Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-27T12:44:04.472Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2012

Stephen Weiner
Affiliation:
Weizmann Institute of Science, Israel
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
Microarchaeology
Beyond the Visible Archaeological Record
, pp. 327 - 372
Publisher: Cambridge University Press
Print publication year: 2010

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

Abelson, P. H. (1955). Organic constituents of fossils. Carnegie Institution of Washington Yearbook 54, 107–109.Google Scholar
Abelson, P. H. (1956). Paleobiochemistry. Journal of American Science 195, 83–88.CrossRefGoogle Scholar
Addadi, L., Berman, A., and Weiner, S. (1991). Intracrystalline proteins from a sea urchin and a mollusk: a comparison. In: Mechanisms and Phylogeny of Mineralization in Biological Systems (ed. Suga, S. and Nakahara, H.), pp. 29–33. Springer, Tokyo.CrossRefGoogle Scholar
Addadi, L., Joester, D., Nudelman, F., and Weiner, S. (2006). Mollusk shell formation: a source of new concepts for understanding biomineralization processes. Chemistry European Journal 12, 980–987.CrossRefGoogle ScholarPubMed
Addadi, L., Raz, S., and Weiner, S. (2003). Taking advantage of disorder: amorphous calcium carbonate and its roles in biomineralization. Advanced Materials 15, 959–970.CrossRefGoogle Scholar
Aitken, M. J. (1958). Magnetic dating. Archaeometry 1, 16–20.CrossRefGoogle Scholar
Aitken, M. J. (1976). Thermoluminescence age evaluation and assessment of error limits: revised system. Archaeometry 18, 233–238.CrossRefGoogle Scholar
Aitken, M. J. (1990). Science-Based Dating in Archaeology. Longman Group, London and New York.Google Scholar
Aitken, M. J., and Valladas, H. (1992). Luminescence dating relevant to human origins. Philosophical Transactions of the Royal Society of London, Series B 337, 139–144.CrossRefGoogle ScholarPubMed
Albert, R. M., and Weiner, S. (2001). Study of phytoliths in prehistoric ash layers using a quantitative approach. In: Phytoliths: Applications in Earth Sciences and Human History (ed. Meunier, J. D. and Colin, F.), pp. 251–266. A. A. Balkema, Exton, PA.Google Scholar
Albert, R. M., Tsatskin, A., Ronen, A., Lavi, O., Estroff, L., Lev-Yadun, S., and Weiner, S. (1999). Mode of occupation of Tabun Cave, Mt Carmel, Israel during the Mousterian Period: A study of the sediments and phytoliths. Journal of Archaeological Science 26, 1249–1260.CrossRefGoogle Scholar
Albert, R. M., Weiner, S., Bar-Yosef, O., and Meignen, L. (2000). Phytoliths in the Middle Paleolithic deposits of Kebara Cave, Mt. Carmel, Israel: study of the plant materials used for fuel and other purposes. Journal of Archaeological Science 27, 931–947.CrossRefGoogle Scholar
Albert, R. M., Bar-Yosef, O., and Weiner, S. (2003a). Use of plant remains in Kebara Cave: phytoliths and mineralogical analyses. In: Kebara Cave, Mt. Carmel, Israel: The Middle and Upper Paleolithic Archaeology, Part 1 (ed. Bar-Yosef, O. and Meignen, L.), pp. 147–164. Peabody Museum of Archaeology and Ethnology, Harvard University, Cambridge, MA.
