. (2012) Rethinking community assembly through the lens of coexistence theory. Annual Review of Ecology, Evolution, and Systematics, 43, 227248.Google Scholar
Hiltner, L. (1904) Über neue Erfahrungen und Probleme auf dem Gebiet der Bodenbakteriologie unter besonderer Berücksichtigung der Gründüngung und Brache. Arbeiten der Deutschen Landwirtschaftlichen Gesellschaft, 98, 5978.Google Scholar
Hobbs, R. J., Mooney, H. A. (1985) Community and population dynamics of serpentine grassland annuals in relation to gopher disturbance. Oecologia, 67, 342351.Google Scholar
Hodda, M. (2007) Phylum Nematoda. Zootaxa, 1668, 265293.Google Scholar
Hodda, M., Nobbs, J. (2008) A review of current knowledge on particular taxonomic features of the Australasian nematode fauna, with special emphasis on plant feeders. Australasian Plant Pathology, 37, 308317.Google Scholar
Hodkinson, I. D., Coulson, S. J., Webb, N. R. (2004) Invertebrate community assembly along proglacial chronosequences in the high Arctic. Journal of Animal Ecology, 73, 556568.Google Scholar
Hodkinson, I. D., Coulson, S. J., Webb, N. R., Block, W. (1996) Can Arctic soil microarthropods survive elevated summer temperatures? Functional Ecology, 10, 314321.Google Scholar
Hoeksema, J. D., Lussenhop, J., Teeri, J. A. (2000) Soil nematodes indicate food web responses to elevated atmospheric CO2. Pedobiologia, 44, 725735.Google Scholar
Hofmann, J., El Ashry, A., Anwar, S., et al. (2010) Metabolic profiling reveals local and systemic responses of host plants to nematode parasitism. The Plant Journal, 62, 10581071.Google Scholar
Hole, D. G., Perkins, A. J., Wilson, J. D., et al. (2005) Does organic farming benefit biodiversity? Biological Conservation, 122, 113130.Google Scholar
Holmstrup, M., Sørensen, J. G., Maraldo, K., et al. (2012) Increased frequency of drought reduces species richness of enchytraeid communities in both wet and dry heathland soils. Soil Biology & Biochemistry, 53, 4349.Google Scholar
Holt, J. A., Lepage, M. (2000) Termites and soil properties. In: Termites: Evolution, Sociality, Systematics, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 389407. Dordrecht, Kluwer Academic Press.Google Scholar
Holterman, M., Karegar, A., Mooijman, P., et al. (2017) Disparate gain and loss of parasitic abilities among nematode lineages. PloS ONE, 12, e0185445.Google Scholar
Hooper, D. U., Adair, C. E., Cardinale, B. J., et al. (2012) A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature, 486, 105108.Google Scholar
Hooper, D. U., Chapin, F. S. I., Ewel, J. J., et al. (2005) Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecological Monographs, 75, 335.Google Scholar
Hopkin, S. P. (1997) Biology of the Springtails (Insecta: Collembola), Oxford, Oxford University Press.Google Scholar
Hopkin, S. P. (2007) A Key to the Collembola (Springtails) of Britain and Ireland, Shropshire, FSC Publications.Google Scholar
House, A. P. N., Burwell, C. J., Brown, S. D., Walters, B. J. (2012) Agricultural matrix provides modest habitat value for ants on mixed farms in eastern Australia. Journal of Insect Conservation, 16, 112.Google Scholar
Howe, A. T., Bass, D., Vickerman, K., Chao, E. E., Cavalier-Smith, T. (2009) Phylogeny, taxonomy, and astounding genetic diversity of Glissomonadida ord. nov., the dominant gliding zooflagellates in soil (Protozoa: Cercozoa). Protist, 160, 159189.Google Scholar
Hughes, K. A., Greenslade, P., Convey, P. (2017) The fate of the non-native Collembolon, Hypogastrura viatica, at the southern extent of its introduced range in Antarctica. Polar Biology, 40, 21272131.Google Scholar
Huguenin, M. T., Leggett, C. G., Paterson, R. W. (2006) Economic valuation of soil fauna. European Journal of Soil Biology, 42, S16S22.Google Scholar
Huhta, V. (2007) The role of soil fauna in ecosystems: A historical review. Pedobiologia, 50, 489495.Google Scholar
Huhta, V., Persson, T., Setälä, H. (1998) Functional implications of soil fauna diversity in boreal forests. Applied Soil Ecology, 10, 277288.Google Scholar
Hunt, H. W., Coleman, D. C., Ingham, E. R., et al. (1987) The detrital food web in a shortgrass prairie. Biology and Fertility of Soils, 3, 5768.Google Scholar
Hunt, H. W., Wall, D. H. (2002) Modelling the effects of loss of soil biodiversity on ecosystem function. Global Change Biology, 8, 3350.Google Scholar
Huston, M. A. (1994) Biological Diversity, Cambridge, Cambridge University Press.Google Scholar
Hyodo, F., Tayasu, I., Konate, S., et al. (2008) Gradual enrichment of 15N with humification of diets in a below-ground food web: Relationship between 15N and diet age determined using 14C. Functional Ecology, 22, 513522.Google Scholar
Hyvönen, R., Persson, T. (1996) Effects of fungivorous and predatory arthropods on nematodes and tardigrades in microcosms with coniferous forest soil. Biology and Fertility of Soils, 21, 121127.Google Scholar
Ingham, R. E. (1988) Interactions between nematodes and vesicular-arbuscular mycorrhizae. Agriculture, Ecosystems and Environment, 24, 169182.Google Scholar
Ingham, R. E., Detling, J. K. (1984) Plant–herbivore interactions in a North American mixed-grass prairie. III. Soil nematode populations and root biomass on Cynomys ludovicianus colonies and adjacent uncolonized areas. Oecologia, 63, 307313.Google Scholar
Ingham, R. E., Trofymow, J. A., Ingham, E. R., Coleman, D. C. (1985) Interactions of bacteria, fungi, and their nematode grazers: Effects on nutrient cycling and plant growth. Ecological Monographs, 55, 119140.Google Scholar
Ingimarsdóttir, M., Caruso, T., Ripa, J., et al. (2012) Primary assembly of soil communities: Disentangling the effect of dispersal and local environment. Oecologia, 170, 745754.Google Scholar
IPCC (2014) Climate Change 2014 – Impacts, Adaptation and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge and New York, Cambridge University Press.Google Scholar
James, S. W. (1995) Systematics, biogeography, and ecology of nearctic earthworms from eastern, central, southern and southwestern United States. In: Earthworm Ecology and Biogeography in North America (ed. Hendrix, P. F.). Boca Raton, FL, Lewis Publishers.Google Scholar
Janion-Scheepers, C., Phillips, L., Sgrò, C. M., et al. (2018) Basal resistance enhances warming tolerance of alien over indigenous species across latitude. Proceedings of the National Academy of Sciences of the United States of America, 115, 145150.Google Scholar
Janion, C., Worland, M. R., Chown, S. L. (2009) Assemblage level variation in springtail lower lethal temperature: The role of invasive species on sub-Antarctic Marion Island. Physiological Entomology, 34, 284291.Google Scholar
Janssen, M. P. M., Heijmans, G. J. S. M. (1998) Dynamics and stratification of protozoa in the organic layer of a Scots pine forest. Biology and Fertility of Soils, 26, 285292.Google Scholar
Jeanne, R. L. (1979) A latitudinal gradient in rates of ant predation. Ecology, 60, 12111224.Google Scholar
Jeffery, S., Gardi, C., Jones, A., et al. (2010) European Atlas of Soil Biodiversity, Luxembourg, European Commission, Publications Office of the European Union.Google Scholar
Jenkins, C. N., Sanders, N. J., Andersen, A. N., et al. (2011) Global diversity in light of climate change: The case of ants. Diversity and Distributions, 17, 652662.Google Scholar
Jenny, H. (1941) Factors of Soil Formation: A System of Quantitative Pedology, New York, McGraw-Hill.Google Scholar
Jensen, T. C., Leinaas, H. P., Hessen, D. O. (2006) Age-dependent shift in response to food element composition in Collembola: Contrasting effects of dietary nitrogen. Oecologia, 149, 583592.Google Scholar
Jentschke, G., Bonkowski, M., Godbold, D. L., Scheu, S. (1995) Soil protozoa and forest tree growth: Non-nutritional effects and interaction with mycorrhizas. Biology and Fertility of Soils, 20, 263269.Google Scholar
Jiménez, J. J., Decaëns, T., Rossi, J. P. (2006) Stability of the spatio-temporal distribution and niche overlap in Neotropical earthworm assemblages. Acta Oecologica, 30, 299311.Google Scholar
Jiménez, J. J., Rossi, J. P. (2006) Spatial dissociation between two endogeic earthworms in the Colombian ‘Llanos’. European Journal of Soil Biology, 42, S218S224.Google Scholar
Johns, A. D. (1992) Species conservation in managed tropical forests. In: Tropical Deforestation and Species Extinction (eds. Whitmore, T. C., Sayer, J. A.) pp. 1550. London, Chapman & Hall.Google Scholar
Johnson, A. W., Nusbaum, C. J. (1968) The activity of Tylenchorhynchus claytoni, Trichodorus christiei, Pratylenchus brachyurus, P. zeae, and Helicotylenchus dihystera in single and multiple inoculations on corn and soybean. Nematologica, 14, 9.Google Scholar
Johnson, D., Krsek, M., Wellington, E. M. H., et al. (2003) Soil invertebrates disrupt carbon flow through fungal networks. Science, 309, 1047.Google Scholar
Johnson, S. N., Clark, K. E., Hartley, S. E., et al. (2012) Aboveground–belowground herbivore interactions: A meta-analysis. Ecology, 93, 22082215.Google Scholar
Johnson, S. N., Rasmann, S. (2015) Root-feeding insects and their interactions with organisms in the rhizosphere. Annual Review of Entomology, 60, 517535.Google Scholar
Jones, C. G., Lawton, J. H., Shachak, M. (1997) Positive and negative effects of organisms as physical ecosystem engineers. Ecology, 78, 19461957.Google Scholar
Jonsson, M., Yeates, G. W., Wardle, D. A. (2009) Patterns of invertebrate density and taxonomic richness across gradients of area, isolation, and vegetation diversity in a lake–island system. Ecography, 32, 963972.Google Scholar
Jørgensen, H. B., Canbäck, B., Hedlund, K., Tunlid, A. (2005) Selective foraging of fungi by collembolans in soil. Biology Letters, 1, 243246.Google Scholar
Jouquet, P., Dauber, J., Lagerlöf, J., Lavelle, P., Lepage, M. (2006) Soil invertebrates as ecosystem engineers: Intended and accidental effects on soil and feedback loops. Applied Soil Ecology, 32, 153164.Google Scholar
Judas, M. (1988) The species–area relationship of European Lumbricidae (Annelida, Oligochaeta). Oecologia, 76, 579587.Google Scholar
Juen, A., Traugott, M. (2006) Amplification facilitators and multiplex PCR: Tools to overcome PCR-inhibition in DNA-gut-content analysis of soil-living invertebrates. Soil Biology & Biochemistry, 38, 18721879.Google Scholar
Kalif, K. A. B., Azevedo-Ramos, C., Moutinho, P., Malcher, S. A. O. (2001) The effect of logging on the ground-foraging ant community in Eastern Amazonia. Studies on Neotropical Fauna and Environment, 36, 215219.Google Scholar
Kambhampati, S., Eggleton, P. (2000) Taxonomy and phylogeny of termites. In: Termites: Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 124. Dordrecht, Kluwer Academic.Google Scholar
Kampichler, C., Bruckner, A. (2009) The role of microarthropods in terrestrial decomposition: A meta-analysis of 40 years of litterbag studies. Biological Reviews, 84, 375389.Google Scholar
Kaneda, S., Kaneko, N. (2004) The feeding preference of a collembolan (Folsomia candida Willem) on ectomycorrhiza (Pisolithus tinctorius (Pers.)) varies with the mycelial growth condition and vitality. Applied Soil Ecology, 27, 15.Google Scholar
Kaneko, N. (1988) Feeding habits and cheliceral size of oribatid mites in cool temperate forest soils in Japan. Revue d'Ecologie et de Biologie du Sol, 25, 353363.Google Scholar
Kaneko, N., Salamanca, E. F. (1999) Mixed leaf litter effects on decomposition rates and soil microarthropod communities in an oak-pine stand in Japan. Ecological Research, 14, 131138.Google Scholar
Kapusta, P., Sobczyk, L., Rozen, A., Weiner, J. (2003) Species diversity and spatial distribution of enchytraeid communities in forest soils: Effects of habitat characteristics and heavy metal contamination. Applied Soil Ecology, 23, 187198.Google Scholar
Kapusta, P., Szarek-Łukaszewska, G. S., Stefanowicz, A. M. (2011) Direct and indirect effects of metal contamination on soil biota in a Zn–Pb post-mining and smelting area (S Poland). Environmental Pollution, 159, 15161522.Google Scholar
Karban, P. (1980) Periodical cicada nymphs impose periodical oak tree wood accumulation. Nature, 287, 326327.Google Scholar
Kardol, P., Bezemer, T. M., van der Putten, W. H. (2006) Temporal variation in plant–soil feedback controls succession. Ecology Letters, 9, 10801088.Google Scholar
Kardol, P., Newton, J. S., Bezemer, T. M., Maraun, M., van der Putten, W. H. (2009) Contrasting diversity patterns of soil mites and nematodes in secondary succession. Acta Oecologica, 35, 603609.Google Scholar
Kardol, P., Reynolds, W. N., Norby, R. J., Classen, A. T. (2011) Climate change effects on soil microarthropod abundance and community structure. Applied Soil Ecology, 47, 3744.Google Scholar
Kardol, P., Throop, H. L., Adkins, J., de Graaff, M.-A. (2016) A hierarchical framework for studying the role of biodiversity in soil food web processes and ecosystem services. Soil Biology & Biochemistry, 102, 3336.Google Scholar
Kardol, P., Wardle, D. A. (2010) How understanding aboveground–belowground linkages can assist restoration ecology. Trends in Ecology and Evolution, 25, 670679.Google Scholar
Kaspari, M., Ward, P. S., Yuan, M. (2004) Energy gradients and the geographic distribution of local ant diversity. Oecologia, 140, 407413.Google Scholar
Kaspari, M., Yuan, M., Alonso, L. (2003) Spatial grain and the causes of regional diversity gradients in ants. The American Naturalist, 161, 459477.Google Scholar
Kaufmann, R. (2001) Invertebrate succession on an alpine glacier foreland. Ecology, 82, 22612278.Google Scholar
Kaufmann, R., Fuchs, M., Gosterxeier, N. (2002) The soil fauna of an alpine glacier foreland: Colonization and succession. Arctic, Antarctic, and Alpine Research, 34, 242250.Google Scholar
Kaya, M., De Smet, W. H., Fontaneto, D. (2010) Survey of moss-dwelling bdelloid rotifers from middle Arctic Spitsbergen (Svalbard). Polar Biology, 33, 833842.Google Scholar
Keilin, D. (1959) The problem of anabiosis or latent life: History and current concept. Proceedings of the Royal Society B, 150, 167173.Google Scholar
Keith, A. M., Brooker, R. W., Ostler, G. H. R., et al. (2009) Strong impacts of belowground tree inputs on soil nematode trophic composition. Soil Biology & Biochemistry, 41, 10601065.Google Scholar
Kerfahi, D., Tripathi, B. M., Porazinska, D. L., et al. (2016) Do tropical rain forest soils have greater nematode diversity than High Arctic tundra? A metagenetic comparison of Malaysia and Svalbard. Global Ecology and Biogeography, 25, 716728.Google Scholar
