Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-28T16:58:35.979Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  18 December 2017

Richard I. Macphail
Affiliation:
University College London
Paul Goldberg
Affiliation:
Boston University
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2017

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

Acott, T. G., Cruise, G. M., and Macphail, R. I., 1997, Soil micromorphology and high resolution images, in Shoba, S., Gerasimova, M., and Miedema, R., eds., Soil Micromorphology: Diversity, Diagnostics and Dynamics: Moscow-Wageningen, International Soil Science Society, pp. 372378.Google Scholar
Adams, A. E., MacKenzie, W. S., and Guilford, C., 1984, Atlas of Sedimentary Rocks Under the Microscope, New York, Longman Scientific & Technical, pp. 104.Google Scholar
Adderley, W. P., Simpson, I. A., and Davidson, D., 2006, Historic landscape management: a validation of quantitative soil thin-section analyses: Journal of Archaeological Science, v. 33, pp. 320334.CrossRefGoogle Scholar
Akeret, Ö., and Rentzel, P., 2001, Micromorphology and plant macrofossil analysis of cattle dung from the Neolithic lake shore settlement of Arbon/Bleiche 3: Geoarchaeology, v. 16, no. 6, pp. 687700.Google Scholar
Albert, R.-M., Shahack-Gross, R., Cabanes, D., Gilboa, A., Lev-Yadun, S., Portillo, M., Sharon, I., Boaretto, E., and Weiner, S., 2008, Phytolith-rich layers from the Late Bronze and Iron Ages at Tel Dor (Israel): mode of formation and archaeological significance: Journal of Archaeological Science, v. 35, pp. 5775.CrossRefGoogle Scholar
Albert, R. M., and Weiner, S., 2001, Study of phytoliths in prehistoric ash layers from Kebara and Tabun caves using a quantitative approach, in Meunier, J. D., and Colin, F., eds., Phytoliths: Applications in Earth Science and Human History: Leiden, Balkema, pp. 239250.Google Scholar
Allen, J. M., 2005, Beaker occupation and development of the downland landscape at Ashcombe Bottom, nr Lewes, East Sussex: Sussex Archaeological Collections, v. 143, pp. 733.Google Scholar
Allen, J. M., 2007, Prehistoric and medieval environment of Old Town, Eastbourne: Sussex Archaeological Collections, v. 145, pp. 1-32.Google Scholar
Allen, J. M., Godden, D., and Matthews, C., 2002, Mesolithic, Late Bronze Age and Medieval activity at Katherine Farm, Avonmouth, 1998, Archaeology in the Severn Estuary 2002: Bath, Severn Estuary Levels Research Committee, pp. 85105.Google Scholar
Allen, M. J., 1988, Archaeological and environmental aspects of colluviation in south-east England, in Waateringe, W. Groenman-van, and Robinson, M., eds., Man-made Soils, BAR International Series 410: Oxford, British Archaeological Reports, pp. 6792.Google Scholar
Allen, M. J. 1990, Magnetic susceptibility, in Bell, M., ed., Excavations at Brean Down, Somerset, Archaeological Report No. 15: London, English Heritage, pp. 197202.Google Scholar
Allen, M. J. 1992, Products of erosion and the Prehistoric land use of the Wessex Chalk., in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 3752.Google Scholar
Allen, M. J. 1994, The Land-Use History of the Southern English Chalklands with an Evaluation of the Beaker Period Using Environmental Data: Colluvial Deposits as Environmental and Cultural Indicators: Southampton University.Google Scholar
Allen, M. J. 1995, Land molluscs, in Wainwright, G. J., and Davis, S. M., eds., Balksbury Camp, Hampshire; Excavations 1973 and 1981, Archaeological Report 4: London, English Heritage, pp. 92100.Google Scholar
Allen, M. J. 2008, Late Palaeolithic (Area 3): environmental evidence for the former environment and possible human activity, in FitzPatrick, A. P., Powell, A. B., and Allen, M. J., eds., Archaeological Excavations on the Route of the A27 Westhampnett Bypass, West Sussex, 1992, Wessex Archaeology Report No. 21: Salisbury, Wessex Archaeology, pp. 3061.Google Scholar
Allen, M. J., and Macphail, R. I., 1987, Micromorphology and magnetic susceptibility studies: their combined role in interpreting archaeological soils and sediments, in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 669676.Google Scholar
Altschul, J. H., Ciolek-Torrello, R., Grenda, D. R., Homburg, J. A., Benaron, S., and Stoll, A. Q., 2005, Ballona Archaeology: A Decade of Multidisciplinary Research: Proceedings of the Society for California Archaeology, v. 18, pp. 283301.Google Scholar
Andrews, P., 1990, Owls, Caves and Fossils. Predation, Preservation and Accumulation of Small Mammal Bones in Caves, with an Analysis of the Pleistocene Cave Faunas from Westbury-sub-Mendip, Somerset, UK, London, The Natural History Museum, Natural History Museum Publications.Google Scholar
Andrews, P., Cook, J., Currant, A., and Stringer, C., 1999, Westbury Cave. The Natural History Museum Excavations 1976–1984., Bristol, CHERUB (Centre for Human Evolutionary Research at the University of Bristol).Google Scholar
Angelucci, D. E., 2008, Geoarchaeological insights from a Roman age incineration feature (ustrinum) at Enconsta de Sant’Ana (Lisbon, Portugal): Journal of Archaeological Science, v. 35, no. 9, pp. 26242633.Google Scholar
Angelucci, D. E., Boschian, G., Fontanals, M., Pedrotti, A., and Vergès, J. M., 2009, Shepherds and karst: the use of caves and rock shelters in the Mediterranean region during the Neolithic: World Archaeology, v. 41, no. 2, pp. 191214.Google Scholar
ApSimon, A. M., Smart, P. L., Macphail, R. I., Scott, K., and Taylor, H., 1992, King Arthur’s Cave, Whitchurh, Herefordshire: a reassessment: Proceedings of the University of Bristol Spelaeological Society, v. 19, pp. 183249.Google Scholar
Armour-Chelu, M., and Andrews, P., 1994, Some effects of bioturbation by earthworms (Oligochaeta) on archaeological sites: Journal of Archaeological Science, v. 21, pp. 433443.Google Scholar
Arpin, T., Macphail, R. I., and Boschian, G., 1998, Summary of the Spring 1998 Meeting of the Working Group on Archaeological Soil Micromorphology, February 27-March 1, 1998: Geoarchaeology, v. 13, no. 6, pp. 645647.Google Scholar
Arpin, T., Mallol, C., and Goldberg, P., 2002, A new method for analyzing and documenting micromorphological thin sections using flatbed scanners: applications in geoarchaeological studies: Geoarchaeology, v. 17, pp. 305313.Google Scholar
Arroyo-Kalin, M., 2010, The Amazonian Formative: Crop Domestication and Anthropogenic Soils: Diversity, v. 2, pp. 473504.Google Scholar
Arroyo-Kalin, M. 2014, Amazonian Dark Earths: Geoarchaeology, 168178. Springer, New York, Encyclopedia of Global Archaeology, Volume 1: New York, Springer, pp. 168–178.Google Scholar
Arroyo-Kalin, M., Neves, E. G., and Woods, W. I., 2008, Anthropogenic Dark Earths of the Central Amazon Region: Remarks on their evolution and polygenetic composition, in Woods, W. I., Teixeira, W. G., Lehmann, J., Steiner, C., WinklerPrins, A. M. G. A., and Rebellato, L., eds., Amazonian Dark Earths: Wim Sombroek’s Vision: New York/Berlin, Springer Science, pp. 99125.Google Scholar
Ashwin, T., and Tester, A., 2014, A Roman Settlement in the Waveney Valley: Excavations at Scole, 1993–4, East Anglian Archaeology, Report No. 152, pp. 431.Google Scholar
Avery, B. W., 1990, Soils of the British Isles, Wallingford, CAB International, p. 463.Google Scholar
Avery, B. W., and Bascomb, C. L., 1974, Soil Survey Laboratory Techniques, Harpenden, Soil Survey of England and Wales, Soil Survey Technical Monograph.Google Scholar
Babel, U., 1975, Micromorphology of soil organic matter, in Giesking, J. E., ed., Soil Components: Organic Components, Volume 1: New York, Springer-Verlag, pp. 369473.CrossRefGoogle Scholar
Bakels, C. C., 1988, Pollen from plaggen soils in the province of North Brabant, the Netherlands, in Groenman-van Waateringe, W., and Robinson, M., eds., Man-made Soils, International Series 410: Oxford, British Archaeological reports, pp. 5566.Google Scholar
Bakels, C. C. 1997a, The beginnings of manuring in western Europe: Antiquity, v. 71, no. 272, pp. 442445.CrossRefGoogle Scholar
Bal, L., 1982, Zoological ripening of soils, Wageningen, Centre for Agricultural Publishing and Documentation, Agricultural Research Report, p. 365.Google Scholar
Balaam, N., Bell, M., David, A., Levitan, B., Macphail, R. I., Robinson, M., and Scaife, R. G., 1987, Prehistoric and Romano-British sites at Westward Ho!, Devon: archaeological and palaeoenvironmental surveys 1983 and 1984, in Balaam, N. D., Levitan, B., and Straker, V., eds., Studies in palaeoeconomy and environment in South West England, British Series 181: Oxford, British Archaeological Reports, pp. 163264.Google Scholar
Ball, D. F., 1975, Processes of soil degradation: a pedological point of view, in Evans, J. G., Limbrey, S., and Cleere, H., eds., The effect of man on the landscape: the Highland Zone, CBA Research Report No. 11: Nottingham, The Council for British Archaeology, pp. 2027.Google Scholar
Banerjea, R. Y., Bell, M., Matthews, W., and Brown, A., 2015, Applications of micromorphology to understanding activity areas and site formation processes in experimental hut floors: Archaeological and Anthropological Sciences, v. 7, no. 1, pp. 89112.Google Scholar
Barclay, A., Lambrick, G., Moore, J., and Robinson, M., 2003, Lines in the Landscape. Cursus Monuments in the Upper Thames Valley: excavations at the Drayton and Lechlade Cursuses, Oxford, Oxford Archaeological Unit.Google Scholar
Barclay, G. T., 1983, Sites of the third millennium BC to the first millennium AD at North Mains, Strathallan, Perthshire: Proceedings of the Society of Antiquaries Scotland, v. 113, pp. 122281.Google Scholar
Barnes, G. L., 1990, Paddy soils now and then, in Thomas, K., ed., Soils and Early Agriculture, World Archaeology 22 (1): London, Routledge, pp. 117.Google Scholar
Barrat, B. C., 1964, A classification of humus forms and microfabrics in temperate grasslands: Journal of Soil Science, v. 15, pp. 342356.Google Scholar
Barton, R. N. E., 1992, Hengistbury Head, Dorset. Volume 2: The Late Upper Palaeolithic & Early Mesolithic sites, Oxford, Oxford University Committee for Archaeology.Google Scholar
Barton, R. N. E. 1997, Fifth Interim Report on the Survey and Excavations in the Wye Valley, 1997: Proceedings of the University of Bristol Spelaeological Society, v. 21, no. 1, pp. 99108.Google Scholar
Barton, R. N. E. 2000, Mousterian hearths and shellfish: Late Neanderthal Activities on Gibraltar, in Stringer, C., Barton, R. N. E., and Finlayson, C., eds., Neanderthals on the Edge: Oxford, Oxbow Books, pp. 211220.Google Scholar
Barton, R. N. E., Ford, S., Collcutt, S. N., Crowther, J., Macphail, R. I., Rhodes, E., and Van Gijn, A., 2009, A Final Upper Palaeolithic site at Nea Farm, Somerley, Hampshire and some reflections on the occupation of Britain in the Late Glacial Interstadial: Quartär, v. 56, pp. 735.Google Scholar
Barton, R. N. E., Stringer, C., and Finlayson, C., 2012, Gibraltar Neanderthals in Context: A report of the 1995–98 excavations at Gorham’s & Vanguards Caves, Gibraltar, in Barton, R. N. E., Stringer, C.B. and Finlayson, C., Eds., Neanderthals in Context. A report of the 1995–98 excavations at Gorham’s & Vanguard Caves, Gibraltar. Oxford Committee for Archaoelogy., ed., Oxford University School of Archaeology: Monograph 75: Oxford, Institute of Archaeology, University of Oxford, p. 328.Google Scholar
Bateman, N., Cowan, C., and Wroe-Brown, R., 2008, London’s Roman Amphitheatre: Guildhall Yard, City of London, London, Museum of London Archaeology Service, p. 241.Google Scholar
Bates, M., and Stafford, E., 2013, Thames Holocene. A Geoarchaeological Approach to the Investigation of the River Floodplain for High Speed 1, 1994–2003, Oxford, Oxford Wessex Archaeology.Google Scholar
Bates, M. R., Bates, C. R., Gibbard, P. L., Macphail, R. I., Owen, F. I., Parfitt, S. A., Preece, R. C., Roberts, M. B., Robinson, J. E., Whittaker, J. E., and Wilkinson, K. N., 2000, Late Middle Pleistocene deposits at Norton Farm on the West Sussex coastal plain, southern England: Journal of Quaternary Science, v. 15, no. 1, pp. 6189.3.0.CO;2-K>CrossRefGoogle Scholar
Bates, R. M., Champness, C., Haggart, A., Macphail, R. I., Parfitt, S. A., and Schwenninger, J.-L., 2014, Early Devensian sediments and palaeovenvironmental evidence from the excavations at the Royal Oak Portal Paddington, West London, UK: Proceedings of the Geologists’ Association, v. 125, pp. 4155.Google Scholar
Beckman, G. G., and Smith, K. J., 1974, Micromorphological changes in surface soils following wetting, drying and trampling, in Rutherford, G. K., ed., Soil Microscopy: Kingston, Ontario, The Limestone Press, pp. 832845.Google Scholar
Becze-Deak, J., Langohr, R., and Verrecchia, E. P., 1997, Small scale secondary CaCO3 accumulations in selected sections of the European loess belt; morphological forms and potential for paleoenvironmental reconstruction: Geoderma, v. 76, no. 3–4, pp. 221252.CrossRefGoogle Scholar
Bell, M., 1983, Valley sediments as evidence of prehistoric land use on the South Downs: Proceedings of the Prehistoric Society, v. 49, pp. 118150.CrossRefGoogle Scholar
Bell, M. 1990, Brean Down Excavations 1983–87, London, English Heritage.Google Scholar
Bell, M. 1992, The prehistory of soil erosion, in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 2135.Google Scholar
Bell, M. 2007, Prehistoric Coastal Communities: The Mesolithic in Western Britain, York, Council for British Archaeology, p. 381.Google Scholar
Bell, M., and Boardman, J., 1992, Past and Present Soil Erosion: Oxford, Oxbow.Google Scholar
Bell, M., Caseldine, A., and Neumann, H., 2000, Prehistoric Intertidal Archaeology in the Welsh Severn Estuary, York, Council for British Archaeology.Google Scholar
Bell, M., Fowler, M. J., and Hillson, S. W., 1996, The Experimental Earthwork Project, 1960–1992., York, Council for British Archaeology, Research Report, p. 267.Google Scholar
Bell, M., and Macphail, R. I., 1990, Auger transect and soil pit survey, in Saville, A., ed., Hazleton North, Gloucestershire, 1979–82: The Excavation of a Neolithic Long Cairn of the Cotswold-Severn Group, Archaeological Report no. 13: London, English Heritage, pp. 226227.Google Scholar
Bergadà, M. M., Villaverde, V., and Romàn, D., 2013, Microstratigraphy of the Magdalenian sequences at Cendres Cave (Teulada-Moraira, Alicante, Spain): formation and diagensis: Quaternary International, v. 315 Site formation processes in archaeology, pp. 5675.CrossRefGoogle Scholar
Berger, J.-F., Salvador, P.-G., Franc, O., Verot-Bourrely, A., and Bravard, J.-P., 2008, La chronologie fluviale postglaciaire du Haut Bassin Rhodanien (The Postglacial fluvial chronology of Upper Rhone basin): Cahiers de Paléoenvironnement, v. Collection EDYTEM – n° 6 -, pp. 117–144.CrossRefGoogle Scholar
Berna, F., Behar, A., Shahack-Gross, R., Berg, J., Boaretto, E., Gilboa, A., Sharon, I., Shalev, S., Shilstein, S., Yahalom-Mack, N., Zorn, J. R., and Weiner, S., 2007, Sediments exposed to high temperatures: reconstructing pyrotechnological processes in Late Bronze Age and Iron Age Strata at Tel Dor (Israel): Journal of Archaeological Science, v. 34, pp. 358373.CrossRefGoogle Scholar
Berna, F., Goldberg, P., Horwitzc, L. K., Brinkd, J., Holtd, S., Bamford, M., and Chazang, M., 2012, Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape province, South Africa: Proceedings of the National Academy of Sciences of the United States of America (PNAS), v. 109 no. 13, pp. 16 (Supporting Information 1–8).Google Scholar
Bethell, P. H., and Máté, I., 1989, The use of soil phosphate analysis in archaeology: A critique., in Henderson, J., ed., Scientific Analysis in Archaeology, Monograph No.19: Oxford, Oxford University Committee, pp. 129.Google Scholar
Biddle, M., Hudson, D., and Heighway, C., 1973, The Future of London’s Past: a survey of the archaeological implications of planning and devlopment in the nation’s capital: Rescue: a trust for British Archaeology, v. Rescue Publication 4.Google Scholar
Biddulph, E., and Brown, L., 2011, Prehistoric and Roman evidence., in Ford, B. M., and Teague, S., eds., Winchester – a City in the Making: Oxford, Oxford Archaeology, pp. 34, 3772.Google Scholar
Biddulph, E., Foreman, S., Stafford, E., Stansbie, D., and Nicholson, R., 2012, London Gateway. Iron Age and Roman salt making in the Thames Estuary; Excavations at Stanford Wharf Nature Reserve, Essex, Oxford, Oxford Archaeology.Google Scholar
Bill, J., and Daly, A., 2012, The plundering of the ship graves from Oseberg and Gokstad: an example of power politics?: Antiquity, v. 86, pp. 808824.Google Scholar
Bill, J., and Rødsrud, C., In Press, Heimdalsjordet - trade, production and communication., in Loftsgarden, K., and Glørstad, A. Z., eds., Viking-Age Transformations: Trade, Craft and Resources in Western Scandinavia: Farnham, Ashgate Publishing.Google Scholar
Binder, D., Brochier, J. E., Duday, H., Helmer, D., Marinval, P., Thiebault, S., and Wattez, J., 1993, L’abri Pendimoun à Castellar (Alpes-Maritimes): nouvelles données sur le complex culturel de la imprimée dans son contexte stratigraphique: Gallia Préhistoire, v. 35, pp. 177251.Google Scholar
Binford, L., 1981, Behavioural archaeology and the “Pompeii premise”: Journal of Anthropological Research, v. 37, pp. 195208.Google Scholar
Birkbeck, V., and Schuster, J., 2009, Living and Working in Roman and Later London. Excavations at 60–63 Fenchurch St, Salisbury, Wessex Archaeology, p. 138.Google Scholar
Blackford, J. J., and Chambers, F. M., 1991, Proxy records of climate from blanket mires: evidence for a Dark Age (1400 BP) climatic deterioration in the British Isles: The Holocene, v. 1, pp. 6367.Google Scholar
Blair, I., Barham, E., and Blackmore, L., 2004, My Lord Essex, British Archaeology, Volume 76, CBA, pp. 1017.Google Scholar
Blake, M. E., 1947, Ancient Roman Construction in Italy from the Prehistoric period to Augustus, Washington, Carnegie Institution of Washington.Google Scholar
Blume, H.-P., and Runge, M., 1978, Genese und Okologie innerstadtischer Boden aus Bauschutt: Zeitschrift fur Pflanzenernahrung und Bodenkunde, v. 141, pp. 727740.Google Scholar
Blume, H. P., 1989, Classification of soils in urban agglomerations: Catena, v. 16, pp. 269275.Google Scholar
Boardman, J., 1992, Current erosion on the South Downs: implications for the past, in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion: Oxford, Oxbow, pp. 919.Google Scholar
Boggs, S., 2001, Principles of Sedimentology and Stratigraphy, Upper Saddle River, NJ, Prentice Hall.Google Scholar
Boiffin, J., and Bresson, L. M., 1987, Dynamique de formation des croutes superficielles: apport de l’analyse microscopique, in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 393399.Google Scholar
Boivin, N. L., 1999, Life rythms and floor sequences: excavating time in rural Rajasthan and Neolithic Çatalhöyük: World Archaeology, v. 31, pp. 367388.Google Scholar
Bolt, A. J. J., Mücher, H. J., Sevink, J., and Verstraten, J. M., 1980, A study on loess-derived colluvia in southern Limbourg (the Netherlands): Netherlands Journal of Agricultural Science, v. 28, pp. 110126.CrossRefGoogle Scholar
Boorman, L., Hazelden, J., and Boorman, M., 2002, New salt marshes for old – salt marsh creation and management, in Eurocoast, , ed., Littoral 2002, The Changing Coast: Porto – Portugal, EUCC, pp. 3545.Google Scholar
Borderie, Q., 2011, L’espace urbain entre Antiquité et Moyen Âge analyse géoarchéologique des terres noires études de cas (unpublished PhD Thesis): Paris 1Panthéon – Sorbonne.Google Scholar
Borderie, Q., Devos, Y., Nicosia, C., Cammas, C., and Macphail, R. I., 2014a, Chapter 18. Dark Earth in the geoarchaeological approach to urban contexts, in Arnaud-Fassetta, G., and Carcaud, N., eds., French geoarchaeology in the 21st century: Paris, CNRS, pp. 245258.Google Scholar
Borderie, Q., Fondrillon, M., Nicosia, C., Devos, Y., and Macphail, R. I., 2014b, Bilan des recherches et nouveaux éclairages sur les terres noires: des processus complexes de stratification aux modalités d’occupation des espaces urbains., in Lorans, E., ed., Archéologie de l’espace urbain – Partie II: Tours, CTHS, pp. 213223.Google Scholar
Boschian, G., 1997, Sedimentology and soil micromorphology of the late Pleistocene and early Holocene deposits of Grotta dell’Edera (Trieste Karst, NE Italy): Geoarchaeology 12, 227–250., v. 12, no. 3, pp. 227250.Google Scholar
Boschian, G., and Montagnari-Kokelji, E., 2000, Prehistoric shepherds and caves in the Trieste Karst (northeastern Italy): Geoarchaeology, v. 15, no. 4, pp. 331371.Google Scholar
Bouma, J., Fox, C. A., and Miedema, R., 1990, Micromorphology of hydromorphic soils: applications for soil genesis and land evaluation, in Douglas, L. A., ed., Soil Micromorphology: A Basic and Applied Science, Developments in Soil Science 19: Amsterdam, Elsevier, pp. 257278.Google Scholar
Bowsher, D., Holder, N., Howell, I., and Dyson, T., 2007, The London Guildhall: The Archaeology and History of the Guildhall Precint from the Medieval Period to the 20th Century: London, Museum of London Archaeological Service, p. 536.Google Scholar
Bradley, R., 1971, Stock Raising and the Origins of the Hill Fort on the South Downs: The Antiquaries Journal, v. 51, no. 1, pp. 829.Google Scholar
Bradley, R., and Keith-Lucas, M., 1975, Excavations and pollen analysis on a bell barrow at Ascot, Berkshire: Journal of Archaeological Science, v. 2, pp. 95108.Google Scholar
Brady, N. C., and Weil, R. R., 2008, The Nature and Properties of Soils, New Jersey, Prentis Education, p. 965.Google Scholar
Brammer, H., 1971, Coatings in seasonal flooded soils: Geoderma, v. 6, pp. 516.CrossRefGoogle Scholar
Breuning-Madsen, H., Holst, M. K., and Rasmussen, M., 2001, The chemical environment in a burial mound shortly after construction – an archaeological-pedological experiment: Journal of Archaeological Science, v. 28, pp. 691697.Google Scholar
Breuning-Madsen, H., Holst, M. K., Rasmussen, M., and Elberling, B., 2003, Preserved within log coffins before and after barrow contruction: Journal of Archaeological Science, v. 30, pp. 343350.Google Scholar
Brewer, R., 1964, Fabric and Mineral Analysis of Soils, New York, Wiley and Sons.Google Scholar
Brewer, R. 1976, Fabric and Mineral Analysis of Soils, Huntington, NY, Robert E. Kreiger Publishing Co.Google Scholar
Briard, J., and Monnier, J. L., 1976, Tumulus Amoricains de l’Age du Bronze et couverture loessique Weichselienne: Bulletin de la Societé Minéralogie Bretagne, v. C, VIII, no. 1/2, pp. 7588.Google Scholar
Briogioli, G. P., Cremaschi, M., and Gelichi, S., 1988, Orocessi di stratificazione in centri urbani dell’Italia settentrionale: Archeologia stratigrafica, v. 1, pp. 2330.Google Scholar
Brochier, J. E., 1983, Bergeries et feux néolithiques dans le Midi de la France, caractérisation et incidence sur la raisonnement sédimentologique: Quatar, v. 33/34, pp. 181193.Google Scholar
Brochier, J. E. 1996, Feuilles ou fumiers? Observations sur le rôle des poussières sphérolitiques dans l’interprétation des dépôts archéologiques Holocènes: Anthrozoologica, v. 24, pp. 1930.Google Scholar
Brochier, J. E., and Thinon, M., 2003, Calcite crystals, starch grains aggregates or – POCC? Comment on ‘calcite crystals inside archaeological plant tissues’: Journal of Archaeological Science, v. 30, pp. 12111214.Google Scholar