Albert, R. M., Bar-Yosef, O., Meignen, L., and Weiner, S. (2003b). Quantitative phytolith study of hearths from the Natufian and Middle Paleolithic levels of Hayonim Cave (Galilee, Israel). Journal of Archaeological Science 30, 461–480.CrossRefGoogle Scholar
Albert, R. M., Bamford, M. K., and Cabanes, D. (2006). Taphonomy of phytoliths and macroplants in different soils from Olduvai Gorge (Tanzania) and the application to Plio-Pleistocene palaeoanthropological samples. Quaternary International 148, 78–94.CrossRefGoogle Scholar
Albert, R. M., Shahack-Gross, R., Cabanes, D., Gilboa, A., 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 35, 57–75.CrossRefGoogle Scholar
Alexander, M. (1977). Introduction to Soil Microbiology. 2nd edition. John Wiley, New York.Google Scholar
Alexandre, A., Meunier, J.-D., Colin, F., and Koud, J.-M. (1997). Plant impact on the biogeochemical cycle of silicon and related weathering processes. Geochimica et Cosmochimica Acta 61, 677–682.CrossRefGoogle Scholar
Alon, D., Mintz, G., Cohen, I., Weiner, S., and Boaretto, E. (2002). The use of Raman spectroscopy to monitor the removal of humic substances from charcoal: quality control for 14C dating of charcoal. Radiocarbon 44, 1–11.CrossRefGoogle Scholar
Al-Shorman, A. (2004). Stable carbon isotopic analysis of human tooth enamel from the Bronze Age cemetery of Ya'amoun in Northern Jordan. Journal of Archaeological Science 31, 1693–1698.CrossRefGoogle Scholar
Ambrose, S. H. (1991). Effects of diet, climate and physiology on nitrogen isotope abundances in terrestrial foodwebs. Journal of Archaeological Science 18, 293–317.CrossRefGoogle Scholar
Ambrose, S. H. (1998). Chronology of the Late Stone Age and food production in East Africa. Journal of Archaeological Science 25, 377–392.CrossRefGoogle Scholar
Amos, G. L. (1952). Silica in timbers. CSIRO Bulletin Melbourne Australia 267, 5–55.Google Scholar
Anderson, S., and Ertug-Yaras, F. (1998). Fuel fodder and faeces: an ethnographic and botanical study of dung fuel use in Central Anatolia. Environmental Archaeology 1, 99–109.CrossRefGoogle Scholar
Andrus, C. F. T., and Crowe, D. E. (2002). Alteration of otolith aragonite: effects of prehistoric cooking methods on otolith chemistry. Journal of Archaeological Science 29, 291–299.CrossRefGoogle Scholar
Angelini, I., and Bellintani, P. (2005). Archaeological ambers from Northern Italy: an FTIR-DRIFT study of provenance by comparison with the geological amber database. Archaeometry 47, 441–454.CrossRefGoogle Scholar
Aoba, T., and Moreno, E. (1990). Changes in the nature and composition of enamel during porcine amelogenesis. Calcified Tissue International 47, 356–364.CrossRefGoogle ScholarPubMed
Arias, J. L., Fink, D. J., Si-Qun, X., Heuer, A. H., and Caplan, A. I. (1993). Biomineralization and eggshells: cell-mediated acellular compartments of mineralized extracellular matrix. International Review of Cytology 145, 217–250.CrossRefGoogle ScholarPubMed
Armitage, P. L. (1975). The extraction and identification of opal phytoliths from the teeth of ungulates. Journal of Archaeological Science 2, 187–197.CrossRefGoogle Scholar
Arnay-de-la-Rosa, M., González-Reimers, E., Gámez-Mendoza, A., and Galindo-Martín, L. (2009). The Ba/Sr ratio, carious lesions, and dental calculus among the population buried in the church La Concepción (Tenerife, Canary Islands). Journal of Archaeological Science 36, 351–358.CrossRefGoogle Scholar
Arnold, D. E. (1985). Ceramic Theory and Cultural Process. Cambridge University Press, Cambridge.Google Scholar
Arnold, D. E. (2005). Linking society with the compositional analysis of pottery: a model from comparative ethnography. In: Pottery Manufacturing Processes: Reconstruction and Interpretation. (ed. Smith, A. Livingstone, Bosquet, D., and Martineau, R.), pp. 15–21. British Archaeological Reports, International Series 1349, Oxford.Google Scholar
Arrhenius, O. (1931). Die bodenanalyse im dienst der archäologie. Zeitschrift für Pflanzenernährung, Düngung und Bodenkunde B 10, 427–439.CrossRefGoogle Scholar
Arsenault, A. L. (1991). Image analysis of collagen-associated mineral distribution in cryogenically prepared turkey leg tendons. Calcified Tissue International 48, 56–62.CrossRefGoogle ScholarPubMed