Kerry, B. R. (2000) Rhizosphere interactions and the exploitation of microbial agents for the biological control of plant-parasitic nematodes. Annual Review of Phytopathology, 38, 423441.Google Scholar
Khalil, M. A., Janssens, T. K. S., Berg, M. P., van Straalen, N. M. (2009) Identification of metal-responsive oribatid mites in a comparative survey of polluted soils. Pedobiologia, 52, 207221.Google Scholar
Khan, M. R., Khan, M. W. (1998) Interactive effects of ozone and root-knot nematode on tomato. Agriculture, Ecosystems and Environment, 70, 97107.Google Scholar
King, K. L., Greenslade, P., Hutchinson, K. J. (1985) Collembolan associations in natural versus improved pastures of the New England tableland, NSW: Distribution of native and introduced species. Australian Journal of Ecology, 10, 421427.Google Scholar
Kitz, F., Steinwandter, M., Traugott, M., Seeber, J. (2015) Increased decomposer diversity accelerates and potentially stabilizes litter decomposition. Soil Biology & Biochemistry, 83, 138141.Google Scholar
Klironomos, J. N. (2002) Feedback with soil biota contributes to plant rarity and invasiveness in communities. Nature, 417, 6770.Google Scholar
Klironomos, J. N., Rillig, M. C., Allen, M. F. (1996) Below-ground microbial and microfauna response to Artemisia tridentata grown under elevated atmospheric CO2. Functional Ecology, 10, 527534.Google Scholar
Klironomos, J. N., Rillig, M. C., Allen, M. F. (1999) Designing belowground field experiments with the help of semi-variance and power analyses. Applied Soil Ecology, 12, 227238.Google Scholar
Konestabo, H. S., Michelsen, A., Holmstrup, M. (2007) Responses of springtail and mite populations to prolonged periods of soil freeze–thaw cycles in a subarctic ecosystem. Applied Soil Ecology, 36, 136146.Google Scholar
Korthals, G. W., Smilauer, P., Van Dijk, C., van der Putten, W. H. (2001) Linking above- and below-ground biodiversity: Abundance and trophic complexity in soil as a response to experimental plant communities on abandoned arable land. Functional Ecology, 15, 506514.Google Scholar
Korthals, G. W., van de Ende, A., van Megen, H., et al. (1996) Short-term effects of cadmium, copper, nickel and zinc on soil nematodes from different feeding and life-history strategy groups. Applied Soil Ecology, 4, 107117.Google Scholar
Krantz, G. W., Walter, D. E. (2009) A Manual of Acarology, 3rd edn, Lubbock, TX, Texas Tech University Press.Google Scholar
Kuikman, P. J., Jansen, A. G., van Veen, J. A., Zehnder, A. J. B. (1990) Protozoan predation and the turnover of soil organic carbon and nitrogen in the presence of plants. Biology and Fertility of Soils, 10, 2228.Google Scholar
Kulmatiski, A., Beard, K. H., Stevens, J. R., Cobbold, S. M. (2008) Plant–soil feedbacks: A meta-analytical review. Ecology Letters, 11, 980992.Google Scholar
Kuntz, M., Berner, A., Gattinger, A., et al. (2013) Influence of reduced tillage on earthworm and microbial communities under organic arable farming. Pedobiologia, 56, 251260.Google Scholar
Kupferschmied, P., Maurhof, M., Keel, C. (2013) Promise for plant pest control: Root-associated pseudomonads with insecticidal activities. Frontiers in Plant Science, 4, 287.Google Scholar
Kusnezov, N. (1957) Numbers of species of ants in faunae of different latitudes. Evolution, 11, 298299.Google Scholar
Kwon, T.-S., Kim, S.-S., Chun, J. H. (2014) Pattern of ant diversity in Korea: An empirical test of Rapoport's altitudinal rule. Journal of Asia-Pacific Entomology, 17, 161167.Google Scholar
Laakso, J., Setälä, H. (1999a) Population- and ecosystem-level effects of predation on microbial-feeding nematodes. Oecologia, 120, 279286.Google Scholar
Laakso, J., Setälä, H. (1999b) Sensitivity of primary production to changes in the architecture of belowground food webs. Oikos, 87, 5764.Google Scholar
Laakso, J., Setälä, H., Palojärvi, A. (2000) Influence of decomposer food web structure and nitrogen availability on plant growth. Plant and Soil, 225, 153165.Google Scholar
Lacey, L. A., Shapiro-Ilan, D. I. (2008) Microbial control of insect pests in temperate orchard systems: Potential for incorporation into IPM. Annual Review of Entomology, 53, 121144.Google Scholar
Lal, R. (2004) Soil carbon sequestration to mitigate climate change. Geoderma, 123, 122.Google Scholar
Laliberté, E., Kardol, P., Didham, R. K., et al. (2017) Soil fertility shapes belowground food webs across a regional climate gradient. Ecology Letters, 20, 12731284.Google Scholar
Lambshead, P. J. D. (1993) Recent developments in marine benthic biodiversity research. Oceanis, 19, 524.Google Scholar
Landesman, W. J., Treonis, A. T., Dighton, J. (2011) Effects of a one-year rainfall manipulation on soil nematode abundances and community composition. Pedobiologia, 54, 8791.Google Scholar
Lang, B., Rall, B. C., Scheu, S., Brose, U. (2014) Effects of environmental warming and drought on size-structured soil food webs. Oikos, 123, 12241233.Google Scholar
Langford, E. A., Nielsen, U. N., Johnson, S. N., Riegler, M. (2014) Susceptibility of Queensland fruit fly, Bactrocera tryoni (Froggatt) (Diptera: Tephritidae), to entomopathogenic nematodes. Biological Control, 69, 3439.Google Scholar
Lara, E., Heger, T. J., Ekelund, F., Lamentowicz, M., Mitchell, E. A. D. (2008) Ribosomal RNA genes challenge the monophyly of the Hyalospheniidae (Amoebozoa: Arcellinida). Protist, 159, 165176.Google Scholar
Larink, O. (1997) Springtails and mites: Important knots in the food web of soils. In: Fauna in Soil Ecosystems: Recycling Processes, Nutrient Fluxes, and Agricultural Production (ed. Benckiser, G.) pp. 225264. New York, Marcel Dekker.Google Scholar
Lavelle, P. (1983) The Structure of Earthworm Communities, London, Chapman & Hall.Google Scholar
Lavelle, P., Blanchart, E., Martin, A., Spain, A. V., Martin, S. (1992) Impact of Soil Fauna on the Properties of Soils in the Humid Tropics, Madison, Soil Science Society of America.Google Scholar
Lavelle, P., Decaëns, T., Aubert, M., et al. (2006) Soil invertebrates and ecosystem services. European Journal of Soil Biology, 42, S3S15.Google Scholar
Lavelle, P., Lapied, E. (2003) Endangered earthworms of Amazonia: And homage to Gilberto Righi. Pedobiologia, 47, 419427.Google Scholar
Lavelle, P., Lattaud, C., Trigo, D., Barois, I. (1995) Mutualism and biodiversity in soils. Plant and Soil, 170, 2333.Google Scholar
Lavelle, P., Pashanasi, B., Charpentier, F., et al. (1998) Influence of earthworms on soil organic matter dynamics, nutrient dynamics and microbiological ecology. In: Earthworm Ecology (ed. Edwards, C. A.) pp. 103122. Boca Raton, FL, CRC Press.Google Scholar
Lavelle, P., Spain, A. V. (2001) Soil Ecology, Amsterdam, Kluwer Scientific.Google Scholar
Lebauer, D. W., Treseder, K. K. (2008) Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology, 89, 371379.Google Scholar
Lee, K. E. (1987) Peregrine species of earthworms. In: On Earthworms (eds. Pagliai, A. M. B., Omodeo, P.) pp. 315328. Modena, Mucchi Editore.Google Scholar
Lee, K. E., Foster, R. C. (1991) Soil fauna and soil structure. Australian Journal of Soil Research, 29, 745775.Google Scholar
Legendre, P., De Cáceres, M. (2013) Beta diversity as the variance of community data: Dissimilarity coefficients and partitioning. Ecology Letters, 16, 951963.Google Scholar
Lehmann, J., Kleber, M. (2015) The contentious nature of soil organic matter. Nature, 60, 6068.Google Scholar
Lehmitz, R., Russell, D., Hohberg, K., Christian, A., Xylander, W. E. R. (2012) Active dispersal of oribatid mites into young soils. Applied Soil Ecology, 55, 1019.Google Scholar
Leinaas, H. P. (1983) Synchronized molting controlled by communication in group-living Collembola. Science, 219, 193195.Google Scholar
Leinaas, H. P., Bengtsson, J., Janion-Scheepers, C., Chown, S. L. (2015) Indirect effects of habitat disturbance on invasion: Nutritious litter from a grazing resistant plant favors alien over native Collembola. Ecology and Evolution, 5, 34623471.Google Scholar
Leinaas, H. P., Fjellberg, A. (1985) Habitat structure and life-history strategies of two partly sympatric and closely related, lichen feeding Collembolan species. Oikos, 44, 448458.Google Scholar
Leniaud, L., Dedeine, F., Pichon, A., Dupont, S., Bagneres, A. G. (2009) Geographical distribution, genetic diversity and social organization of a new European termite, Reticulitermes urbis (Isoptera: Rhinotermitidae). Biological Invasions, 12, 13891402.Google Scholar
Lenoir, L., Persson, T., Bengtsson, J., Wallander, H., Wirén, A. (2007) Bottom-up or top-down control in forest soil microcosms? Effects of soil fauna on fungal biomass and C/N mineralisation. Biology and Fertility of Soils, 43, 281294.Google Scholar
Lessaard, J. P., Fordyce, J. A., Gotelli, N. J., Sanders, N. J. (2009) Invasive ants alter the phylogenetic structure of ant communities. Ecology, 90, 26642669.Google Scholar
Letourneau, D. K., Jedlicka, J. A., Bothwell, S. G., Moreno, C. R. (2009) Effects of natural enemy biodiversity on the suppression of arthropod herbivores in terrestrial ecosystems. Annual Review of Ecology, Evolution and Systematics, 40, 573592.Google Scholar
Li, Q., Bao, X., Lu, C., et al. (2012) Soil microbial food web responses to free-air ozone enrichment can depend on the ozone-tolerance of wheat cultivars. Soil Biology & Biochemistry, 47, 2735.Google Scholar
Liang, W., Jiang, Y., Liu, Q. L. Y., Wen, D. (2005) Spatial distribution of bacterivorous nematodes in a Chinese Ecosystem Research Network (CERN) site. Ecological Research, 20, 491486.Google Scholar
Liao, C., Peng, R., Luo, Y., et al. (2008) Altered ecosystem carbon and nitrogen cycles by plant invasion: A meta-analysis. New Phytologist, 177, 706714.Google Scholar
Liiri, M., Häsä, M., Haimi, J., Setälä, H. (2012) History of land-use intensity can modify the relationship between functional complexity of the soil fauna and soil ecosystem services – A microcosm study. Applied Soil Ecology, 55, 5361.Google Scholar
Liiri, M., Setälä, H., Haimi, J., Pennanen, T., Fritze, H. (2002) Relationship between soil microarthropod species diversity and plant growth does not change when the system is disturbed. Oikos, 96, 137149.Google Scholar
Lindo, Z., Winchester, N. N. (2009) Spatial and environmental factors contributing to patterns in arboreal and terrestrial oribatid mite diversity across spatial scales. Oecologia, 160, 817825.Google Scholar
Liu, M., Chen, X., Griffiths, B. S., et al. (2012) Dynamics of nematode assemblages and soil function in adjacent restored and degraded soils following disturbance. European Journal of Soil Biology, 49, 3746.Google Scholar
Liu, T., Chen, X., Ran, W., Shen, Q., Li, H. (2016) Carbon-rich organic fertilizers to increase soil biodiversity: Evidence from a meta-analysis of nematode communities. Agriculture, Ecosystems and Environment, 232, 199207.Google Scholar
Lobry de Bruyn, L., Conacher, A. J. (1990) The role of termites and ants in soil modification: A review. Australian Journal of Soil Research, 28, 5593.Google Scholar
Long, S. P., Ainsworth, E. A., Rogers, A., Ort, D. R. (2004) Rising atmospheric carbon dioxide: Plants face the future. Annual Review of Plant Biology, 55, 591628.Google Scholar
Loof, P. A. A. (1971) Freeliving and plant parasitic nematodes from Spitzbergen, collected by H. van Rossen. Meded. Landbouwhogeschool Wageningen, 71, 186.Google Scholar
Loranger, G., Bandyopadhyaya, I., Razaka, B., Ponge, J. F. (2001) Does soil acidity explain altitudinal sequences in collembolan communities? Soil Biology & Biochemistry, 33, 381393.Google Scholar
Loranger, G., Ponge, J. F., Blanchart, E., Lavelle, P. (1998) Impact of earthworms on the diversity of microarthropods in a vertisol (Martinique). Biology and Fertility of Soils, 27, 2126.Google Scholar
Loranger, G. I., Pregitzer, K. S., King, J. S. (2004) Elevated CO2 and O3t concentrations differentially affect selected groups of the fauna in temperate forest soils. Soil Biology & Biochemistry, 36, 15211524.Google Scholar
Loreau, M. (1987) Vertical distribution of activity of carabid beetles in a beech forest floor. Pedobiologia, 30, 173178.Google Scholar
Loreau, M., Naeem, S., Inchausti, P., et al. (2001) Biodiversity and ecosystem functioning: Current knowledge and future challenges. Science, 294, 804808.Google Scholar
Luxton, M. (1967) The zonation of saltmarsh Acarina. Pedobiologia, 7, 257277.Google Scholar
Luxton, M. (1972) Studies on the oribatid mites of a Danish beech wood soil. I. Nutritional biology. Pedobiologia, 12, 434463.Google Scholar
Luxton, M. (1982) The biology of mites from beech woodland soil. Pedobiologia, 23, 18.Google Scholar
Lyford, W. H. (1975) Overland migration of Collembola (Hypogastrura nivicola Fitch) colonies. American Midland Naturalist, 94, 205209.Google Scholar
Lynch, J. M., Whipps, J. M. (1990) Substrate flow in the rhizosphere. Plant and Soil, 129, 110.Google Scholar
Maaß, S., Caruso, T., Rillig, M. C. (2015) Functional role of microarthropods in soil aggregation. Pedobiologia, 58, 5963.Google Scholar
Maaß, S., Migliorini, M., Rillig, M. C., Caruso, T. (2014) Disturbance, neutral theory, and patterns of beta diversity in soil communities. Ecology and Evolution, 4, 47664774.Google Scholar
Maboreke, H. R., Graf, M., Grams, T. E. E., et al. (2017) Multitrophic interactions in the rhizosphere of a temperate forest tree affect plant carbon flow into the belowground food web. Soil Biology & Biochemistry, 115, 526536.Google Scholar
Macarthur, R. H., Wilson, E. O. (1967) The Theory of Island Biogeography, Princeton, NJ, Princeton University Press.Google Scholar
Macfadyen, A. (1962) Soil arthropod sampling. Advances in Ecological Research, 1, 134.Google Scholar
Macfadyen, A. (1963) The contribution of the fauna to the total soil metabolism. In: Soil Organisms (eds. Doeksen, J., Van Der Drift, J.) pp. 317. Amsterdam, North Holland Publishing.Google Scholar
Macnamara, C. (1924) The food of Collembola. Canadian Entomology, 56, 99105.Google Scholar
Madej, G., Stodółka, A. (2008) Successional changes and diversity of mesostigmatid mite communities (Acari: Mesostigmata) on reclaimed power plant waste dumps. Annales Zoologici, 58, 267278.Google Scholar
Magurran, A. E. (2004) Measuring Biological Diversity, Oxford, Blackwell.Google Scholar