Brochier, J. E., Villa, P., and Giacomarra, M., 1992, Shepherds and Sediments: geo-ethnoarchaeology of pastoral sites: Journal of Anthropological Archaeology, v. 11, pp. 47102.Google Scholar
Brown, A. G., 1997, Alluvial Geoarchaeology. Floodplain archaeology and environmental change, Cambridge, Cambridge University Press, Manuals in Archaeology.CrossRefGoogle Scholar
Brown, G., 1988, Whittington Ave (WIV88): London, Museum of London Archaeological Service.Google Scholar
Brown, G. E., 1990, Testing of concretes, plasters, and stuccos: Archaeomaterials, v. 4, pp. 185191.Google Scholar
Brunborga, L. A., Julshamna, K., Nortvedta, R., and Frøylanda, L., 2006, Nutritional composition of blubber and meat of hooded seal (Cystophora cristata) and harp seal (Phagophilus groenlandicus) from Greenland: Food Chemistry, v. 96, no. 4, pp. 524531.Google Scholar
Bullock, P., 1974a, Micromorphology, in Avery, B. W., and Bascomb, C. L., eds., Soil Survey Laboratory Methods, Technical Monograph No. 6: Harpenden, Soil Survey, pp. 7081.Google Scholar
Bullock, P. 1974b, The use of soil micromorphology in the new system of soil classification for England and Wales, in Rutherford, G. K., ed., Soil Microscopy: Kingston, Ontario, The Limestone Press, pp. 607631.Google Scholar
Bullock, P., Fedoroff, N., Jongerius, A., Stoops, G., and Tursina, T., 1985, Handbook for Soil Thin Section Description, Wolverhampton, Waine Research Publications, p. 152.Google Scholar
Bullock, P., and Gregory, P. J., 1991, Soils in the Urban Environment: Oxford, Blackwell Scientific Publications.Google Scholar
Bullock, P., and Murphy, C. P., 1979, Evolution of a paleo-argillic brown earth (Paleudalf) from Oxfordshire, England: Geoderma, v. 22, pp. 225252.Google Scholar
Bullock, P. 1983, Soil Micromorphology (Volumes 1 and 2): Berkhamsted, A B Academic Publishers.Google Scholar
Burch, M., Treveil, P., and Keene, D., 2011, The development of early medieval and later Poultry and Cheapside. Excavations at 1 Poultry and vicinity, City of London, London, Museum of London Archaeology, p. 365.Google Scholar
Burnham, C. P., and Pitman, J. I., 1986, Soil Erosion, The Journal of the South East Soils Discussion Group, Volume 3: Ashford, South East Soils Discussion Group, p. 105.Google Scholar
Butzer, K. W., 1982, Archaeology as Human Ecology: Method and Theory for a Contextual Approach, Cambridge, Cambridge University Press.Google Scholar
Cammas, C., 1994, Approche micromorphologique de la stratigraphie urbaine à Lattes: premiers résultats, Lattara 7: Lattes, A R A L O, pp. 181202.Google Scholar
Cammas, C. 2004, Les “terre noires” urbaines du Nord de la France: première typologie pédo-sédimentaire., in Verslype, L., and Brulet, R., eds., Terres Noires – Dark Earth: Louvain-la-Neuve, Université Catholique de Louvain, pp. 4355.Google Scholar
Cammas, C., David, C., and Guyard, L., 1996a, La question des terre noires dans les sites tardo-antiques et médiéval: le cas du Collège de France (Paris, France), Proceedings XIII International Congress of Prehistoric and Protohistoric Sciences, Colloquim 14: Forlì, ABACO, pp. 8993.Google Scholar
Cammas, C., Wattez, J., and Courty, M.-A., 1996b, L’enregistrement sédimentaire des modes d’occupation de l’espace, in Castelletti, L., and Cremaschi, M., eds., Paleoecology; Colloquium a3 of XIII International Congress of Prehistoric and Protohistoric Sciences, Volume 3: Forlì, ABACO, pp. 8186.Google Scholar
Campbell, G., and Robinson, M., 2007, Environment and Land Use in the Valley Bottom, in Healy, F., and Harding, J., eds., The Raunds Area Project. A Neolithic and Bronze Age Landscape in Northamptonshire Swindon, English Heritage, pp. 1836.Google Scholar
Canti, M., 1997, An investigation into microscopic calcareous spherulites from herbivore dung: Journal of Archaeological Science, v. 24, pp. 435444.Google Scholar
Canti, M. 1998a, Origin of calcium carbonate granules found in buried soils and Quaternary deposits: Boreas, v. 27, pp. 275288.Google Scholar
Canti, M. 1999, The production and preservation of faecal spherulites: animals, environment and taphonomy: Journal of Archaeological Science, v. 26, pp. 251258.CrossRefGoogle Scholar
Canti, M. 2000, Mineralogical analysis of samples from Buildings 1, 6, and surroundings., in Bell, M., Caseldine, A., and Neumann, H., eds., Prehistoric Intertidal Archaeology in the Welsh Severn Estuary, Research Report 120: York, Council for British Archaeology, pp. 269270.Google Scholar
Canti, M. 2003, Aspects of the chemical and microscopic characteristics of plant ashes found in archaeological soils, in Stoops, G., ed., Catena, Special Issue 54, 3. Achievements in Micromorphology: Amsterdam, pp. 339361.Google Scholar
Canti, M. G., 1998b, The micromorphological identification of feacal spherulites from archaeological and modern materials: Journal of Archaeological Science, v. 25, pp. 435444.Google Scholar
Carneiro, Â., and Mateiciucová, I., 2007, Daub fragments and the question of structures, in Whittle, A., ed., The Early Neolithic on the Great Hungarian Plain: investigations of the Körös culture site of Ecsegfalva 23, Co. Békés, Volume I: Budapest, Institute of Archaeology, pp. 255288.Google Scholar
Carpentier, F., 2015, Minoans under the microscope. Archaeological soil micromorphology at the Cretan Bronze Age site of Sissi: KU Leuven, p. 336.Google Scholar
Carruthers, W. J., 2000, Mineralised plant remains, in Lawson, A. J., ed., Potterne 1982–5. Animal Husbandry in Later Prehistoric Wiltshire, Wessex Archaeology Report No. 17: Salisbury, Wessex Archaeology, pp. 7284.Google Scholar
Carter, S., 1987, The Reconstruction of land-Snail Death Assemblages: University of London.Google Scholar
Carter, S. 1998a, Soil micromorphology, in Lowe, C., ed., St. Boniface Church, Orkney: coastal erosion and archaeological assessment: Stroud, Sutton Publishing/Historic Scotland, pp. 172186.Google Scholar
Carter, S. 1998b, The use of peat and other organic sediments as fuel in northern Scotland: identifications derived from soil thin sections, in Coles, G. and Mills, C.M. eds., Life on the Edge: Human Settlement and Marginality, Monograph 100: Oxford, Oxbow, pp. 99104.Google Scholar
Carter, S., and Davidson, D., 1998, An evaluation of the contribution of soil micromorphology to the study of ancient arable cultivation: Geoarchaeology, v. 13, pp. 535547.3.0.CO;2-#>CrossRefGoogle Scholar
Carter, S. P., 1990, The stratification and taphonomy of shells in calcareous soils: implications for landsnail analysis in archaeology: Journal of Archaeological Science, v. 17, pp. 495507.Google Scholar
Carver, M. O. H., 1987, The nature of urban deposits, in Schofield, J., and Leech, R., eds., Urban Archaeology in Britain, CBA Research Report 61: York, Council for British Archaeology, pp. 926.Google Scholar
Carver, M. O. H. 1998, Sutton Hoo. Burial ground of Kings?, London, British Museum Press.Google Scholar
Carver, R. E., 1971, Procedures in Sedimentary Petrology, New York, Wiley-Interscience, p. 653.Google Scholar
Catt, J. A., 1986, Soils and Quaternary Geology. A Handbook for Field Scientists, Oxford, Clarendon Press.Google Scholar
Catt, J. A. 1999, Particle size distribution and mineralogy of the deposits, in Roberts, M. B., and Parfitt, S. A., eds., Boxgrove. A Middle Pleistocene Hominid Site at Eartham Quarry, Boxgrove, West Sussex, Archaeological Report 17: London, English Heritage, pp. 111118.Google Scholar
Catt, J. A., and Staines, S. J., 1998, Petrography of sediments and soils, in Preece, R. C., and Bridgland, D. R., eds., Late Quaternary Environmental Change in North-West Europe: Excavations at Holywell Coombe, South-east England: London, Chapman & Hall, pp. 6985.Google Scholar
Chepstow-Lusty, A., 2011, Agro-pastoralism and social change in the Cuzco heartland of Peru: a brief history using environmental proxies, Antiquity, v. 85, no. 328, pp. 570582.Google Scholar
Ciezar, P., Gonzalez, V., Pieters, M., Rodet-Belarbi, I., and Van-Ossel, P., 1994, In suburbano – new data on the immediate surroundings of Roman and early medieval Paris, in Hall, A. R., and Kenward, H. K., eds., Urban-Rural Connexions: Perspectives from Environmental Archaeology: Oxford, Oxford: Oxbow Books, pp. 137146.Google Scholar
Clark, A., 2000, Seeing beneath the Soil: Prospecting Methods in Archaeology. New edition London, Routledge.Google Scholar
Collins, J. E., and Larney, F. J., 1987, Micromorphological observations of compacted horizons (cultivation pans) from various horizons in Irish tillage soils., in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 451457.Google Scholar
Connell, E. R., J., E. K., and Hall, A. M., 1982, Evidence for two pre-Flandrian palaeosols in Buchan, Scotland: Nature, v. 297, pp. 570572.Google Scholar
Connell, E. R., and Romans, J. C. C., 1984, Kirkhill. Palaeosols., in Hall, A. M., ed., Buchan Field Guide: Cambridge, Quaternary Research Association, pp. 7076.Google Scholar
Conry, M. J., 1971, Irish Plaggen soils, their distribution, origin and properties: Journal of Soil Science, v. 22, pp. 401416.Google Scholar
Conway, J. S., 1983, An investigation of soil phosphorus distribution within occupation deposits from a Romano-British hut group: Journal of Archaeological Science, v. 10, pp. 117128.Google Scholar
Cooke, N., Brown, F., and Phillpotts, C., 2008, From Hunter Gatherers to Huntsmen. A History of the Stansted Landscape, Oxford/Salisbury, Framework Archaeology.Google Scholar
Cornwall, I. W., 1953, Soils for the Archaeologist, New York, The Macmillan Company.Google Scholar
Courty, M.-A., 1992, Soil micromorphology in archaeology, in Pollard, M., ed., New Developments in Archaeological Science: Oxford, Oxford University Press, pp. 3962.Google Scholar
Courty, M.-A., Cachier, H., Hardy, M., and Ruellan, S., 1998, Soil record of exceptional wild-fires linked to climatic anomalies (Inter-Tropical and Mediterranean regions), in N. Fedoroff, and Catt, J. A., eds., Proceedings World Congress of Soil Science August 1998, CD-ROM: Montpellier, International Society of Soil Science, Montpellier.Google Scholar
Courty, M.-A., Macphail, R. I., and Wattez, J., 1991, Soil micromorphological indicators of pastoralism: with special reference to Arene Candide, Finale Ligure, Italy: Revista di Studi di Liguri, v. 57, pp. 127150.Google Scholar
Courty, M. A., 1984, Formation et évolution des accumulations cendreuses: approche micromorphologique, Actes du Colloque Interrégional sur le Néolithique, 1981: Le Puy, pp. 341353.Google Scholar
Courty, M. A. 2001, Microfacies analysis assisting archaeological stratigraphy, in Goldberg, P., Holliday, V. T., and Ferring, C. R., eds., Earth Sciences and Archaeology: New York, Kluwer, pp. 205239.Google Scholar
Courty, M. A. 2012, Ancestral processing of exceptional organo-mineral materials: microfacies and multi-analytical study, in Poch, R. M., Casamitjana, M., and Francis, M. L., eds., Proceedings of the 14th International Working Meeting on Soil Micromorphology – Lleida 8–14 July 2012: Lleida, Universitat of Lleida and International Union of Soil Sciences, pp. 321325.Google Scholar
Courty, M. A., and Fedoroff, N., 1982, Micromorphology of a Holocene dwelling, Proceedings Nordic Archaeometry, PACT 7, pp. 257277.Google Scholar
Courty, M. A., Goldberg, P., and Macphail, R. I., 1989, Soils and Micromorphology in Archaeology, Cambridge, Cambridge University Press, Cambridge Manuals in Archaeology, p. 344.Google Scholar
Courty, M. A. 1994, Ancient people – lifestyles and cultural patterns, in Etchevers, J. D., ed., Transactions of the 15th World Congress of Soil Science, International Society of Soil Science, Mexico, Volume 6a: Acapulco, International Society of Soil Science, pp. 250269.Google Scholar
Courty, M. A., and Nørnberg, P., 1985, Comparison between buried uncultivated and cultivated Iron Age soils on the west coast of Jutland, Denmark, in Edgren, T., and Jungner, H., eds., Proceedings of the Third Nordic Conference on the Application of Scientific Methods in Archaeology: Helsinki, The Finnish Antiquarian Society, pp. 5769.Google Scholar
Courty, M. A., and Vallverdu, J., 2001, The microscopic record of Abrupt Climate Changes in cave sediments of the Western Mediterranean: Geoarchaeology, v. 16, no. 5, pp. 467500.Google Scholar
Couts, J., Osborn, A., and Edwards, A., 1930, The Complete Book of Gardening (Royal Botanic Gardens, Kew), London, Ward, Lock & Co, p. 768.Google Scholar
Cowan, C., 2003, Urban development in north-west Roman Southwark: Excavations 1974–90, Monograph 16: London, MOLAS, p. 209.Google Scholar
Cowgill, J., 2003, The iron production industry and its extensive demand upon woodland resources: a case study from Creeton Quarry, Lincolnshire, in Murphy, P., and Wiltshire, P. E. J., eds., The Environmental Archaeology of Industry, Symposia of the Association for Environmental Archaeology No. 20: Oxford, Oxbow, pp. 4857.Google Scholar
Crampton, C. B., 1963, The development and morphology of iron pan podzols in Mid and South Wales: Journal of Soil Science, v. 14, pp. 282302.Google Scholar
Cremaschi, M., Di Lernia, S., and Trombino, L., 1996, From taming to pastoralism in a drying environment. Site formation processes in the shelter of the Tadrat Massif (Libya, Central Sahara), in Castelletti, L., and Cremaschi, M., eds., Paleoecology, Proceedings of Int. Union of Prehistoric and Protohistoric Sciences: Forlì, ABCO, pp. 87106.Google Scholar
Crombé, P., 1993, Tree-fall features on Final Palaeolithic and Mesolithic sites situated on sandy soils: how to deal with it: Helenium, v. XXXIII, no. I, pp. 5066.Google Scholar
Crombé, P., Langohr, R., and Louwagie, G., 2015, Mesolithic hearth-pits: fact or fantasy? A reassessment based on the evidence from the sites of Doel and Verrebroek (Belgium): Journal of Archaeological Science, 61, pp. 158–171.Google Scholar
Crowther, J., 1996a, Phosphate migration around buried bones, in Bell, M., Fowler, P. J., and Hillson, S. W., eds., The Experimental Earthwork Project 1960–1992, Research Report 100: York, Council for British Archaeology, pp. 195196.Google Scholar
Crowther, J. 1996b, Report on sediments from Building 5, Old Market Street “86,” in Marvell, A. G., ed., “Excavations at Usk 1986–1988 (Brittania XXVII, 51–110): Britannia, v. XXVII, pp. 9299.Google Scholar
Crowther, J. 1996c, Soil chemistry, in Bell, M., Fowler, P. J, and Hillson, S. W.., eds., The Experimental Earthwork Project 1960–1992, Research report 100: York, Council for British Archaeology, pp. 107118.Google Scholar
Crowther, J. 1997, Soil phosphate surveys: critical approaches to sampling, analysis and interpretation: Archaeological Prospection, v. 4, pp. 93102.Google Scholar
Crowther, J. 2000, Phosphate and magnetic susceptibility studies, in Bell, M., Caseldine, A., and Neumann, H., eds., Prehistoric Intertidal Archaeology in the Welsh Severn Estuary, York, CBA Research report 120, pp. 5758 (and CD).Google Scholar
Crowther, J. 2003, Potential magnetic susceptibility and fractional conversion studies of archaeological soils and sediments: Archaeometry, v. 45, no. 4, pp. 685701.Google Scholar
Crowther, J. 2007, Chemical and magnetic properties of soils and pit fills, in Whittle, A., ed., The Early Neolithic on the Great Hungarian Plain: investigations of the Körös culture site of Ecsegfalva 23, Co. Békés., Volume I: Budapest, Institute of Archaeology, pp. 227254.Google Scholar
Crowther, J. 2014, Chemistry, magnetic susceptibility and particle size of various contexts from the Marco Gonzalez Mayan site, Belize (report for E. Graham, Institute of Archaeology, University College London): Archaeological Services, University of Wales: Trinity Saint David.Google Scholar
Crowther, J., and Barker, P., 1995, Magnetic susceptibility: distinguishing anthropogenic effects from the natural: Archaeological Prospection, v. 2, pp. 207215.Google Scholar
Crowther, J., Macphail, R. I., and Cruise, G. M., 1996, Short-term burial change in a humic rendzina, Overton Down Experimental Earthwork, Wiltshire, England: Geoarchaeology, v. 11(2), pp. 95117.Google Scholar
Cruise, G. M., 1990, Holocene peat initiation in the Ligurian Appenines, northern Italy: Review of Palaeobotany and Palynology, v. 63, pp. 173182.Google Scholar
Cruise, G. M. 2000, Microstratigraphical Signatures of Experimental Rural Occupation Deposits and Archaeological Sites, in Roskams, S., ed., Interpreting Stratigraphy, Volume 9: York, University o f York, pp. 183191.Google Scholar
Cruise, G. M., Macphail, R. I., Linderholm, J., Maggi, R., and Marshall, P. D., 2009, Lago di Bargone, Liguria, N. Italy: a reconstruction of Holocene environmental and land-use history: The Holocene, v. 19, no. 7, pp. 9871003.Google Scholar
Crummy, P., 1992, Excavations at Culver St, the Giberd School and and other sites in Colchester., Colchester Archaeological Report, Volume No 6: Colchester, Colchester Archaeological Trust, pp. 295307.Google Scholar
Dalwood, H., and Edwards, R., 2004, Excavations at Deansway, Worcester, 1988–89: Romano-British small town to late medieval city, CBA Research Report No 139: York, Council for British Archaeology, p. 605.Google Scholar
Dammers, K., and Joergensen, R. G., 1996, Progressive loss of Carbon and Nitrogen from simulated daub on heating: Journal of Archaeological Science, v. 23, pp. 639648.Google Scholar
Davidson, D. A., Carter, S., Boag, B., Long, D., Tipping, R., and Tyler, A., 1999, Analysis of pollen in soils: processes of incorporation and redistribution of pollen in five soil profile types: Soil Biology & Chemistry, v. 31, pp. 643653.Google Scholar
De Coninck, F., 1980, Major mechanisms in formation of spodic horizons: Geoderma, v. 24, pp. 101128.Google Scholar
De Coninck, F., and Righi, D., 1983, Podzolisation and the spodic horizon, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology, Volume 2: Soil Genesis: Berkhamsted, A B Academic Publishers, pp. 389417.Google Scholar
Deer, W. A., Howie, R. A., and Zussman, J., 1992, An Introduction to the Rock-Forming Minerals (2nd Edition), Harlow, Longman, p. 696.Google Scholar
Dejmal, M., Lisá, L., Fišáková Nývltová, M., Bajer, A., and Petr, L., 2014, Medieval Horse Stable; The Results of Multi Proxy Interdisciplinary Research, PLoS ONE Volume 9.Google Scholar
Dennell, R., 2008, The Palaeolithic Settlement of Asia, Cambridge, Cambridge University Press, Cambridge World Archaeology, p. 572.Google Scholar
Denny, C. S., and Goodlett, J. C., 1956, Microrelief resulting from fallen trees, in Dennny, C. S., ed., Surficial Geology and Geomorphology of Potter County, Volume 288, US Geological Survey professional Paper, pp. 5965.Google Scholar
Denyer, P., 1996–97, Geology of Naval Station Guantanamo Bay, Cuba (A Supplemental technical report to the Rapid Ecological ssessment): University of Costa Rica.Google Scholar
Devos, Y., Nicosia, C., Vrydaghs, L., and Modrie, S., 2013a, Studying urban stratigraphy: Dark Earth and a microstratified sequence on the site of the Court of Hoogstraeten (Brussels, Belgium). Integrating archaeopedology and phytolith analysis: Quaternary International, v. 315, pp. 147166.Google Scholar
Devos, Y., Vrydaghs, L., A., D., and Fechner, K., 2009, An archaeopedological and phytolitarian study of the “Dark Earth” on the Site of Rue de Dinant (Brussels, Belgium): Catena, v. 78, no. 3, pp. 270284.Google Scholar
Devos, Y., Wouters, B., Vrydaghs, L., Tys, D., Bellens, T., and Schryvers, A., 2013, A soil micromorphological study on the origins of the early medieval trading centre of Antwerp (Belgium): Quaternary International, v. 315, pp. 167183.Google Scholar
Dexter, A. R., 1976, Internal structure of tilled soil: Journal of Soil Science, v. 27, pp. 267278.Google Scholar
Dexter, A. R. 1979, Prediction of soil structures produced by tillage: Journal of Terramechanics, v. 16, no. 3, pp. 117127.Google Scholar
Dimbleby, G. W., 1962, The Development of British Heathlands and their Soils, Oxford, Clarendon Press, p. 121.Google Scholar
Dimbleby, G. W. 1985, The Palynology of Archaeological Sites, London, Academic Press.Google Scholar
Dimbleby, G. W., and Evans, J. G., 1974, Pollen and land snail analysis of calcareous soils: Journal of Archaeological Science, v. 1, pp. 117133.Google Scholar
Dimbleby, G. W., and Gill, J. M., 1955, The occurrence of podzols under deciduous woodland in the New Forest: Forestry, v. 8, pp. 95106.Google Scholar
Dinç, U., Miedema, R., Bal, L., and Pons, L. J., 1976, Morphological and physio-chemical aspects of three soils developed in peat in The Netherlands and their classification: Netherlands Journal of Agricultural Science, v. 24, pp. 247265.Google Scholar
Dockrill, S. J., Bond, J. M., Milles, A., Simpson, I. I., and Ambers, J., 1994, Tofts Ness, Sanday, Orkney. An integrated study of a buried Orcadian landscape, in Luff, R., and Rowley-Conwy, P., eds., Whither Environmental Archaeology, Oxbow Monograph 38: Oxford, Oxbow, pp. 115132.Google Scholar
Dockrill, S. J., and Simpson, I., 1994, The identification of prehistoric anthropogenic soils in the Northern Isles using an integrated sampling methodology: Archaeological Prospection, v. 1, pp. 7792.Google Scholar
Douglas, L. A., 1990, Soil Micromorphology: A Basic and Applied Science, Proceedings of the VIIIth International Working Meeting of Soil Micromorphology, San Antonio, Texas – July 1988: Amsterdam, Elsevier.Google Scholar
Douglas, L. A., and Thompson, M. L., 1985, Soil micromorphology and soil classification, Soil Science Society of America, p. 216.Google Scholar
Drescher, H. E., Harms, U., and Huschenbeth, E., 1977, Organochlorines and heavy metals in the harbour seal Phoca vitulina from the German North Sea Coast, Marine Biology, v. 1, pp. 99106.Google Scholar
Drewett, P., 1976, The excavation of four round barrows of the second millenium BC at West Heath, harting, 1973–75: Sussex Archaeological Collections, v. 14, pp. 126150.Google Scholar
Drewett, P. L., 1989, Anthropogenic soil erosion in prehistoric Sussex: excavations at West Heath and Ferring, 1984: Sussex Archaeological Collections, v. 127, pp. 1129.Google Scholar
Duchaufour, P., 1982, Pedology, London, Allen and Unwin, p. 448.Google Scholar
Dunn, R. K., and Mazzullo, S. J., 1993, Holocene paleocoastal reconstruction and its relationship to Marco Gonzalez, Ambergris Caye, Belize: Journal of Field Archaeology v. 20, no. 2, pp. 121131.Google Scholar
Durand, N., Monger, H. C., and Canti, M., 2010, Calcium carbonate features, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 149194.Google Scholar
Ellis, J. C., and Rawlings, M., 2001, Excavations at Balksbury Camp, Andover 1995–97: Hampshire Field Club and Archaeological Society, v. 56, pp. 2193.Google Scholar
Engelmark, R., 1985, Carbonized seeds in postholes – a reflection of human activity, in Edgren, T., and Jungner, H., eds., Proceedings of the Third Nordic Conference on the Application of Scientific Methods in Archaeology. 57–69: Helsinki, The Finnish Antiquarian Society, pp. 205209.Google Scholar
Engelmark, R., and Linderholm, J., 1996, Prehistoric land management and cultivation. A soil chemical study, in Mejdahl, V., and Siemen, P., eds., Proceedings from the 6th Nordic Conference on the Application of Scientific Methods in Archaeology, Esbjerg 1993, Arkaeologiske Rapporter Number 1: Esbjerg, Esbjerg Museum, pp. 315322.Google Scholar