Ascough, P. L., Cook, G. T., Dugmore, A. J., Scott, E. M., and Freeman, S. P. H. T. (2005). Influence of mollusk species on marine ΔR determinations. Radiocarbon 47, 433–440.CrossRefGoogle Scholar
Aufderheide, A. C., Muñoz, I., and Arriaza, B. (2005). Seven Chinchorro mummies and the prehistory of northern Chile. American Journal of Physical Anthropology 91, 189–201.CrossRefGoogle Scholar
Ayliffe, L. K., and Chivas, A. R. (1990). Oxygen isotopic composition of the bone phosphate of Australian kangaroos: potential as a palaeoenvironmental recorder. Geochimica et Cosmochimica Acta 54, 2603–2609.CrossRefGoogle Scholar
Baillie, M. G. L. (1995). A Slice Through Time: Dendrochronology and Precision Dating. Routledge, London.Google Scholar
Baker, G., Jones, L. H. P., and Wardrop, L. D. (1959). Cause of ware in sheep's teeth. Nature 184, 1583–1584.CrossRefGoogle Scholar
Ball, T. B., Gardner, J. S., and Anderson, N. (1999). Identifying inflorescence phytoliths from selected species of wheat (Triticum monococcum, T. dicoccum, T. dicoccoides, and T. aestivum) and barley (Hordeum vulgare and H. spontaneum) (Gramineae). American Journal of Botany 86, 1615–1623.CrossRefGoogle Scholar
Bamber, R. K., and Lanyon, J. W. (1960). Silica deposition in several woods of New South Wales. Tropical Woods 113, 48–53.Google Scholar
Bamford, M. (1999). Pliocene fossil woods from an early hominid cave deposit, Sterkfontein, South Africa. South African Journal of Science 95, 231–238.Google Scholar
Bamford, M. K., Albert, R. M., and Cabanes, D. (2006). Plio-Pleistocene macroplant fossil remains and phytoliths from Lowermost Bed II in the eastern paleolake margin of Olduvai Gorge, Tanzania. Quaternary International 148, 95–112.CrossRefGoogle Scholar
Baraldi, P., Bonazzi, A., Giordani, N., Paccagnella, F., and Zannini, P. (2006). Analytical characterization of Roman plasters of the “Domus Farini” in Modena. Archaeometry 48, 481–499.CrossRefGoogle Scholar
Barba, L. (2007). Chemical residues in lime-plastered archaeological floors. Geoarchaeology 22, 439–452.CrossRefGoogle Scholar
Barber, M., Field, D., and Topping, P. (1999). Neolithic Flint Mines in England. Royal Commission on the Historical Monuments of England, Swindon.
Barton, H. (2007). Starch residues on museum artifacts: implications for determining tool use. Journal of Archaeological Science 34, 1752–1762.CrossRefGoogle Scholar
Bar-Yosef, O. (2001). The world around Cyprus: from Epi-Paleolithic foragers to the collapse of the PPNB civilization. In: The Earliest Prehistory of Cyprus from Colonization to Exploitation (ed. Swiny, S.), pp. 129–164. American Schools of Oriental Research, Boston.Google Scholar
Bar-Yosef Mayer, D. E. (2000). The economic importance of mollusks in the Levant. In: Archaeozoology of the Near East, vol. IV A (ed. Mashkour, M., Choyke, A. M., Buitenhuis, H., and Poplin, F.), pp. 218–227. Archeological Research and Consultancy, Groningen, Netherlands.Google Scholar
Bar-Yosef Mayer, D. E. (2005a). Archaeomalacology: Mollusks in Former Environments of Human Behavior. Oxbow, Oxford.Google Scholar
Bar-Yosef Mayer, D. E. (2005b). The exploitation of shells as beads in the Paleolithic and Neolithic of the Levant. Paléorient 31, 176–185.CrossRefGoogle Scholar
Bar-Yosef Mayer, D. E. (2008). Archaeomalacological research in Israel: the current state of research. Israel Journal of Earth Sciences 56, 191–206.CrossRefGoogle Scholar
Battarbee, R. W. (1988). The use of diatom analysis in archaeology: a review. Journal of Archaeological Science 15, 621–644.CrossRefGoogle Scholar
Baxter, M. S., and Walton, A. (1970). Radiocarbon dating of mortars. Nature 225, 937.CrossRefGoogle ScholarPubMed
Beck, C. W., Wilbur, E., Meret, S., Kossove, D., and Kermani, K. (1965). The infrared spectra of amber and the identification of Baltic amber. Archaeometry 8, 96–109.CrossRefGoogle Scholar
Behrensmeyer, A. K. (1978). Taphonomic and ecologic information from bone weathering. Paleobiology 4, 150–162.CrossRefGoogle Scholar
Behrensmeyer, A. K., and Hill, A. P., eds. (1988). Fossils in the Making: Vertebrate Taphonomy and Paleoecology. University of Chicago Press, Chicago.