Majer, J. D., Brennan, K. E. C., Moir, M. L. (2007) Invertebrates and the restoration of a forest ecosystem: 30 years of research following bauxite mining in Western Australia. Restoration Ecology, 15, S104S115.Google Scholar
Mando, A., Brussaard, L., Stroosnijder, L. (1999) Termite- and mulch-mediated rehabilitation of vegetation on crusted soil in West Africa. Restoration Ecology, 7, 3341.Google Scholar
Mando, A., Stroosnijder, L., Brussaard, L. (1996) Effects of termites on infiltration into crusted soil. Geoderma, 74, 107113.Google Scholar
Manning, P. (2012) The impact of nitrogen enrichment on ecosystems and their services. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 256269. Oxford, Oxford University Press.Google Scholar
Manning, P., Newington, J. E., Robson, H. R., et al. (2006) Decoupling the direct and indirect effects of nitrogen deposition on ecosystem function. Ecology Letters, 9, 10151024.Google Scholar
Maraun, M., Alphei, J., Bonkowski, M., et al. (1999) Middens of the earthworm Lumbricus terrestris (Lumbricidae): Microhabitats for micro- and mesofauna in forest soil. Pedobiologia, 43, 276287.Google Scholar
Maraun, M., Schatz, H., Scheu, S. (2007) Awesome or ordinary? Global diversity patterns of oribatid mites. Ecography, 30, 209216.Google Scholar
Marichal, R., Grimaldi, M., Feijoo, M. A., et al. (2014) Soil macroinvertebrate communities and ecosystem services in deforested landscapes of Amazonia. Applied Soil Ecology, 83, 177185.Google Scholar
Mariotte, P., Le Bayon, R.-C., Eisenhauer, N., Guenat, C., Buttler, A. (2016) Subordinate plant species moderate drought effects on earthworm communities in grasslands. Soil Biology & Biochemistry, 96, 119127.Google Scholar
Mariotte, P., Mehrabi, Z., Bezemer, T. M., et al. (2018) Plant–soil feedback: Bridging natural and agricultural sciences. Trends in Ecology and Evolution, 33, 129142.Google Scholar
Maron, J. L., Klironomos, J. N., Waller, L., Callaway, R. A. (2014) Invasive plants escape from suppressive soil biota at regional scales. Journal of Ecology, 102, 1927.Google Scholar
Marshall, D. J., Pugh, P. J. A. (1996) Origin of the inland Acari of Continental Antarctica with particular reference to Dronning Maud Land. Zoological Journal of the Linnean Society, 118, 101118.Google Scholar
Martius, C., Höfer, H., Garcia, M. V. B., et al. (2004) Microclimate in agroforestry systems in central Amazonia: Does canopy closure matter to soil organisms? Agroforestry Systems, 60, 291304.Google Scholar
Masters, G. J., Brown, V. K., Gange, A. C. (1993) Plant mediated interactions between above- and below-ground insect herbivores. Oikos, 66, 148151.Google Scholar
Masters, G. J., Jones, T. H., Rogers, M. (2001) Host-plant mediated effects of root herbivory on insect seed predators and their parasitoids. Oecologia, 127, 246250.Google Scholar
Mathieu, J., Davies, T. J. (2014) Glaciation as an historical filter of below-ground biodiversity. Journal of Biogeography, 41, 12041214.Google Scholar
Mathieu, J., Grimaldi, M., Jouquet, P., et al. (2009) Spatial patterns of grasses influence soil macrofauna biodiversity in Amazonian pastures. Soil Biology & Biochemistry, 41, 586593.Google Scholar
Mathieu, J., Rossi, J. P., Mora, P., et al. (2005) Recovery of soil macrofauna communities after forest clearance in Eastern Amazonia, Brazil. Conservation Biology, 19, 15981605.Google Scholar
Matthews, J. A., Vater, A. E. (2015) Pioneer zone geo-ecological change: Observations from a chronosequence on the Storbreen glacier foreland, Jotunheimen, southern Norway. Catena, 135, 219230.Google Scholar
McCary, M. A., Martínez, J.-C., Umek, L., Heneghan, L., Wise, D. H. (2015) Effects of woodland restoration and management on the community of surface-active arthropods in the metropolitan Chicago region. Biological Conservation, 190, 154166.Google Scholar
McCary, M. A., Mores, R., Farfan, M. A., Wise, D. H. (2016) Invasive plants have different effects on trophic structure of green and brown food webs in terrestrial ecosystems: A meta-analysis. Ecology Letters, 19, 328335.Google Scholar
McGaughran, A., Stevens, M. I., Holland, B. A. (2010) Biogeography of circum-Antarctic springtails. Molecular Phylogenetics and Evolution, 57, 4858.Google Scholar
McGill, B. J., Enquist, B. J., Weiher, E., Westoby, M. (2006) Rebuilding community ecology from functional traits. Trends in Ecology and Evolution, 21, 178185.Google Scholar
McInnis, S. J., Pugh, P. J. A. (2007) An attempt to revisit the global biogeography of limno-terrestrial Tardigrada. Journal of Limnology, 66 (Suppl. 1), 9096.Google Scholar
Meehan, T. D., Crossley, M. S., Lindroth, R. L. (2010) Impacts of elevated CO2 and O3 on aspen leaf litter chemistry and earthworm and springtail productivity. Soil Biology & Biochemistry, 42, 11321137.Google Scholar
Meloni, F., Varanda, E. M. (2015) Litter and soil arthropod colonization in reforested semi-deciduous seasonal Atlantic forests. Restoration Ecology, 23, 690697.Google Scholar
Middleton, E. L., Bever, J. D. (2012) Inoculation with a native soil community advances succession in a grassland restoration. Restoration Ecology, 20, 218226.Google Scholar
Migliorini, M., Pigino, G., Caruso, T., et al. (2005) Soil communities (Acari Oribatida; Hexapoda Collembola) in a clay pigeon shooting range. Pedobiologia, 49, 113.Google Scholar
Mikola, J., Bardgett, R. D., Hedlund, K. (2002) Biodiversity, ecosystem functioning and soil decomposer food webs. In: Biodiversity and Ecosystem Functioning – Synthesis and Perspectives (eds. Loreau, M., Naeem, S., Inchausti, P.) pp. 169180. Oxford, Oxford University Press.Google Scholar
Mikola, J., Setälä, H. (1998) No evidence of trophic cascades in an experimental microbial-based soil food web. Ecology, 79, 153164.Google Scholar
Milchunas, D. G., Mosier, A. R., Morgan, J. A., et al. (2005) Root production and tissue quality in a shortgrass steppe exposed to elevated CO2: Using a new ingrowth method. Plant and Soil, 268, 111122.Google Scholar
Milcu, A., Manning, P. (2011) All size classes of soil fauna and litter quality control the acceleration of litter decay in its home environment. Oikos, 120, 13661370.Google Scholar
Minor, M. A., Cianciolo, J. M. (2007) Diversity of soil mites (Acari: Oribatida, Mesostigmata) along a gradient of land use types in New York. Applied Soil Ecology, 35, 140153.Google Scholar
Mitchell, E. A. D., Borcard, D., Buttler, A. J., et al. (2000) Horizontal distribution patterns of testate amoeba (Protozoa) in a Sphagnum magellanicum carpet. Microbial Ecology, 39, 290300.Google Scholar
Mitchell, E. A. D., Meisterfeld, R. (2005) Taxonomic confusion blurs the debate on cosmopolitanism versus local endemism of free-living protists. Protist, 156, 263267.Google Scholar
Mitchell, R. J., Urpeth, H. M., Britton, A. J., Black, H., Taylor, A. R. (2016) Relative importance of local- and large-scale drivers of alpine soil microarthropod communities. Oecologia, 913924.Google Scholar
Mittelbach, G. G., Steiner, C. F., Scheiner, S. M., et al. (2001) What is the observed relationship between species richness and productivity? Ecology, 82, 23812396.Google Scholar
Monroy, F., van der Putten, W. H., Yergeau, E., et al. (2012) Community patterns of soil bacteria and nematodes in relation to geographic distance. Soil Biology & Biochemistry, 45, 17.Google Scholar
Moreau, C. S., Bell, C. D., Vila, R., Archibald, S. B., Pierce, N. E. (2006) Phylogeny of the ants: Diversification in the age of angiosperms. Science, 312, 101104.Google Scholar
Moroenyane, I., Dong, K., Singh, D., Chimphango, S. B. M., Adams, J. M. (2016) Deterministic processes dominate nematode community structure in the Fynbos Mediterranean heathland of South Africa. Evolutionary Ecology, 30, 685701.Google Scholar
Morriën, E., Hannula, S. E., Snoek, L. B., et al. (2017) Soil networks become more connected and take up more carbon as nature restoration progresses. Nature Communications, 8, 14349, doi:10.1038/ncomms14349.Google Scholar
Mulder, C., Ahrestani, F. S., Bahn, M., et al. (2013) Connecting the green and brown worlds: Allometric and stoichiometric predictability of above- and below-ground networks. Advances in Ecological Research, 49, 69175.Google Scholar
Mulder, C., Baerselman, R., Posthuma, C. (2007) Empirical maximum lifespan of earthworms is twice that of mice. Age (Dordr), 29, 229231.Google Scholar
Mulder, C., Boit, A., Bonkowski, M., et al. (2011) A belowground perspective on Dutch agroecosystems: How soil organisms interact to support ecosystem services. Advances in Ecological Research, 44, 277357.Google Scholar
Mulder, C., Hettelingh, J.-P., Montanarella, L., et al. (2015) Chemical footprints of anthropogenic nitrogen deposition on recent soil C:N ratios in Europe. Biogeosciences, 12, 41134119.Google Scholar
Murphy, D. H. (1955) Long-term changes in collembolan populations with special reference to moorland soils. In: Soil Zoology (ed. Kevan, D. K. McE) pp. 157166. London, Buttersworth.Google Scholar
Murray, P. J., Hatch, D. J. (1994) Sitona weevils (Coleoptera, Curculionidae) as agents for rapid transfer of nitrogen from white clover (Trifolium repens L.) to perennial ryegrass (Lolium perenne L.). Annals of Applied Biology, 125, 2933.Google Scholar
Neher, D. A., Weicht, T. R. (2013) Nematode genera in forest soil respond differentially to elevated CO2. Journal of Nematology, 45, 214222.Google Scholar
Neidig, N., Jousset, A., Nunes, F., et al. (2010) Interference between bacterial feeding nematodes and amoebae relies on innate and inducible mutual toxicity. Functional Ecology, 24, 11331138.Google Scholar
Neufeld, J. D., Wagner, M., Murrell, J. C. (2009) Who eats what, where and when? Isotope-labelling experiments are coming of age. The ISME Journal, 1, 103110.Google Scholar
Newington, J. E., Setälä, H., Bezemer, T. M., Jones, T. H. (2004) Potential effects of earthworms on leaf-chewer performance. Functional Ecology, 18, 746751.Google Scholar
Ngosong, C., Gabriel, E., Ruess, L. (2014) Collembola grazing on arbuscular mycorrhiza fungi modulates nutrient allocation in plants. Pedobiologia, 57, 171179.Google Scholar
Ngosong, C., Raupp, J., Richnow, H.-H., Ruess, L. (2011) Tracking Collembola feeding strategies by the natural C-13 signal of fatty acids in an arable soil with different fertilizer regimes. Pedobiologia, 54, 225233.Google Scholar
Nicholas, W. L., Viswanathan, S. (1975) A study of the nutrition of Caenorhabditis briggsae (Rhabditidae) fed Oo 14C and 32P-labelled bacteria. Nematologica, 21, 385400.Google Scholar
Nielsen, U. N., Ayres, E., Wall, D. H., Bardgett, R. D. (2011a) Soil biodiversity and carbon cycling: A synthesis of studies examining diversity–function relationships. European Journal of Soil Science, 62, 105116.Google Scholar
Nielsen, U. N., Ayres, E., Wall, D. H., et al. (2014) Global-scale patterns of assemblage structure of soil nematodes in relation to climate and ecosystem properties. Global Ecology and Biogeography, 23, 968978.Google Scholar
Nielsen, U. N., Ball, B. A. (2015) Impacts of altered precipitation regimes on soil communities and biogeochemistry in arid and semi-arid ecosystems. Global Change Biology, 21, 14071421.Google Scholar
Nielsen, U. N., Gilarte, P., Ochoa-Hueso, R., et al. (2016) Effects of altered precipitation patterns on soil fauna in an Australian grassland. In: Invertebrate Ecology of Australasian Grasslands. Proceedings of the Ninth ACGIE (ed. Johnson, S. N.). Penrith, Western Sydney University.Google Scholar
Nielsen, U. N., Osler, G. H. R., Campbell, C. D., Burslem, D. F. R. P., van der Wal, R. (2010a) The influence of vegetation type, soil properties and precipitation on the composition of soil mite and microbial communities at the landscape scale. Journal of Biogeography, 37, 13171328.Google Scholar
Nielsen, U. N., Osler, G. H. R., Campbell, C. D., Burslem, D. F. R. P., van der Wal, R. (2012a) Predictors of fine-scale spatial variation in soil mite and microbe community composition differ between biotic groups and habitats. Pedobiologia, 55, 8391.Google Scholar
Nielsen, U. N., Osler, G. H. R., Campbell, C. D., et al. (2010b) The enigma of soil animal species diversity revisited: The role of small-scale heterogeneity. PloS ONE, 5, e11567.Google Scholar
Nielsen, U. N., Osler, G. H. R., van der Wal, R., Campbell, C. D., Burslem, D. F. R. P. (2008) Soil pore volume and the abundance of soil mites in two contrasting habitats. Soil Biology & Biochemistry, 40, 15381541.Google Scholar
Nielsen, U. N., Prior, S., Delroy, B., et al. (2015a) Response of belowground communities to short-term phosphorus addition in a phosphorus-limited woodland. Plant and Soil, 391, 321331.Google Scholar
Nielsen, U. N., Wall, D. H. (2013) The future of soil invertebrate communities in polar regions: Different climate change responses in the Arctic and Antarctic? Ecology Letters, 16, 409419.Google Scholar
Nielsen, U. N., Wall, D. H., Adams, B. J., Virginia, R. A. (2011b) Antarctic nematode communities: Observed and predicted responses to climate change. Polar Biology, 34, 17011711.Google Scholar
Nielsen, U. N., Wall, D. H., Adams, B. J., et al. (2012b) The ecology of pulse events: Insights from an extreme climatic event in a polar desert ecosystem. Ecosphere, 3, Article 17.Google Scholar
Nielsen, U. N., Wall, D. H., Li, G., Toro, M., Adams, B. J., Virginia, R. A. (2011c) Nematode communities of Byers Peninsula, Livingston Island, maritime Antarctica. Antarctic Science, 23, 349357.Google Scholar
Nielsen, U. N., Wall, D. H., Six, J. (2015b) Soil biodiversity and the environment. Annual Review of Environment and Resources, 40, 6390.Google Scholar
Niemelä, J., Haila, Y., Halme, E., Pajunen, T., Punttila, P. (1992) Small-scale heterogeneity in the spatial distribution of carabid beetles in the southern Finnish taiga. Journal of Biogeography, 19, 173181.Google Scholar
Niklaus, P. A., Alphei, J., Ebersberger, D., et al. (2003) Six years of in situ CO2 enrichment evoke changes in soil structure and soil biota of nutrient-poor grassland. Global Change Biology, 9, 585600.Google Scholar
Nkem, J. N., Wall, D. H., Virginia, R. A., et al. (2006) Wind dispersal of soil invertebrates in the McMurdo Dry Valleys, Antarctica. Polar Biology, 29, 346352.Google Scholar
Noble, J. C., Whitford, W. G., Kaliszweski, M. (1996) Soil and litter microarthropod populations from two contrasting ecosystems in semi-arid eastern Australia. Journal of Arid Environments, 32, 329346.Google Scholar