Engelmark, R., and Viklund, K., 1986, Järnålders-jordbruk in Norrland – Teori och praktik: Populär Arkeologi, v. 4, no. 2, pp. 2224.Google Scholar
Engen, T., and Bill, J., 2015, Rapport Arkeologisk utgravning. Viking Museum, University of Oslo.Google Scholar
Evans, E. E., 1957, Irish Folk Ways, London, Routledge and Kegan Paul.Google Scholar
Evans, J. G., 1971, Habitat changes on the calcareous soils of Britain: the impact of Neolithic man, in Simpson, D. D. A., ed., Economy and Settlement in Neolithic and Early Bronze Age Britain and Europe: Leicester, Leicester University Press, pp. 2774.Google Scholar
Evans, J. G. 1972, Land Snails in Archaeology, London, Seminar Press, p. 436.Google Scholar
Evans, J. G. 1990, Notes on some Late Neolithic and Bronze Age events in long barrow ditches in southern and eastern England: Proceedings of the Prehistoric Society, v. 56, pp. 111116.Google Scholar
Evans, J. G., and Limbrey, S., 1974, The experimental earthwork on Morden Bog, Wareham, Dorset, England: 1963–1972: Proceedings of the Prehistoric Society, v. 40, pp. 170202.Google Scholar
Evans, J. G., Limbrey, S., and Macphail, R. I., 2007, The Environmental setting, in Benson, D., and Whittle, A., eds., Building Memories. The Neolithic Cotswold Long Barrow at Ascot-Under-Wychwood, Oxfordshire: Oxford, Oxbow, pp. 5577.Google Scholar
Evans, R., 1992, Erosion in England and Wales – the present the key to the past, in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 5366.Google Scholar
Evershed, R. P., Bethell, P. H., and Walsh, N. J., 1997, 5ß-stigmastanol and related 5ß-stanols as biomarkers of manuring: analysis of modern experimental material and assessment of the archaeological potential: Journal of Archaeological Science, v. 24, pp. 485495.Google Scholar
Farres, P. J., Wood, S. J., and Seeliger, S., 1992, A conceptual model of soil deposition and its implications for environmental reconstruction, in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 217226.Google Scholar
Feathers, J. K., Rhodes, E. J., Huot, S., and McAvoy, J. M., 2006, Luminescence dating of Pre-Clovis strata at the Cactus Hill site, Virginia, USA: Quaternary Geochronology, v. 1, pp. 167187.CrossRefGoogle Scholar
Fechner, K., Baes, R., Louwagie, G., and Gebhardt, A., 2014, Relic Holocene buried colluvial and alluvial deposition in the basins of the Scheldt, the Meuse, the Seine and the Rhine (Belgium, Luxembourg and Northern France). A prospective state of research in rescue excavations avec la collaboration de Deschodt L., Bécu B., Schartz E. Dans: Meylemans E., Poesen J., In’t Ven I., eds., 2014. The Archaeology of Erosion, the Erosion of Archaeology. Conference Brussels, 28–30 avril 2008. Relicta Monographien 9 (VIOE, Brussels): pp. 147–190.Google Scholar
Fedoroff, N., 1974, classification of accumulations of translocated particles, in Rutherford, G. K., ed., Soil Microscopy: Kingston, Ontario, The Limestone Press, pp. 695714.Google Scholar
Fedoroff, N. 1982, Soil fabric at the microscopic level, in Bonneau, M., and Souchier, B., eds., Constituents and Properties of Soils: London, Academic Press, pp. 288303.Google Scholar
Fedoroff, N., Bresson, L. M., and Courty, M.-A., 1987, Micromorphologie des Sols – Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol.Google Scholar
Fedoroff, N., Courty, M. A., and Guo, Z., 2010a, Palaeosols and Relict Soils, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 623662.Google Scholar
Fedoroff, N., Courty, M. A., Guo, Z., and Rue, M., 2010b, Soil reactions to extreme environmental stress: lessons from the past records, 19th World Congress of Soil Science, Soil Solutions for a Changing World: Brisbane, Australia, World Congres of Soil Science, pp. 2326.Google Scholar
Fedoroff, N., Courty, M. A., and Thompson, M. L., 1990, Micromorphological evidence of palaeoenvironmental change in Pleistocene and Holocene Paleosols, in Douglas, L. A., ed., Soil Micromorphology: a Basic and Applied Science, Developments in Soil Science 19: Amsterdam, Elsevier, pp. 653666.Google Scholar
Fedoroff, N., De Kimpe, C. R., Page, F., and Bourbeau, G., 1981, Essai d’interpretation des transferts sous forme figurée dans les podzols du Québec meridional à partir de l’étude micromorphologique des profils: Geoderma, v. 26, pp. 2545.Google Scholar
Fedoroff, N., and Goldberg, P., 1982, Comparative micromorphology of two late Pleistocene palaeosols (in the Paris basin): Catena, v. 9, pp. 227251.Google Scholar
Fenton, A., 1968, Alternating stone and turf – an obsolete building practice: Folk Life, v. 6, pp. 94103.Google Scholar
Fenton, A. J., 1981, Early manuring techniques, in Mercer, R., ed., Farming Practice in Britsh Prehistory: Edinburgh, Edinburgh University Press, pp. 210217.Google Scholar
Ferring, C. R., 2001, Geoarchaeology in alluvial landscapes, in Goldberg, P., Holliday, V. T., and Ferring, C. R., eds., Earth Sciences and Archaeology: New York, Kluwer, pp. 77106.Google Scholar
Findlay, D. C., Colborne, G. J. N., Cope, D. W., Harrod, T. R., Hogan, D. V., and Staines, S. J., 1984, Soils and their use in South West England, Harpenden, Lawes Agricultural Trust, Soil Survey of England and Wales.Google Scholar
Findlay, D. C., Colbourne, G. J. N., cope, D. W., Harrod, T. R., Hogan, D. V., and Staines, S. J., 1983, Soils of England and Wales. Sheet 5, South West England: Ordnance Survey.Google Scholar
Fisher, P. F., and Macphail, R. I., 1985, Studies of archaeological soils and deposits by micromorphological techniques, in Fieller, N. R. J., Gilbertson, D. D., and Ralph, N. G. A., eds., Palaeoenvironmental Investigations: Research Design, Methods and Data Analysis, 258: Oxford, British Archaeological Reports International Series, pp. 92112.Google Scholar
FitzPatrick, E. A., 1956, An indurated soil horizon formed by permafrost: Journal of Soil Science, v. 7, pp. 248254.Google Scholar
FitzPatrick, E. A. 1984, Micromorphology of Soils, London, Chapman and Hall.Google Scholar
FitzPatrick, E. A., 1993, Soil Microscopy and Micromorphology, Chichester, J. Wiley and Sons.Google Scholar
Fondrillon, M., 2007, La formation du sol urbain: étude archéologique des terres noires à Tours (4e-12e siècle): Université François Rabelais Tours, p. 320.Google Scholar
Fondrillon, M., Devos, Y., Graz, Y., Laurent, C., Macphail, R., and Vrydaghs, L., 2009, Processus d’urbanisation au Moyen-Age: recherches interdisciplinaires sur les terres noires à Bruxelles et à Tours (5e-14e siècle)., XVIIe colloque du GMPCA, Archéométrie: Resources, sociétés, biodiversité, Montpellier, France; 05/2009.Google Scholar
Ford, B. M., and Teague, S., 2011, Winchester – a City in the Making, Oxford Archaeology Monograph No 12: Oxford, Oxford Archaeology, p. 402.Google Scholar
Foreman, S., Hiller, J., and Petts, D., 2002, Gathering the People, Settling the Land. The Archaeology of a Middle Thames Landscape. Anglo-Saxon to Post-Medieval, Thames Valley Landscapes monograph No. 14: Oxford, Oxford Archaeology.Google Scholar
Fowler, P. J., 2002, Farming in the first millennium AD: British agriculture between Julius Caesar and William the Conqueror, Cambridge, Cambridge University Press.Google Scholar
Fox, C. A., 1985, Micromorphological characterisation of histosols, in Douglas, L. A., and Thompson, R., eds., Soil Micromorphology and Soil Classification, Special Publication Number 15: Madison, Wisconsin, Soil Science Society of America, pp. 85104.Google Scholar
Francheschi, V. R., and Horner, H. T., 1980, Calcium oxalate crystals in plants: The Botanical review, v. 46, no. 4, pp. 239250.Google Scholar
French, C., 1998, Soils and sediments, in Pryor, F., ed., Etton. Excavations at a Neolithic causewayed enclosure near Maxey, Cambridgeshire, 1982–7, Archaeological Report 18: London, English Heritage, pp. 311331.Google Scholar
French, C. 2001, The development of the prehistoric landscape in the Flag Fen Basin, in Prior, F., ed., The Flag Fen Basin. Archaeology and environment of a Fenland landscape, Archaeological Reports: London, English Heritage, pp. 400404.Google Scholar
French, C. 2003, Geoarchaeology in Action. Studies in soil micromorphology and landscape evolution, London, Routledge.Google Scholar
French, C. 2015, A Handbook of Geoarchaeological Approaches to Settlement Sites and Landscapes, Oxford, Oxbow, Studying Scientific Archaeology I, p. 118.Google Scholar
French, C., and Milek, K., 2012, The geoarchaeological evidence, in Tipper, J., ed., Experimental Archaeology and Fire: the investigation of a burnt reconstruction at West Stow Anglo-Saxon village, East Anglian Archaeology 146: Bury St Edmunds, Archaeological Service, Suffolk County Council, pp. 7789.Google Scholar
Friesem, D. E., Boaretto, E., Eliyahu-Behar, A., and Shahack-Gross, R., 2011, Degradation of mud brick houses in an arid environment: a geoarchaeological model: Journal of Archaeological Science, v. 38, no. 5, pp. 11351147.Google Scholar
Friesem, D. E., Karkanas, P., Tsartsidou, G., and Shahack-Gross, R., 2014a, Sedimentary processes involved in mud brick degradation in temperate environments: a micromorphological approach in an ethnoarchaeological context in northern Greece: Journal of Archaeological Science, v. 41, pp. 556567.Google Scholar
Friesem, D. E., Tsartsidou, G., Karkanas, P., and Shahack-Gross, R., 2014b, Where are the roofs? A geo-ethnoarchaeological study of mud brick structures and their collapse processes, focusing on the identification of roofs: Archaeological and Anthropological Sciences, v. 6, pp. 7392.Google Scholar
Fulford, M., and Wallace-Hadrill, A., 1995–6, The House of Amarantus at Pompeii (I, 9, 11–12): an interim report on survey and excavations in 1995–96: Revista di Studi Pompeiani, v. VII, pp. 77113.Google Scholar
Fuller, D. Q., and Qin, L., 2009, Water management and labour in the origins and dispersal of Asian rice: World Archaeology, v. 41, no. 1, pp. 88111.Google Scholar
Gale, S. J., and Hoare, P. G., 1991, Quaternary Sediments. Petrographic methods for the study of unlithified rocks, London, Belhaven Press (John Wiley).Google Scholar
Galinié, H., 2000, Terres Noires – 1, Documents. Sciences de la Ville, Volume No. 6: Tours, La Maison des Sciences de la Ville, p. 119.Google Scholar
Galinié, H. 2004, L’expression terres noires, un concept d’attente, in Verslype, L., ed., Terres Noires Dark Earth: Louvain-La-Neuve, Université Catholique de Louvain., pp. 111.Google Scholar
Galinié, H. 2007, Tours, antique et médiéval. Lieux de vie Temps de la ville, Tours, Revue Archéologique du Centre de la France (FERACF), p. 440.Google Scholar
Galinié, H., Lorans, E., Macphail, R. I., Seigne, J., Fondrillon, M., Laurent, A., and Moreau, A., 2007, Chapter 53. La fouille du square Prosper-Mérimée. The excavation in Prosper-Mérimée Square, in Galinié, H., ed., Tours, antique et médiéval. Lieux de vie Temps de la ville, Volume 30th Supplément: spécial de la collection Recherches sur Tours: Tours, Revue Archéologique du Centre de la France (FERACF), pp. 171180.Google Scholar
, T., Courty, M. A., Matthews, W., and Wattez, J., 1993, Sedimentary formation processes of occupation surfaces, in Goldberg, P., Nash, D. T., and Petraglia, M. D., eds., Formation Proceses in Archaeological Contexts, Monographs in World Archaeology No. 17: Madison, Wisconsin, Prehistory Press, pp. 149163.Google Scholar
Gebhardt, A., 1990, Evolution du Paléopaysage Agricole dans Le Nord-Ouest de la France: apport de la micromorphologie: L’Université de Rennes I.Google Scholar
Gebhardt, A. 1992, Micromorphological analysis of soil structural modification caused by different cultivation implements, in Anderson, P. C., ed., Préhistoire de l’Agriculture: nouvelles approaches expérimentales et ethnographiques, Monogaphie de CRA No. 6: Paris, Centre Nationale de la Recherche Scientifique, pp. 373392.Google Scholar
Gebhardt, A. 1993, Micromorphological evidence of soil deterioration since the mid-Holocene at archaeological sites in Brittany, France: The Holocene, v. 3, no. 4, pp. 331341.Google Scholar
Gebhardt, A. 1995, Soil micromorphological data from traditional and experimental agriculture, in Barham, A. J., and Macphail, R. I., eds., Archaeological Sediments and Soils: Analysis, Interpretation and management: London, Institute of Archaeology, pp. 2540.Google Scholar
Gebhardt, A. 2007, Impact of charcoal production activities on soil profiles: the micromorphological point of view: ARCHAEOSCIENCES, revue d’archéométrie, v. 31, pp. 127136.Google Scholar
Gebhardt, A. 2008, La pédologie d’un billon, in Burnouf, J., ed., Archéologie médiévale en France, le second Moyen-Age (XIIe-XVIe siècle), La Découverte, pp. 43.Google Scholar
Gebhardt, A., Fechner, K., and Occhietti, S., 2014, Grandes phases de pédogenèse, d’érosion et d’anthropisation des sols au cours de la seconde moitié de l’Holocène en Lorraine (France): Archéo-Science, v. 38, no. 1, pp. 729.Google Scholar
Gebhardt, A., and Langohr, R., 1999, Micromorphological study of construction materials and living floors in the medieval motte of Werken (West Flanders, Belgium): Geoarchaeology, v. 14, no. 7, pp. 595620.Google Scholar
Gebhardt, A. 2015, Traces de roulage ou de labour? Le diagnostic micromorphologique: Archéo Science, v. 39, pp. 3138.Google Scholar
Gebhardt, A., and Marguerie, D., 1989, Les fouilles archéologiques d’Er Grah and La Tables-Des-Marchand (Locmariaquer, Morbihan); synthèse des raports d’ètudes sèdimentologiques 1986, 1987, 1988: Laboratoire d’Anthropogie, Université de Rennes 1.Google Scholar
Gerasimova, M., and Lebedeva-Verba, M., 2010, Topsoils – Mollic, Takyric and Yermic Horizons, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 351–368.Google Scholar
Gilbertson, D. D., 1995, Studies of lithostratigraphy and lithofacies: a selective review of research developments in the last decade and their applications to geoarchaeology, in Barham, A. J., and Macphail, R. I., eds., Archaeological Sediments and Soils: analysis, interpretation and management: London, Institute of Archaeology, pp. 99145.Google Scholar
Gillieson, D., 1996, Caves: Processes, Development, Management, Oxford, Blackwell.Google Scholar
Glanville-Wallis, F., 2015Of Crabs and Men”: Artefact analysis of residual waste deposits and a preliminary investigation into crab bioturbation at Marco Gonzalez, Belize. [BSc Archaeology Dissertation]: University College London.Google Scholar
Glaser, B., and Woods, W. I., 2004, Amazonian dark earths: exploration in space and time: New York, Springer, p. 193.Google Scholar
Goldberg, P., 1979a, Geology of Late Bronze Agemudbrick from Tel Lachish: Tel Aviv, Journal of the Tel Aviv Institute of Archaeology, v. 6, pp. 6071.Google Scholar
Goldberg, P. 1979b, Micromorphology of Pech-de-l’Azé: Journal of Archaeological Science, v. 6, pp. 1747.Google Scholar
Goldberg, P. 1980, Micromorphology in archaeology and prehistory: Paléorient, v. 6, pp. 159164.Google Scholar
Goldberg, P. 2000a, Micromorphological aspects of site formation at Keatley Creek, in Hayden, B., ed., The Ancient Past of Keatley Creek: Simon Fraser University, Burnaby, B.C., Archaeology Press, pp. 7995.Google Scholar
Goldberg, P. 2000b, Micromorphology and site formation at Die Kelders Cave 1, South Africa: Journal of Human Evolution, v. 38, pp. 4390.Google Scholar
Goldberg, P. 2005, Thin Section Observation of samples collected from Officer’s Club, Presidio, California, February, 2005: Boston University.Google Scholar
Goldberg, P., and Berna, F., 2010, Micromorphology and context: Quaternary International, v. 214, pp. 5662.Google Scholar
Goldberg, P., Berna, F., and Macphail, R. I., 2009a, Comment on “DNA from Pre-Clovis Human Coprolites in Oregon, North America”: Science, v. 325, p. 148-c.Google Scholar
Goldberg, P., and Byrd, B. F., Fall, 1999, The interpretive potential of micromorphological analysis at prehistoric shell midden sites on Camp Pendleton: Pacific Coast Archaeological Society Quarterly, v. 35(4), pp. 123.Google Scholar
Goldberg, P., and Guy, J., 1996, Micromorphological observations of selected rock ovens, Wilson-Leonard site, Central Texas, in Castelletti, L., and Cremaschi, M., eds., Paleoecology; Colloquium 3 of XIII International Congress of Prehistoric and Protohistoric Sciences: Forlì, ABACO, pp. 115122.Google Scholar
Goldberg, P., Lev-Yadun, S., and Bar-Yosef, O., 1994, Petrographic thin sections of archaeological sediments: a new method for palaeobotanical studies: Geoarchaeology, v. 9, no. 3, pp. 243257.Google Scholar
Goldberg, P., and Macphail, R. I., 2000, Micromorphology of sediments from Gibraltar caves: some preliminary results from Gorham’s Cave and Vanguard Cave, in Finlayson, C., Finlayson, G., and Fa, D., eds., Gibraltar during the Quaternary, Volume 1: Gibraltar, Gibraltar Government Heritage Publications Monographs, pp. 93108.Google Scholar
Goldberg, P., and Macphail, R. I. 2003, Short contributions: strategies and techniques in collecting micromorphology samples: Geoarchaeology, v. 18, no. 5, pp. 571578.Google Scholar
Goldberg, P., and Macphail, R. I. 2006a, Playa Vista Archaeological and Historical Project (California): soil micromorphology – preliminary report: Statistical Research Inc.Google Scholar
Goldberg, P., and Macphail, R. I. 2006b, Practical and Theoretical Geoarchaeology, Oxford, Blackwell Publishing, p. 455.Google Scholar
Goldberg, P., and Macphail, R. I. 2012, Gorham’s Cave sediment micromorphology, in Barton, R. N. E., Stringer, C., and Finlayson, C., eds., Neanderthals in Context. A report of the 1995–1998 excavations at Gorham’s and Vanguard Caves, Gibraltar, Monograph 75: Oxford, Oxford University School of Archaeology, pp. 5061; Appendix 52: 314–321; Color figures at: www.arch.ox.ac.uk/gibraltar.Google Scholar
Goldberg, P., and Macphail, R. I., 2016, Microstratigraphy, in Gilbert, A. S., ed., Encyclopedia of Geoarchaeology: Dordrecht, Springer Scientific, pp. 532-537.Google Scholar
Goldberg, P., and Macphail, R. I. Forthcoming, Chapter 3. Geomorphology of the ancient landscape, in Byrd, B. F., ed., Hunter-gatherers of the Southern California Coast, New Approaches to Anthropological Archaeology: Oxford, Equinox Publishers.Google Scholar
Goldberg, P., Miller, C. E., Schiegl, S., Ligouis, B., Berna, F., Conard, N. J., and Wadley, L., 2009b, Bedding, hearths, and site maintenance in the Middle Stone Age of Sibudu Cave, KwaZulu-Natal, South Africa: Archaeol Anthropol Sci, v. 1, pp. 95122.Google Scholar
Goldberg, P., Weiner, S., Bar-Yosef, O., Xu, Q., and Liu, J., 2001, Site formation processes at Zhoukoudian, China: Journal of Human Evolution, v. 41, pp. 483530.Google Scholar
Goldberg, P., and Whitbread, I., 1993, Micromorphological study of a Bedouin tent floor, in Goldberg, P., Nash, D. T., and Petraglia, M. D., eds., Formation Processes in Archaeological Context, Monographs in World Archaeoogy No. 17: Madison, Prehistory Press, pp. 165188.Google Scholar
Gollwitzer, M., 2012, Brønner, graver og bosetningsspor fra bronsealder til middelalder på Hesby (lok. 13), in Mjærum, A., and Gjerpe, L.-E., eds., Dyrking, bosetninger og graver i Stokke og Sandefjord, Volume 1: Bergen, Fagbokforlaget, pp. 107215.Google Scholar
Grabowski, R., and Linderholm, J., 2013, Functional interpretation of Iron Age longhouses at Gedved Vest, East Jutland, Denmark: multiproxy analysis of house functionality as a way of evaluating carbonised botanical assemblages, Archaeological and Anthropological Sciences, Volume 10: Berlin, Springer.Google Scholar
Graham, E., 2006, A Neotropical Framework for Terra Preta, in Balée, W., and Erickson, C. L., eds., Time and Complexity in Historical Ecology: New York, Columbia University press, pp. 5786.Google Scholar
Graham, E., Macphail, R. I., Turner, S., Crowther, J., Stegemann, J., Arroyo-Kalin, M., Duncan, L., Whittet, R., Rosique, C., and Austin, P., 2017, The Marco Gonzalez Maya site, Ambergris Caye, Belize: Assessing the impact of human activities by examining diachronic processes at the local scale, Quaternary International, 437 B, pp. 115–142.Google Scholar
Graham, E., and Pendergast, D. M., 1989, Excavations at the Marco Gonzalez Site, Abergris Cay, Belize, 1986: Journal of Field Archaeology, v. 16, pp. 116.Google Scholar
Grant, M. J., Stevens, C. J., Whitehouse, N. J., Norcott, D., Macphail, R. I., Langdon, C., Cameron, N., Barnett, C., Langdon, P. G., Crowther, J. N., Mulhall, N., Attree, K., Leivers, M., Greatorex, R., and Ellis, C., 2014, A palaeoenvironmental context for Terminal Upper Palaeolithic and Mesolithic activity in the Colne Valley: Offsite records contemporary with occupation at Three Ways Wharf, Uxbridge: Environmental Archaeology, v. 19, no. 2, pp. 131152.Google Scholar
Greenland, D. J., 1981, Recent progress in studies of soil structure, and its relation to properties and management of paddy soils, in Institute of Soil Science, Proceedings of Symposium on Paddy Soil: Berlin Heidelberg, Science Press: Springer-Verlag, pp. 4258.Google Scholar
Gregory, D. A., Nials, F. L., and Hill, J. B., 2008, Early Agricultural period settlement strategies in the southern Southwest in Mabry, J. B., ed., Las Capas: Early Irrigation and Sedentism in a Southwestern Floodplain (AP 28): Tucson, Archaeology Southwest.Google Scholar
Greig, J., 1988, The interpretation of some Roman well-fills from the Midlands of England., in Küster, H., ed., Der Prähistorische Mensch und Seine Umwelt: Baden-Württemberg, Forschungen und berichte sur Vor-Und Frühgeschichte.Google Scholar
Greig, J. 2004, Buried soil pollen, in Dalwood, H., and Edwards, R., eds., Excavations at Deansway, Worcester, 1988–89: Romano-British small town to late medieval city., CBA Research Report No 139: York, Council for British Archaeology, pp. 556558.Google Scholar
Grieve, I. C., 1980, Some contrasts in soil development between grassland and deciduous woodland sites: Journal of Soil Science, v. 31, pp. 137145.Google Scholar
Grimes, W. F., 1968, The Excavations of Roman and Medieval London, London, Routledge and Kegan Paul.Google Scholar
Grimm, P., 1968, Tilleda. Eine Königspfalz am Kyffhäusser. Teil 1: Die Hauptburg, Berlin, Deutsche Akademie der Wissenschaften zu Berlin, Schriften der Sektion Ur- und Frühgeschichte.Google Scholar
Grindkåsa, L., 2012, Boplasspor og grav fra romertid-merovingertid på Jarlsberg og Tem (lok. 8, 9 og 10), in Gjerpe, L.-E., and Mjærum, A., eds., E18-prosjektet Gulli-Langåker. Dyrking, bosetninger og graver i Stokke og Sandefjord, Bind 1: Bergen, Fagbokforlaget, pp. 43105.Google Scholar
Grøn, O., and Kuznetsov, O., 2004, What is a hunter-gatherer settlement? An ethno-archaeological and interdisciplinary approach, Section 7 The Mesolithic. Acts of the XIVth UISPP Congress, University of Liège, Belgium, 2–8 September 2001, BAR International Series 1302: Oxford, British Archaeological Reports, pp. 4753.Google Scholar
Guélat, M., and Federici-Schenardi, M., 1999, Develier-Courtételle (Jura) L’histoire d’une cabane en fosse reconstituée grâce à la micromorphologie: helvetica archaeologica, v. 118/119, pp. 5863.Google Scholar
Guillet, B., 1982, Study of turnover of soil organic matter using radio-isotopes, in Bonneau, M., and Souchier, B., eds., Constituents and Properties of Soils: London, Academic Press, pp. 238257.Google Scholar
Gundersen, I. M., 2016, Gård og utmark i Gudbrandsdalen. Arkeologiske undersøkelser i Fron 2011-2012: Kristiansand, Portal forlag.Google Scholar