Belfiore, C. M., Day, P. M., Hein, A., Kilikoglou, V., Rosa, V., Mazzoleni, P., and Pezzino, A. (2007). Petrographic and chemical characterization of pottery production of the Late Minoan I kiln at Haghia Triada, Crete. Archaeometry 49, 621–653.CrossRefGoogle Scholar
Bell, I. M., Clark, R. J., and Gibbs, P. J. (1997). Raman spectroscopic library of natural and synthetic pigments (pre- approximately 1850 AD). Spectrochimica Acta A 53, 2159–2179.CrossRefGoogle Scholar
Bellomo, R. V. (1993). A methodological approach for identifying archaeological evidence of fire resulting from human activities. Journal of Archaeological Science 20, 525–553.CrossRefGoogle Scholar
Bellemo, R. V., and Harris, J. W. K. (1990). Preliminary reports of actualistic studies of fire within Virunga National Park, Zaire: towards an understanding of archaeological occurrences. In: Evolution of Environments and Hominidae in the African Western Rift Valley, vol. 1 (ed. Boaz, N. T.), pp. 317–338. Virginia Museum of Natural History Memoir, Martinsville.Google Scholar
Benayas, J. (1963). Disolución parcial de sílice organica en suelos. Anales de Edafologia y Agrobiologia 22, 623–626.Google Scholar
Benecke, N. (1987). Studies on early dog remains from northern Europe. Journal of Archaeological Science 14, 31–49.CrossRefGoogle Scholar
Beniash, E., Aizenberg, J., Addadi, L., and Weiner, S. (1997). Amorphous calcium carbonate transforms into calcite during sea urchin larval spicule growth. Proceedings of the Royal Society of London, Series B 264, 461–465.CrossRefGoogle Scholar
Bentley, R. A., Krause, R., and Price, T. D. (2003). Human mobility at the early Neolithic settlement of Vaihingen, Germany: evidence from strontium isotope analysis. Archaeometry 45, 471–486.CrossRefGoogle Scholar
Benyon, A. D., and Wood, B. A. (1987). Patterns and rates of enamel growth in the molar teeth of early hominids. Nature 326, 493–496.CrossRefGoogle Scholar
Berman, A., Addadi, L., and Weiner, S. (1988). Interactions of sea urchin skeleton macromolecules with growing calcite crystals – a study of intracrystalline proteins. Nature 331, 546–548.Google Scholar
Berman, A., Hanson, J., Leiserowitz, L., Koetzle, T. F., Weiner, S., and Addadi, L. (1993). Biological control of crystal texture: a widespread strategy for adapting crystal properties to function. Science 259, 776–779.CrossRefGoogle Scholar
Berna, F., and Goldberg, P. (2008). Assessing Paleolithic pyrotechnology and associated hominin behavior in Israel. Israel Journal of Earth Sciences 56, 107–121.CrossRefGoogle Scholar
Berna, F., Matthews, A., and Weiner, S. (2003). Solubilities of bone mineral from archaeological sites: the recrystallization window. Journal of Archaeological Science 31, 867–882.CrossRefGoogle Scholar
Berna, F., Behar, A., Shahack-Gross, R., Berg, J., Boaretto, E., Gilboa, A., Sharon, I., et al. (2007). Sediments exposed to high temperatures: reconstructing pyrotechnological processes in Late Bronze and Iron Age Strata at Tel Dor (Israel). Journal of Archaeological Science 34, 358–373.CrossRefGoogle Scholar
Bernal, M. P., Sánchez-Monedero, M. A., Paredes, C., and Roig, A. (1998). Carbon mineralization from organic wastes at different composting stages during their incubation with soil. Agriculture, Ecosystems, and Environment 69, 175–189.CrossRefGoogle Scholar
Bertram, B. C. R. (1992). The Ostrich Communal Nesting System. Princeton University Press, Princeton, NJ.CrossRefGoogle Scholar