Norton, D. C. (1978) Ecology of Plant-Parasitic Nematodes, New York, John Wiley.Google Scholar
Nowak, E. (2001) Enchytraeid communities in successional habitats (from meadow to forest). Pedobiologia, 45, 497508.Google Scholar
Nowrouzi, S., Andersen, A. N., Macfadyen, S., et al. (2016) Ant diversity and distribution along elevation gradients in the Australian wet tropics: The importance of seasonal moisture stability. PloS ONE, 11 (4), e0153420, doi:10.1371/journal.pone.0153420Google Scholar
O'Connor, F. B. (1967) The Enchytraeidae. In: Soil Biology (eds. Burges, A., Raw, F.) pp. 213257. London and New York, Academic Press.Google Scholar
O'Dowd, D. J., Green, P. T., Lake, P. S. (2003) Invasional meltdown on an oceanic island. Ecology Letters, 6, 812817.Google Scholar
Oerke, E.-C. (2006) Crop losses to pests. Journal of Agricultural Science, 144, 3143.Google Scholar
Oliver, I., Garden, D., Greenslade, P. J., et al. (2005) Effects of fertilizer and grazing on the arthropod communities of a native grassland in South-Eastern Australia. Agriculture, Ecosystems & Environment, 109, 323334.Google Scholar
Orgiazzi, A., Bardgett, R. D., Barrios, E., et al. (2016) Global Soil Biodiversity Atlas, Luxembourg, Publications Office of the European Union.Google Scholar
Orgiazzi, A., Dunbar, M. B., de Groot, G. A., Lemanceau, P. (2015) Soil biodiversity and DNA barcodes: Opportunities and challenges. Soil Biology & Biochemistry, 80, 244250.Google Scholar
Osler, G. H. R., Beattie, A. J. (1999) Taxonomic and structural similarities in soil oribatid communities. Ecography, 22, 567574.Google Scholar
Osler, G. H. R., Beattie, A. J. (2001) Contribution of oribatid and mesostigmatid soil mites in ecologically based estimates of global species richness. Austral Ecology, 26, 7079.Google Scholar
Osler, G. H. R., Sommerkorn, M. (2007) Toward a complete soil C and N cycle: Incorporating the soil fauna. Ecology, 88, 16111621.Google Scholar
Ostle, N., Briones, M. J. I., Ineson, P., et al. (2007) Isotopic detection of recent photosynthate carbon flow into grassland rhizosphere fauna. Soil Biology & Biochemistry, 39, 768777.Google Scholar
Ostle, N. J., Smith, P., Fisher, R., et al. (2009) Integrating plant–soil interactions into global carbon cycle models. Journal of Ecology, 97, 851863.Google Scholar
Ouédraogo, E., Mando, A., Brussaard, L. (2006) Soil macrofauna affect crop water and nitrogen use efficiencies in semi-arid West Africa. European Journal of Soil Biology, 42 (Suppl. 1), S275S277.Google Scholar
Paoletti, M. G. (1999) The role of earthworms for assessment of sustainability and as bioindicators. Agriculture, Ecosystems and Environment, 74, 137155.Google Scholar
Parmelee, R. W., Bohlen, P. J., Blair, J. M. (1998) Earthworms and nutrient cycling processes: Integrating across the ecological hierarchy. In: Earthworm Ecology (ed. Edwards, C. A.) pp. 123143. Boca Raton, FL, St. Lucie Press.Google Scholar
Parnikoza, I., Convey, P., Dykyy, I., et al. (2009) Current status of the Antarctic herb tundra formation in the Central Argentine Islands. Global Change Biology, 15, 16851693.Google Scholar
Parr, C. L., Robertson, H. G., Biggs, H. C., Chown, S. L. (2004) Response of African savanna ants to long-term fire regimes. Journal of Applied Ecology, 41, 630642.Google Scholar
Patten, B. C., Witkamp, M. (1967) Systems analysis of 134cesium kinetics in terrestrial microcosms. Ecology, 48, 813824.Google Scholar
Pauli, N., Barrios, E., Conacher, A. J. (2011) Soil macrofauna in agricultural landscapes dominated by the Quesungual slash-and mulch agroforestry system, western Honduras. Applied Soil Ecology, 47, 119132.Google Scholar
Paustian, K., Collins, H. P., Paul, E. A. (1997) Management controls on soil carbon. In: Soil Organic Matter in Temperate Agroecosystems (ed. Paul, E. A.) pp. 1549. Boca Raton, FL, CRC Press.Google Scholar
Pavao-Zuckerman, M. A., Sookhdeo, C. (2017) Nematode community response to green infrastructure design in a semiarid city. Journal of Environmental Quality, 46, 687694.Google Scholar
Peltzer, D. A., Bellingham, P. J., Kurokawa, H., et al. (2009) Punching above their weight: Low-biomass non-native plant species alter soil properties during primary succession. Oikos, 118, 10011014.Google Scholar
Peltzer, D. A., Wardle, D. A., Allison, V. J., et al. (2010) Understanding ecosystem retrogression. Ecological Monographs, 80, 509529.Google Scholar
Pendall, E., Bridgham, S., Hanson, P. J., et al. (2004a) Below-ground process responses to elevated CO2 and temperature: A discussion of observations, measurement methods, and models. New Phytologist, 162, 311322.Google Scholar
Pendall, E., Mosier, A. R., Morgan, J. A. (2004b) Rhizodeposition stimulated by elevated CO2 in a semiarid grassland. New Phytologist, 162, 447458.Google Scholar
Penev, L. D. (1992) Qualitative and quantitative spatial variation in soil wire-worm assemblages in relation to climate and habitat factors. Oikos, 63, 180192.Google Scholar
Peñuelas, J., Sardans, J., Rivas-Ubach, A., Janssens, I. A. (2012) The human-induced imbalance between C, N and P in Earth's life system. Global Change Biology, 18, 36.Google Scholar
Perry, R. N. (1996) Chemoreception in plant parasitic nematodes. Annual Review of Phytopathology, 34, 181199.Google Scholar
Persson, T., Bååth, E., Clarholm, M., et al. (1980) Trophic structure, biomass dynamics and carbon metabolism in a Scots pine forest. Ecological Bulletins, 32, 419459.Google Scholar
Petersen, H., Krogh, P. H. (1987) Effects of perturbing microarthropod communities of a permanent pasture and a rye field by an insecticide and a fungicide. In: Soil Fauna and Soil Fertility (ed. Strignova, B. R.) pp. 217229. Moscow, Proceedings of the 9th International Colloquium on Soil Zoology.Google Scholar
Petersen, H., Luxton, M. (1982) A comparative-analysis of soil fauna populations and their role in decomposition processes. Oikos, 39, 287388.Google Scholar
Petzold-Maxwell, J. L., Jaronski, S. T., Clifton, E. H., et al. (2013) Interactions among Bt maize, entomopathogens, and rootworm species (Coleoptera: Chrysomelidae) in the field: Effects on survival, yield, and root injury. Journal of Economic Entomology, 106, 622632.Google Scholar
Pey, B., Nahmani, J., Auclerc, A., et al. (2014) Current use of and future needs for soil invertebrate functional traits in community ecology. Basic and Applied Ecology, 15, 194206.Google Scholar
Pfeiffer, M., Chimedregzen, L., Ulykpan, K. (2003) Community organization and species richness of ants (Hymenoptera/Formicidae) in Mongolia along an ecological gradient from steppe to Gobi desert. Journal of Biogeography, 30, 19211935.Google Scholar
Phillips, R. P., Finzi, A. C., Bernhardt, E. S. (2011) Enhanced root exudation induces microbial feedbacks to N cycling in a pine forest under long-term CO2 fumigation. Ecology Letters, 14, 187194.Google Scholar
Pieterse, C. M. J., Zamioudis, R. L. B., Weller, D. M., Van Wees, S. C. M., Bakker, P. A. H. M. (2014) Induced systemic resistance by beneficial microbes. Annual Review of Phytopathology, 52, 347375.Google Scholar
Pilato, G., Binda, M. G. (2001) Biogeography and limno-terrestrial tardigrades: Are they truly incompatible binomials? Zoologisher Anzeigler, 240, 511516.Google Scholar
Poff, N. (1997) Landscape filters and species traits: Towards mechanistic understanding and prediction in stream ecology. Journal of the North American Benthological Society, 16, 391409.Google Scholar
Poll, J., Marhan, S., Haase, S., et al. (2007) Low amounts of herbivory by root-knot nematodes affect microbial community dynamics and carbon allocation in the rhizosphere. FEMS Microbiology Ecology, 62, 268279.Google Scholar
Pollierer, M. M., Langel, R., Körner, C., Maraun, M., Scheu, S. (2007) The underestimated importance of belowground carbon input for forest soil animal food webs. Ecology Letters, 10, 729736.Google Scholar
Pollierer, M. M., Langel, R., Scheu, S., Maraun, M. (2009) Compartmentalization of the soil animal food web as indicated by dual analysis of stable isotope ratios (15N/14N and 13C/12C). Soil Biology & Biochemistry, 41, 12211226.Google Scholar
Ponge, J.-F. (2003) Humus forms in terrestrial ecosystems: A framework to biodiversity. Soil Biology & Biochemistry, 35, 935945.Google Scholar
Ponge, J.-F., Dubs, F., Gillet, S., Sousa, J. P., Lavelle, P. (2006) Decreased biodiversity in soil springtail communities: The importance of dispersal and landuse history in heterogeneous landscapes. Soil Biology & Biochemistry, 38, 11581161.Google Scholar
Ponge, J.-F., Pérès, G., Guernion, M., et al. (2013) The impact of agricultural practices on soil biota: A regional study. Soil Biology & Biochemistry, 67, 271284.Google Scholar
Pop, V. V. (1998) Earthworm biology and ecology – A case study: The genus Octodrilus omodeo, 1956 (Oligochaeta, Lumbricidae), from the Carpathians. In: Earthworm Ecology (ed. Edwards, C. A.) pp. 65103. Boca Raton, FL, St. Lucie Press.Google Scholar
Popp, E. (1962) Semiaquatile Lebensräume (Bülten) im Hoch- und Niedermooren II. Internationale Revue der gesamten Hydrobiologie, 47, 533579.Google Scholar
Porazinska, D. L., Giblin-Davis, R. M., Faller, L., et al. (2009) Evaluating high-throughput sequencing as a method for metagenomic analysis of nematode diversity. Molecular Ecology Resources, 9, 14391450.Google Scholar
Porazinska, D. L., Giblin-Davis, R. M., Sequivel, A., et al. (2010) Ecometagenetics confirm high tropical rainforest nematode diversity. Molecular Ecology, 19, 55215530.Google Scholar
Porco, D., Decaëns, T., Deharveng, L., et al. (2013) Biological invasions in soil: DNA barcoding as a monitoring tool in a multiple taxa survey targeting European earthworms and springtails in North America. Biological Invasions, 15, 899910.Google Scholar
Post, E., Forchhammer, M. C., Bret-Harte, M. S., et al. (2009) Ecological dynamics across the Arctic associated with recent climate change. Science, 325, 13551358.Google Scholar
Postma-Blaauw, M. B., Bloem, J., Faber, J. H., et al. (2006) Earthworm species composition affects the soil bacterial community and net nitrogen mineralization. Pedobiologia, 50, 243256.Google Scholar
Postma-Blaauw, M. B., de Goede, R. G. M., Bloem, J., Faber, J. H., Brussaard, L. (2010) Soil biota community structure and abundance under agricultural intensification and extensification. Ecology, 91, 460473.Google Scholar
Postma-Blaauw, M. B., de Goede, R. G. M., Bloem, J., Faber, J. H., Brussaard, L. (2012) Agricultural intensification and de-intensification differentially affect taxonomic diversity of predatory mites, earthworms, enchytraeids, nematodes and bacteria. Applied Soil Ecology, 57, 3949.Google Scholar
Postma-Blaauw, M. B., de Vries, F. T., de Goede, R. G. M., et al. (2005) Within-trophic group interactions of bacterivorous nematode species and their effects on the bacterial community and nitrogen mineralization. Oecologia, 142, 428439.Google Scholar
Poveda, K., Steffan-Dewenter, I., Scheu, S., Tscharntke, T. (2005) Effects of decomposers and herbivores on plant performance and aboveground plant–insect interactions. Oikos, 108, 503510.Google Scholar
Poveda, K., Steffan-Dewenter, I., Scheu, S., Tscharntke, T. (2006) Belowground effects of organic and conventional farming on aboveground plant–herbivore and plant–pathogen interactions. Agriculture, Ecosystems & Environment, 113, 162167.Google Scholar
Powers, L. E., Ho, M. C., Freckman, D. W., Virginia, R. A. (1998) Distribution, community structure, and microhabitats of soil invertebrates along an elevational gradient in Taylor Valley, Antarctica. Arctic and Alpine Research, 30, 133141.Google Scholar
Prasse, I. (1989) Indications of structural changes in the communities of microarthropods of the soil in an agro-ecosystem after applying herbicides. Agriculture, Ecosystems & Environment, 13, 205215.Google Scholar
Procter, D. L. C. (1984) Towards a biogeography of free-living soil nematodes. 1. Changing species richness, diversity and densities with latitude. Journal of Biogeography, 11, 103117.Google Scholar
Procter, D. L. C. (1990) Global overview of the functional roles of soil-living nematodes in terrestrial communities and ecosystems. Journal of Nematology, 22, 17.Google Scholar
Prot, J.-C. (1980) Migration of plant parasitic nematodes toward plant roots. Revue Nématology, 3, 305318.Google Scholar
Puppe, D., Ehrmann, O., Kaczorek, D., Wanner, M., Sommer, M. (2015) The protozoic Si pool in temperate forest ecosystems – Quantification, abiotic controls and interactions with earthworms. Geoderma, 243–244, 196204.Google Scholar
Purtauf, T., Roschewitz, I., Dauber, J., et al. (2005) Landscape context of organic and conventional farms: Influences on carabid beetle diversity. Agriculture, Ecosystems & Environment, 108, 165174.Google Scholar
Qiu, J. J., Westerdahl, B. B., Pryor, A. (2009) Reduction of root-knot nematode, Meloidogyne javanica, and ozone mass transfer in soil treated with ozone. Journal of Nematology, 41, 241246.Google Scholar
Qiu, Q., Wu, J., Liang, G., et al. (2015) Effects of simulated acid rain on soil and soil solution chemistry in a monsoon evergreen broadleaved forest in southern China. Environmental Monitoring and Assessment, 187, 272.Google Scholar
Quist, C. W., Gort, G., Mulder, C., et al. (2017) Feeding preference as a main determinant of microscale patchiness among terrestrial nematodes. Molecular Ecology Resources, 17, 12571270.Google Scholar
Quist, C. W., Smant, G., Helder, J. (2015) Evolution of plant parasitism in the Phylum Nematoda. Annual Review of Phytopathology, 53, 289310.Google Scholar
Rahbek, C. (1995) The elevational gradient of species richness: A uniform pattern? Ecography, 18, 200205.Google Scholar
Rahbek, C. (2005) The role of spatial scale and the perception of large-scale species-richness patterns. Ecology Letters, 8, 224239.Google Scholar
Rahman, L., Whitelaw-Weckert, M. A., Orchard, B. (2014) Impact of organic soil amendments, including poultry-litter biochar, on nematodes in a Riverina, New South Wales, vineyard. Soil Research, 52, 604619.Google Scholar
Raw, F. (1967) Arthropods (except Acari and Collembola). In: Soil Biology (eds. Burges, A., Raw, F.) pp. 323362. London and New York, Academic Press.Google Scholar
Reay, F., Wallace, H. R. (1981) Plant nematodes associated with native vegetation in South Australia. Nematologica, 27, 319329.Google Scholar