Gur-Arieh, S., Shahack-Gross, R., Maeir, A. M., Lehmann, G., Hitchcock, L. A., and Boaretto, E., 2014, The Taphonomy and Preservation of Wood and Dung Ashes Found in Archaeological Cooking Installations: Case Studies From Iron Age Israel: Journal of Archaeological Science, v. 46, pp. 5067.Google Scholar
Gustavs, S., 1998, Spätkaiserzeitliche Baubefunde von Klein Köris, Lkr. Dahme-Spreewald, in Henning, J., and Leube, A., eds., Haus und Hof im östlichen Germanien (Berlin, 1994), Universtätsforschungen zur prähistorischen Archäologie Band 50 and Schriften zur Archäologie der germanischen und slawischen Frühgeschichte Band 2: Bonn, Dr Rudolf Habelt GmbH, pp. 4066.Google Scholar
Haită, C., 2003, Micromorphology. Inhabited space disposition and uses. Analysis of an occupation zone placed outside the dwellings, in Popovici, D., ed., Archaeological Pluridisciplinary Researches at Borduşani-Popină, Pluridisciplinary Researches series VI: Bucharest, National Museum of Romanian History, Ialomita County Museum, pp. 5174.Google Scholar
Haită, C. 2012, Sedimentologie şi Micromorfologie. Aplicaţii în Arheologie., Târgovişte, Cetatea de Scaun, p. 216.Google Scholar
Hall, A. M., and Jarvis, J., 1993, Chapter 8: North-east Scotland. Site: Kirkhill, in Gordon, J. E., and Sutherland, D. G., eds., Quaternary of Scotland, Geological Conservation Review Series, No. 6: London, Chapman and Hall, London, p. 695.Google Scholar
Halstead, P., and Tierney, J., 1998, Leafy Hay: an ethnoarchaeological study in NW Greece: Environmental Archaeology, v. 1, pp. 7180.Google Scholar
Harding, J., and Healy, F., 2011, The Raunds Area Project. A Neolithic and Bronze Age Landscape in Northamptonshire., Vol 2. Supplementary Studies: Swindon, English Heritage www.english-heritage.org.uk/publications/neolithic-and-bronze-age-landscape-vol2/, p. 978.Google Scholar
Harris, C., and Ellis, S., 1980, Micromorphology of soils in soliflucted materials, Okstindan, northern Norway: Geoderma, v. 23, pp. 1129.Google Scholar
Hatch, F. H., and Rastall, R. H., 1965, Petrology of Sedimentary Rocks, London, Thomas Murby & Co, p. 408.Google Scholar
Hayden, B., 2004, The ancient past of Keatley Creek. Volume III: Excavations: Burnaby, BC, Archaeology Press, Simon Fraser University.Google Scholar
Hazelden, J., and Boorman, L. A., 2001, Soils and managed retreat in South East England: Soil Use and Management, v. 17, pp. 150154.Google Scholar
Hazelden, J., Sturdy, R. G., and Loveland, P. J., 1987, Saline soils in North Kent, in Jarvis, M. G., ed., SEESOIL, Volume 4: Bedford, South East Soils Discussion Group, pp. 223.Google Scholar
Healy, F., and Harding, J., 2007, The Raunds Area Project. A Neolithic and Bronze Age Landscape in Northamptonshire: Swindon, English Heritage, p. 324.Google Scholar
Healy, F., Harding, J., and Bayliss, A., 2007, The Development of the Monuments, in Healy, F., and Harding, J., eds., The Raunds Area Project. A Neolithic and Bronze Age Landscape in Northamptonshire: Swindon, English Heritage, pp. 37119.Google Scholar
Henning, J., 2009, Revolution or Relapse? Technology, Agriculture and Early Medieval Archaeology in Germanic Central Europe, in Ausenda, G., Delogu, P., and Wickham, C., eds., The Langobards Before the Frankish Conquest; an ethnographic perspective, Studies in Historical Archaeoethnology 8: San Marino, The Boydell Press, pp. 151175.Google Scholar
Henning, J., and Macphail, R. I., 2004, Das karolingische Oppidum Büraburg: Archälogische und mikromorphologische. Stedien zur Funktion einer frümittelaterlichen Bergbefestigung in Nordhessen (The Carolingian times oppidum Büraburg: archaeological and soil investigations on the function of an early medieval hillfort in North Hesse), in Hänsel, B., ed., Parerga Praehistorica. Jubiläumsschrift zur Prähistorischen Archäologie 15 Jahre UPA, Band 100: Bonn, Verlag Dr Rudolf habelt GmbH, pp. 221252.Google Scholar
Henning, J., McCormick, M., and Fischer, T., 2012a, Decempagi at the end of antiquity and the fate of the Roman road system in eastern Gaul, in Bidwell, P., ed., Proceedings of the XXIst International Limes (Roman Frontiers) Congress, 2009 at Newcastle upon Tyne: Oxford, British Archaeological Reports.Google Scholar
Heppel, E., 2004, Wallsea Island; the history and archaeology of a marshland: Essex Archaeology and History, v. 35, pp. 98113.Google Scholar
Heron, C., 2001, Geochemical prospecting, in Brothwell, D., and Pollard, A. M., eds., Handbook of Archaeological Sciences: Chichester, Wiley.Google Scholar
Hewitson, C., Ramsey, E., Shaw, M., Hislop, M., and Cuttler, R., 2010, Archaeological Investigations at Old Hall Street, Wolverhampton, 2000-2007, British Archaeological Reports, New Series 5: Birmingham, Birmingham Archaeology.Google Scholar
Hillson, S. W., (Ed), 1996, The earthwork on Morden Bog, Wareham, Dorset 1972 (10th year) – 1990 (27th year), in Bell, M., Fowler, P. J., and Hillson, S. W., eds., The Experimental Earthwork Project 1960–1992: York, Council for British Archaeology, pp. 201224.Google Scholar
Hillson, S. W., Parfitt, S. A., Bello, S. M., Roberts, M. B., and Stringer, C. B., 2010, Two hominin incisor teeth from the middle Pleistocene site of Boxgrove, Sussex, England: Journal of Human Evolution, v. 59 pp. 493503.Google Scholar
Hilton, M. R., 2002, Evaluating Site Formation Processes at a Higher Resolution: An Archaeological Case Study in Alaska Using Micromorphology and Experimental Techniques: University of California.Google Scholar
Hodge, C., A. H., Burton, R. G. O., Corbett, W. M., Evans, R., George, H., Heaven, F. W., Robson, J. D., and Seale, R. S., 1983, Soils of England and Wales, Sheet 4 Eastern England, Southampton, Ordnance Survey, Soils of England and Wales.Google Scholar
Holden, T. G., 1990, Taphonomic and methodological problems in reconstructing diet from ancient human gut and faecal remains: University College London.Google Scholar
Holden, T. G. 1994, Dietary evidence from the intestinal content of ancient humans with particular reference to dessicated remains from northern Chile, in Hather, J. G., ed., Tropical Archaeobotany. Applications and new developments: London, Routledge, pp. 6585.Google Scholar
Holliday, V. T., 2004, Soils in Archaeological Research, Oxford, Oxford University Press, p. 448.Google Scholar
Holliday, V. T., Hoffecker, J., Goldberg, P., Macphail, R. I., Forman, S., Anikovich, M. V., and Sinitsyn, A. A., 2007, Geoarchaeology of the Kostenki-Borshchevo sites, Don River Valley, Russia: Geoarchaeology, v. 22, no. 2, pp. 181228.Google Scholar
Hollocher, K. T., Hollocher, T. C., and Keith Rigby, j., J., 2010, A phosphatic coprolite lacking diageneitic permineralalization from the Upper Cretaceous Hell Creek Formation, Northeastern Montana: importance of dietary calcium phosphate in preservation: PALAIOS v. 25, no. Research Note, pp. 132140.Google Scholar
Homburg, J. A., Macphail, R. I., Goldberg, P., and Mayer, J., Poster: Geochemical and soil micromorphological properties of archaeological deposits in coastal southern California, in Proceedings GSA-Tri-Society meetings Oct. 5–9, Houston, 2008.Google Scholar
Hopkins, D. W., Wiltshire, P. E. J., and Turner, B. D., 2000, Microbial characteristics of soils from graves: an investigation at the interface of soil microbiology and forensic science: Applied Soil Ecology, v. 14, pp. 283288.Google Scholar
Horwitz, L. K., and Goldberg, P., 1989, A Study of Pleistocene and Holocene Hyaena Coprolites: Journal of Archaeological Science, v. 16, pp. 7194.Google Scholar
Huisman, D. J., 2009, Degradation of archaeological remains: Den Haag, Sdu Uitgevers b.v., p. 245.Google Scholar
Hume, P. D., and Shirriffs, J., 1985, Pollen analysis of a radio-carbon dated core from North Mains, Strathallan, Perthshire: Proceedings of the Society of Antiquaries Scotland, v. 115, pp. 105113.Google Scholar
Imeson, A. C., and Jungerius, P. D., 1976, Aggregate stability and colluviation in the Luxembourg Ardennes: an experimental and micromorphological study: Earth Surface Processes, v. 1, pp. 259271.Google Scholar
Imeson, A. C., Kwaad, F. J. P. M., and Mucher, H. J., 1980, Hillslope processes and deposits in forested areas of Luxembourg, in Cullingford, R. A., Davidson, D. A., and Lewin, J., eds., Timescales in Geomorphology: Chichester, John Wiley and Sons, pp. 3142.Google Scholar
Ismail-Meyer, K., and Rentzel, P., 2004, Mikromorphologische Untersuchung der Schichtabfolge, in Jacomet, S., Leuzinger, U., and Schibler, J., eds., Die jungsteinzeitliche Seeufersiedlung Arbon/Bleiche 3. Umwelt und Wirtschaft, Archäologie im Thurgau/Band 12: Kanton Thurgau, Departement für Erziehung und Kultur des Kantons Thurgau, pp. 6680.Google Scholar
Ismail-Meyer, K., Rentzel, P., and Wiemann, P., 2013, Neolithic Lakeshore Settlements in Switzerland: New Insights on Site Formation Processes from Micromorphology: Geoarchaeology, v. 28, pp. 317339.Google Scholar
Jacobi, R., Undated, The Late Pleistocene archaeology of Somerset.Google Scholar
Jacobson, G. L. J., and Bradshaw, R. H. W., 1981, A selection of sites for paleovegetational studies: Quaternary Research, v. 16, pp. 8096.Google Scholar
Jacomet, S., Leuzinger, U., and Schibler, J., 2004, Die jungsteinzeitliche Seeufersiedlung Arbon/Bleiche 3. Umwelt und Wirtschaft, Archäologie im Thurgau/Band 12: Kanton Thurgau, Departement für Erziehung und Kultur des Kantons Thurgau, p. 458.Google Scholar
Jarvis, M. G., Allen, R. H., Fordham, S. J., Hazleden, J., Moffat, A. J., and Sturdy, R. G., 1983, Soils of England and Wales. Sheet 6. South East England: Ordnance Survey, scale 1:250,000.Google Scholar
Jarvis, M. G., Allen, R. H., Fordham, S. J., Hazleden, J., Moffat, A. J., and Sturdy, R. G. 1984, Soils and Their Use in South-East England, Harpenden, Soil Survey of England and Wales, p. 405.Google Scholar
Jenny, H., 1941, Factors of Soil Formation, New York, McGraw-Hill.Google Scholar
Jones, A. K. G., 1985, Trichurid ova in archaeological deposits: their value as indicators of ancient faeces, in Fieller, N. R. J., Gilbertson, D. D., and Ralph, N. G. A., eds., Palaeoenvironmental Investigations: Research Design, Methods and Data Analysis, Volume 258: Oxford, British Archaeological Reports International Series, pp. 105115.Google Scholar
Jones, A. W., 1958, The flora of the city of London bombed sites: The London Naturalist, v. 37, pp. 189210.Google Scholar
Jongerius, A., 1957, Morfologische onder zoekingen over de bodemstructur., Wageningen, Mededelingen van de Stichting voor Bodemkartering, p. 93.Google Scholar
Jongerius, A. 1962, Soil genesis in organic soils: Boor en Spade, v. 12, pp. 156168.Google Scholar
Jongerius, A. 1970, Some morphological aspects of regrouping phenomena in Dutch soils: Geoderma, v. 4, pp. 311331.Google Scholar
Jongerius, A. 1983, The role of micromorphology in agricultural research, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology, Volume Volume 1: Techniques and Applications: Berkhamsted, A B Academic Publishers, pp. 111138.Google Scholar
Jongerius, A., and Jager, A., 1964, The morphology of humic gley soils (Orthic Haplaquolls) under different land use, in Jongerius, A., ed., Soil Micromorphology: Amsterdam, Elsevier, pp. 491503.Google Scholar
Karkanas, K., Rigaud, J.-P., Simek, J., Albert, R., and Weiner, S., 2002, Ash, bones and guano: a study of the minerals and phytoliths in the sediments of Grotte XVI, Dordogne, France: Journal of Archaeological Science, v. 29, pp. 721732.Google Scholar
Karkanas, P., 2006, Late Neolithic household activities in marginal areas: the micromorphological evidence from the Kouveleiki caves, Peloponnese, Greece: Journal of Archaeological Science, v. 33, pp. 16281641.Google Scholar
Karkanas, P. 2007, Identification of lime plaster in prehistory using petrographic methods: a review and reconsideration of the data on the basis of experimental and case studies: Geoarchaeology, v. 22, no. 7, pp. 775796.Google Scholar
Karkanas, PP., Shahack-Gross, R., Ayalon, A., Bar-Matthews, M., Barkai, R., Gopher, A., and Stiner, M., 2007, Evidence of habitual use of fire at the end of the Lower Palaeolithic: Site-formation processes at Qesem Cave, Israel: Journal of Human Evolution, v. 53, pp. 197212.Google Scholar
Karkanas, P., Dabney, M. K., Smith, A. K., and Wright, J. C., 2012, The geoarchaeology of Mycenaean chamber tombs: Journal of Archaeological Science, v. 39, pp. 27222732.Google Scholar
Karkanas, P., and Goldberg, P., 2010, Phosphatic features, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 521541.Google Scholar
Karkanas, P., Kyparissi-Apostolika, N., Bar-Yosef, O., and Weiner, S., 2000, Mineral assemblages in Theopetra, Greece: a framework for understanding diagenesis in a prehistoric cave: Journal of Archaeological Science, v. 26, pp. 11711180.Google Scholar
Karkanas, P., Pavlopoulos, K., Kouli, K., Ntinou, M., Tsartsidou, G., Facorellis, Y., and Tsourou, T., 2011, Palaeoenvironments and Site Formation Processes at the Neolithic Lakeside Settlement of Dispilio, Kastoria, Northern Greece: Geoarchaeology, v. 26, no. 1, pp. 83117.Google Scholar
Karkanas, P., and Van de Moortel, A., 2014, Micromorphological analysis of sediments at the Bronze Age site of Mitrou, central Greece: patterns of floor construction and maintenance: Journal of Archaeological Science, v. 43, pp. 198213.Google Scholar
Karlsson, N., 2006, Bosättning och resursutnyttjande: miljöarkeologiska studier av bopplatser med hädar från perioden 600–1900 e. Kr inom skogssamiskt område., Umeå, University of Umeå.Google Scholar
Keeley, H. C. M., Hudson, G. E., and Evans, J., 1977, Trace element contents of human bones in various states of preservation: Journal of Archaeological Science, v. 4, pp. 1924.Google Scholar
Keeley, H. C. M., and Macphail, R. I., 1981, A soil survey of part of Shaugh Moor, Devon, in The Shaugh Moor Project: Third Report – Settlement and environmental investigations, edited by Smith, K., Coppen, J., Wainwright, G. J., Beckett, S.: Proceedings of the Prehistoric Society, v. 47, pp. 240245.Google Scholar
Keeley, H. C. M., and Macphail, R. I. 1982, Soils of the Saddlesbourough Reave area, Shaugh Moor, in The Shaugh Moor Project: Fourth Report – Environment, context and conclusion: Proceedings of the Prehistoric Society, v. 48, pp. 219220, Microfiche 214–220.Google Scholar
Keevill, G., 2004, The Tower of London Moat. Archaeological Excavations 1995–9: Oxford, Oxford Archaeology with Historic Royal Palaces, p. 315.Google Scholar
Kelley, J. T., Belknap, D. F., and Claesson, S., 2010, Drowned coastal deposits with associated archeological remains from a sea-level “slowstand”: northwestern Gulf of Maine, USA: Geology, v. 38, pp. 695698.Google Scholar
Kelly, J., and Wiltshire, P. E. J., 1996, Microbiological report, in Bell, M., Fowler, P. J., and Hillson, S. W., eds., The Experimental Earthwork Project 1960–1992, CBA Research Report: York, Council for British Archaeology, pp. 148155.Google Scholar
Kemp, R. A., 1985, The decalcified Lower Loam at Swanscombe, Kent: a buried Quaternary soil: Proceedings of the Geological Association, v. 96, pp. 343354.Google Scholar
Kemp, R. A. 1986, Pre-Flandrian Quaternary soils and pedogenic processes in Britain, in Wright, V. P., ed., Paleosols. Their Recognition and Interpretation: Oxford, Blackwell Scientific Publications, pp. 242262.Google Scholar
Kemp, R. A., Jerz, H., Grottemthaler, W., and Preece, R. C., 1994, Pedosedimentary fabrics of soils within loess and colluvium in southern England and southern Germany, in Ringrose-Voase, A. J., and Humphreys, G. S., eds., Soil Micromorphology: studies in management and genesis: Amsterdam, Elesevier, pp. 207219.Google Scholar
Kenward, H. F., and Hall, A. R., 1997, Enhancing bioarchaeological interpretation using indicator groups: stable manure as a paradigm: Journal of Archaeological Science, v. 24, pp. 663673.Google Scholar
Kenward, H. K., and Hall, A. R., 1995, Biological Evidence from Anglo-Scandinavian Deposits at 16–22 Coppergate, York, York Archaeological Trust, p. 797.Google Scholar
Kerney, M. P., Brown, E. H., and Chandler, T. J., 1963, The Late Glacial and Post Glacial history of the Chalk escarpment near Brook, Kent: Philosophical Transactions of the Royal Society, London B, v. 745, no. 248, pp. 135204.Google Scholar
Kerney, M. P., Preece, R. C., and Turner, C., 1980, Molluscan and plant biostratigraphy of some Late Devensian and Flandrian deposits in Kent: Philosophical Transactions of the Royal Society, London, v. B 291, pp. 143.Google Scholar
Kerr, P. F., 1959, Optical Mineralogy, New York, McGraw-Hill Book Company.Google Scholar
Kidder, T. R., Liu, H., Xu, Q., and Li, M., 2012, The alluvial geoarchaeology of the Sanyangzhuang Site on the Yellow River floodplain, Henan Province, China: Geoarchaeology, v. 27, no. 4, pp. 324343.Google Scholar
Kobylínski, Z., 1989, An ethnic change or a socio-eonomic one? The 5th and 6th centuries AD in the Polish lands, Shennan, S. (ed.), Archaeoplogical Approaches to Cultural Identity: One World Archaeology, v. 10, pp. 303310.Google Scholar
Kooistra, M., and Pulleman, M. M., 2010, Features Related to Faunal Activity, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 397418.Google Scholar
Kooistra, M. J., 1978, Soil Development in Recent Marine Sediments of the Intertidal Zone in the Oosterschelde – the Netherlands: a Soil Micromorphological Approach, Wageningen, Soil Survey Institute, Soil Survey Papers.Google Scholar
Kooistra, M. J. 1987, The effects of deep compaction and deep tillage on soil structure in a Dutch sandy loam soil, in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 445450.Google Scholar
Kramer, C., 1982, Village Ethnoarchaeology: rural Iran in archaeological perspective, New York, Academic Press, p. 302.Google Scholar
Kresten, P., and Hjärthner-Holdar, E., 2001, Analyses of the Swedish ancient iron reference slag W-25:R: Historical Metallurgy, v. 35, no. 1, pp. 4851.Google Scholar
Kubiena, W. L., 1938, Micropedology, Ames, Iowa, Collegiate Press.Google Scholar
Kubiena, W. L. 1953, The Soils of Europe, London, Murby.Google Scholar
Kühn, P., Aguilar, J., and Miedema, R., 2010, Textural pedofeatures and related horizons, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 217250.Google Scholar
Kukal, Z., 1971, The Geology of Recent Sediments, Prague, Science Press, p. 490.Google Scholar
Kwaad, F. J. P. M., and Mücher, H. J., 1977, The evolution of of soils and slope deposits in the Luxembourg Ardennes near Wilts: Geoderma, v. 17, pp. 137.Google Scholar
Kwaad, F. J. P. M., and Mücher, H. J. 1979, The formation and evolution of colluvium on arable land in northern Luxembourg: Geoderma, v. 22, no. 2, pp. 173192.Google Scholar
Lambrick, G., 1992, Alluvial archaeology of the Holocene in the Upper Thames Basin 1971–1991: a review, in Needham, S., and Macklin, M. G., eds., Alluvial Archaeology in Britain, Monograph 27: Oxford, Oxbow, pp. 209228.Google Scholar
Landuydt, C. J., 1990, Micromorphology of ironminerals from bog ores of the Belgian Campine Area, in Douglas, L. A., ed., Soil Micromorphology: a Basic and Applied Science: Amsterdam, Elsevier, pp. 289301.Google Scholar
Langohr, R., 1991, Soil characteristcis of the Motte of Werken (West Flanders, Belgium), in Tauber, J., ed., Methoden und Perspektiven der Archaologie des Mittelalters, Tagungsberichte zum interdisziplinaren Kolloquium, September 1989, Liestal, Switzerland, 209–223: Heft, Berichte aus der Arbeit des Amtes fur Museen und Archaologie des Kantons Baseeland.Google Scholar
Langohr, R. 1993, Types of tree windthrow, their impact on the environment and their importance for the understanding of archaeological excavation data: Helenium, v. XXXIII, no. 1, pp. 3649.Google Scholar
Lanting, J. N., and Mook, W. G., 1977, Pre- and Proto-history of the Netherlands in Terms of Radiocarbon dates, Groningen, Isotope Physics Laboratory.Google Scholar
Larkin, N., Alexander, J., and Lewis, M. D., 2000, Using experimental studies of recent faecal material to examine hyena coprolites from the West Runton Freshwater bed, Norfolk, UK: Journal of Archaeological Science, v. 27, pp. 1931.Google Scholar
Laurent, A., and Fondrillon, M., 2010, Mesurer la ville par l’évaluation et la caractérisation du sol urbain: l’exemple de Tours: Revue archéologique du Centre de la France, v. 49, pp. 307343.Google Scholar
Lawson, A. J., 2000, Potterne 1982–5: Animal Husbandry in Later Prehsitoric Wiltshire, Salisbury, Wessex Archaeology.Google Scholar
Leary, J., Canti, M., Field, D., Fowler, P., Marshall, P., and Cambell, G., 2013, The Marlborough Mound, Wiltshire. A Further Neolithic Monumental Mound by the River Kennet: Proceedings of the Prehistoric Society, v. 79, pp. 137163.Google Scholar
Lechtman, H. N., and Hobbs, L. W., 1983, Roman concrete and Roman architectural revolution: Journal of Archaeological Science, v. 13, pp. 81128.Google Scholar
Lee, H., French, C., and Macphail, R. I., 2014, Microscopic Examination of Ancient and Modern Irrigated Paddy Soils in South Korea, with Special Reference to the Formation of Silty Clay Concentration Features: Geoarchaeology, v. 29, pp. 326348.Google Scholar
Lewis, H. A., 2012, Investigating Ancient Tillage. An experimental and soil micromorphological study, Oxford, British Archaeological Reports, p. 119.Google Scholar
Lewis, J., Leivers, M., Brown, L., Smith, A., Cramp, K., Mepham, L., and Phillpotts, C., 2010, Landscape Evolution in the Middle Thames, Valley. Heathrow Terminal 5 Excavation Volume 2, Framework Archaeology Monograph 13: Oxford/Salisbury, Framework Archaeology.Google Scholar
Lewis, J. S., Wiltshire, P., and Macphail, R. I., 1992, A Late Devensian/Early Flandrian site at Three Ways Wharf, Uxbridge: environmental implications, in Needham, S., and Macklin, M. G., eds., Alluvial Archaeology in Britain, Monograph 27: Oxford, Oxbow, pp. 235248.Google Scholar
Lewis, J. S. C., and Rackham, J., 2011, Three Ways Wharf, Uxbridge. A Lateglacial and Early Holocene hunter-gatherer site in the Colne Valley, London, Museum of London, p. 228.Google Scholar
Lewis, J. S. C., and Welsh, K., 2004, Perry Oaks – Neolithic inhabitation of a west London landscape, in Cotton, J., and Field, D., eds., Towards a New Stone Age, Research Report 137: York, Council for British Archaeology, pp. 105109.Google Scholar
Lewis, M., 1997, An analysis of the Early Middle Pleistocene hyena fauna coprolites from Boxgrove, Sussex: University College London, p. 60.Google Scholar
Limbrey, S., 1975, Soil Science and Archaeology, London, Academic Press.Google Scholar
Linderholm, J., 1998, Fähus, markanalys och arkeologi – att studera yngan i tiden och rummet., in Viklund, K., Engelmark, R., and Linderholm, J., eds., Fähus från bronsålder till i dag, Volume Skrifter om skogs- och lantbrukshistoria: Lund, Nordiska museet, pp. 2227.Google Scholar
Linderholm, J. 2007, Soil chemical surveying: a path to a deeper understanding of prehistoric sites and societies in Sweden: Geoarchaeology, v. 22, no. 4, pp. 417438.Google Scholar
Linderholm, J. 2010, Soil prospection: chemical and magnetic susceptibility attributes from off-site and intra-site perspectives. A case study from a late Mesolithic-early Neolithic dwelling in northern Sweden, in Linderholm, J., ed., The Soil as a Source Material in Archaeology. Theoretical Considerations and Pragmatic Applications, Archaeology and Environment 25: Umeå, Umeå University, pp. 143.Google Scholar