Bethell, P. H., and Máté, I. (1989). The use of soil phosphate analysis in archaeology: a critique. In: Scientific Analysis in Archaeology, Monograph 19 (ed. Henderson, J.), pp. 1–29. Oxford University Committee for Archaeology, Oxford.Google Scholar
Bieber, A. M., Brooks, D. W., Harbottle, G., and Sayre, E. V. (1976). Application of multivariate techniques to analytical data on Aegean ceramics. Archaeometry 18, 59–74.CrossRefGoogle Scholar
Binford, L. R. (1978). Nunamiut Ethnoarchaeology. Academic Press, New York.Google Scholar
Binford, L. R. (1981). Bones: Ancient Men and Modern Myths. Academic Press, New York.Google Scholar
Binford, L. R. (1983). In Pursuit of the Past. Thames and Hudson, New York.Google Scholar
Bird, M. I., Turney, C. S. M., Fifield, K., Jones, R., Ayliffe, L. K., Palmer, A., Cresswell, R., and Robertson, S. (2002). Radiocarbon analysis of the early archaeological site of Nauwalabila I, Arnhem Land, Australia: implications for sample suitability and stratigraphic integrity. Quaternary Science Review 21, 1061–1075.CrossRefGoogle Scholar
Bird, M. I., Turney, C. S. M., Fifield, L. K., Smith, M. A., Miller, G. H., Roberts, R. G., and Magee, J. W. (2003). Radiocarbon dating of organic- and carbonate-carbon in Genyornis and Dromaius eggshell using stepped combustion and stepped acidification. Quaternary Science Reviews 22, 15–17.CrossRefGoogle Scholar
Black, D. (1931). Evidences of the use of fire by Sinanthropus. Bulletin of the Geological Society of China 11, 107–108.CrossRefGoogle Scholar
Black, D., Teilhard de Chardin, P., and Young, C. C. (1933). Fossil man in China. Memoirs of the Geological Survey of China A, 1–166.Google Scholar
Blackman, E. B. (1969). Observations on the development of the silica cells of the leaf sheath of wheat (Triticum aestivum). Canadian Journal of Botany 47, 827–838.CrossRefGoogle Scholar
Blake, R. E., O'Neil, J. R., and Garcia, G. A. (1997). Oxygen isotope systematics of biologically mediated reactions of phosphate: I. Microbial degradation of organophosphorus compounds. Geochimica et Cosmochimica Acta 61, 4411–4422.CrossRefGoogle Scholar
Blow, M. J., Zhang, T., Woyke, T., Speller, C. F., Krivoshapkin, A., Yang, D. Y., Derevianko, A., and Rubin, E. M. (2008). Identification of ancient remains through genomic sequencing. Genome Research 18, 1347–1353.CrossRefGoogle ScholarPubMed
Blum, J. D., Taliaferro, H. E., Weisse, M. T., and Holmes, R. T. (2000). Changes in Sr/Ca, Ba/Ca and 87Sr/86Sr ratios between trophic levels in two forest ecosystems in the northeastern U.S.A. Biogeochemistry 49, 87–101.CrossRefGoogle Scholar
Boardman, S., and Jones, G. (1990). Experiments on the effect of charring on cereal plant components. Journal of Archaeological Science 17, 1–11.CrossRefGoogle Scholar
Boaretto, E. (2008). Determining the chronology of an archaeological site using radiocarbon: minimizing uncertainty. Israel Journal of Earth Sciences 56, 207–216.CrossRefGoogle Scholar
Boaretto, E. (2009a). Dating materials in good archaeological contexts: the next challenge for radiocarbon analysis. Radiocarbon 51, 275–282.CrossRefGoogle Scholar
Boaretto, E., Barkai, R., Gopher, A., Berna, F., Kubic, P. W., and Weiner, S. (2009b). Specialized flint procurement strategies for hand axes, scapersa and blades in the Late Lower Paleolithic: a 10Be study at Qesem Cave, Israel. Human Evolution.