Reinhart, K. O., Callaway, R. A. (2006) Soil biota and invasive plants. New Phytologist, 170, 445457.Google Scholar
Reynolds, H. L., Packer, A., Bever, J. D., Clay, K. (2003) Grassroots ecology: Plant–microbe–soil interactions as drivers of plant community structure and dynamics. Ecology, 84, 22812291.Google Scholar
Reynolds, J. F., Smith, D. M., Lambin, E. F., et al. (2007) Global desertification: Building a science for dryland development. Science, 316, 847851.Google Scholar
Ritz, K., Black, H. I. J., Campbell, C. D., Harris, J. A., Wood, C. (2009) Selecting biological indicators for monitoring soils: A framework for balancing scientific and technical opinion to assist policy development. Ecological Indicators, 9, 12121221.Google Scholar
Roberts, D. W., Humber, R. A. (1981) Entomogenous fungi In: Biology of Conidial Fungi (eds. Cole, G. T., Kendrick, W. B.) pp. 201236. New York, Academic.Google Scholar
Robertson, G. P., Freckman, D. W. (1995) The spatial distribution of nematode trophic groups across a cultivated ecosystem. Ecology, 76, 14251432.Google Scholar
Robeson, M. S., King, A. J., Freeman, K. R., et al. (2011) Soil rotifer communities are extremely diverse globally but spatially autocorrelated locally. Proceedings of the National Academy of Sciences of the United States of America, 108, 44064410.Google Scholar
Roger-Estrade, J., Anger, C., Bertrand, M., Richard, G. (2010) Tillage and soil ecology: Partners for sustainable agriculture. Soil Tillage Research, 111, 3340.Google Scholar
Rogers, H. H., Prior, S. A., Runion, G. B., Mitchell, R. J. (1996) Root to shoot ratio of crops as influenced by CO2. Plant and Soil, 187, 229248.Google Scholar
Rønn, R., Griffiths, B. S., Ekelund, F., Christensen, S. (1996) Spatial distribution and successional pattern of microbial activity and micro-faunal populations on decomposing barley roots. Journal of Applied Ecology, 33, 662672.Google Scholar
Rønn, R., Mccaig, A., Griffiths, B. S., Prosser, I. (2002) Impact of protozoan grazing on bacterial community structure in soil microcosms. Applied and Environmental Microbiology, 68, 60946105.Google Scholar
Rouatt, J. W., Katznelson, H., Payne, T. M. B. (1960) Statistical evaluation of the rhizosphere effect. Proceedings – Soil Science Society of America, 24, 271273.Google Scholar
Ruess, L., Chamberlain, P. M. (2010) The fat that matters: Soil food web analysis using fatty acids and their carbon stable isotope signature. Soil Biology & Biochemistry, 42, 18981910.Google Scholar
Ruf, A., Beck, L. (2005) The use of predatory soil mites in ecological soil classification and assessment concepts, with perspectives for oribatid mites. Ecotoxicology and Environmental Safety, 62, 290299.Google Scholar
Ruiz-Jean, M. C., Aide, T. M. (2005) Restoration success: How is it being measured? Restoration Ecology, 13, 569577.Google Scholar
Rusek, J. (1985) Soil microstructures – Contributions on specific soil organisms. Quaestiones Entomologicae, 21, 497514.Google Scholar
Rusek, J. (1998) Biodiversity of the Collembola and their functional role in ecosystems. Biodiversity and Conservation, 7, 12071219.Google Scholar
Rusek, J., Úlehlová, B., Unar, J. (1975) Soil biological features of some alpine grasslands in Czechoslovakia. In: Progress in Soil Zoology (ed. Vanek, J.) pp. 199215. Praha, Academia.Google Scholar
Rusek, J., Weyda, F. (1981) Morphology, ultrastructure and function of pseudocelli in Onychiurus armatus (Collembola, Onychiuridae). Revue d’Écologie et de Biologie du Sol, 18, 127133.Google Scholar
Rutgers, M., Orgiazzi, A., Gardi, C., et al. (2014) Mapping earthworm communities in Europe. Applied Soil Ecology, 97, 98111.Google Scholar
Ryalls, J. M. W., Moore, B. D., Riegler, M., Johnson, S. N. (2016) Above-belowground herbivore interactions in mixed plant communities are influenced by altered precipitation patterns. Frontiers in Plant Science, 7, Article 345.Google Scholar
Sackett, T. E., Classen, A. T., Sanders, N. S. (2010) Linking soil food web structure to above- and belowground ecosystem processes: A meta-analysis. Oikos, 119, 19841992.Google Scholar
Saetre, P. (1998) Decomposition, microbial community structure, and earthworm effects along a birch-spruce soil gradient. Ecology, 79, 834846.Google Scholar
Sala, O. E., Chapin, F. S. I., Armesto, J. J., et al. (2000) Global biodiversity scenarios for the year 2100. Science, 287, 17701774.Google Scholar
Salmon, J. T. (1941) The collembolan fauna of New Zealand, including a discussion of its distribution and affinities. Transactions of the Royal Society of New Zealand, 70, 282431.Google Scholar
Salmon, S., Ponge, J.-F., Gachet, S., et al. (2014) Linking species, traits and habitat characteristics of Collembola at European scale. Soil Biology & Biochemistry, 75, 7385.Google Scholar
San-Blas, E. (2013) Progress on entomopathogenic nematology research: A bibliometric study of the last three decades; 1980–2010. Biological Control, 66, 102124.Google Scholar
Sánchez-Moreno, S. (2010) Suppressive service of the soil food web: Effects of environmental management. Agriculture, Ecosystems & Environment, 119, 7587.Google Scholar
Sánchez-Moreno, S., Camargo, J. A., Navas, A. (2006) Ecotoxicological assessment of the impact of residual heavy metals on soil nematodes in the Guadiamar River Basin (Southern Spain). Environmental Monitoring and Assessment, 116, 245262.Google Scholar
Sanders, N. J., Gotelli, N. J., Gordon, D. M. (2003) Community disassembly by an invasive species. Proceedings of the National Academy of Sciences of the United States of America, 100, 24742477.Google Scholar
Sanders, N. J., Lessard, J.-P., Fitzpatrick, M. C., Dunn, R. R. (2007) Temperature, but not productivity or geometry, predicts elevational diversity gradients in ants across spatial grains. Global Ecology and Biogeography, 16, 640649.Google Scholar
Sanderson, R. A., Rushton, S. P., Cherrill, A. J., Byrne, J. P. (1995) Soil, vegetation and space: An analysis of their effects on the invertebrate communities of a moorland in north-east England. Journal of Applied Ecology, 32, 506518.Google Scholar
Santos, P. F., Phillips, J., Whitford, W. G. (1981) The role of mites and nematodes in early stages of buried litter decomposition in a desert. Ecology, 62, 664669.Google Scholar
Satchell, J. E. (1967) Lumbricidae. In: Soil Biology (eds. Burges, A., Raw, F.) pp. 259322. London and New York, Academic Press.Google Scholar
Schädler, M., Jung, G., Brandl, R., Auge, H. (2004) Secondary succession is influenced by belowground insect herbivory on a productive site. Oecologia, 138, 242252.Google Scholar
Schaefer, M., Schauermann, J. (1990) The soil fauna of beech forests: Comparison between a mull and a moder soil. Pedobiologia, 34, 299314.Google Scholar
Scherber, C., Eisenhauer, N., Weisser, W. W., et al. (2010) Bottom-up effects of plant diversity on multitrophic interactions in a biodiversity experiment. Nature, 468, 553556.Google Scholar
Scheu, S., Albers, D., Alphei, J., et al. (2003) The soil fauna community in pure and mixed stands of beech and spruce of different age: Trophic structure and structuring forces. Oikos, 101, 225238.Google Scholar
Scheu, S., Ruess, L., Bonkowski, M. (2005) Interactions between microorganisms and soil micro- and mesofauna. In: Soil Biology, Microorganisms in Soils: Roles in Genesis and Functions (eds. Buscot, F., Varma, A.) pp. 253275. Berlin, Springer-Verlag.Google Scholar
Scheu, S., Schulz, E. (1996) Secondary succession, soil formation and development of a diverse community of oribatids and saprophagous soil macro-invertebrates. Biodiversity and Conservation, 5, 235250.Google Scholar
Scheu, S., Theenshaus, A., Hefin Jones, T. (1999) Links between the detritivore and the herbivore system: Effects of earthworms and Collembola on plant growth and aphid development. Oecologia, 119, 541551.Google Scholar
Schlegel, J., Riesen, M. (2012) Environmental gradients and succession patterns of carabid beetles (Coleoptera: Carabidae) in an Alpine glacier retreat zone. Journal of Insect Conservation, 16, 657675.Google Scholar
Schmelz, R. M., Niva, C. C., Römbke, J., Collado, R. (2013) Diversity of terrestrial Enchytraeidae (Oligochaeta) in Latin America: Current knowledge and future research potential. Applied Soil Ecology, 69, 1320.Google Scholar
Schmidt, M. H., Roschewitz, I., Theis, C., Tscharntke, T. (2005) Differential effects of landscape and management on diversity and density of ground-dwelling farmland spiders. Journal of Applied Ecology, 42, 281287.Google Scholar
Schneider, K., Migge, S., Norton, R. A., et al. (2004) Trophic niche differentiation in soil microarthropods (Oribatida, Acari): Evidence from stable isotope ratios (15N/14N). Soil Biology & Biochemistry, 36, 17691774.Google Scholar
Schoener, T. W. (1974) Resource partitioning in ecological communities. Science, 185, 2739.Google Scholar
Schon, N. L., Mackay, A. D., Gray, R. A., van Koten, C., Dodd, M. B. (2017) Influence of earthworm abundance and diversity on soil structure and the implications for soil services throughout the season. Pedobiologia – Journal of Soil Ecology, 62, 4147.Google Scholar
Schrader, S., Bender, J., Weigel, H. J. (2009) Ozone exposure of field-grown winter wheat affects soil mesofauna in the rhizosphere. Environmental Pollution, 157, 33573362.Google Scholar
Schuldt, A., Assmann, T. (2009) Environmental and historical effects on richness and endemism patterns of carabid beetles in the western Palaearctic. Ecography, 32, 705714.Google Scholar
Schultz, J. C., Appel, H. M., Ferrieri, A. P., Arnold, T. M. (2013) Flexible resource allocation during plant defense responses. Frontiers in Plant Science, 4, 324.Google Scholar
Schultz, T. R., Brady, S. G. (2008) Major evolutionary transitions in ant agriculture. Proceedings of the National Academy of Sciences of the United States of America, 105, 54355440.Google Scholar
Schutter, M. E., Sandeno, J. M., Dick, R. P. (2001) Seasonal, soil type, and alternative management influences on microbial communities of vegetable cropping systems. Biology and Fertility of Soils, 34, 397410.Google Scholar
Schuurman, G. W. (2012) Ecosystem influences of fungus-growing termites in the dry paleotropics. In: Soil Ecology and Ecosystem Services (ed. Wall, D. H.) pp. 173188. Oxford, Oxford University Press.Google Scholar
Schwarz, B., Barnes, A. D., Thakur, M. P., et al. (2017) Warming alters energetic structure and function but not resilience of soil food webs. Nature Climate Change, 7, 895900.Google Scholar
Scott, W. A., Anderson, R. (2003) Temporal and spatial variation in carabid assemblages from the United Kingdom Environmental Change Network. Biological Conservation, 110, 197210.Google Scholar
Seastedt, T. (1984) The role of microarthropods in decomposition and mineralization processes. Annual Review of Entomology, 29, 2546.Google Scholar
Segers, H. (2008) Global diversity of rotifers (Rotifera) in freshwater. Hydrobiologia, 595, 4959.Google Scholar
Segers, H., De Smet, W. H. (2008) Diversity and endemism in Rotifera: A review, and Keratella Bory de St Vincent. Biodiversity Conservation, 17, 303316.Google Scholar
Sell, P., Kuo-Sell, H. L. (1990) Influence of infestation of oats by root-knot nematodes (Meloidogyne sp.) on the performance of the cereal aphid, Metopolophium dirhodum (Walk) (Hom Aphididae). Journal of Applied Entomology, 109, 3743.Google Scholar
Selonen, S., Liiri, M., Setälä, H. (2014) Can the soil fauna of boreal forests recover from lead-derived stress in a shooting range area? Ecotoxicology, 23, 437448.Google Scholar
Serreze, M. C., Walsh, J. E., Chapin, F. S. I., et al. (2000) Observational evidence of recent changes in the northern high-latitude environment. Climate Change, 46, 159207.Google Scholar
Setälä, H. (2000) Reciprocal interactions between Scots pine and soil food web structure in the presence and absence of ectomycorrhiza. Oecologia, 125, 109118.Google Scholar
Shao, Y., Zhang, W., Shen, J., et al. (2008) Nematodes as indicators of soil recovery in tailings of a lead/zinc mine. Soil Biology & Biochemistry, 40, 20402046.Google Scholar
Sharpley, A. N., Syers, J. K., Springett, J. A. (1979) Effect of surface-casting earthworms on the transport of phosphorus and nitrogen in surface runoff from pasture. Soil Biology & Biochemistry, 11, 459462.Google Scholar
Shaw, A. E., Adams, B. J., Barrett, J. E., et al. (2018) Stable C and N isotope ratios reveal soil food web structure and identify the nematode Eudorylaimus antarcticus as an omnivore-predator in Taylor Valley, Antarctica. Polar Biology, 41, 10131018.Google Scholar
Sherlock, E. (2012) Key to the Earthworms of the UK and Ireland, Shrewsbury, Field Studies Council.Google Scholar
Shmida, A., Wilson, M. V. (1985) Biological determinants of species diversity. Journal of Biogeography, 12, 120.Google Scholar
Siepel, H. (1994) Life history tactics of microarthropods. Biology and Fertility of Soils, 18, 263278.Google Scholar
Siepel, H., de Ruiter-Dijkman, E. M. (1993) Feeding guilds of oribatid mites based on their carbohydrase activities. Soil Biology & Biochemistry, 25, 14911497.Google Scholar
Siepel, H., Maaskamp, F. (1994) Mites of different feeding guilds affect decomposition of organic matter. Soil Biology & Biochemistry, 26, 13891394.Google Scholar
Sieriebriennikov, B., Ferris, H., De Goede, R. G. M. (2014) NINJA: An automated calculation system for nematode-based biological monitoring. European Journal of Soil Biology, 61, 9093.Google Scholar
Sikora, R. A., Malek, R. B., Taylor, D. P., Edwards, D. I. (1979) Reduction of Meloidogyne naasi infection of creeping bentgrass by Tylenchorhynchus agri and Paratrichodorus minor. Nematologica, 25, 179183.Google Scholar
Sileshi, G., Kenis, M., Ogol, C. K. P. O., Sithanantham, S. (2001) Predators of Mesoplatys ochroptera Stål in sesbania-planted fallows in eastern Zambia. BioControl, 46, 289310.Google Scholar
Simaiakis, S. M., Tjørve, E., Gentile, G., Minelli, A., Mylonas, M. (2012) The species–area relationship in centipedes (Myriapoda: Chilopoda): A comparison between Mediterranean island groups. Biological Journal of the Linnean Society, 105, 146159.Google Scholar
Simberloff, D. S., Wilson, E. O. (1970) Experimental zoogeography of islands: A two year record of colonization. Ecology, 51, 934937.Google Scholar
Simmons, B. L., Wall, D. H., Adams, B. J., et al. (2009) Long-term experimental warming reduces soil nematode populations in the McMurdo Dry Valleys, Antarctica. Soil Biology & Biochemistry, 41, 20522060.Google Scholar