Linderholm, J., and Lundberg, E., 1994, Chemical characterisation of various archaeological soil samples using main and trace elements determined by inductively coupled plasma atomic emission spectrometry: Journal of Archaeological Science, v. 21, pp. 303314.Google Scholar
Linford, N. T., and Canti, M. G., 2001, Geophysical Evidence for Fires in Antiquity: Preliminary Results from an Experimental Study. Paper Given at the EGS XXIV General Assembly in The Hague, April 1999: Archaeological Prospection, v. 8, pp. 211225.Google Scholar
Liu, L., Chen, X., and Jiang, L., 2004, A study of Neolithic water buffalo remains from Zhejiang, China: Bulletin of the Indo-Pacific Prehistory Association, v. 24 (Indo-Pacific Prehistory: the Taipei Papers, Volume 2), pp. 113120.Google Scholar
Liu, L., Chen, X., Lee, Y. K., Wright, H., and Rosen, A., 20022004, Settlement patterns and developing of social complexity in the Yiluo Region, north China: Journal of Field Archaeology, v. 29, no. 1 and 2, pp. 75100.Google Scholar
Loaiza, J., Stoops, G., Poch, R. M., and Casamitjana, M., 2015, Manual de micromorfología de suelos y técnicas complementarias: Medellín, Colombia, Fondo Editorial Pascual Bravo p. 384.Google Scholar
Locock, M., Robinson, S., and Yates, A., 1998, Late Bronze Age sites at Cabot Park, Avonmouth: Archaeology in the Severn Estuary, v. 9, pp. 3136.Google Scholar
Loveland, P. J., 1981, Weathering of a soil glauconite in southern England: Geoderma, v. 25, pp. 3554.Google Scholar
Loveland, P. J., and Findlay, D. C., 1982, Composition and development of some soils on glauconitic Cretaceous (Upper Greensand) rocks in southern England: Journal of Soil Science, v. 33, pp. 279294.Google Scholar
Lutz, H. J., and Griswold, F. S., 1939, The influence of tree roots on soil morphology: American Journal Science, v. 237, pp. 389400.Google Scholar
MacKenzie, W. S., and Guilford, C., 1980, Atlas of Rock-forming Minerals in Thin Sections, New York, Longman Scientific & Technical, p. 98.Google Scholar
Mackney, D., 1961, A podzol development sequence in oakwoods and heath in central England: Journal of Soil Science, v. 12, pp. 2340.Google Scholar
MacLeod, C. L., Scrimshaw, M. D., Emmerson, R. H. L., Chang, Y. H., and Lester, J. N., 1999, Geochemical changes in metal and nutrient loading at Orplands Fram managed Retreat site, Essex, UK (April 1995–1997): Marine Pollution Bulletin, v. 38, pp. 11151125.Google Scholar
Macphail, R. I., 1981, Soil and botanical studies of the “Dark Earth,” in Jones, M., and Dimbleby, G. W., eds., The Environment of Man: the Iron Age to the Anglo-Saxon Period, British Series 87: Oxford, British Archaeological Reports, pp. 309331.Google Scholar
Macphail, R. I. 1983, The micromorphology of dark earth from Gloucester, London and Norwich: an analysis of urban anthropogenic deposits from the Late Roman to Early Medieval periods in England, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology, Volume 1: Techniques and Applications: Berkhamsted, A B Academic Publishers, pp. 245252.Google Scholar
Macphail, R. I. 1987, A review of soil science in archaeology in England, in Keeley, H. C. M., ed., Environmental Archaeology: A Regional Review Vol. II, Occasional paper No. 1: London, Historic Buildings & Monuments Commission for England, pp. 332379.Google Scholar
Macphail, R. I. 1990a, The soils, in Saville, A., ed., Hazleton North, Gloucestershire, 1979–82: The Excavation of a Neolithic Long Cairn of the Cotswold-Severn Group, Archaeological Report no. 13: London, English Heritage, pp. 223226.Google Scholar
Macphail, R. I. 1990b, Soil report on Drayton Cursus, near Abingdon, Oxfordshire: English Heritage, 57/90.Google Scholar
Macphail, R. I. 1990c, Soil report on Carn Brea, Redruth, Cornwall, with some reference to similar sites in Brittany, France: English Heritage, Ancient Monuments Laboratory Report 55/90.Google Scholar
Macphail, R. I. 1990d, Soil history and micromorphology, in Bell, M., ed., Brean Down Excavations 1983–1987, Archaeological report No. 15: London, English Heritage, pp. 187196.Google Scholar
Macphail, R. I. 1990e, Soil micromorphological evidence of the impact of ancient agriculture, Transactions of the 14th International Congress of Soil Science, Kyoto, Japan, August 1990, Volume VII: Kyoto, International Society of Soil Science, pp. 264269.Google Scholar
Macphail, R. I. 1991, The archaeological soils and sediments, in Sharples, N. M., ed., Maiden Castle: Excavations and field survey 1985–6, Archaeological Report no 19: London, English Heritage, pp. 106118.Google Scholar
Macphail, R. I. 1992a, Late Devensian and Holocene soil formation, in Barton, R. N. E., ed., Hengistbury Head, Dorset. Volume 2: The Late Upper Palaeolithic & Early Mesolithic sites, Monograph No. 34: Oxford, Oxford University Committee for Archaeology, pp. 4451.Google Scholar
Macphail, R. I. 1992b, Soil micromorphological evidence of ancient soil erosion, in Bell, M., and Boardmand, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 197216.Google Scholar
Macphail, R. I. 1993, Soil micromorphology, in A Neolithic downland monument in its environmentL: excavations at the Easton Down Long Barrow, Bishops Canning, North Wiltshire, edited by Whittle, A., Rouse, A. J. and Evans, J. G.: Proceedings of the Prehistoric Society, v. 59, pp. 218219, 234–235.Google Scholar
Macphail, R. I. 1994a, The reworking of urban stratigraphy by human and natural processes, in Hall, A. R., and Kenward, H. K., eds., Urban-Rural Connexions: Perspectives from environmental Archaeology, Monograph 47: Oxford, Oxbow, pp. 1343.Google Scholar
Macphail, R. I. 1994b, Soil micromorphological investigations in archaeology, with special reference to drowned coastal sites in Essex, in Cook, H. F., and Favis-Mortlock, D. T., eds., SEESOIL, Volume 10: Wye, South East Soils Discussion Group, pp. 1328.Google Scholar
Macphail, R. I. 1995, Soils, in Wainwright, G. J., and Davies, S. M., eds., Balksbury Camp, Hampshire: Excavations 1973 and 1981, Archaeological Report 4: London, English Heritage, pp. 100104.Google Scholar
Macphail, R. I. 1998, A reply to Carter and Davidson’s “An evaluation of the contribution of soil micromorphology to the study of ancient arable agriculture”: Geoarchaeology, v. 13, no. 6, pp. 549564.Google Scholar
Macphail, R. I. 1999a, Sediment micromorphology, in Roberts, M. B., and Parfitt, S. A., eds., Boxgrove. A Middle Pleistocene hominid site at Eartham Quarry, Boxgrove, West Sussex, Archaeological Report 17: London, English Heritage, pp. 118148.Google Scholar
Macphail, R. I. 1999, Soils, in Whittle, A., Pollard, J., and Grigson, c., eds., The harmony of symbols: the Windmill Hill causewayed enclosure, Wiltshire: Oxford, Oxbow, pp. 121126.Google Scholar
Macphail, R. I. 2000, Soils and microstratigraphy: a soil micromorphological and micro-chemical approach, in Lawson, A. J., ed., Potterne 1982–5: Animal Husbandry in Later Prehistoric Wiltshire, Volume Archaeology Report No. 17: Salisbury, Wessex Archaeology, pp. 4770.Google Scholar
Macphail, R. I. 2003a, Industrial Activities – Some Suggested Microstratigraphic Signatures: ochre, building materials and iron-working, in Wiltshire, P. E. J., and Murphy, P., eds., The Environmental Archaeology of Industry, AEA Symposia No. 20: Oxford, Oxbow, pp. 94106.Google Scholar
Macphail, R. I. 2003b, Soil microstratigraphy: a micromorphological and chemical approach, in Cowan, C., ed., Urban development in north-west Roman Southwark Excavations 1974–90, Monograph 16: London, MoLAS, pp. 89105.Google Scholar
Macphail, R. I. 2004, Soil micromorphology, in Dalwood, H., and Edwards, R., eds., Excavations at Deansway, Worcester, 1988–89: Romano-British small town to late medieval city. CBA Research Report No 139: York, Council for British Archaeology, pp. 558567.Google Scholar
Macphail, R. I. 2005, Soil micromorphology and chemistry, in Shelley, A., ed., Dragon Hall, King Street, Norwich: Excavation and Survey of a Late Medieval Merchant’s Trading Complex, Report No. 112: Norwich, East Anglian Archaeology, pp. 175178.Google Scholar
Macphail, R. I. 2007a, Soil micromorphology (Chapter 11), in Whittle, A., and Kovács, G., eds., The Early Neolithic on the Great Hungarian Plain: investigations of the Körös culture site of Ecsegfalva 23, Co. Békés: Budapest, Institute of Archaeology, pp. 189226.Google Scholar
Macphail, R. I. 2007b, Soil micromorphology in Prehistoric and Medieval environment in Old Town, Eastbourne: studies of hillwash in the Bourne Valley, Star Brewery site, by M. Allen: Sussex Archaeological Collections, v. 145, p. 40.Google Scholar
Macphail, R. I. 2008, Soil micromorphology of the Late Glacial soil and marls, in FitzPatrick, A. P., Powell, A. B., and Allen, M. J., eds., Archaeological Excavations on the Route of the A27 Westhampnett Bypass, West Sussex, 1992, Wessex Archaeology Report No. 21: Salisbury, Wessex Archaeology, pp. 3843.Google Scholar
Macphail, R. I. 2009, Marine inundation and archaeological sites: first results from the partial flooding of Wallasea Island, Essex, UK., 2009, Antiquity Project Gallery; antiquity.ac.uk/projgall/macphail/.Google Scholar
Macphail, R. I. 2010a, Chongokni, Hantan River, Imjin River Valley, Korea (2009): soil micromorphology: Seoul National University.Google Scholar
Macphail, R. I. 2010b, Dark earth and insights into changing land use of urban areas, in Speed, G., and Sami, D., eds., Debating Urbanism: Within and Beyond the Walls c. AD 300 to c. AD 700 (Conference Proceedings Leicester University Nov 15th 2008), Leicester Archaeology Monograph 17: Leicester, Leicester Archaeology, pp. 145165.Google Scholar
Macphail, R. I. 2011a, CD Table 11 Micromorphology – summarised soil data and interpretation; CD Table 12 Micromorphology – facies types (soil microfabric types and associated data), in Burch, M., Treveil, P., and Keene, D., eds., The development of early medieval and later Poultry and Cheapside: excavations at 1 Poultry and vicinity, City of London, MOLA Monograph 38: London, Museum of London.Google Scholar
Macphail, R. I. 2011b, Soil micromorphology, in Fulford, M., and Rippon, S., eds., Pevensey Castle, Sussex. Excavations in the Roman fort and Medieval keep, 1993–95, Wessex Archaeology Report No. 26: Salisbury, Wessex Archaeology and University of Reading, pp. 109121.Google Scholar
Macphail, R. I. 2011c, Soils and sediments, in Harding, J., and Healy, F., eds., The Raunds Area Project. A Neolithic and Bronze Age Landscape in Northamptonshire. Volume 2 Supplementary Studies: Swindon, English Heritage www.english-heritage.org.uk/publications/neolithic-and-bronze-age-landscape-vol2/, pp. 737–838.Google Scholar
Macphail, R. I. 2013, Site formation processes in archaeology: soil and sediment micromorphology. Proceedings of the 14th IWMSM Session 5, Lleida, Spain, July 2012 (Guest Editorial): Quaternary International, v. 315, pp. 12.Google Scholar
Macphail, R. I. 2014a, Archaeological Soil Micromorphology Working Group, in Smith, C., ed., Encyclopedia of Global Archaeology: New York, Springer, pp. 364365.Google Scholar
Macphail, R. I. 2014b, Reconstructing past land use from dark earth: examples from England and France, in Lorans, E., and Rodier, X., eds., Archéologie de l’espace urbain – Partie II, Volume 21/17: Tours, CTHS, pp. 251–261.Google Scholar
Macphail, R. I. 2014c, Archaeological Soil Micromorphology, Encyclopedia of Global Archaeology. Heidelberg, Springer Science, 356–364.Google Scholar
Macphail, R. I. 2016a, Privies and latrines, in Gilbert, A. S., ed., Encyclopedia of Geoarchaeology: Dordrecht, Springer Scientific, pp. 682–687.Google Scholar
Macphail, R. I. 2016b, House Pits & Grubenhausen, in Gilbert, A. S., ed., Encyclopedia of Geoarchaeology: Dordrecht, Springer Scientific, pp. 425–432.Google Scholar
Macphail, R. I., Allen, M. J., Crowther, J., Cruise, G. M., and Whittaker, J. E., 2010, Marine inundation: effects on archaeological features, materials, sediments and soils: Quaternary International, no. 214, pp. 4455.Google Scholar
Macphail, R. I., Bill, J., Cannell, R., Linderholm, J., and Rødsrud, C. L., 2013, Integrated microstratigraphic investigations of coastal archaeological soils and sediments in Norway: the Gokstad ship burial mound and its environs including the Viking harbour settlement of Heimdaljordet, Vestfold: Quaternary International, v. 14th IWMSM, Site formation (R. I. Macphail, Ed.), no. 315, pp. 131146.Google Scholar
Macphail, R. I., and Courty, M. A., 1985, Interpretation and significance of urban deposits, in Edgren, T., and Jungner, H., eds., Proceedings of the Third Nordic Conference on the Application of Scientific Methods in Archaeology: Helsinki, The Finnish Antiquarian Society, pp. 7183.Google Scholar
Macphail, R. I., Courty, M. A., and Gebhardt, A., 1990, Soil micromorphological evidence of early agriculture in north-west Europe: World Archaeology, v. 22, no. 1, pp. 5369.Google Scholar
Macphail, R. I., Courty, M. A., Hather, J., and Wattez, J., 1997, The soil micromorphological evidence of domestic occupation and stabling activities, in Maggi, R., ed., Arene Candide: a Functional and Environmental Assessment of the Holocene Sequence (Excavations Bernabò Brea-Cardini 1940–50): Roma, Memorie dell’Istituto Italiano di Paleontologia Umana, pp. 5388.Google Scholar
Macphail, R. I., and Crowther, J., 2003, Mesolithic, Late Bronze Age and Medieval activity at Katherine Farm, Avonmouth, 1998 (microstratigraphy), in Allen, J. M., Godden, D., Matthews, C., and Powell, A. B., eds., Archaeology in the Severn Estuary 2002, Volume 13 (ed. P. Davies): Bath, Severn Estuary Levels Research Committee, pp. 89105.Google Scholar
Macphail, R. I., and Crowther, J. 2004, Tower of London Moat: sediment micromorphology, particle size, chemistry and magnetic properties, in Keevil, G., ed., Tower of London Moat Excavation, Historic Royal Palaces Monograph 1: Oxford, Oxford Archaeology, pp. 4143, 48–50, 78–79, 82–83, 155, 183–186, 202–204 and 271–284.Google Scholar
Macphail, R. I., and Crowther, J. 2007, Soil micromorphology, chemistry and magnetic susceptibility studies at Huizui (Yiluo region, Henan Province, northern China), with special focus on a typical Yangshao floor sequence: Bulletin of the Indo-Pacific Prehistory Association, v. 27, pp. 93113.Google Scholar
Macphail, R. I., and Crowther, J. 2008a, Illustrations from soil micromorphology and complementary investigations, in Thiemeyer, H., ed., Archaeological Soil Micromorphology – Contributions to the Archaeological Soil Micromorphology Working Group Meeting 3rd to 5th April 2008, Volume D30: Frankfurt A.M, Frankfurter Geowiss. Arb., pp. 8187.Google Scholar
Macphail, R. I., and Crowther, J. 2008b, Soil micromorphology and chemistry, in Ellis, C., and Powell, A. B., eds., An Iron Age settlement outside Battlesbury hillfort, Warminster and Sites along the Southern Range Road, Wessex Archaeology Report 22: Salisbury, Wessex Archaeology and Defence Estates, pp. 125132.Google Scholar
Macphail, R. I., and Crowther, J. 2009, Soil micromorphology, chemistry and magnetic susceptibility, in Phillips, M., Duncan, H., and Mallows, C., eds., Four Millenia of Human Activity Along the A505 Baldock Bypass, Hertfordshire, East Anglian Archaeology Report No. 128: Bedford, Albion Archaeology, pp. 8485, 122–123, CD-ROM Appendix 124 IV Soils.Google Scholar
Macphail, R. I., and Crowther, J. 2009, A2/A282/M25 improvement scheme, Dartford District, Kent: soil micromorphology, chemistry and magnetic susceptibility: Oxford Archaeology.Google Scholar
Macphail, R. I., and Crowther, J. 2010, Terminal 5: soil micromorphology, chemistry, magnetic susceptibility and particle size analyses (report for Framework Archaeology), www.framearch.co.uk/t5/evidence/: Oxford, Framework Archaeology, T5_Volume_2_Section_19.Google Scholar
Macphail, R. I., and Crowther, J. 2011a, Experimental pig husbandry: soil studies from West Stow Anglo-Saxon Village, Suffolk, UK, Antiquity Project Gallery, Volume 85, 330, (antiquity.ac.uk/projgall/macphail330/).Google Scholar
Macphail, R. I., and Crowther, J. 2011b, Soil micromorphology, chemistry and magnetic susceptibility, in Simmonds, A., Wenban-Smith, F., Bates, M., Powell, K., Sykes, D., Devaney, R., Stansbie, D., and Score, D., eds., Excavations in North-West Kent 2005–2007. One hundred thousand years of human activity in and around the Darent Valley, Oxford Archaeology Monograph No. 11: Oxford, Oxford Archaeology, pp. 180184.Google Scholar
Macphail, R. I., and Crowther, J. 2012, Micromorphology and bulk analyses, in Stafford, E., Goodburn, D., and Bates, M., eds., Landscape and Prehistory of the East London Wetlands. Investigations along the A13 DBFO Roadscheme, Tower Hamlets, Newham and Barking and Dagenham, 2000–2003, Oxford Archaeology Monograph No. 17: Oxford, Oxford Archaeology, pp. 270290.Google Scholar
Macphail, R. I., and Crowther, J. 2013, Appendix B: Soil micromorphology Tables, in Bates, M., and Stafford, E., eds., Thames Holocene. A Geoarchaeological Approach to the Investigation of the River Floodplain for High Speed 1, 1994–2003, pp. 191200.Google Scholar
Macphail, R. I., and Crowther, J. Forthcoming, Soil Analysis, in Fraser, B., ed., Oxford Archaeology North, From an ancient Eden to a new frontier: an archaeological journey along the Carlisle Northern Development Route. Lancaster, Lancaster Imprints.Google Scholar
Macphail, R. I., Crowther, J., Acott, T. G., Bell, M. G., and Cruise, G. M., 2003, The Experimental Earthwork at Wareham, Dorset after 33 years: changes to the buried LFH and Ah horizon: Journal of Archaeological Science, v. 30, pp. 7793.Google Scholar
Macphail, R. I., Crowther, J., and Berna, F., 2012, Soil micromorphology, microchemistry, chemistry, magnetic susceptibility and FTIR, in Biddulph, E., Foreman, S., Stafford, E., Stansbie, D., and Nicholson, R., eds., London Gateway. Iron Age and Roman salt making in the Thames Estuary; Excavations at Stanford Wharf Nature Reserve, Essex (library.thehumanjourney.net/909), Oxford Archaeology Monograph No. 18: Oxford, Oxford Archaeology, p. 193.Google Scholar
Macphail, R. I., Crowther, J., and Cruise, G. M., 1995, The soils, in Thurley, S., ed., The King’s Privy Garden at Hampton Court Palace 1689–1995: London, Apollo, pp. 116118.Google Scholar
Macphail, R. I., Crowther, J., and Cruise, G. M. 2007a, Micromorphology and post-Roman town research: the examples of London and Magdeburg., in Henning, J., ed., Post-Roman Towns and Trade in Europe, Byzantium and the Near-East. New methods of structural, comparative and scientific methods in archaeology: Berlin, Walter de Gruyter & Co. KG, pp. 303317.Google Scholar
Macphail, R. I., Crowther, J., and Cruise, G. M. 2007b, Microstratigraphy: soil micromorphology, chemistry and pollen, in Bowsher, D., Dyson, T., Holder, N., and Howell, I., eds., The London Guildhall. An archaeological history of a neighbourhood from early medieval to modern times, MoLAS Monograph 36: London, Museum of London Archaeological Service, pp. 18, 25–16, 35, 39, 55–16, 57, 59, 76, 90, 97, 98, 134, 154–135, 428–430.Google Scholar
Macphail, R. I., Crowther, J., and Cruise, G. M. 2008, Microstratigraphy, in Bateman, N., Cowan, C., and Wroe-Brown, R., eds., London’s Roman Amphitheatre: Guildhall Yard, City of London, MoLAS Monograph 35: London, Museum of London Archaeology Service, pp. 16, 95, 160–164.Google Scholar
Macphail, R. I., Crowther, J., and Macphail, G. M., 2011, Soil micromorphology, chemistry and magnetic susceptibility, in Ford, B. M., and Teague, S., eds., Winchester – A City in the Making. Archaeological investigations between 2002 and 2007 on the sites of Northgate House, Staple Gardens and the former Winchester Library, Jewry St., Oxford Archaeology Monograph No 12: Oxford, Oxford Archaeology, p. 376, CD Part 373.317 (Soil Micromorphology Report.pdf).Google Scholar
Macphail, R. I., Crowther, J., and Macphail, G. M. 1996a, Soil micromorphology, in Bell, M., Fowler, P. J., and Hillson, S. W., eds., The Experimental Earthwork Project 1960–1992, Research Report 100: York, Council for British Archaeology, pp. 95107.Google Scholar
Macphail, R. I., Crowther, J., and Macphail, G. M. 1996b, soil micromorphology, in A moated site in Tempsford Park, Tempsford: edited by Shotliff, D.: Bedfordshire Archaeology, v. 22, pp. 123124.Google Scholar
Macphail, R. I., and Cruise, G. M., 1995, Imjin River Basin Project, Korea: geoarchaeological and environmental report on Chuoli, Kawoli, Chongokni, Kado, KD and PS: Department of Archaeology, Seoul National University.Google Scholar
Macphail, R. I., and Cruise, G. M., 2000a, Rescuing our urban archaeological soil heritage: a multidisciplinary microstratigraphical approach, in Burghardt, W., and Dornauf, C., eds., Proceedings of the First International Conference on Soils of Urban, Industrial, Traffic and Mining Areas., Volume 1: Essen, IUSS/IBU, pp. 914.Google Scholar
Macphail, R. I., and Cruise, G. M., 2000b, Soil micromorphology on the Mesolithic site, in Bell, M., Caseldine, A., and Neumann, H., eds., Prehistoric Intertidal Archaeology in the Welsh Severn Estuary, Research Report 120: York, Council for British Archaeology, pp. 5557 and CD-ROM.Google Scholar
Macphail, R. I., and Cruise, G. M. 2001, The soil micromorphologist as team player: a multianalytical approach to the study of European microstratigraphy, in Goldberg, P., Holliday, V., and Ferring, R., eds., Earth Science and Archaeology: New York, Kluwer Academic/Plenum Publishers, pp. 241267.Google Scholar
Macphail, R. I., Cruise, G. M., Allen, M. J., and Linderholm, J., 2006, A rebuttal of the views expressed in “Problems of unscientific method and approach in Archaeological soil and pollen analysis of experimental floor deposits; with special reference to Butser Ancient Farm, Hampshire, UK by R. I. Macphail, G.M. Cruise, M. Allen, J. Linderholm and P. Reynolds” by Matthew Canti, Stephen Carter, Donald Davidson and Susan Limbrey: Journal of Archaeological Science, v. 33, pp. 299305.Google Scholar
Macphail, R. I., Cruise, G. M., Allen, M. J., Linderholm, J., and Reynolds, P., 2004, Archaeological soil and pollen analysis of experimental floor deposits; with special reference to Butser Ancient Farm, Hampshire, UK: Journal of Archaeological Science, v. 31, pp. 175191.Google Scholar
Macphail, R. I., Cruise, G. M., Courty, M. A., Crowther, J., and Linderholm, J., 2016, 27. E6 Gudbrandsdalen Valley Project (Brandrud, Fryasletta, Grytting and Øybrekka), Oppland, Norway: soil micromorphology (with selected microchemistry, bulk soil chemistry, carbon polymer, particle size and pollen analyses), in Gundersen, I. M., ed., Gård og utmark i Gudbrandsdalen. Arkeologiske undersøkelser i Fron 2011–2012: Kristiansand, Portal forlag, pp. 304–317.Google Scholar
Macphail, R. I., Graham, E., Crowther, J., and Turner, S., 2017, Marco Gonzalez, Ambergris Caye, Belize: A geoarchaeological record of ground raising associated with surface soil formation and the presence of a Dark Earth. Journal of Archaeological Science, v. 77, Geoarchaeology in the Humid Tropics: Practice, Problems, Prospects, pp. 35–51.Google Scholar
Macphail, R. I., Cruise, G. M., Engelmark, R., and Linderholm, J., 2000, Integrating soil micromorphology and rapid chemical survey methods: new developments in reconstructing past rural settlement and landscape organization, in Roskams, S., ed., Interpreting Stratigraphy, Volume 9: York, University of York, pp. 7180.Google Scholar