Boaretto, E., Wu, X., Yuan, J., Bar-Yosef, O., Chu, V., Pan, Y., Liu, K., et al. (2009c). Radiocarbon dating of the deposits and pottery at Yuchanyan Cave, Hunan Province, China. Proceedings of the National Academy of Sciences of the United States of America. 106, 9595–9600.CrossRefGoogle Scholar
Bocherens, H., Koch, P. L., Mariotti, A., Geraads, D., and Jaeger, J.-J. (1996). Isotopic biogeochemistry (13C, 18O) of mammalian enamel from African Pleistocene hominid sites. Palaios 11, 306–318.CrossRefGoogle Scholar
Boggild, O. (1930). The shell structure of the mollusks. K. Dan. Vidensk. Selsk. Skr. Naturvidensk. Math. Afd. 9, 233–326.Google Scholar
Bolan, N. S., Hedley, M. J., and White, R. E. (1991). Processes of soil acidification during nitrogen cycling with emphasis on legume based pastures. Plant and Soil 134, 53–63.CrossRefGoogle Scholar
Bonani, G., Ivy, S., Hajdas, I., Niklaus, T. R., and Suter, M. (1994). AMS 14C age determinations of tissue, bone and grass samples from the Ötztal Ice Man. Radiocarbon 36, 247–250.CrossRefGoogle Scholar
Borrelli, E. (1999). Binders. International Centre for the Study of the Preservation and Restoration of Cultural Property, Rome.
Bowman, H. R., Asaro, F., and Perlman, I. (1973). Composition variations in obsidian sources and the archaeological implications. Archaeometry 15, 123–127.CrossRefGoogle Scholar
Bowman, S. (1990). Radiocarbon Dating: Interpreting the Past. University of California Press, Berkeley.Google Scholar
Boyde, A. (1963). Estimation of age at death of young human skeletal remains from incremental lines in the dental enamel. Ecerpta Medica International Congress Series 80, 36–46.Google Scholar
Boyde, A. (1997). Microstructure of enamel. In: Dental Enamel, CIBA Foundation Symposium 205 (ed. Chadwick, D. J. and Cardew, G.), pp. 18–31. John Wiley, Chichester, UK.Google Scholar
Boynton, R. S. (1980). Chemistry and Technology of Lime and Limestone. John Wiley, New York.Google Scholar
Bozarth, S. (1987). Diagnostic opal phytoliths from rinds of selected Cucurbita species. American Antiquity 52, 607–615.CrossRefGoogle Scholar
Brain, C. K. (1967a). Hottentot food remains and their bearings on the interpretation of fossil bone assemblages. Scientific Papers of the Namib Desert Research Station 32, 1–11.Google Scholar
Brain, C. K. (1967b). New light on old bones. South African Museum Association Bulletin 9, 22–27.Google Scholar
Brain, C. K. (1981). The Hunters or the Hunted? An Introduction to African Cave Taphonomy. University of Chicago Press, Chicago.Google Scholar
Brain, C. K., and Sillen, A. (1988). Evidence from the Swartkrans cave for the earliest use of fire. Nature 336, 464–466.CrossRefGoogle Scholar
Branda, F., Luciani, G., Constantini, A., and Piccioli, C. (2001). Interpretation of the thermogravimetric curves of ancient pozzolanic concretes. Archaeometry 43, 447–453.CrossRefGoogle Scholar
Bray, H. J., and Redfern, S. A. T. (1999). Kinetics of dehydration of Ca-montmorillonite. Physical Chemistry of Minerals 26, 591–600.CrossRefGoogle Scholar
Brill, R. H., and Wampler, J. M. (1965). Isotope ratios in archaeological artifacts of lead. In: Applications of Science in the Examination of Works of Art, pp. 155–166. Museum of Fine Arts, Boston.