Six, J., Frey, S. D., Thiet, R. K., Batten, K. M. (2006) Bacterial and fungal contributions to carbon sequestration in agroecosystems. Soil Science Society of America Journal, 70, 555569.Google Scholar
Slabber, S., Worland, M. R., Leinaas, H. P., Chown, S. L. (2007) Acclimation effects on thermal tolerances of springtails from sub-Antarctic Marion Island: Indigenous and invasive species. Journal of Insect Physiology, 53, 113125.Google Scholar
Smith, H. G. (1996) Diversity of Antarctic terrestrial protozoa. Biodiversity and Conservation, 5, 13791394.Google Scholar
Smith, H. G., Bobrov, A., Lara, E. (2008a) Diversity and biogeography of testate amoebae. Biodiversity and Conservation, 17, 329343.Google Scholar
Smith, H. G., Wilkinson, D. M. (2007) Not all free-living microorganisms have cosmopolitan distributions – The case of Nebela (Apodera) vas Certes (Protozoa: Amoebozoa: Arcellinida). Journal of Biogeography, 34, 18221831.Google Scholar
Smith, R. G., Mcswiney, C. P., Grandy, A. S., et al. (2008b) Diversity and abundance of earthworms across an agricultural land-use intensity gradient. Soil & Tillage Research, 100, 8388.Google Scholar
Smolik, J. D., Dodd, J. L. (1983) Effect of water and nitrogen, and grazing on nematodes in a shortgrass prairie. Journal of Range Management, 36, 744748.Google Scholar
Snyder, B. A., Hendrix, P. F. (2008) Current and potential roles of soil macroinvertebrates (earthworms, millipedes, and isopods) in ecological restoration. Restoration Ecology, 16, 629636.Google Scholar
Sohlenius, B. (1980) Abundance, biomass and contribution to energy flow by nematodes in terrestrial ecosystems. Oikos, 34, 186194.Google Scholar
Sohlenius, B., Boström, S. (2005) The geographic distribution of metazoan microfauna on East Antarctic nunataks. Polar Biology, 28, 439448.Google Scholar
Sohlenius, B., Wasilewska, L. (1984) Influence of irrigation and fertilization on the nematode community in a Swedish pine forest soil. Journal of Applied Ecology, 21, 327342.Google Scholar
Soininen, J. (2012) Macroecology of unicellular organisms – Patterns and processes. Environmental Microbiology Reports, 4, 1022.Google Scholar
Soininen, J., Heino, J. (2005) Relationships between local population persistence, local abundance and regional occupancy of species: Distribution patterns of diatoms in boreal streams. Journal of Biogeography, 32, 19711978.Google Scholar
Soininen, J., Korhonen, J. J., Karhu, J., Vetterli, A. (2011) Disentangling the spatial patterns in community composition of prokaryotic and eukaryotic lake plankton. Limnology & Oceanography, 56, 508520.Google Scholar
Soler, R., Bezemer, T. M., van der Putten, W. H., Vet, L. E. M., Harvey, J. A. (2005a) A multitrophic approach linking below and aboveground insects: The effects of root herbivory on the performance of an aboveground herbivore, its parasitoid and hyperparasitoid. Journal of Animal Ecology, 74, 11211130.Google Scholar
Soler, R., Bezemer, T. M., van der Putten, W. H., Vet, L. E. M., Harvey, J. A. (2005b) Root herbivore effects on above-ground herbivore, parasitoid and hyperparasitoid performance via changes in plant quality. Journal of Animal Ecology, 74, 11211130.Google Scholar
Soler, R., Harvey, J. A., Kamp, A. F. D., et al. (2007) Root herbivores influence the behaviour of an aboveground parasitoid through changes in plant volatile signals. Oikos, 116, 367376.Google Scholar
Soler, R., Schaper, S., Harvey, J. A., et al. (2009) Influence of presence and spatial arrangement of belowground insects on host-plant selection of aboveground insects: A field study. Ecological Entomology, 34, 339345.Google Scholar
Sommer, M., Jochheim, H., Höhn, A., et al. (2013) Si cycling in a forest biogeosystem – The importance of transient state biogenic Si pools. Biogeosciences, 10, 49915007.Google Scholar
Soong, J., Nielsen, U. N. (2016) The role of microarthropods in emerging models of soil organic matter. Soil Biology & Biochemistry, 102, 3739.Google Scholar
Soong, J. L., Vandegehuchte, M. L., Horton, A J., et al. (2016) Soil microarthropods support ecosystem productivity and soil C accrual: Evidence from a litter decomposition study in the tallgrass prairie. Soil Biology & Biochemistry, 92, 230238.Google Scholar
Spain, A. V., McIvor, J. G. (1988) The nature of herbaceous vegetation associated with termitaria in north-eastern Australia. Journal of Ecology, 76, 181191.Google Scholar
St John, M. G., Bellingham, P. J., Walker, L. R., et al. (2012) Loss of a dominant nitrogen-fixing shrub in primary succession: Consequences for plant and below-ground communities. Journal of Ecology, 100, 10741084.Google Scholar
Staley, J. T., Johnson, S. N. (2008) Climate change impacts on root herbivores. In: Root Feeders – An Ecosystem Perspective (eds. Johnson, S. N., Murray, P. J.) pp. 192213. Wallingford, CABI Publishing.Google Scholar
Staley, J. T., Mortimer, S. R., Morecroft, M. D., Brown, V. K., Masters, G. J. (2007) Summer drought alters plant-mediated competition between foliar- and root-feeding insects. Global Change Biology, 13, 866877.Google Scholar
Standen, V. (1978) The influence of soil fauna of decomposition by micro-organisms in blanket bog litter. Journal of Animal Ecology, 47, 2538.Google Scholar
Standen, V. (1984) Production and diversity of enchytraeids, earthworms and plants in fertilized hay meadow plots. Journal of Applied Ecology, 21, 293312.Google Scholar
Stanton, N. L. (1979) Patterns of species diversity in temperate and tropical litter mites. Ecology, 60, 295304.Google Scholar
Stanton, N. L., Tepedino, V. J. (1977) Island habitats in soil communities. Ecological Bulletins, 25, 511514.Google Scholar
Stary, J., Block, W. (1998) Distribution and biogeography of oribatid mites (Acari: Oribatida) in Antarctica, the sub-Antarctic islands and nearby land areas. Journal of Natural History, 32, 861894.Google Scholar
Steinacker, D. F., Wilson, S. D. (2008) Scale and density dependent relationship among roots, mycorrhizal fungi and collembola on grassland and forest. Oikos, 117, 703710.Google Scholar
Stevens, G. C. (1989) The latitudinal gradient in geographical range: How so many species coexist in the tropics. American Naturalist, 133, 240256.Google Scholar
Stevnbak, K., Maraldo, K., Georgieva, S., et al. (2012) Suppression of soil decomposers and promotion of long-lived, root herbivorous nematodes by climate change. European Journal of Soil Biology, 52, 17.Google Scholar
Sticht, C., Schrader, S., Geisemann, A., Weigel, H.-J. (2009) Sensitivity of nematode feeding types in arable soil to free air CO2 enrichment (FACE) is crop specific. Pedobiologia, 52, 337349.Google Scholar
Stirling, G. R. (2011) Suppressive biological factors influence populations of root lesion nematode (Pratylenchus thornei) on wheat in vertosols from the northern grain-growing region of Australia. Australasian Plant Pathology, 40, 416429.Google Scholar
Stout, J. D. (1963) Some observations on the Protozoa of soil beechwood soils on the Chiltern Hills. Journal of Animal Ecology, 32, 281287.Google Scholar
Stout, J. D. (1968) The significance of the protozoan fauna in distinguishing mull and mor of beech (Fagus sylvatica L.). Pedobiologia, 8, 387400.Google Scholar
Stout, J. D., Heal, O. W. (1967) Protozoa. In: Soil Biology (eds. Burges, A., Raw, F.) pp. 149195. London, Academic Press.Google Scholar
Strickland, M. S., Callaham, M. A. Jr., Gardiner, E. S., et al. (2017) Response of soil microbial community composition and function to bottomland forest restoration. Applied Soil Ecology, 119, 317326.Google Scholar
Strickland, M. S., Wickings, K., Bradford, M. A. (2012) The fate of glucose, a low molecular weight compound of root exudates, in the belowground foodweb of forests and pastures. Soil Biology & Biochemistry, 49, 2329.Google Scholar
Strong, D. T., De Wever, H., Merckx, R., Recous, S. (2004) Spatial location of carbon decomposition in the soil pore system. European Journal of Soil Science, 55, 739750.Google Scholar
Stuart, R. J., Barbercheck, M. E., Grewal, P. S., Taylor, R. A. J., Hoy, C. W. (2006) Population biology of entomopathogenic nematodes: Concepts, issues, and models. Biological Control, 38, 80102.Google Scholar
Suarez, A. V., Holway, D. A., Case, T. J. (2001) Patterns of spread in biological invasions dominated by long-distance jump dispersal: Insights from Argentine ants. Proceedings of the National Academy of Sciences of the United States of America, 98, 10951100.Google Scholar
Suding, K. N., Collins, S. L., Gough, L., et al. (2005) Functional- and abundance-based mechanisms explain diversity loss due to N fertilization. Proceedings of the National Academy of Sciences of the United States of America, 102, 43874392.Google Scholar
Sugimoto, A., Bignell, D. E., MacDonald, J. A. (2000) Global impact of termites on the carbon cycle and atmospheric trace gases. In: Termites: Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 409435. Dordrecht, Kluwer Academic Press.Google Scholar
Sun, X., Zhang, X., Zhang, S., et al. (2013) Soil nematode responses to increases in nitrogen deposition and precipitation in a temperate forest. PloS ONE, 8, e82468.Google Scholar
Susilo, F. X., Neutel, A. M., van Noordwijk, M., et al. (2004) Soil biodiversity and food webs. In: Below-ground Interactions in Tropical Agroecosystems (eds. van Noordwijk, M., Cadisch, G., Ong, C. K.) pp. 285302. Wallingford, CAB International.Google Scholar
Susoy, V., Sommer, R. J. (2016) Stochastic and conditional regulation of nematode mouth-form dimorphisms. Frontiers in Ecology and Evolution, 4, 23. doi:10.3389/fevo.2016.00023.Google Scholar
Swift, M. J., Heal, O. W., Anderson, J. M. (1979) Decomposition in Terrestrial Ecosystems, Berkeley, University of California Press.Google Scholar
Swift, M. J., Izac, A. M. N., van Noordwijk, M. (2004) Biodiversity and ecosystem services in agricultural landscapes – Are we asking the right questions? Agriculture, Ecosystems & Environment, 104, 113134.Google Scholar
Tarnokai, C., Canadell, J. G., Schuur, E. A. G., et al. (2009) Soil organic carbon pools in the northern circumpolar permafrost region. Global Biogeochemical Cycles, 23, GB2023, doi:10.1029/2008GB003327.Google Scholar
Taylor, B. R., Parkinson, D., Parsons, W. F. J. (1989) Nitrogen and lignin content as predictors of litter decay rates: A microcosm test. Ecology, 70, 97104.Google Scholar
Terauds, A., Chown, S. L., Fraser, M., et al. (2012) Conservation biogeography of the Antarctic. Diversity and Distribution, 18, 726741.Google Scholar
Teuben, A., Roelofsma, T. A. P. J. (1990) Dynamic interactions between functional groups of soil arthropods and microorganisms during decomposition of coniferous litter in microcosm experiments. Biology and Fertility of Soils, 9, 145151.Google Scholar
Teuben, A., Verhoef, H. A. (1992) Direct contribution by soil arthropods to nutrient availability through body and faecal nutrient content. Biology and Fertility of Soils, 15, 7175.Google Scholar
Thakur, M. P., Könne, T., Griffin, J. N., Eisenhauer, N. (2017a) Warming magnifies predation and reduces prey coexistence in a model litter arthropod system. Proceedings of the Royal Society B, 284, 20162570.Google Scholar
Thakur, M. P., Reich, P. B., Fisichelli, N. A., et al. (2014) Nematode community shifts in response to experimental warming and canopy conditions are associated with plant community changes in the temperate–boreal forest ecotone. Oecologia, 175, 713723.Google Scholar
Thakur, M. P., Reich, P. B., Hobbie, S. E., et al. (2017b) Reduced feeding activity of soil detritivores under warmer and drier conditions. Nature Climate Change, 8, 7578.Google Scholar
The C. elegans Sequencing Consortium (1998) Genome sequence of the nematode C. elegans: A platform for investigating biology. Science, 282, 20122018.Google Scholar
Thoden, T. C., Korthals, G. W., Termorshuizen, A. J. (2011) Organic amendments and their influences on plant-parasitic and free-living nematodes: A promising method for nematode management. Nematology, 13, 133153.Google Scholar
Thompson, L. J., Hoffmann, A. A. (2007) Effects of ground cover (straw and compost) on the abundance of natural enemies and soil macro invertebrates in vineyards. Agricultural and Forest Entomology, 9, 173179.Google Scholar
Thorne, B. L., Grimaldi, D. A., Krishna, K. (2000) Early fossil history of the termites. In: Termites: Evolution, Sociality, Symbioses, Ecology (eds. Abe, T., Bignell, D. E., Higashi, M.) pp. 7794. Dordrecht, Kluwer Academic.Google Scholar
Tiemann, L. K., Grandy, A. S., Atkinson, E. E., Marin-Spiotta, E., Mcdaniel, M. D. (2015) Crop rotational diversity enhances belowground communities and functions in an agroecosystem. Ecology Letters, 18, 761771.Google Scholar
Tietze, F. (1968) Untersuchungen über die Beziehungen zwischen Bodenfeuchte und Carabiden-besiedlung in Wiesengesellschaften. Pedobiologia, 8, 5058.Google Scholar
Tilman, D. (1982) Resource Competition and Community Structure, Princeton, NJ, Princeton University Press.Google Scholar
Tingey, D. T., Johnson, M. G., Lee, E. H., et al. (2006) Effects of elevated CO2 and O3 on soil respiration under ponderosa pine. Soil Biology & Biochemistry, 38, 17641778.Google Scholar
Tixier, M.-S., Kreiter, S., De Moraes, G. J. (2008) Biogeographic distribution of the Phytoseiidae (Acari: Mesostigmata). Biological Journal of the Linnean Society, 93, 845856.Google Scholar
Todd, T. C., Blair, J. M., Milliken, G. A. (1999) Effects of altered soil–water availability on a tallgrass prairie nematode community. Applied Soil Ecology, 13, 4555.Google Scholar
Toepfer, S., Haye, T., Erlandson, M., et al. (2009) A review of the natural enemies of beetles in the subtribe Diabroticina (Coleoptera: Chrysomelidae): Implications for sustainable pest management. Biocontrol Science and Technology, 19, 165.Google Scholar
Topoliantz, S., Ponge, J. F., Viaux, P. (2000) Earthworm and enchytraeid activity under different arable farming systems, as exemplified by biogenic structures. Plant and Soil, 225, 3951.Google Scholar
Torode, M. D., Barnett, K. L., Facey, S. L., et al. (2016) Altered precipitation impacts on above and below-ground grassland invertebrates: Summer drought leads to outbreaks in spring. Frontiers in Plant Science, 7, 1468.Google Scholar