Macphail, R. I., Cruise, G. M., Gebhardt, A., and Linderholm, J., Forthcoming, West Heslerton: soil micromorphology and chemistry of the Roman and Saxon deposits, in Tipper, J., ed., West Heslerton Anglo-Saxon Settlement: Yedingham, Landscape Research Centre.Google Scholar
Macphail, R. I., Cruise, G. M., and Linderholm, J., 2001, Soil micromorphology and chemistry in Ellis, J.C. and Rawlings, M., (eds.), Excavations at Balksbury Camp, Andover 1995–97: Hampshire Field Club and Archaeological Society, 56: 2193., v. 56, pp. 67–70.Google Scholar
Macphail, R. I., Cruise, G. M. 2014, Soil micromorphology and chemistry, in Ashwin, T., and Tester, A., eds., A Roman Settlement in the Waveney Valley: Excavations at Scole, 1993–4, Report No. 152, Chapter 9, Section VI, pp. 422–431 (CD).Google Scholar
Macphail, R. I., Cruise, G. M., and Mellalieu, S. J., 1999, Soil micromorphological and diatom analysis of thin sections and microchemical analysis of polished blocks, in Niblett, R., ed., The Excavation of a Ceremonial site at Folly Lane, Verulamium, Britannia Monograph No. 14: London, Britannia, pp. 365384.Google Scholar
Macphail, R. I., Cruise, G. M., Mellalieu, S. J., and Niblett, R., 1998, Micromorphological interpretation of a “Turf-filled” funerary shaft at St. Albans, United Kingdom: Geoarchaeology, v. 13, no. 6, pp. 617644.Google Scholar
Macphail, R. I., Galinié, H., and Verhaeghe, F., 2003b, A future for dark earth?: Antiquity, v. 77, no. 296, pp. 349358.Google Scholar
Macphail, R. I., and Goldberg, P., 1990a, The micromorphology of Tree Subsoil Hollows: their significance to soil science and archaeology, in Douglas, L. A., ed., Soil Micromorphology: A Basic and Applied Science: Amsterdam, Elsevier, pp. 425429.Google Scholar
Macphail, R. I., and Goldberg, P. 1995, Recent advances in micromorphological interpretations of soils and sediments from archaeological sites, in Barham, A. J., and Macphail, R. I., eds., Archaeological Sediments and Soils: Analysis, Interpretation and Management: London, Institute of Archaeology, pp. 1–24e.Google Scholar
Macphail, R. I., and Goldberg, P. 1999, The soil micromorphological investigation of Westbury Cave, in Andrews, P., Cook, J., Currant, A., and Stringer, C., eds., Westbury Cave. The Natural History Museum Excavations 1976–1984: Bristol, CHERUB (Centre for Human Evolutionary Research at the University of Bristol), pp. 5986.Google Scholar
Macphail, R. I., and Goldberg, P. 2000, Geoarchaeological investigations of sediments from Gorham’s and Vanguard Caves, Gibraltar: microstratigraphical (soil micromorphological and chemical) signatures, in Stringer, C. B., Barton, R. N. E., and Finlayson, C., eds., Neanderthals on the Edge: Oxford, Oxbow, pp. 183200.Google Scholar
Macphail, R. I., and Goldberg, P. 2003, Gough’s Cave, Cheddar, Somerset: Microstratigraphy of the Late Pleistocene/earliest Holocene -sediments: Bulletin of Natural History Museum London (Geology). v. 58 (supp), pp. 5158.Google Scholar
Macphail, R. I., and Goldberg, P. 2004, Geoarchaeology of Site GTMO-1 and Cuzco Beach, U.S. Naval Station Guantanomo Bay: Thin Section and Bulk Samples Studies, in Sara, T., and Keegan, W., eds., Archaeological Survey and Palaeoenvironmental Investigations of Portions of U.S. Naval Station Guantanomo Bay, Cuba, Miscellaneous Reports of Investigations No. 310: Newport News, Geo-Marine Inc, pp. D1-D-35.Google Scholar
Macphail, R. I., and Goldberg, P. 2010, Archaeological materials, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 589622.Google Scholar
Macphail, R. I., and Goldberg, P. 2012, Soil Micromorphology of Gibraltar Coprolites, in Barton, R. N. E., ed., Gibraltar Neanderthals in Context: A report of the 1995–98 excavations at Gorham’s & Vanguards Caves, Gibraltar, Volume Oxford University School of Archaeology: Monograph 75: Oxford, Oxford University Press, pp. 240243.Google Scholar
Macphail, R. I., Goldberg, P., and Barton, R. N. E., 2012, Vanguard Cave sediments and soil micromorphology, in Barton, R. N. E., Stringer, C., and Finlayson, C., eds., Neanderthals in Context. A report of the 1995–1998 excavations at Gorham’s and Vanguard Caves, Gibraltar, Monograph 75: Oxford, Oxford University School of Archaeology, pp. 193210; Color figures at: www.arch.ox.ac.uk/gibraltar.Google Scholar
Macphail, R. I., Bill, J., Crowther, J., Haită, C., Linderholm, J., Popovici, D., and Rødsrud, C. L., 2016b, European ancient settlements – a guide to their composition and morphology based on soil micromorphology and associated geoarchaeological techniques; introducing the contrasting sites of Chalcolithic Borduşani-Popină, Borcea River, Romania and the Viking Age Heimdaljordet, Vestfold, Norway. Quaternary International, http://dx.doi.org/10.1016/j.quaint.2016.08.049.Google Scholar
Macphail, R. I., Haita, C., Bailey, D. W., Andreescu, R., and Mirea, P., 2008b, The soil micromorphology of enigmatic Early Neolithic pit-features at Magura, southern Romania: Romanian Association of Archaeology. Studies of Prehistory, v. 5, pp. 6177.Google Scholar
Macphail, R. I., Hather, J., Hillson, S. W., and Maggi, R., 1994, The Upper Pleistocene deposits at Arene Candide: soil micromorphology of some samples from the Cardini 1940–42 excavations: Quaternaria Nova, v. IV, pp. 79100.Google Scholar
Macphail, R. I., and Linderholm, J., 2004a, ‘Dark earth’: recent studies of ‘dark earth’ and ‘dark earth-like’ microstratigraphy in England, in Verslype, L., and Brulet, R., eds., Terres Noire; Dark Earth: Louvain-la-Neuve, Université Catholique de Louvain, pp. 3542.Google Scholar
Macphail, R. I., and Linderholm, J. 2004b, Neolithic land use in south-east England: a brief review of the soil evidence, in Cotton, J., and Field, D., eds., Towards a New Stone Age, Research Report 137: York, CBA, pp. 2937.Google Scholar
Macphail, R. I., and Linderholm, J. 2011, Micromorphology in Burch, M., Treveil, P., and Keene, D., eds., The development of early medieval and later Poultry and Cheapside: excavations at 1 Poultry and vicinity, City of London, MOLA Monograph 38: London, Museum of London Archaeology, CD Tables 11–12.Google Scholar
Macphail, R. I., and Linderholm, J., 2013, B13, Oslo Wreck Microstratigraphy Report 1: soil micromorphology, chemistry and magnetic susceptibility (report for Cultural History Museum, University of Oslo): Institute of Archaeology, University College London.Google Scholar
Macphail, R. I., and Linderholm, J. In Press/2017, Avaldsnes: Scientific Analyses – Microstratigraphy (soil micromorphology and microchemistry, soil chemistry and magnetic susceptibility), in Skre, D., Bauer, E. M., and Østmo, M. A., eds., Avaldsnes Royal Manor: Berlin, De Gruyter.Google Scholar
Macphail, R. I., Linderholm, J., and Karlsson, N., 2006, Scanian pithouses; interpreting fills of grubenhäusser: examples from England and Sweden, in Engelmark, R., and Linderholm, J., eds., Proceedings from the 8th Nordic Conference on the Application of Scientific Methods in Archaeology in Ümea 2001, Archaeology and Environment 21: Umeå, Umeå University, pp. 119127.Google Scholar
Macphail, R. I., and McAvoy, J. M., 2008, A micromorphological analysis of stratigraphic integrity and site formation at Cactus Hill, an Early Paleoindian and hypothesized pre-Clovis occupation in south-central Virginia, USA: Geoarchaeology, v. 23, no. 5, pp. 675694.Google Scholar
Macphail, R. I., Romans, J. C. C., and Robertson, L., 1987, The application of micromorphology to the understanding of Holocene soil development in the British Isles; with special reference to cultivation, in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 669676.Google Scholar
MacWilliams, A. C., Kuehn, D. D., Homburg, J. A., Murrell, M. L., Leckman, P. O., Macphail, R. I., Adams, K. R., Smith, S., and Yosti, C., In press, Investigations of an El Paso Phase Water Reservoir in the Northern Chihuahuan Desert: Journal of Archaeological Science.Google Scholar
Maggi, R., 1990, Archeologia Dell’Appennino Ligure. Gli scavi del Castellaro di Uscio: un insediamento di crinale occupato dal Neolitico alla conquista Romana, Istituto Internazionale Di Studi Liguri, Collezinoe Monographie preistoriche ed Archaeologiche, Volume VIII: Chiavari, Bordighera.Google Scholar
Maggi, R. 1997, Arene Candide: a Functional and Environmental Assessment of the Holocene Sequence (Excavations Bernabò Brea-Cardini 1940–50), Roma, Memorie dell’Istituto Italiano di Paleontologia Umana, p. 643.Google Scholar
Maggi, R., and Nisbet, R., 2000, Alberi da foraggio scalvatura Neolitica: Nuovi data dalle Arene Candide (Fodder trees and Neolithic shredding: new data from Arene Candide), in Pessina, A., and Musicio, G., eds., La Neolitizzazione Tra Oriente e Occidente: Udine, Comune di Udine; Museo Fruilano di Storia Naturale, pp. 289308.Google Scholar
Malim, T., and Hayes, L., 2011, An engineered Iron Age road, associated Roman use (Margary Route 64), and Bronze Age activity recorded at Sharpstone Hill, 2009: Transactions of the Shropshire Archaeological and Historical Society, v. 85, pp. 780.Google Scholar
Mallol, C., and Carbonell i Roura, E., 2007, The Collapse of Gran Dolina Cave,Sierra de Atapuerca, Spain: SiteFormation Processes of Layer TD10-1: Geoarchaeology, v. 23, no. 1, pp. 1341.Google Scholar
Mallol, C., Hernández, C. M., Cabanes, D., Machado, J., Sistiaga, A., Pérez, L., and Galván, B., 2013a, Human actions performed on simple combustion structures: An experimental approach to the study of Middle Palaeolithic fire: Quaternary International, v. 315, pp. 315.Google Scholar
Mallol, C., Hernández, C. M., Cabanes, D., Sistiaga, A., Machado, J., Rodriguez, Å., and Pérez, L., 2013b, The Black Layer of Middle Palaeolithic combustion structures. Interpretation and archaeostratigraphic implications: Journal of Archaeological Science, v. 40, pp. 25152537.Google Scholar
Mallol, C., Marlowe, F. W., Wood, B. M., and Porter, C. C., 2007, Earth, wind, and fire: ethnoarchaeological signals of Hadza fires: Journal of Archaeological Science, v. 34, pp. 20352052.Google Scholar
Maltby, E., and Caseldine, C. J., 1982, Prehistoric soil and vegetation development on Bodmin Moor, Southwestern England: Nature, v. 297, pp. 397400.Google Scholar
Manby, T. G., 1976, The excavation of the Kilham long barrow, East Riding of Yorkshire: Proceedings of the Prehistoric Society, v. 42, pp. 111159.Google Scholar
Mannino, M. A., and Thomas, K., 2004–6, New radiocarbon dates for hunter gatherers and early farmers in Sicily: Accordia Research Papers, v. 10, pp. 1333.Google Scholar
Maslin, S. P., 2015, The taphonomy and micromorphology of sunken-featured buildings from Lyminge, Kent: A comparative mixed-method analysis: Environmental Archaeology, v. 20, no. 2, pp. 202220.Google Scholar
Mateu, M., Bergadà, M. M., and Garcia i Rubert, D., 2013, Manufacturing technical differences employing raw earth at the protohistoric site of Saint Jaume (Alcanar, Tarragona, Spain): construction and furniture elements: Quaternary International, v. 315, pp. 7686.Google Scholar
Matsui, A., Hiraya, R., Mijaji, A., and Macphail, R. I., 1996a, Availability of soil micromorphology in archaeology in Japan (in Japanese): Archaeology Society of Japan, v. 38, pp. 149152.Google Scholar
Matthews, W., 1995, Micromorphological characterisation and interpretation of occupation deposits and microstratigraphic sequences at Abu Salabikh, Southern Iraq, in Barham, A. J., and Macphail, R. I., eds., Archaeological Sediments and Soils: Analysis, Interpretation and Management: London, Institute of Archaeology, pp. 4174.Google Scholar
Matthews, W. 2005, Micromorphological and microstratigraphic traces of uses and concept of space., in Hodder, I., ed., Inhabiting Çatalhöyük: Reports from the 1995–99 Seasons: Cambridge, McDonald Institute Monographs, pp. 355398.Google Scholar
Matthews, W. 2010, Geoarchaeology and taphonomy of plant remains and microarchaeological residues in early urban environments in the Ancient Near East: Quaternary International, v. 214, no. 1–2, pp. 98113.Google Scholar
Matthews, W., French, C. A. I., Lawrence, T., and Cutler, D., 1996, Multiple Surfaces: the Micromorphology, in Hodder, I., ed., On the Surface: Çatalhöyük 1993–95: Cambridge, McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, pp. 301342.Google Scholar
Matthews, W., French, C. A. I., Lawrence, T., Cutler, D. F., and Jones, M. K., 1997, Microstratigraphic traces of site formation processes and human activity: World Archaeology, v. 29, pp. 281308.Google Scholar
Matthews, W., French, C. A. I., Lawrence, T., Cutler, D. F., and Jones, M. K., 1998, Microstratigraphy and micromorphology of depositional sequences, in Oates, D., Oates, J., and McDonald, H., eds., Excavations at Tell Brak., The Mitani and Old Babylonian periods, British School of Archaeology in Iraq, pp. 135140.Google Scholar
Matthews, W., Hastorf, C. A., and Ergenekon, B., 2000, Ethnoarchaeology: studies in local villages aimed at understanding aspects of the Neolithic site, in Hodder, I., ed., Towards Reflexive Method in Archaeology: the Example at Çatalhöyük: Cambridge, McDonald Institute for Archaeological Research and British Institute of Archaeology at Ankara, pp. 177188.Google Scholar
Matthews, W., and Postgate, J. N., 1994, The imprint of living in a Mesopotamian City: questions and Answers, in Luff, R., and Rowley Conwy, P., eds., Whither Environmental Archaeology?: Monograph 38: Oxford, Oxbow Books, pp. 171212.Google Scholar
Mayer, J., and Homburg, J. A., Micromorphology of Intermediate Period “Middens” in the Ballona of Coastal California, in Proceedings Archaeological Sciences of the Americas Symposium, September 13–16, 2006, Tucson, Arizona, 2006, Schedule of Events and Abstracts, The University of Arizona., pp. 39–40.Google Scholar
McAuliffe, J. R., 2000, Desert Soils, in Philipps, S. J., and Comus, P. W., eds., A Natural History of the Sonoran Desert: Tucson, Arizona-Sonora Desert Museum Press, pp. 87104.Google Scholar
McCann, J. M., Woods, W. I., and Meyer, D. W., 2001, Organic matter and anthrosols in Amazonia: Interpreting the Amerindian legacy, in Rees, R. M., Ball, B. C., Campbell, C. D., and Watson, C. A., eds., Sustainable Management of Soil Organic Matter: Wallingford, CAB International, pp. 180189.Google Scholar
McCormick, M., 2007, Towards a molecular history of the Justinian pandemic, in Little, L. K., ed., Plague and the End of Antiquity. The Pandemic of 541–750: Cambridge, Cambridge University Press/The American Academy in Rome, pp. 290312.Google Scholar
McIntyre, D. S., 1958a, Permeability measurements of soil crusts formed by raindrop impact: Soil Science, v. 85, pp. 185189.Google Scholar
McIntyre, D. S. 1958b, Soil splash and the formation of surface crusts by raindrop impact: Soil Science, v. 85, pp. 261266.Google Scholar
McKeague, J. A., 1983, Clay skins and argillic horizons, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology, Volume 2: Soil Genesis: Berkhamsted, AB Academic Publishers, pp. 367388.Google Scholar
McOmish, D., Field, D., and Brown, G., 2010, The Late Bronze Age and Early Iron Age Midden Site at East Chisenbury: Wiltshire Archaeological and Natural History Magazine, v. 103, pp. 35101.Google Scholar
McOmish, D., 1996, East Chisenbury: ritual and rubbish at the Bronze Age-Iron Age transition: Antiquity, v. 267, pp. 6876.Google Scholar
Mees, F., and Stoops, G., 2010, Sulphidic and sulphuric materials, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 543568.Google Scholar
Mees, F., and Tursina, T. V., 2010, Salt minerals in saline soils and salt crusts, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 441469.Google Scholar
Merkel, J. F., 1997, Geochemical and stratigraphic sampling strategies for ancient tin smelting. In, in Budd, P., and Gale, D., eds., Prehistoric Extractive Metallurgy in Cornwall. Conference Proceedings from Camborne School of Mines, Redruth, 11 July 1992: Truro, Cornwall Archaeological Unit, pp. 41–44.Google Scholar
Merriman, N., 1992, Predicting the unexpected: prehistoric sites recently discovered under alluvium in central London, in Needham, S., and Macklin, M. G., eds., Alluvial Archaeology in Britain: Oxford, Oxbow, pp. 261267.Google Scholar
Mikkelsen, J. H., Langohr, R., Boas, N. A., and Macphail, R. I., 2006, Land use and environmental degradation in Bronze Age settlements, Eastern Jutland, Denmark, in Engelmark, R., and Linderholm, J., eds., Proceedings from the 8th Nordic Conference on the Application of Scientific Methods in Archaeology in Umeå 2001, Archaeology and Environment 21: Ümea, Ümea University, pp. 8192.Google Scholar
Mikkelsen, J. H., Langohr, R., and Macphail, R. I., 2007, Soilscape and land-use evolution related to drift sand movements since the Bronze Age in Eastern Jutland, Denmark: Geoarchaeology, v. 22, no. 2, pp. 155180.Google Scholar
Mikkelsen, J. H., Langohr, R., Macphail, R. I., and Vanwesenbeeck, V., 2003, Roman postal byres, a case study from the coversand belt of northern Belgium, in Boschian, G., ed., Second International Conference on Soils and Archaeology, Pisa, 12th-15th May, 2003. Extended Abstracts: Pisa, Dipartimento di Scienze Archeologiche, Università di Pisa, pp. 118119.Google Scholar
Mikkelsen, P. H., and Bartholin, T. S., 2013, Vedanatomiske analyser fra E18-prosjektet Gulli–Langåker, in Gerpe, L.-E., ed., E18-prosjektet Gulli-Langåker. Oppsummering og arkeometriske analyser, Varia, Bind 3: Oslo, Fagbokforlaget, pp. 85110.Google Scholar
Milek, K., 2004, Aoalstraeti, Reykjavik, 2001: geoarchaeological report on deposits within the house and the soils immediately pre- and post-dating its occupation., in Roberts, H., ed., Excavations at Aoalstraeti, Reykjavik, 2003: Reykavik, Fornleifastofnum Íslands, pp. 73114.Google Scholar
Milek, K. 2005, Soil micromorphology, in Sharples, N. M., ed., A Norse farmstead in the Outer Hebrides: Excavations at Mound 3, Bornais, South Uist: Oxford, Oxbow, pp. 98104.Google Scholar
Milek, K. 2006, Houses and Households in Early Icelandic Society: Geoarchaeology and the Interpretation of Social Space, PhD thesis: University of Cambridge, p. 413.Google Scholar
Milek, K. 2012, Floor formation processes and the interpretation of site activity areas: An ethnoarchaeological study of turf buildings at Thverá, northeast Iceland: Journal of Anthropological Archaeology, v. 31, pp. 119137.Google Scholar
Milek, K., and French, C., 2007, Soils and sediments in the settlement and harbour at Kaupang., in Skre, D., ed., Kaupang in Skiringssal: Aarhus, Aarhus University Press, pp. 321360.Google Scholar
Mjærum, A., 2013, Boplasspor fra mellommesolitikum og bosetnings- og dyrkningsspor fra eldre jernalder på Unnerstvedt og Ragnhildrød (lok. 35), in Gerpe, L.-E., ed., E18-prosjektet Gulli-Langåker. Jordbruksbosetnining og graver i Tønsberg og Stokke, Bind 2: Bergen, Fagbokforlaget, pp. 1980.Google Scholar
Moe, D., 1983, Palynology of sheep’s faeces: relationship between pollen content, diet and local pollen rain: Grana, v. 22, pp. 105113.Google Scholar
Mokma, D. L., and Buurman, P., 1982, Podzols and Podzolisation in Temperate Regions, Wageningen, International Soil Museum.Google Scholar
Moormann, F. R., and van Breeman, N., 1978, Rice: Soil, Water, Land, Manila, International Rice Research Institute, p. 185.Google Scholar
Mücher, H. J., 1974, Micromorphology of slope deposits: the necessity of a classification, in Rutherford, G. K., ed., Soil Microscopy: Kingston, Ontario, The Limestone Press, pp. 553556.Google Scholar
Mücher, H. J., and de Ploey, J., 1977, Experimental and micromorphological investigation of erosion and redeposition of loess by water: Earth Surface Processes and Landforms, v. 2, pp. 117124.Google Scholar
Mücher, H. J., and Morozova, T. D., 1983, The application of soil micromorphology in Quaternary geology and geomorphology, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology: Berkhamsted, A B Academic Publishers, pp. 151194.Google Scholar
Mücher, H. J., Slotboom, R. T., and ten Veen, W. J., 1990, Palynology and micromorphology of a man-made soil. A reconstruction of the agricultural history since Late-medieval times of the Posteles in the Netherlands: Catena, v. 17, pp. 5567.Google Scholar
Mücher, H. J., van Steijn, H., and Kwaad, F. J. P. M., 2010, Colluvial and mass wasting deposits, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 3748.Google Scholar
Murphy, C. P., Bullock, P., and Turner, R. H., 1977, The measurement and characterisation of voids in thin sections by image analysis. Part 1. Principles and techniques: Journal of Soil Science, v. 29, pp. 498508.Google Scholar
Murphy, C. P., and Kemp, R. A., 1984, The over-estimation of clay and the under-estimation of pores in soil thin sections: Journal of Soil Science, v. 35, pp. 481496.Google Scholar
Murphy, C. P., McKeague, J. A., Bresson, L. M., Bullock, P., Kooistra, M. J., Miedema, R., and Stoops, G., 1985, Description of soil thin sections: an international comparison: Geoderma, v. 35, pp. 1537.Google Scholar
Murphy, P., and Fryer, V., 1999, The plant macrofossils, in Niblett, R., ed., The Excavation of a Ceremonial site at Folly Lane, Verulamium., Britannia Monograph No. 14: London, Society for the Promotion of Roman Studies, pp. 384388.Google Scholar
Myaji, A., 2003, Analysis of cultivated fields and their soils in Japanese archaeology: soil micromorphology, in Boschian, G., ed., Second International Conference on Soils and Archaeology, Pisa, 12th-15th May, 2003. Extended Abstracts: Pisa, Dipartimento di Scienze Archeologiche, Università di Pisa, p. 10.Google Scholar
Myhre, B., 2004, Agriculture, landscape and society ca. 4000 BC-AD 800, in Almås, R., ed., Norwegian Agricultural History: Trondheim, Tapir Academic Press, pp. 1477.Google Scholar
Newell, R. R., 1980, Mesolithic dwelling structures: fact and fantasy: Veröffentlichungen des Museums für Ur- und Frühgeschichte Potsdam, Band 14/15, pp. 235284.Google Scholar
Nicolaysen, N., 1882, Langskibet fra Gokstad ved Sandefjord, Kristiania, Cammermeyer.Google Scholar
Nicosia, C., 2006, Archaeopedological Study of Medieval “Dark Earth”From Firenze, Italy [MSc]: Gent University, p. 112.Google Scholar
Nicosia, C., Devos, Y., and Macphail, R. I., In press/2017, European ‘Dark Earth’, in Nicosia, C., and Stoops, G., eds., Encyclopedia of Archaeological Soil and Sediment Micromorphology: Chichester, Wiley.Google Scholar
Nicosia, C., and Stoops, G., In press/2017, Archaeological Soil and Sediment Micromorphology, Chichester, Wiley.Google Scholar
Nisbet, R., 1997, Arene Candide: charcoal remains and prehistoric woodland use, in Maggi, R., ed., Arene Candide: a Functional and Environmental Assessment of the Holocene Sequence (Excavations Bernabò Brea-Cardini 1940–50): Roma, Memorie dell’Istituto Italiano di Paleontologia Umana, pp. 103112.Google Scholar
Nodarou, E., Frederick, C., and Hein, A., 2008, Another (mud)brick in the wall: scientific analysis of Bronze Age earthen construction materials from East Crete: Journal of Archaeological Science, v. 35, pp. 29973015.Google Scholar
Norman, P., and Reader, F. W., 1912, Further discoveries relating to Roman London, 1906–12: Archaeologia, v. LXIII, pp. 257344.Google Scholar
Nørnberg, P., and Courty, M. A., 1985, Standard geological methods used on archaeological problems, in Edgren, T., and Jungner, H., eds., Proceedings of the Third Conference on the Application of Scientific Methods in Archaeology, ISKOS 5: Helsinki, The Finnish Antiquarian Society, pp. 107118.Google Scholar
O’Brien, R. M. G., Romans, J. C. C., and Robertson, L., 1979, Three profiles from Elephant Island, South Shetland Islands: British Antarctic Survey Bulletin, v. 47, pp. 112.Google Scholar
Oades, J. M., and Townsend, W. N., 1963, The detection of ferromagnetic minerals in soils and clays: Journal of Soil Science, v. 14, no. 2, pp. 177187.Google Scholar
Oldfield, F., and Crowther, J., 2007, Establishing fire incidence in temperate soils using magnetic measurements: PALAEO, v. 249, pp. 362369.Google Scholar
Oldfield, F., Krawiecki, A., Maher, B., Taylor, J. J., and Twigger, S., 1985, The role of mineral magnetic measurements in archaeology, in Fieller, N. R. J., Gilbertson, D. D., and Ralph, N. G. A., eds., Palaeoenvironmental investigations: research design, methods and data analysis, International Series 266: Oxford, British Archaeological Reports, pp. 2943.Google Scholar
Pagliai, M., and Stoops, G., 2010, Physical and Biological Surface Crusts, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, pp. 419440.Google Scholar
Pape, J. C., 1970, Plaggen soils in the Netherlands: Geoderma, v. 4, pp. 229255.Google Scholar
Parker Pearson, M., Bevins, R., Ixer, R., Pollard, J., Richards, C., Welham, K., Chan, B., Edinborough, K., Hamilton, D., Macphail, R., Schlee, D., Simmons, E., and Smith, M., 2015, Craig Rhos-y-felin: a Welsh bluestone megalith quarry for Stonehenge: Antiquity, v. 89 348, pp. 13311352.Google Scholar
Patto, P. M., Clement, C. R., and Forbes, T. J., 1974, Permanent Grassland Studies. 2, Grassland Poaching in England and Wales: Maidenhead, The Permanent Pasture Group.Google Scholar
Payton, R. W., 1983, The micromorphology of some fragipans and related horizons in British soils with particular reference to their consistence characteristics, in Bullock, P., and Murphy, C. P., eds., Soil Micromorphology, Volume 1: Techniques and Applications: Berkhamsted, A B Academic Publishers, pp. 317336.Google Scholar
Pearce, J., and Luff, R., 1994, The Taphonomy of Cooked Bone, in Luff, R., and Rowley Conwy, P., eds., Whither Environmental Archaeology, Oxbow Monograph 38: Oxford, Oxbow, pp. 5156.Google Scholar
Perrin, R. M. S., Willis, E. H., and Hodge, C. A. H., 1964, Dating of humus podzols by residual radiocarbon activity: Nature, v. 202, pp. 165166.Google Scholar
Peterken, G. F., 1996, Natural Woodland. Ecology and Conservation in Northern Regions, Cambridge, Cambridge University Press.Google Scholar
Pettijohn, F. J., 1975, Sedimentary Rocks, New York, Harper & Row, Publishers.Google Scholar
Pettijohn, F. J., Potter, E. P. and Siever, R. 1973. Sand and Sandstone. New York, Springer-Verlag.Google Scholar
Phillips, S. J., and Comus, P. W., 2000, A Natural History of the Sonoran Desert: Tucson, Arizona-Sonora Desert Museum, p. 628.Google Scholar
Pirrie, D., Butcher, A. R., Power, M. R., Gottlieb, P., and Miller, G. L., 2004, Rapid quantitative mineral and phase analysis using an automated scanning electron microscope (QemScan); potential applications in forensic science., in Pye, K., and Croft, D. J., eds., Forensic Geoscience: Principles, Techniques and Applications, Special Publications 232: London, The Geological Society of London, pp. 123136.Google Scholar
Pope, M., Bates, M., Conneller, C., Reubens, K., Underhill, D., Shaw, A., and Wragg-Sykes, R., 2012, Quaternary Environments and Archaeology of Jersey: A new multidisciplinary project looking at the early prehistoric occupation of the English Channel Region., in Reubens, K., Romanowska, I., and Bynoe, R., eds., Unravelling the Palaeolithic: 10 years of research at the Centre for the Archaeology of Human Origins, British Archaeological Reports 2400: Oxford, Oxbow.Google Scholar
Popovici, D., Haită, C., Bălăşescu, A., Radu, V., Vlad, F., and Tomescu, I., 2003, Archaeological Pluridiscipinary Researches at Borduşani-Popină, Bucharest, National Museum of Romanian History, Ialomita County Museum, p. 196.Google Scholar
Powlesland, D., 1998, The West Heslerton Assessment, Internet Archaeology, Volume 5.Google Scholar
Powlesland, D., Haughton, C., and Hanson, J., 1986, Excavations at Heslerton, North Yorkshire 1978–82: Archaeology Journal, v. 143, pp. 53173.Google Scholar
Preece, R. C., and Bridgland, D. R., 1998, Late Quaternary Environmental Change in North-West Europe: Excavations at Holywell Coombe, South-east England: London, Chapman & Hall, p. 422.Google Scholar
Preece, R. C., Kemp, R. A., and Hutchinson, J. N., 1995, A Late-glacial colluvial sequence at Watcombe Bottom, Ventnor, Isle of Wight, England: Journal of Quaternary Science, v. 10, no. 2, pp. 107121.Google Scholar
Prentice, C., 1985, Pollen representation, source area, and basin size: toward a unified theory of pollen analysis: Quaternary Research, v. 23, pp. 7686.Google Scholar
Prentiss, A. M., Foor, T. A., Cross, G., Harris, L. E., and Wanzenried, M., 2012, The Cultural Evolution of Material Wealth Based Inequality at Bridge River, British Columbia: American Antiquity, v. 77, pp. 542564.Google Scholar
Pye, E., 2000/2001, Wall painting in the Roman empire: colour, design and technology: Archaeology International, pp. 2427.Google Scholar
Quinn, P. S., 2013, Ceramic Petrography. The Interpretation of Archaeological Pottery and Related Artefacts in Thin Section, Oxford, Archeopress, p. 234.Google Scholar
Ragg., J. M., Beard, G. R., Hollis, J. M., Jones, R. J. A., Palmer, R. C., Reeve, M. J., and Whitfield, W. A. D., 1983, Soils of England and Wales. Sheet 3. Midland and Western England: Ordnance Survey.Google Scholar
Rasmussen, P., 1993, Analysis of goat/sheep faeces from Egolzwil 3, Switzerland: evidence for branch and twig foddering of livestock in the Neolithic: Journal of Archaeological Science, v. 20, pp. 479502.Google Scholar
Reineck, H. E., and Singh, I. B., 1986, Depositional Sedimentary Environments, Berlin, Springer-Verlag.Google Scholar
Rellini, I., Firpo, M., Martino, G., Riel-Salvatore, J., and Maggi, R., 2013, Climate and environmental changes recognized by micromorphology in Palaeolithic deposits at Arene Candide (Liguria, Italy): Quaternary International, v. 315 Site formation processes in archaeology, pp. 4255.Google Scholar
Renfrew, C., and Bahn, P., 2001, Archaeology: Theories, Methods and Practice, London, Thames and Hudson.Google Scholar
Rentzel, P., 1997, Geoarchäologische beobachtungen an der Römischen Wasserleitung von Leistal nach Augst, in Ewald, J., Hartman, M., and Rentzel, P., eds., Die Römischen Wasserleitung von Leistal nach Augst: Liestal, Schweiz, Berichte aus Archäologie und Kantonmuseum Baselland, pp. 3762.Google Scholar
Rentzel, P. 1998, Ausgewähite Grubenstrukturen aus spätlatènezeitlichen Fundstelle Basel-Gasfabrik: Geoarchäologische interpretation der Grubenfüllungen, Jahresbericht der archäologischen Bodenforschung des Kantons Basel-Stadt 1995: Basel, pp. 3579.Google Scholar
Rentzel, P. 2009, The arena floor of the amphitheater of Augst-Nine towers. Geo-archaeological investigations., in Hufschmid, T., ed., Amphitheatrum in Provincia et Italia. Architecture and use Roman amphitheater of Augusta Raurica to Puteoli (with contributions by Ph. Rentzel, N. Frésard and ME Fox), Volume 43: Augst, Research in Augst pp. 569578.Google Scholar
Rentzel, P. 2011, Spuren der nutzung in Mithraeum von Biesheim. Mikromorphologische Untersuchungen (Traces of the use of the Biesheim Mithraeum. Micromorphological investigations) in Chapter 13: Le Mithraeum, une Fouille ancienne Revisiteé, Fortuné, C (ed), pp. 250257, 294.Google Scholar
Rentzel, P., and Narten, G.-B., 2000, Zur Entstehung von Gehniveaus in sandig-lehmigen Ablagerungen – Experimente und archäologische Befunde (Activity surfaces in sandy-loamy deposits – experiments and archaeological examples), Jahresbericht 1999: Basel, Archäologische Bodenforschung des Kantons Basel-Stadt, pp. 107127.Google Scholar
Reynolds, P., 1979, Iron Age Farm. The Butser Experiment, London, British Museum Publications Ltd., p. 112.Google Scholar
Reynolds, P. 1981, Deadstock and livestock, in Mercer, R., ed., Farming Practice in British Prehistory: Edinburgh, Edinburgh University Press, pp. 97122.Google Scholar
Reynolds, P. 1987, Ancient Farming, Aylesbury, Shire Publications Ltd. Archaeology.Google Scholar
Reynolds, P. 1995, The life and death of a post-hole, in Shepherd, E., ed., Interpreting Stratigraphy 5: Bawdeswell, pp. 2125.Google Scholar
Reynolds, P., and Shaw, C., 2000, Butser Ancient Farm. The Open Air Laboratory for Archaeology, Waterlooville, Butser Ancient Farm.Google Scholar
Righi, D., Van Ranst, E., De Coninck, F., and Guillet, B., 1982, Microprobe study of a Placohumod in the Antwerp Campine (North Belgium): Pedologie, v. 32, no. 2, pp. 117134.Google Scholar
Roberts, M. B., and Parfitt, S. A., 1999, Boxgrove. A Middle Pleistocene hominid site at Eartham Quarry, Boxgrove, West Sussex, London, English Heritage.Google Scholar
Roberts, M. B., and Pope, M. I., in press/2017, The Boxgrove Wider Area Project: Mapping the early Middle Pleistocene Deposits of the Slindon Formation Across the Coastal Plain of West Sussex and eastern Hampshire., Brighton: Spoilheap Press.Google Scholar
Roberts, M. B., Pope, M. I., and Parfitt, S. A., Forthcoming, Boxgrove: An early Middle Pleistocene hominid site at the Q1/B waterhole, Boxgrove, West Sussex. UK. Excavations and research, 1993–2007. Brighton: Spoilheap Press.Google Scholar
Roberts, M. B., Stringer, C. B., and Parfitt, S. A., 1994, A hominid tibia from Middle Pleistocene sediments at Boxgrove, UK: Nature, v. 369, pp. 311313.Google Scholar
Robinson, D., and Rasmussen, P., 1989, Botanical investigations at the Neolithic lake village at Weier, N. E. Switzerland: leaf hay and cereals as animal fodder, in Milles, A., Williams, D., and Gardner, N., eds., The Beginnings of Agriculture, International Series 496: Oxford, British Archaeological Reports, pp. 137148.Google Scholar
Robinson, M., 1992, Environment, archaeology and alluvium on the river gravels of the South Midlands floodplains, Alluvial Archaeology in Britain: Oxford, Oxbow, pp. 197208.Google Scholar
Rødsrud, C. L., 2007, Graver og Bosetningsspor på Bjørnstad (Lokalitet 44), in Bårdseth, G. A., ed., Hus, gard og graver langs E6 i Sarpsborg kommune. E6-prosjeket Østfold, Varia, Band 2: Oslo, Kulturhistorisk Museum Fornminneseksjonen, pp. 91181.Google Scholar
Rødsrud, C. L., 2012, Markeds-/Produksjonplass og Gravfelt; Gokstad Nedre, 48/2, 46/4, Samdefjord, Vestfold: Cultural History Museum, University of Oslo.Google Scholar
Romans, J. C. C., and Robertson, L., 1974, Some aspects of the genesis of alpine and upland soils in the British Isles, in Rutherford, G. K., ed., Soil Microscopy: Kingston, Ontario, The Limestone Press, pp. 498510.Google Scholar
Romans, J. C. C., and Robertson, L. 1975a, Soils and Archaeology in Scotland, in Evans, J. G., Limbrey, S., and Cleere, H., eds., The Effect of Man on the Landscape: the Highland Zone, Research Report No. 11: York, The Council for British Archaeology, pp. 3739.Google Scholar
Romans, J. C. C., and Robertson, L. 1975b, Some genetic characteristics of the freely drained soils of the Ettrick Association in east Scotland: Geoderma, v. 14, pp. 297317.Google Scholar
Romans, J. C. C., and Robertson, L. 1983a, An account of the soils at North Mains, in ‘Sites of the third millenium bc to the first millenium ad at North mains, Strathallan, Perthshire’, by Barclay, G. J.: Proceedings of the Society of Antiquities Scotland, v. 113, pp. 260269.Google Scholar
Romans, J. C. C., and Robertson, L. 1983b, The environment of north Britain: soils, in Chapman, J. C., and Mytum, H. C., eds., Settlement in north Britain 1000 BC to AD 1000: Oxford, British archaeological reports, British Series 118, pp. 5580.Google Scholar
Romans, J. C. C., and Robertson, L. 1983c, The general effects of early agriculture on soil, in Maxwell, G. S., ed., The Impact of Aerial Reconnaissance on Archaeology, Research Report No. 49: London, Council for British Archaeology, pp. 136141.Google Scholar
Rose, E. P. F., 2015, The Battle of Waterloo, 18 June 1815: some geological reflections to mark the bicentenary: Geology Today, v. 31, no. 3, pp. 103108.Google Scholar
Rosen, A., 1999, Phytoliths as indicators of prehistoric irrigation farming., in Anderson, P. C., ed., Prehistory of Agriculture: New Experimental and Ethnographic Approaches: Los Angeles, UCLA Institute of Archaeology, pp. 193198.Google Scholar
Rosen, A. 2007, The role of environmental change in the development of complex societies in China: a study from the Huizui site: Bulletin of the Indo-Pacific Prehistory Association, v. 27, pp. 3948.Google Scholar
Rosen, A. M., 1986, Cities of clay. The geoarchaeology of tells, Chicago, University Press of Chicago.Google Scholar
Rosen, A. M., 2008, The impact of environmental change and human land use on the alluvial valleys in the Loess Plateau of China during the Middle Holocene: Geomorphology, v. 101, pp. 298307.Google Scholar
Rosen, A. M., Macphail, R., Liu, L., Chen, X., and Weisskopf, A., 2017, Rising Social Complexity, Agricultural Intensification, and the Earliest Rice Paddies on the Loess Plateau of Northern China, Quaternary International, v. 37, pp. 50–59.Google Scholar
Rosen, S. A., 1988, Notes on the origins of pastoral nomadism: a case study from the Negev and Sinai: Current Anthropology, v. 29, no. 3, pp. 498506.Google Scholar
Rosen, S. A., Savinetsky, A. B., Plakht, J., Kisseleva, N. K., Khassanov, B. F., Pereladov, A. M., and Haiman, M., 2005, Dung in the desert: preliminary results from the Negev Ecology Project: Current Anthropology, v. 46, pp. 317327.Google Scholar
Rowley Conwy, P., 1981, Slash and burn in the Temperate neolithic, in Mercer, R., ed., Farming Practice in Britsh Prehistory: Edinburgh, Edinburgh University Press, pp. 8596.Google Scholar
Rowsome, P., 2000, Heart of the City. Roman, Medieval and Modern London Revealed by Archaeology at 1 Poultry, London, Museum of London Archaeology Service.Google Scholar
Rudeforth, C. C., Hartnup, R., Lea, J. W., Thompson, T. R. E., and Wright, P. S., 1984, Soils and their Use in Wales, Harpenden, Soil Survey of England and Wales.Google Scholar
Runia, L. T., 1988, So-called secondary podzolisation in barrows, in Groenman-van Waateringe, W., and Robinson, M., eds., Man-made Soils, International Series 410: Oxford, British Archaeological Reports, pp. 129142.Google Scholar
Sageidet, B. M., 2005, Sub-local differences in late Holocene land use at Orstad, Jæren in SW Norway, revealed by soil pollen stratigraphy: Environmental Archaeology, v. 10, pp. 5171.Google Scholar
Salquea, M., Radi, G., Tagliacozzo, A., Pino Uriac, B., Wolfram, S., Hohled, I., Stäuble, H., Hofmann, D., Whittleg, A., Pechtl, J., Schade-Lindigi, S., Eisenhauer, U., and Evershedk, R. P., 2012, New insights into the Early Neolithic economy and management of animals in Southern and Central Europe revealed using lipid residue analyses of pottery vessels: Anthropozoologica v. 47, no. 2, pp. 4562.Google Scholar
Šamonil, P., Krála, K., and Horta, L., 2010, The role of tree uprooting in soil formation: A critical literature review, Geoderma, v. 157, no. 3–4, pp. 65252.Google Scholar
Sankey, D., and McKenzie, M., 1998, 7–11 Bishopsgate: a hole in the heart of London’s business district: The London Archaeologist, v. 16, pp. 171179.Google Scholar
Sara, T. R., Macphail, R., Goldberg, P., and Larson, B., 2006, Archaeological Context and Geoarchaeological Study, Camp Lemonier, Djibouti, Africa: Geo-Marine Inc, Miscellaneous Reports of Investigation Number 349.Google Scholar
Saville, A., 1990, Hazleton North. The Excavation of a Neolithic Long Cairn of the Cotswold-Severn Group, London, English Heritage, p. 281.Google Scholar
Scaife, R. G., 1980, Pollen analysis of some dark earth samples: Ancient Monuments Laboratory, Ancient Monuments Laboratory Report 3001.Google Scholar
Scaife, R. G. 1992, The vegetation history, in Barton, R. N. E., ed., Hengistbury Head, Dorset. Volume 2: The Late Upper Palaeolithic & Early Mesolithic sites, Monograph No. 34: Oxford, Oxford University Committee for Archaeology, pp. 5159.Google Scholar
Scaife, R. G. 1995, Pollen Analysis: Intertidal sites (Blackwater Sites, 3, 18 and 28; Crouch Sites 9 and 8), in Wilkinson, T. J., and Murphy, P. L., eds., The Archaeology of the Essex Coast, Volume I: The Hullbridge Survey, Report No. 71: Chelmsford, East Anglian Archaeology, pp. 4351.Google Scholar
Scaife, R. G., and Macphail, R. I., 1983, The post-Devensian development of heathland soils and vegetation, in Burnham, P., ed., Soils of the Heathlands and Chalklands, Volume 1: Wye, South-East Soils Discussion Group, pp. 7099.Google Scholar
Schaetzl, R. J., 1990, Effects of treethrow microtopography on the characteristics and genesis of spodosols: Catena, v. 17, pp. 111126.Google Scholar
Schiegl, S. G., Goldberg, P., Bar-Yosef, O., and Weiner, S., 1996, Ash deposits in Hayonim and Kebara Caves, Israel: macroscopic, microscopic and mineralogical observations, and their archaeological implications: Journal of Archaeological Science, v. 23, pp. 763781.Google Scholar
Schiffer, M. B., 1987, Formation Processes in the Archaeological Record, Alburquerque, University of New Mexico Press.Google Scholar
Schofield, D., and Hall, D. M., 1985, A method to measure the susceptibility of pasture soils to poaching by cattle: Soil Use and Management, v. 1, pp. 134138.Google Scholar
Schoute, J. F. T., 1984, Vegetation Horizons and related Phenomena. A palaeocological-micromorphological study in the younger coastal Holocene of the northern Netherlands (Schildmeer Area), Hirschenberg, Strauss & Cramer, p. 270.Google Scholar
Schoute, J. F. T. 1987, Micromorphology of soil horizons developed during semi-terrestrial phases in transgressive and regressive sedimentary sequences in the Northern Netherlands, in Fedoroff, N., Bresson, L. M., and Courty, M. A., eds., Soil Micromorphology: Plaisir, Association Française pour l’Étude du Sol, pp. 661667.Google Scholar
Schuldenrein, J., 1995, Geochemistry, phosphate fractionation, and the detection of activity areas at prehistoric North American sites, in Collins, M. E., Carter, B. J., Gladfelter, B. G., and Southard, R. J., eds., Pedological Perspectives in Archaeological Research, SSSA Special Publication Number 44: Madison, Soil Science Society of America, pp. 107132.Google Scholar
Schulz, E., Biester, H., Bogenrieder, A.,Eckmeier, E., Ehrmann, O., Gerlach, R., Hall, M., Hartktopf-Fröder, C., Herrmann, L., Kury, B., Rösch, M., and Schier, W., 2011, How long will it take? Regeneration of vegetation and soil after clearing, burning and cultivation. The Forchtenberg-Experiment., in Sauer, D., Jahn, R., and Stahr, K., eds., Landscape & Soils through time: Hohenheim, IUSS, pp. 112113.Google Scholar
Segerström, U., 1991, Soil pollen analysis – an application for tracing ancient arable patches: Journal of Archaeological Science, v. 18, pp. 165175.Google Scholar
Shahack-Gross, R., 2010, Herbivorous livestock dung: formation, taphonomy, methods for identification, and archaeological significance: Journal of Archaeological Science, v. 38, pp. 205218.Google Scholar
Shahack-Gross, R. 2011, Herbivorous Livestock Dung: Formation, Taphonomy, Methods For Identification, and Archaeological Significance: Journal of Archaeological Science, v. 38, pp. 205218.Google Scholar
Shahack-Gross, R. 2015, Archaeological Micromorphology Self-Evaluation Exercise: Geoarchaeology, v. 31, pp. 4957.Google Scholar
Shahack-Gross, R., Albert, R.-M., Gilboa, A., Nagar-Hillman, O., Sharon, I., and Weiner, S., 2005, Geoarchaeology in an urban context: the uses of space in a Phoenician monumental building at Tel Dor (Israel): Journal of Archaeological Science, v. 32, pp. 14171431.Google Scholar
Shahack-Gross, R., Ayalon, A., Goldberg, P., Goren, Y., Boaz, O., Rabinovich, R., and Hovers, E., 2008a, Formation processes of cemented features in karstic cave sites revealed using stable Oxygen and Carbon Isotopic analyses: a case study at Middle Palaeolithic Amud Cave, Israel: Geoarchaeology, v. 23, pp. 4362.Google Scholar
Shahack-Gross, R., Berna, F., Karkanas, P., and Weiner, S., 2004a, Bat guano and preservation of archaeological remains in cave sites: Journal of Archaeological Science, v. 31, pp. 12591272.Google Scholar
Shahack-Gross, R., and Finkelstein, I., 2008, Subsistence practices in an arid environment: a geoarchaeological investigation in an Iron Age site, the Negev Highlands, Israel: Journal of Archaeological Science, v. 35, pp. 965982.Google Scholar
Shahack-Gross, R., Gafri, M., and Finkelstein, I., 2009, Identifying Threshing Floors in the Archaeological Record: A Test Case at Iron Age Tel Megiddo, Israel: Journal of Field Archaeology, v. 34(2), pp. 171184.Google Scholar
Shahack-Gross, R., Marshall, F., Ryan, K., and Weiner, S., 2004, Reconstruction of spatial organisation in abandoned Maasai settlements: implications for site structure in Pastoral Neolithic of East Africa: Journal of Archaeological Science, v. 31, pp. 13951411.Google Scholar
Shahack-Gross, R., Marshall, F., and Weiner, S., 2003, Geo-ethnoarchaeology of pastoral sites: the identification of livestock enclosures in abandoned Maasai settlements: Journal of Archaeological Science, v. 30, pp. 439459.Google Scholar
Shahack-Gross, R., Simons, A., and Ambrose, S. H., 2008, Identification of pastoral sites using stable nitrogen and carbon isotopes from bulk sediment samples: a case study in modern and archaeological pastoral settlements in Kenya: Journal of Archaeological Science, v. 35, pp. 983990.Google Scholar
Shahack-Gross, R., Berna, F., Karkanas, PP., Lemorini, C., Gopher, A., and Barkai, R., 2014, Evidence for the repeated use of a central hearth at Middle Pleistocene (300 ky ago) Qesem Cave, Israel: Journal of Archaeological Science, v. 44, pp. 1221.Google Scholar
Sharples, N. M., 1991, Maiden Castle. Excavations and field survey 1985–6, Archaeological Report no 19: London, English Heritage, p. 284.Google Scholar
Shelley, A., 2005, Dragon Hall, King Street, Norwich: Excavation and Survey of a Late Medieval Merchant’s Trading Complex, Report No. 112: Norwich, East Anglian Archaeology, p. 206.Google Scholar
Sherwood, S., and Kidder, T. R., 2010, The DaVincis of dirt: Geoarchaeological perspectives on Native American mound building in the Mississippi River basin: Journal of Antrhopological Archaeology, v. 30, pp. 6987.Google Scholar
Shillito, L. M., and Ryan, P., 2013, Surfaces and streets: phytoliths, micromorphology and changing use of space at Neolithic Çatalhöyük (Turkey): Antiquity, v. 87, pp. 684700.Google Scholar
Sidell, E. J., 2000, Dark earth and obscured stratigraphy, in Huntley, J. P., and Stallibrass, S., eds., Taphonomy and Interpretation, Symposia of the Association for Environmental Archaeology No. 14: Oxford, Oxbow, pp. 3542.Google Scholar
Sidell, E. J. 2003, The London Thames: a decade of research into the river and its floodplain, in Howard, A. J., Macklin, M. G., and Passmore, D. G., eds., Alluvial Archaeology in Europe: Lisse, A. A. Balkema Publishers, pp. 133143.Google Scholar