Traniello, J. F. A., Levings, S. C. (1986) Intra- and intercolony patterns of nest dispersion in the ant Lasius neoniger: Correlations with territoriality and foraging ecology. Oecologia, 69, 413419.Google Scholar
Trap, J., Bonkowski, M., Plassard, C., Villenave, C., Blanchart, E. (2016) Ecological importance of soil bacterivores for ecosystem functions. Plant and Soil, 398, 124.Google Scholar
Traunspurger, W., Reiff, N., Krashevska, V., Majdi, N., Scheu, S. (2017) Diversity and distribution of soil micro-invertebrates across an altitudinal gradient in a tropical montane rainforest of Ecuador, with focus on freeliving nematodes. Pedobiologia, 62, 2835.Google Scholar
Treonis, A. M., Austin, E. E., Buyer, J. S., et al. (2010) Effects of organic amendment and tillage on soil microorganisms and microfauna. Applied Soil Ecology, 46, 103110.Google Scholar
Treseder, K. K. (2004) A meta-analysis of mycorrhizal responses to nitrogen, phosphorus, and atmospheric CO2 in field studies. New Phytologist, 164, 347355.Google Scholar
Treseder, K. K. (2008) Nitrogen additions and microbial biomass: A meta-analysis of ecosystem studies. Ecology Letters, 11, 11111120.Google Scholar
Tsiafouli, M. A., Bhusal, D. R., Sgardelis, S. P. (2017) Nematode community indices for microhabitat type and large scale landscape properties. Ecological Indicators, 73, 472479.Google Scholar
Tsiafouli, M. A., Thébault, E., Sgardelis, S. P., et al. (2015) Intensive agriculture reduces soil biodiversity across Europe. Global Change Biology, 21, 973985.Google Scholar
Tsyganov, A. N., Nijs, I., Beyens, L. (2011) Does climate warming stimulate or inhibit soil protist communities? A test on testate Amoebae in high-Arctic tundra with free-air temperature increase. Protist, 162, 237248.Google Scholar
Tuck, S. L., Winqvist, C., Mota, F., et al. (2014) Land-use intensity and the effects of organic farming on biodiversity: A hierarchical meta-analysis. Journal of Applied Ecology, 51, 746755.Google Scholar
Tullgren, A. (1918) Ein sehr einfacher Ausleseapparat für terricole Tierfaunen. Zeitschrift für angewandte Entomologie, 4, 149150.Google Scholar
Turner, J., Bindschadler, R., Convey, P., et al. (2009) Antarctic Climate Change and the Environment, Cambridge, Scientific Committee for Antarctic Research.Google Scholar
Turner, J., Colwell, S. R., Marshall, G. J., et al. (2005) Antarctic climate change during the last 50 years. International Journal of Climatology, 25, 279294.Google Scholar
Tyler, A. N., Carter, S., Davidson, D. A., Long, D. J., Tipping, R. (2001) The extent and significance of bioturbation on 137Cs distributions in upland soils. Catena, 43, 8199.Google Scholar
Ulrich, W., Fiera, C. (2009) Environmental correlates of species richness of European springtails (Hexapoda: Collembola). Acta Oecologica, 35, 4552.Google Scholar
Ulyshen, M. D. (2016) Wood decomposition as influenced by invertebrates. Biological Reviews, 91, 7085.Google Scholar
UN (2014) United Nations Sustainable Development Goals. Open Working Group Proposal for Sustainable Development Goals. www.sustainabledevelopment.un.org/focussdgs.html.Google Scholar
UN (2015) World PopulationProspects: The 2015 Revision, Volume I: Comprehensive Tables (ST/ESA/SER.A/379). New York, Population Division of the Department of Economic and Social Affairs of the United Nations Secretariat.Google Scholar
Urbášek, F., Chalupský, J. (1992) Effects of artificial acidification and liming on biomass and on the activity of digestive enzymes in Enchytraeidae (Oligochaeta): Results of an ongoing study. Biology and Fertility of Soils, 14, 6770.Google Scholar
Urbášek, F., Rusek, J. (1994) Activity of digestive enzymes in seven species of Collembola (Insecta, Entognatha). Pedobiologia, 38, 400406.Google Scholar
Urich, T., Lanzén, A., Qi, J., Huson, D. H., Schleper, C., Schuster, C. H. (2008) Simultaneous assessment of soil microbial community structure and function through analysis of the meta-transcriptome. PloS ONE, 3, e2527.Google Scholar
Usher, M. B., Davis, P., Harris, J., Longstaff, B. (1979) A profusion of species? Approaches towards understanding the dynamics of the populations of microarthropods in decomposer communities. In: Population Dynamics (eds. Anderson, R. M., Turner, B. D., Taylor, L. R.) pp. 359384. Oxford, Blackwell Scientific.Google Scholar
van den Bosch, R., Stern, V. M. (1962) The integration of chemical and biological control of arthropod pests. Annual Review of Entomology, 7, 367386.Google Scholar
van der Putten, W. H. (2003) Plant defense belowground and spatiotemporal processes in natural vegetation. Ecology, 84, 22692280.Google Scholar
van der Putten, W. H., Bardgett, R. D., Bever, J. D., et al. (2013) Plant–soil feedbacks: The past, the present and future challenges. Journal of Ecology, 101, 265276.Google Scholar
van der Putten, W. H., Bardgett, R. D., De Ruiter, P. C., et al. (2009) Empirical and theoretical challenges in aboveground–belowground ecology. Oecologia, 161, 114.Google Scholar
van der Putten, W. H., Klironomos, J. N., Wardle, D. A. (2007) Microbial ecology of biological invasions. The ISME Journal, 1, 2837.Google Scholar
van der Putten, W. H., Yeates, G. W., Duyts, H., Reis, C. S., Karssen, G. (2005) Invasive plants and their escape from root herbivory: A worldwide comparison of the root-feeding nematode communities of the dune grass Ammophila arenaria in natural and introduced ranges. Biological Invasions, 7, 733746.Google Scholar
van Dooremalen, C., Berg, M. P., Ellers, J. (2013) Acclimation responses to temperature vary with vertical stratification: Implications for vulnerability of soil-dwelling species to extreme temperature events. Global Change Biology, 19, 975984.Google Scholar
van Eekeren, N., de Boer, H., Hanegraaf, M., et al. (2010) Ecosystem services in grassland associated with biotic and abiotic soil parameters. Soil Biology & Biochemistry, 42, 14911504.Google Scholar
van Elsas, J. D., Chiurazzi, M., Mallon, C. A., et al. (2012) Microbial diversity determines the invasion of soil by a bacterial pathogen. Proceedings of the National Academy of Sciences of the United States of America, 109, 11591164.Google Scholar
van Groeningen, J. W., Lubbers, I. M., Vos, H. M. J., et al. (2014) Earthworms increase plant production: A meta-analysis. Scientific Reports, 4, 6365.Google Scholar
Van Gundy, S. D., Stolzy, L. H. (1961) Influence of soil oxygen concentrations on the development of Meloidogyne javanica. Science, 134, 665666.Google Scholar
van Vliet, P. C. J. (1998) Hydraulic conductivity and pore size distribution in small microcosms with and without enchytraeids (Oligochaeta). Applied Soil Ecology, 9, 277282.Google Scholar
Van Wensem, J., Verhoef, H. A., Van Straalen, N. M. (1993) Litter degradation stage as a prime factor for isopod interaction with mineralization processes. Soil Biology & Biochemistry, 25, 11751183.Google Scholar
Vanbergen, A. J., Watt, A. D., Mitchell, R., et al. (2007) Scale-specific correlations between habitat heterogeneity and soil fauna diversity along a landscape structure gradient. Oecologia, 153, 713725.Google Scholar
Vandegehuchte, M. L., Sylvain, Z. A., Reichmann, L. G., et al. (2015) Responses of a desert nematode community to changes in water availability. Ecosphere, 6, e44.Google Scholar
Vanek, J. (1967) Industrie exhalate und Moosmilben gemeinschaften in Nordböhmen. In: Progress in Soil Biology (eds. Graff, O., Satchell, J. E.) pp. 331339. North Holland, Amsterdam.Google Scholar
Vanfleteren, J. R., Blaxter, M. L., Tweedie, S. A. R., et al. (1994) Molecular genealogy of some nematode taxa as based on cytochrome c and globin amino acid sequences. Molecular Phylogenetics and Evolution, 3, 92101.Google Scholar
Vannier, G. (1987) The porosphere as an ecological medium emphasized in Professor Ghilarov's work on soil animal adaptations. Biology and Fertility of Soils, 3, 3944.Google Scholar
Vasconcelos, H. L., Maravalhas, J. B., Feitosa, R. M., et al. (2018) Neotropical savanna ants show a reversed latitudinal gradient of species richness, with climatic drivers reflecting the forest origin of the fauna. Journal of Biogeography, 45, 259268.Google Scholar
Vegter, J. J., Huyer-Brugman, F. A. (1983) Comparative water relations in Collembola: Transpiration, desiccation tolerance and effects of body size. In: New Trendsin Soil Biology (eds. Lebrun, P., André, H. M., De Medts, A., Wauthy, G.) pp. 411416. Ottignies, Diet-Brichart.Google Scholar
Velasco-Castrillón, A., Page, T. J., Gibson, J. A. E., Stevens, M. I. (2014a) Surprisingly high levels of biodiversity and endemism amongst Antarctic rotifers uncovered with mitochondrial DNA. Biodiversity, 15, 130142.Google Scholar
Velasco-Castrillón, A., Schultz, M. B., Colombo, F., et al. (2014b) Distribution and diversity of soil microfauna from East Antarctica: Assessing the link between biotic and abiotic factors. PloS ONE, 9, e87529. doi:10.1371/journal.pone.0087529.Google Scholar
Veresoglou, S. D., Halley, J. M., Rillig, M. C. (2015) Extinction risk of soil biota. Nature Communications, 6,8862, doi:10.1038/ncomms9862.Google Scholar
Verhoef, H. A., Brussaard, L. (1990) Decomposition and nitrogen mineralization in natural and agri-ecosystems: The contribution of soil animals. Biogeochemistry, 11, 175211.Google Scholar
Verhoef, H. A., Nagelkerke, C. J., Joose, E. N. G. (1977) Aggregation pheromones in Collembola. Journal of Insect Physiology, 23, 10091013.Google Scholar
Verschoor, B. C., Pronk, T. E., de Goede, R. G. M, Brussaard, L. (2002) Could plant-feeding nematodes affect the competition between grass species during succession in grasslands under restoration management? Journal of Ecology, 90, 753761.Google Scholar
Vervoort, M. T. W., Vonk, J. A., Mooijman, P. J. W., et al. (2012) SSU ribosomal DNA-based monitoring of nematode assemblages reveals distinct seasonal fluctuations within evolutionary heterogeneous feeding guilds. PloS ONE, 7 (10), e47555. doi:10.1371/journal.pone.0047555.Google Scholar
Viketoft, M., Bengtsson, J., Sohlenius, B., et al. (2009) Long-term effects of plant diversity and composition on soil nematode communities in model grasslands. Ecology, 90, 9099.Google Scholar
Viketoft, M., van der Putten, W. (2015) Top-down control of root-feeding nematodes in range-expanding and congeneric native plant species. Basic and Applied Ecology, 16, 260268.Google Scholar
Virginia, R. A., Wall, D. H. (1999) How soils structure communities in the Antarctic Dry Valleys. Bioscience, 49, 973983.Google Scholar
Vitousek, P. M., Aber, J. D., Howarth, R. W., et al. (1997a) Human alteration of the global nitrogen cycle: Sources and consequences. Ecological Applications, 7, 737750.Google Scholar
Vitousek, P. M., Mooney, H. A., Lubchenco, J., Melillo, J. M. (1997b) Human domination of earth's ecosystems. Science, 277, 494499.Google Scholar
Vos, C., Schouteden, N., van Tuinen, D., et al. (2013) Mycorrhiza-induced resistance against the rooteknot nematode Meloidogyne incognita involves priming of defense gene responses in tomato. Soil Biology & Biochemistry, 60, 4554.Google Scholar
Vreeken-Bruijs, M. J., Hassink, J., Brussaard, L. (1998) Relationships of soil microarthropod biomass with organic matter and pore size distribution in soils under different land use. Soil Biology & Biochemistry, 30, 97106.Google Scholar
Wagg, C., Bendera, S. F., Widmerc, F., van der Heijden, M. G. A. (2014) Soil biodiversity and soil community composition determine ecosystem multifunctionality. Proceedings of the National Academy of Sciences of the United States of America, 111, 52665270.Google Scholar
Walker, M. D., Walker, D. A., Welker, J. M., et al. (1999) Long-term experimental manipulation of winter snow regime and summer temperature in arctic and alpine tundra. Hydrological Processes, 13, 23152330.Google Scholar
Wall, D. H. (2007) Global change tipping points: Above- and belowground biotic interactions in a low diversity ecosystem. Philosophical Transactions of the Royal Society B: Biological Sciences, 362, 22912306.Google Scholar
Wall, D. H., Bradford, M. A., St John, M. G., et al. (2008) Global decomposition experiment shows soil animal impacts on decomposition are climate-dependent. Global Change Biology, 14, 26612677.Google Scholar
Wall, D. H., Nielsen, U. N., Six, J. (2015) Soil biodiversity and human health. Nature, 528, 6976.Google Scholar
Wall, J. W., Skene, K. R., Neilson, R. (2002) Nematode community and trophic structure along a sand dune succession. Biology and Fertility of Soils, 35, 293301.Google Scholar
Wallace, A. R. (1853) On the insects used for food in the Indians of the Amazon. Transactions of the Royal Entomological Society of London, 2, 241244.Google Scholar
Wallace, R. L., Snell, T. W., Ricci, C., Nogrady, T. (2006) Rotifera Volume 1: Biology, Ecology and Systematics, 2nd edn, Gent: Kenobi Productions and The Hague: Backhyus Academic Publishing BV.Google Scholar
Wallwork, J. A. (1970) Ecology of Soil Animals, New York, McGraw-Hill.Google Scholar
Wallwork, J. A. (1976) The Diversity and Distribution of Soil Fauna, London, Academic Press.Google Scholar
Walsh, C. L., Johnson-Maynard, J. L. (2016) Earthworm distribution and density across a climatic gradient within the Inland Pacific Northwest cereal production region. Applied Soil Ecology, 104, 104110.Google Scholar
Wang, J. G., Bakken, L. R. (1997) Competition for nitrogen during decomposition of plant residues in soil: Effect of spatial placement of N-rich and N-poor plant residues. Soil Biology & Biochemistry, 29, 153162.Google Scholar
Wang, L., Chen, Z., Shang, H., Wang, J., Zhang, P. Y. (2014) Impact of simulated acid rain on soil microbial community function in Masson pine seedlings. Electronic Journal of Biotechnology, 17, 199203.Google Scholar
Wang, X., Nielsen, U. N., Yang, X., et al. (2018) Grazing induces direct and indirect shrub effects on soil nematode communities. Soil Biology & Biochemistry, 121, 193201.Google Scholar
Wardle, D. A. (1992) A comparative assessment of factors which influence microbial biomass carbon and nitrogen levels in soil. Biology Reviews, 67, 321358.Google Scholar
Wardle, D. A. (2002) Communities and Ecosystems: Linking the Aboveground and Belowground Components, Princeton, NJ, Princeton University Press.Google Scholar
Wardle, D. A. (2006) The influence of biotic interactions on soil biodiversity. Ecology Letters, 9, 870886.Google Scholar
Wardle, D. A., Giller, K. E. (1996) The quest for a contemporary ecological dimension to soil biology. Soil Biology & Biochemistry, 28, 15491554.Google Scholar
Wardle, D. A., Peltzer, D. A. (2017) Impacts of invasive biota in forest ecosystems in an aboveground–belowground context. Biological Invasions, 19, 33013316.Google Scholar
Wardle, D. A., Yeates, G. W., Baker, G. M., Bonner, K. I. (2006) The influence of plant litter diversity on decomposer abundance and diversity. Soil Biology & Biochemistry, 38, 10521062.Google Scholar
Wardle, D. A., Yeates, G. W., Watson, R. N., Nicholson, K. S. (1995) Development of the decomposer food-web, trophic relationships, and ecosystem properties during a three-year primary succession in sawdust. Oikos, 73, 155166.Google Scholar
Wasilewska, L., Bienkowski, P. (1985) Experimental study on the occurrence and activity of soil nematodes in decomposition of plant material. Pedobiologia, 28, 4157.Google Scholar
Watanabe, H., Tokuda, G. (2010) Cellulolytic systems in insects. Annual Review of Entomology, 55, 609632.Google Scholar
Wauthy, G. (1982) Synecology of forest soil oribatid mites of Belgium (Acari, Oribatida). III. Ecological groups. Acta Oecologia, 3, 469494.Google Scholar
Wauthy, G., Noti, M.-I., Dufrêne, M. (1989) Geographic ecology of soil oribatid mites in deciduous forests. Pedobiologia, 33, 399416.Google Scholar
Weaver, H. J., Hawdon, J. M., Hoberg, E. P. (2010) Soil-transmitted helminthiases: Implications of climate change and human behavior. Trends in Parasitology, 26, 574581.Google Scholar
Webb, C. O. (2000) Exploring the phylogenetic structure of ecological communities: An example for rain forest trees. American Naturalist, 156, 145155.Google Scholar
Webb, N. R., Hoeting, J. A., Ames, G. M., Pyne, M. I., Poff, N. L. (2010) A structured and dynamic framework to advance traits-based theory and prediction in ecology. Ecology Letters, 13, 267283.Google Scholar
Wei, H., Liu, W., Zhang, J., Qin, Z. (2017) Effects of simulated acid rain on soil fauna community composition and their ecological niches. Environmental Pollution, 220, 460468.Google Scholar
Weis-Fogh, T. (1948) Ecological investigations on mites and collemboles in the soil. Natura Jutlandica, 1, 309330.Google Scholar
Weiser, M. D., Michaletz, S. T., Buzzard, V., et al. (2018) Toward a theory for diversity gradients: The abundance–adaptation hypothesis. Ecography, 41, 255264.Google Scholar
Wells, T. C. E., Sheail, J., Ball, D. F., Ward, L. K. (1976) Ecological studies in the Porton Ranges. Relationships between vegetation, soils and land-use history. Journal of Ecology, 64, 589624.Google Scholar
Weronika, E., Łukasz, K. (2017) Tardigrades in space research – Past and future. Astrobiology, 47, 545533.Google Scholar
Wertheim, B., van Baalen, E.-J. A., Dicke, M., Vet, L. E. M. (2005) Pheromone-mediated aggregation in nonsocial arthropods: An evolutionary ecological perspective. Annual Review of Entomology, 50, 321346.Google Scholar
Wharton, D. A. (1986) A Functional Biology of Nematodes, Baltimore, MD, The Johns Hopkins University Press.Google Scholar
Wharton, D. A., Fern, D. J. (1995) Survival of intracellular freezing by the Antarctic nematode Panagrolaimus davidi. The Journal of Experimental Biology, 198, 13811387.Google Scholar
Wickings, K., Grandy, A. S. (2011) The oribatid mite Scheloribates moestus (Acari: Oribatida) alters litter chemistry and nutrient cycling during decomposition. Soil Biology & Biochemistry, 43, 351358.Google Scholar
Wickings, K., Grandy, A. S., Reed, S. C. (2012) The origin of litter chemical complexity during decomposition. Ecology Letters, 15, 11801188.Google Scholar
Widenfalk, L. A., Bengtsson, J., Berggren, Å., et al. (2015) Spatially structured environmental filtering of collembolan traits in late successional salt marsh vegetation. Oecologia, 179, 537549.Google Scholar
Widenfalk, L. A., Leinaas, H. P., Bengtsson, J., Birkemoe, T. (2018) Age and level of self-organization affect the small-scale distribution of springtails (Collembola). Ecosphere, 9, e02058.Google Scholar
Widenfalk, L. A., Malmström, A., Berg, M. P., Bengtsson, J. (2016) Small-scale Collembola community composition in a pine forest soil – Overdispersion in functional traits indicates the importance of species interactions. Soil Biology & Biochemistry, 103, 5262.Google Scholar
Wilkinson, D. M. (2001) What is the upper size limit for cosmopolitan distribution in free living microorganisms? Journal of Biogeography, 28, 285291.Google Scholar
Wilkinson, D. M., Creevy, A. L., Valentine, J. (2012) The past, present and future of soil protist ecology. Acta Protozoologica, 51, 189199.Google Scholar
Wilkinson, D. M., Mitchell, E. A. D. (2010) Testate amoebae and nutrient cycling with particular reference to soils. Geomicrobiology Journal, 27, 520533.Google Scholar
Williams, B. L., Griffiths, B. S. (1989) Enhanced nutrient mineralization and leaching from decomposing Sitka spruce litter by enchytraeid worms. Soil Biology & Biochemistry, 21, 18831888.Google Scholar
Williamson, W. M., Wardle, D. A., Yeates, G. W. (2005) Changes in soil microbial and nematode communities during ecosystem decline across a long-term chronosequence. Soil Biology & Biochemistry, 37, 12891301.Google Scholar
Willis, K. J., Jeffers, E. S., Tovar, C. (2018) What makes a terrestrial ecosystem resilient? Science, 359, 988989.Google Scholar
Wilson, E. O. (1974) The Insect Societies, Cambridge, MA, Belknap Press of Harvard University Press.Google Scholar
Wissuwa, J., Salamon, J.-A., Frank, T. (2012) Effects of habitat age and plant species on predatory mites (Acari, Mesostigmata) in grassy arable fallows in Eastern Austria. Soil Biology & Biochemistry, 50, 96107.Google Scholar
Wolters, V. (2000) Invertebrate control of soil organic matter stability. Biology and Fertility of Soils, 31, 119.Google Scholar
Wolters, V. (2001) Biodiversity of soil animals and its function. European Journal of Soil Biology, 37, 221227.Google Scholar
Womersley, H. (1939) Primitive Insects of South Australia: Silverfish, Springtails and their Allies, Adelaide, Government Printer.Google Scholar
Wood, T. G. (1971) The distribution and abundance of Folsomides deserticola (Collembola: Isotomidae) and other micro-arthropods in arid and semi-arid soils in southern Australia, with a note on nematode populations. Pedobiologia, 11, 446468.Google Scholar
Wood, T. G. (1976) The role of termites (Isoptera) in decomposition processes. In: The Role of Terrestrial and Aquatic Organismsin Decomposition Processes (eds. Anderson, J. M., Macfadyen, A.) pp. 145168. Oxford, Blackwell Scientific.Google Scholar
Wood, T. G. (1978) Food and feeding habits of termites. In: Production Ecology of Ants and Termites (ed. Brian, M. V.) pp. 5580. Cambridge, Cambridge University Press.Google Scholar
Wood, T. G., Sands, W. A. (1978) The role of termites in ecosystems. In: Production Ecology of Ants and Termites (ed. Brian, M. V.) pp. 245293. Cambridge, Cambridge University Press.Google Scholar
Woods, L. E., Cole, C. V., Elliott, E. V., Anderson, R. V., Coleman, D. C. (1982) Nitrogen transformation in soils as affected by bacterial–microfaunal interaction. Soil Biology & Biochemistry, 14, 9398.Google Scholar
Wu, T., Ayres, E., Bardgett, R. D., Wall, D. H., Garey, J. R. (2011a) Molecular study of worldwide distribution and diversity of soil animals. Proceedings of the National Academy of Sciences of the United States of America, 108, 1772017725.Google Scholar
Wu, X., Duffy, J. E., Reich, P. B., Sun, S. (2011b) A brown-world cascade in the dung decomposer food web of an alpine meadow: Effects of predator interactions and warming. Ecological Monographs, 81, 313328.Google Scholar
Wu, Z., Dijkstra, P., Koch, G. W., Peñuelas, J., Hungate, B. A. (2011c) Responses of terrestrial ecosystems to temperature and precipitation change: A meta-analysis of experimental manipulation. Global Change Biology, 17, 927942.Google Scholar
Wubs, E. R. J., van der Putten, W. H., Bosch, M., Bezemer, T. M. (2016) Soil inoculation steers restoration of terrestrial ecosystems. Nature Plants, 2,16107, doi:10.1038/NPLANTS.2016.1107.Google Scholar
Wurst, S., Gebhardt, K., Rillig, M. C. (2011) Independent effects of arbuscular mycorrhiza and earthworms on plant diversity and newcomer plant establishment. Journal of Vegetation Science, 22, 10211030.Google Scholar
Wurst, S., Langel, R., Reineking, A., Bonkowski, M., Scheu, S. (2003) Effects of earthworms and organic litter distribution on plant performance and aphid reproduction. Oecologia, 137, 9096.Google Scholar
Wurst, S., Ohgushi, T. (2015) Do plant- and soil-mediated legacy effects impact future biotic interactions? Functional Ecology, 29, 13731382.Google Scholar
Wurst, S., van Dam, N. M., Monroy, F., Biere, A., Van Der Putten, W. H. (2008) Intraspecific variation in plant defense alters effects of root herbivores on leaf chemistry and aboveground herbivore damage. Journal of Chemical Ecology, 34, 13601367.Google Scholar
Wurst, S., van der Putten, W. H. (2007) Root herbivore identity matters in plant-mediated interactions between root and shoot herbivores. Basic and Applied Ecology, 8, 491499.Google Scholar
Wurst, S., Wagenaar, R., Biere, A., van der Putten, W. H. (2010) Microorganisms and nematodes increase levels of secondary metabolites in roots and root exudates of Plantago lanceolata. Plant and Soil, 329, 117126.Google Scholar
Xia, J. Y., Wan, S. Q. (2008) Global response patterns of terrestrial plant species to nitrogen addition. New Phytologist, 179, 428439.Google Scholar
Xiao, H. F., Schaefer, D. A., Lei, Y. B., et al. (2013) Influence of invasive plants on nematode communities under simulated CO2 enrichment. European Journal of Soil Biology, 58, 9197.Google Scholar
Yamada, A., Inoue, T., Wiwatwitaya, D., et al. (2005) Carbon mineralization by termites in tropical forests, with emphasis on fungus combs. Ecological Research, 20, 453460.Google Scholar
Yeates, G. W. (1974) Studies on a climosequence of soils in tussock grasslands. New Zealand Journal of Zoology, 1, 171177.Google Scholar
Yeates, G. W. (1998) Soil nematode assemblages: Regulators of ecosystem productivity. Phytoparasitica, 26, 97100.Google Scholar
Yeates, G. W., Bongers, T., De Goede, R. G. M., Freckman, D. W., Georgieva, S. S. (1993) Feeding habits in soil nematode families and genera – An outline for soil ecologists. Jounal of Nematology, 25, 315331.Google Scholar
Yeates, G. W., Hawke, M. F., Rijkse, W. C. (2000) Changes in soil fauna and soil conditions under Pinus radiata agroforestry regimes during a 25-year tree rotation. Biology and Fertility of Soils, 31, 391406.Google Scholar
Yeates, G. W., Newton, P. C. D. (2009) Long-term changes in topsoil nematode populations in grazed pasture under elevated atmospheric carbon dioxide. Biology and Fertility of Soils, 45, 799808.Google Scholar
Yeates, G. W., Saggar, S., Denton, C. S., Mercer, C. F. (1998) Impact of clover cyst nematode (Heterodera trifolia) infection on soil microbial activity in the rhizophere of white clover (Trifolium repens) – A pulse-labelling experiment. Nematologica, 44, 8190.Google Scholar
Yeates, G. W., Wardle, D. A. (1996) Nematodes as predators and prey: Relationships to biological control and soil processes. Pedobiologia, 40, 4350.Google Scholar
Young, I. M., Crawford, J. W. (2004) Interactions and self-organization in the soil–microbe complex. Science, 304, 16341637.Google Scholar
Young, I. M., Ritz, K. (2000) Tillage, habitat space and function of soil microbes. Soil & Tillage Research, 53, 201213.Google Scholar
Young, M. R., Behan-Pelletier, V. M., Hebert, P. D. N. (2012) Revealing the hyperdiverse mite fauna of subarctic Canada through DNA barcoding. PloS ONE, 7, e48755. doi:10.1371/journal.pone.0048755.Google Scholar
Zak, D. R., Pregitzer, K. S., Curtis, P. S., et al. (1993) Elevated atmospheric CO2 and feedback between carbon and nitrogen cycles. Plant and Soil, 151, 105117.Google Scholar
Zaller, J. G., Arnone III, J. A. (1999) Earthworm responses to plant species’ loss and elevated CO2 in calcareous grassland. Plant and Soil, 208, 18.Google Scholar
Zeng, G., Pyle, J. A., Young, P. J. (2008) Impact of climate change on tropospheric ozone and its global budgets. Atmospheric Chemistry and Physics, 8, 369387.Google Scholar
Zhang, J., Yu, J., Ouyang, Y. (2015) Activity of earthworm in latosol under simulated acid rain stress. Bulletin of Environmental Contamination and Toxicology, 94, 108111.Google Scholar
Zhang, J. E., Yu, J. Y., Ouyang, Y., Xu, H. Q. (2014) Impact of simulated acid rain on trace metals and aluminum leaching in latosol from Guangdong Province, China. International Journal of Soil and Sediment Contamination, 23, 725735.Google Scholar
Zhang, X., Ferris, H., Mitchell, J., Liang, W. (2017) Ecosystem services of the soil food web after long-term application of agricultural management practices. Soil Biology & Biochemistry, 111, 3643.Google Scholar
Zhao, C., Griffin, J. N., Wu, X., Sun, S. (2013) Predatory beetles facilitate plant growth by driving earthworms to lower soil layers. Journal of Animal Ecology, 82, 749758.Google Scholar
Zhu, T., Yang, C., Wang, J., et al. (2018) Bacterivore nematodes stimulate soil gross N transformation rates depending on their species. Biology and Fertility of Soils, 54, 107118.Google Scholar
Zullini, A., Peretti, E. (1986) Lead pollution and moss-inhabiting nematodes of an industrial area. Water, Air, & Soil Pollution, 27, 403410.Google Scholar
Zwart, K. B., Kuikman, P. J., van Veen, J. A. (1994) Rhizosphere protozoa: Their significance in nutrient dynamics. In: Soil Protozoa (ed. Darbyshire, J. F.) pp. 93122. Wallingford, CAB International.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.

  • Bibliography
  • Uffe N. Nielsen, Western Sydney University
  • Book: Soil Fauna Assemblages
  • Online publication: 18 March 2019
  • Chapter DOI: https://doi.org/10.1017/9781108123518.011
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.

  • Bibliography
  • Uffe N. Nielsen, Western Sydney University
  • Book: Soil Fauna Assemblages
  • Online publication: 18 March 2019
  • Chapter DOI: https://doi.org/10.1017/9781108123518.011
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.

  • Bibliography
  • Uffe N. Nielsen, Western Sydney University
  • Book: Soil Fauna Assemblages
  • Online publication: 18 March 2019
  • Chapter DOI: https://doi.org/10.1017/9781108123518.011
Available formats
×