Sidell, E. J., Wilkinson, K. N., Scaife, R. G., and Cameron, N., 2000, The Holocene Evolution of the London Thames. Archaeological Investigations (1991–1998) in advance of the London Underground Limited Jubilee Line Extension, London, Museum of London Archaeology Service.Google Scholar
Simmons, I. G., 1975, The ecological setting of mesolithic man in the highland zone, in Evans, J. G., Limbrey, S., and Cleere, H., eds., The Effect of Man on the Landscape: the Highland Zone, Research Report 11: York, CBA, pp. 5763.Google Scholar
Simpson, I. A., 1998, Early land management at Tofts Ness, Sanday, Orkney: the evidence of thin section micromorphology, in Mills, C. M., and Coles, G., eds., Life on the Edge. Human Settlement and Marginality, Monograph 100: Oxford, Oxbow, pp. 9198.Google Scholar
Simpson, I. A., Milek, K. B., and Guðmundsson, G., 1999, A reinterpretation of the great pit at Hofstaðir, Iceland using sediment thin-section micromorphology: Geoarchaeology, v. 14, pp. 511530.Google Scholar
Skaarup, J., and Grøn, O., 2004, Møllegabet II. A submerged Mesolithic settlement in southern Denmark., Oxford, Archaeopress.Google Scholar
Slager, S., and Van der Wetering, H. T. J., 1977, Soil formation in archaeological pits and adjacent loess soils in Southern Germany: Journal of Archaeological Science, v. 4, pp. 259267.Google Scholar
Smith, G., Macphail, R. I., Mays, S. A., Nowakowski, J., Rose, P., Scaife, R. G., Sharpe, A., Tomalin, D. J., and Williams, D. F., 1996, Archaeology and environment of a Bronze Age Cairn and Romano-British field system at Chysauster, Gulval, near Penzance, Cornwall: Proceedings of the Prehistoric Society, v. 62, pp. 167219.Google Scholar
Soil Survey Staff, 1999, Soil Taxonomy, Washington D. C., U. S. Department of Agriculture, U. S. Government Printing Office.Google Scholar
Sordoillet, D., 1997, Formation des dépôts archéologiques en grotte: la Grotte du Gardon (Ain) durant le Néolithique, in Bravard, J.-F., and Prestreau, M., eds., Dynamique du Paysage. Entretiens de Géoarchéologie, Documents d’Archéologiques en Rhône-Alpes No. 15: Lyon, Ministère de la Culture, pp. 3957.Google Scholar
Sordoillet, D. 2009, Géoarchéologie de sites prèhistoriques; Le Gardon (Ain), Montou (Pyrénées-Orientales) et Saint-Alban (Isère), Paris, Éditions de la Maison des Sciences de l’Homme, 188 pp.Google Scholar
Sørensen, R., Henningsmoen, K. E., Høeg, H., Stabell, B., and Bukholm, K. M., 2007a, Geology, soils, vegetation and sea levels in the Kaupang area., in Skre, D., ed., Kaupang in Skiringssal, Aarhus, Denmark, Aarhus University Press.Google Scholar
Sørensen, R., Wu, Q., Bukholm, K. M., and Pilø, L., 2007b, Analysis of carbon fractions in anthropomorphic soil materials, from a Viking Age town to recent municipal waste deposits: Atti Soc. tosc. Sci. nat. mem. Serie A, v. 122, no. Boschian, G. (ed), pp. 5560.Google Scholar
Spataro, M., 2002, The First Farming Communities of the Adriatic: Pottery Production and Circulation in the Early and Middle Neolithic, Trieste, Società per la Preistoria e Protostoria della Regione Friuli-Venezia Giulla.Google Scholar
Speed, G., 2010, The excavation of an enclosed Iron Age Settlement at Hallam Fields, Birstall, Leicestershire: Transactions of the Leicestershire Archaeological and Historical Society, v. 84, pp. 2775.Google Scholar
Stahischmidt, M. C., Miller, C. E., Ligouis, B., Hambach, U., Goldberg, P., and Berna, F., 2015, On the evidence for human use and control of fire at Schöningen: Journal of Human Evolution, v. 89, 181–201.Google Scholar
Stein, J. K., 1992, Deciphering a Shell Midden, San Diego, Academic Press.Google Scholar
Stephens, E. P., 1956, The uprooting of trees: a forest process: Soil Science Society of America Proceedings, v. 20, pp. 113116.Google Scholar
Stewart, H., 1995, Cedar: Tree of life to the northwest coast Indians, Seattle, University of Washington Press.Google Scholar
Stolt, M. H., and Lindbo, D. L., 2010, Soil organic matter, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths, Amsterdam, Elsevier, pp. 369396.Google Scholar
Stoltman, J. B., 2001, The role of petrology in the study of archaeological ceramics, in Goldberg, P., Holliday, V. T., and Ferring, C. R., eds., Earth Sciences and Archaeology, New York, Kluwer Academic/Plenum Publishers, pp. 297326.Google Scholar
Stoops, G., 1984a, The environmental physiography of Pessinus in function of the study of the archaeological stratigraphy and natural building materials, in ‘Les Fouilles de la Rijksuniversiteit te Gent a Pessiononte 1967–73 (Devreker, J. and Waelkens, M.) eds.: Dissertationes Archaeologicae Gardenses, v. XXII, pp. 3850.Google Scholar
Stoops, G., 1984b, Petrographic study of mortar and plaster samples, in ‘Les Fouilles de la Rijksuniversiteit te Gent a Pessiononte 1967–73 (J. Devreker and M. Waelkens) eds.: Dissertationes Archaeologicae Gardenses, v. XXII, pp. 164170.Google Scholar
Stoops, G. 1996, Complementary techniques for the study of thin sections of archaeological materials, in Castelletti, L., and Cremaschi, M., eds., XIII International Congress of Prehistoric and Protohistoric Sciences Forlì-Italia-8/14 September 1996, Volume 3 Paleoecology: Forlì, A.B.A.C.O., pp. 175182.Google Scholar
Stoops, G., 2003, Guidelines for Analysis and Description of Soil and Regolith Thin Sections, Madison, WI, Soil Science Society of America, p. 184.Google Scholar
Stoops, G., 2014, The “Fabric” of soil micromorphological research in the 20th century – A bibliometric analysis: Geoderma, v. 213, pp. 193202.Google Scholar
Stoops, G., and Jongerius, A., 1975, Proposal for a micromorphological classification in soil materials. I. A classification of the related distributions of coarse and fine particles: Geoderma, v. 13, pp. 189200.Google Scholar
Stoops, G., Marcelino, V., and Mees, F., 2010, Interpretation of Micromorphological Features of Soils and Regoliths: Amsterdam, Elsevier, p. 720.Google Scholar
Stoops, G., and Nijs, R., 1986, Micromorphological characteristics of some tell materials from Mesopotamia: Pédologie, v. XXXVI-3, pp. 329336.Google Scholar
Straulino, L., Sedov, S., Michelet, D., and Balanzario, S., 2013, Weathering of carbonated materials in ancient Maya constructions (Río Bec and Dzibanché): limestone and stucco deterioration patterns: Quaternary International, v. Site formation processes in archaeology, no. 315, pp. 87100.Google Scholar
Stringer, C., 2000, The Gough’s Cave human fossils: an introduction: Bulletin of the British Museum (Natural History): Geology, v. 56, pp. 135139.Google Scholar
Stringer, C. B., Barton, R. N. E., Currant, J. C., Finlayson, J. C., Goldberg, P., Macphail, R., and Pettit, P. B., 1999, Gibraltar Palaeolithic revisited: new excavations at Gorham’s and Vanguard Caves, in Davies, W., and Charles, R., eds., Dorothy Garrod and the Progress of the Palaeolithic (Studies in the Prehistoric Archaeology of the Near East and Europe), Oxford, Oxbow Books, pp. 8496.Google Scholar
Stringer, C. B., Trinkhaus, E., Roberts, M. B., Macphail, R. I., and Parfitt, S. A., 1998, The Middle Pleistocene human tibia from Boxgrove: Journal of Human Evolution, v. 34, pp. 509547.Google Scholar
Stromberg, C. A. M., 2009, Methodological concerns for analysis pf phytolith assemblages: Does count size matter?: Quaternary International, v. 193, pp. 124140.Google Scholar
Styles, C. A., Foss, J. E., and Lewis, R. J., 1995, Lead fractions in soils from Hadrian’s Villa, Italy, in Collins, M. E., Carter, B. J., Gladfelter, B. G., and Southard, R. J., eds., Pedological Perspectives in Archaeological Research, SSSA Special Publication Number 4: Madison, Soil Science Society of America, Inc, pp. 151157.Google Scholar
Sukopp, H., Blume, H.-P., and Kunick, W., 1979, The soil, flora and vegetation of Berlin’s waste lands,’ in Laurie, I. C., ed., Nature in Cities, Chichester, John Wiley, pp. 115132.Google Scholar
Sveinbjarnardóttir, G., Erlendsson, E., Vickers, K., McGovern, T. H., Milek, K. B., K.J., E., Simpson, D. D. A., and Cook, G., 2007, The palaeoecology of a high status Icelandic farm: Environmental Archaeology, v. 12, pp. 187206.Google Scholar
Terrible, F., Wright, R., and FitzPatrick, E. A., 1997, Image analysis in soil micromorphology: from univariate approach to multivariate solution, in Shoba, S., Gerasimova, M., and Miedema, R., eds., Soil Micromorphology: Studies on Soil Diversity, Diagnostics, Dynamics, Moscow – Wageningen, International Society of Soil Science, pp. 397417.Google Scholar
Tesch, S., 1992, House, farm and village in the Kopinge area from the early Neolithic to the early middle ages., in Larsson, L., Callmer, J., and Stjarnquist, B., eds., The Archaeology of the Cultural Landscape. Fieldwork and research in a south Swedish rural region, Lund, Acta Archaeologica Ludensia. Series 4, No. 19.Google Scholar
Thirly, M., Galbois, J., and Schmitt, J.-M., 2006, Unusual phosphate concretions related to groundwater flow in a continental environment: Journal of Sedimentary Research, v. 76, pp. 866877.Google Scholar
Thomas, G., 2010, “Life before the Anglo-Saxon minster”: Interim Report on University of Reading Excavations at Lyminge, 2010, www.reading.ac.uk/archaeology/research/Lyminge/arch-lyminge2010.aspx: Reading, University of Reading.Google Scholar
Thurley, S., 1995, The King’s Privy Garden at Hampton Court Palace 1689–1995, London, Apollo, p. 118.Google Scholar
Tipper, J., 2004, The Grubenhauss in Anglo-Saxon England: an analysis and interpretation of the evidence from a most distinctive building type., Yeddingham, North Yorkshire, Lanscape Research Centre, p. 208.Google Scholar
Tipper, J. Forthcoming, West Heslerton Anglo-Saxon Settlement: Yeddingham, Landscape Research Centre.Google Scholar
Tite, M. S., 1972, The influence of geology on the magnetic susceptibility of soils on archaeological sites: Archaeometry, v. 14, pp. 229236.Google Scholar
Tite, M. S., and Mullins, C. E., 1971, Enhancement of magnetic susceptibility of soils on archaeological sites: Archaeometry, v. 13, pp. 209219.Google Scholar
Tovey, N. K., and Sokolov, V. N., 1997, Image analysis applications in soil micromorphology, in Shoba, S., Gerasimova, M., and Miedema, R., eds., Soil Micromorphology: Studies on Soil Diversity, Diagnostics, Dynamics, Moscow – Wageningen, International Society of Soil Science, pp. 345356.Google Scholar
Troels-Smith, J., 1984, Stall feeding and field manuring in Switzerland about 6000 years ago: Tools and Tillage, v. 5, no. 1, pp. 1325.Google Scholar
Trow-Smith, R., 1957, A History of British Livestock Farming to 1700, London, Routledge.Google Scholar
Turner, B. D., 1987, Forensic entomology: insects against crime. Science Progress: Science Progress, v. 71, pp. 169180.Google Scholar
Turner, B. D., and Wiltshire, P., 1999, Experimental validation of forensic evidence: a study of the decomposition of buried pigs in a heavy clay soil: Forensic Science International, v. 101, pp. 113112.Google Scholar
Turner, K., 2012, 19. Plant macrofossils, in Biddulph, E., Foreman, S., Stafford, E., Stansbie, D., and Nicholson, R., eds., London Gateway. Iron Age and Roman salt making in the Thames Estuary; Excavations at Stanford Wharf Nature Reserve, Essex (library.thehumanjourney.net/909), Oxford Archaeology Monograph No. 18: Oxford, Oxford Archaeology.Google Scholar
Tylecote, R. F., 1986, The Early History of Metallurgy in Europe, London, Longman.Google Scholar
Usai, M. R., 2001, Textural pedofeatures and pre-Hadrian’s wall ploughed palaeosols at Stanwix, Carlisle, Cumbira, UK: Journal of Archaeological Science, v. 28, pp. 541553.Google Scholar
Valentin, C., 1983, Effects of grazing and trampling around recently drilled water holes on the soil deterioration in the Sahelian zone: Soil Erosion and Conservation (Soil Conservation Society of America), no. 51–65.Google Scholar
Valentin, C. 1991, Surface crusting in two alluvial soils of northern Niger: Geoderma, v. 48, pp. 201222.Google Scholar
van de Westeringh, W., 1988, Man-made soils in the Netherlands. especially in sandy areas (“Plaggen Soils”), in Groenman-van Waateringe, W., and Robinson, M., eds., Man-made Soils, International Series 410: Oxford, British Archaeological Reports, pp. 519.Google Scholar
Van Vliet-Lanoë, B., 1998, Frost and soils: implications for paleosols, paleoclimates and stratigraphy: Catena, v. 34, no. 1–2, pp. 157183.Google Scholar
Van Vliet-Lanoë, B., 2010, Frost action, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths, Amsterdam, Elsevier, pp. 81108.Google Scholar
Van Vliet-Lanöe, B., 1985, Frost effects in soils, in Boardman, J., ed., Soils and Quaternary Landscape Evolution: Chichester, John Wiley & Sons, pp. 117158.Google Scholar
Van Vliet-Lanoë, B., Helluin, M., Pellerin, J., and Valadas, B., 1992, Soil erosion in Western Europe: from the last interglacial to the present, in Bell, M., and Boardman, J., eds., Past and Present Soil Erosion, Monograph 22: Oxford, Oxbow, pp. 101114.Google Scholar
Van Vliet, B., 1982, Structures and microstructures associées à la formation de glace de ségrégation: leurs conséquences., Proceedings of the Fourth Canadian Permafrost Conference, Ottawa, pp. 116122.Google Scholar
Veneman, P. I. M., Jacke, P. V., and Bodine, S. M., 1984, Soil formation as affected by pit and mound relief in Massachsetts, USA: Geoderma, v. 33, pp. 8999.Google Scholar
Viklund, K., 1998, Cereals, Weeds and Crop Processing in Iron Age Sweden. Methodological and interpretive aspects of archaeobotanical evidence, Umeå, Archaeology and Environment 14, p. 192.Google Scholar
Viklund, K., Engelmark, R., and Linderholm, J., 1998, Fåhus från bronsålder till idag, Skrifter om skogs- och lantbrukshistoria 12: Lund, Nordiska Museet.Google Scholar
Viklund, K., Linderholm, J., and Macphail, R. I.,2013, Integrated Palaeoenvironmental Study: Micro- and Macrofossil Analysis and Geoarchaeology (soil chemistry, magnetic susceptibility and micromorphology), in Gerpe, L.-E., ed., E18-prosjektet Gulli-Langåker. Oppsummering og arkeometriske analyser, Bind 3: Bergen, Fagbokforlaget, pp. 2583.Google Scholar
Villagran, X. S., Giannini, P. C. F., and DeBlasis, P., 2009, Archaeofacies analysis: using depositional attributes to identify anthropogenic processes of deposition in a monumental shell mound of Santa Catarina State, southern Brazil: Geoarchaeology, v. 24, no. 3, pp. 311335.Google Scholar
Villagran, X. S., Schaefer, C. E. G. R., and Ligouis, B., 2013, Living in the cold: Geoarchaeology of sealing sites from Byers Peninsula (Livingston Island, Antarctica): Quaternary International, v. 315 – Site formation processes in archaeology, pp. 184199.Google Scholar
Villaseñor, I., and Graham, E., 2010, The use of volcanic materials for the manufacture of pozzolanic plasters in the Maya lowlands: a preliminary report: Journal of Archaeological Science, v. 37, pp. 13391347.Google Scholar
Vissac, C., 2002, Les Terres Raportées dans les Jardins du XVIe au XIXe Siècles Caracterisations de l’Impact Anthropique a Differentes Echèlles d’Organisation du Sol: Angers, p. 237.Google Scholar
Voroney, R. P., Van Veen, J. A., and Paul, E. A., 1981, Organic C dynamics in grassland soils 2. Model validation and simulation of long term effects of cultivation and rainfall erosion: Canadian Journal of Soil Science, v. 61, pp. 211224.Google Scholar
Wagner, D. P., and McAvoy, J. M., 2004, Pedoarchaeology of Cactus Hill, a sandy palaeoindian site in southeastern Virginia: Geoarchaeology, v. 19, pp. 297322.Google Scholar
Wainwright, G. J., and Davies, S. M., 1995, Balksbury Camp, Hampshire: Excavations 1973 and 1981, London, English Heritage, Archaeological report 4.Google Scholar
Wattez, J., 1992, Dynamique de formation des structures de combustion de la fin du Paléolithique au Néolithique Moyen. Approche méthodologiques et implications culturelles. Thiverval-Grignon, Institut National Agronomique.Google Scholar
Wattez, J., Courty, M. A., and Macphail, R. I., 1990, Burnt organo-mineral deposits related to animal and human activities in prehistoric caves, in Douglas, L. A., ed., Soil Micromorphology: a Basic and Applied Science, Developments in Soil Science 19: Amsterdam, Elsevier, pp. 431439.Google Scholar
Wegener, O., 2009, Soil micromorphological investigations on trampling floors in pit houses (Grubenhäuser) of the deserted medieval town Marsleben (Saxony-Anhalt), in Thiemeyer, H., ed., Archaeological Soil Micromorphology – Contributions to the Archaeological Soil Micromorphology Working Group Meeting 3rd to 5th April 2008, Volume D30: Frankfurt A M, Frankfurter Geowiss. Arb., pp. 133141.Google Scholar
Weiner, S., 2010, Microarchaeology. Beyond the Visible Archaeological Record, Cambridge, Cambridge University Press, p. 396.Google Scholar
Weiner, S., Schiegl, S. G., Goldberg, P., and Bar-Yosef, O., 1995, Mineral assemblages in Kebara and Hayonim Caves, Israel: excavation strategies, bone preservation, and wood ash remnants: Israel Journal of Chemistry, v. 35, pp. 143154.Google Scholar
Weiner, S., Xu, Q., Goldberg, P., Liu, J., and Bar-Yosef, O., 1998, Evidence for the use of fire at Zhoukoudian, China: Science, v. 281, pp. 251253.Google Scholar
Wenban-Smith, F. F., Allen, P., Bates, M. R., Parfitt, S. A., Preece, R. C., Stewart, J. R., Turner, C., and Whittaker, J. E., 2006, The Clactonian elephant butchery site at Southfleet Road, Ebbsfleet, UK: Journal of Quaternary Science, v. 21, no. 5, pp. 471483.Google Scholar
Wenbann-Smith, F. F., 2013, The Ebbsfleet Elephant. Excavations at Southfleet Road, Swnscombe in advance of High Speed 1, 2003–4, Oxford Archaeology Monograph No. 20: Oxford, Oxford Archaeology, p. 562.Google Scholar
West, S. E., 1985, West Stow: the Anglo-Saxon village: East Anglian Archaeology, v. 24, no. 1 and 2.Google Scholar
White, R. E., 1987, Introduction to Principles and Practice of Soil Science, Oxford, Blackwell.Google Scholar
Whittle, A., 2007, The Early Neolithic on the Great Hungarian Plain: investigations of the Körös culture site of Ecsegfalva 23, Co. Békés: Budapest, Institute of Archaeology, p. 809.Google Scholar
Whittle, A., Pollard, J., and Grigson, c., 1999, The harmony of symbols: the Windmill Hill causewayed enclosure, Wiltshire, Oxford, Oxbow.Google Scholar
Whittle, A., Rouse, A. J., and Evans, J. G., 1993, A Neolithic downland monument in its environment: excavations at the Easton Down Long Barrow, Bishops Canning, North Wiltshire: Proceedings of the Prehistoric Society, v. 59, pp. 197239.Google Scholar
Wilkinson, T. J., 1990, Soil development and early land use in the Jazira region, Upper Mesopotamia, in Thomas, K., ed., Soils and Early Agriculture, World Archaeology 22, No. 1: London, Routledge, pp. 87103.Google Scholar
Wilkinson, T. J. 1997, Holocene environments of the High Plateau, Yemen. Recent geoarchaeological investigations: Geoarchaeology, v. 12, no. 8, pp. 833864.Google Scholar
Wilkinson, T. J., and Murphy, P. L., 1995, The Archaeology of the Essex Coast, Volume I: The Hullbridge Survey, Chelmsford, Essex County Council, East Anglian Archaeology report No. 71.Google Scholar
Wilkinson, T. J., Murphy, P. L., Brown, N., and Heppell, E., 2012a, The Archaeology of the Essex Coast Volume II. Excavations at the Prehistoric Site of The Stumble., East Anglian Archaeology 142: Chelmsford, Essex County Council.Google Scholar
Wilkinson, T. J., Murphy, P. L., Brown, N., and Heppell, E. M., 2012b, The Archaeology of the Essex Coast, Volume II: Excavations at the Prehistoric Site of the Stumble, Chelmsford, Historic Environment Essex County Council, p. 160.Google Scholar
Williams, T., 2007, The city of Sultan Kala, Merv, Turkmenistan: communities, neighbourhoods and urban planning from the eighth to the thirteenth century., in Bennison, A. K., and Gascoigne, A., eds., Cities in the pre-modern Islamic world: the urban impact of religion, state and society, London, Routledge, pp. 42–62.Google Scholar
Wilson, M. A., and Righi, D., 2010, Spodic materials, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths, Amsterdam, Elsevier, pp. 251273.Google Scholar
Wiltshire, P., 1999, Palynological analysis of filling in the funerary shaft, in Niblett, R., ed., The Excavation of a Ceremonial Site at Folly Lane, Verulamium, Brtannia Monograph Series No. 14: London, Society for the Promotion of Roman Studies, pp. 346365.Google Scholar
Wiltshire, P. E. J., Edwards, K. J., and Bond, S., 1994, Microbially-derived metallic sulphide spherules, pollen, and the waterlogging of archaeological sites: Proceedings of the American Association of Sedimentary Palynologists, v. 29, pp. 207-221.Google Scholar
Woods, W. I., and McCann, J. M., 1999, The Anthropogenic origin and persistence of Amazonian Dark Earths: Conference of Latin Americanist Geographers, v. 25, pp. 714.Google Scholar
Wootten, G., 2012, Waterloo 1815: The Birth of Modern Europe, Oxford, Osprey Publishing.Google Scholar
Wright, V. P., 1986, Paleosols. Their recognition and Interpretation, Oxford, Blackwell Scientific Publications.Google Scholar
Yi, S., 1988, Quaternary geology and paleoecology of hominid occupation of Imjin Basin: The Korean Journal of Quaternary Research, v. 2, no. 1, pp. 2530.Google Scholar
Yi, S. 2005, New Data on the Formation of the Basalt Plain in the Imjin River Basin: Journal of the Korean Geomorphological Association, v. 12, no. 3, pp. 2138.Google Scholar
Yi, S. 2015, Hand Axes in the Imjin River Basin, Korea – Implications for Late Pleistocene Hominin Evolution in East Asia, in Kaifu, Y., Izuho, M., Goebel, T., Sato, H., and Ono, A., eds., Emergence and Diversity of Modern Human Behavior in Paleolithic Asia: College Station, Texas A&M University Press, pp. 259269.Google Scholar
Yoo, Y., 2007, Long-Term Changes in the Organization of Lithic Technology: A Case Study from the Imjin-Hantan River Area, Korea, McGill University, p. 369.Google Scholar
Yule, B., 1990, The “dark earth” and late Roman London: Antiquity, v. 64, pp. 620628.Google Scholar
Zauyah, S., Schaefer, C. E. G. R., and Simas, F. N. B., 2010, Saprolites, in Stoops, G., Marcelino, V., and Mees, F., eds., Interpretation of Micromorphological Features of Soils and Regoliths, Amsterdam, Elsevier, pp. 4968.Google Scholar
Zhuang, Y., Bao, W., and French, C., 2013, River floodplain aggradation history and culture activities: Geoarchaeological investigation at the Yuezhuang site, Lower Yellow River, China: Quaternary International, v. 315, pp. 101115.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.

  • References
  • Richard I. Macphail, University College London, Paul Goldberg, Boston University
  • Book: Applied Soils and Micromorphology in Archaeology
  • Online publication: 18 December 2017
  • Chapter DOI: https://doi.org/10.1017/9780511895562.016
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.

  • References
  • Richard I. Macphail, University College London, Paul Goldberg, Boston University
  • Book: Applied Soils and Micromorphology in Archaeology
  • Online publication: 18 December 2017
  • Chapter DOI: https://doi.org/10.1017/9780511895562.016
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.

  • References
  • Richard I. Macphail, University College London, Paul Goldberg, Boston University
  • Book: Applied Soils and Micromorphology in Archaeology
  • Online publication: 18 December 2017
  • Chapter DOI: https://doi.org/10.1017/9780511895562.016
Available formats
×