Britton, K., Grimes, V., Dau, J., and Richards, M. P. (2009). Reconstructing faunal migrations using intra-tooth sampling and strontium and oxygen isotope analyses: a case study of modern caribou (Rangifer tarandus granti). Journal of Archaeological Science 36, 1163–1172.CrossRefGoogle Scholar
Brochier, J. É. (1983). Combustions et parcage des herbivored domestiques. Bulletin de la Societé Préhistorique Française 80, 143–145.Google Scholar
Brochier, J. É., and Thinon, M. (2003). Calcite crystals, starch grains aggregates or…POCC? Comment on “calcite crystals inside archaeological plant tissues.” Journal of Archaeological Science 30, 1211–1214.CrossRefGoogle Scholar
Brochier, J. É., Villa, P., and Giacomarra, M. (1992). Shepherds and sediments: geo-ethnoarchaeology of pastoral sites. Journal of Anthropological Archaeology 11, 47–102.CrossRefGoogle Scholar
Bronk Ramsey, C (1995). Radiocarbon calibration and analysis of stratigraphy: the OxCal Program. Radiocarbon 37, 425–430.CrossRefGoogle Scholar
Bronk Ramsey, C (2001). Development of the radiocarbon program OxCal. Radiocarbon 43, 355–363.CrossRefGoogle Scholar
Bronk Ramsey, C. (2003). OxCal Version 3.9. Oxford Radiocarbon Accelerator Unit, Oxford.
Bronk Ramsey, C., and Higham, T. (2003). Towards high precision AMS: progress and limitations. Paper presented at the 18th International Radiocarbon Conference, Wellington, New Zealand.
Bronk Ramsey, C., Plicht, J., and Weninger, B. (2001). “Wiggle matching” radiocarbon dates. Radiocarbon 43, 381–389.CrossRefGoogle Scholar
Bronk Ramsey, C., Buck, C. E., Manning, S. W., Reimer, P., and Plicht, J. (2006). Developments in radiocarbon calibration for archaeology. Antiquity, 783–798.CrossRefGoogle Scholar
Brooks, A. S., Hare, P. E., Kokis, J. E., Miller, G. H., Ernst, R. D., and Wendorf, F. (1990). Dating archaeological sites by protein diagenesis in ostrich eggshell. Science 248, 60–64.CrossRefGoogle Scholar
Brothwell, D. R., and Pollard, A. M., eds. (2001). Handbook of Archaeological Sciences. John Wiley, Chichester, UK.
Brown, A. (1973). Bone strontium content as a dietary indicator in human populations. University of Michigan. PhD Dissertation.
Brown, T. A. (2001). Ancient DNA. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 301–311. John Wiley, Chichester, UK.Google Scholar
Bryant, J. D., Luz, B., and Froelich, P. N. (1994). Oxygen isotopic composition of fossil horse tooth phosphate as a record of continental paleoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 107, 303–316.CrossRefGoogle Scholar
Bryant, J. D., Koch, P., Froelich, P. N., Showers, W. J., and Genna, B. J. (1996). Oxygen isotope partitioning between phosphate and carbonate in mammalian bone. Geochimica et Cosmochimica Acta 60, 145–148.Google Scholar
Buchardt, B., and Weiner, S. (1981). Diagenesis of aragonite from Upper Cretaceous ammonites, a geochemical case-study. Sedimentology 28, 423–438.CrossRefGoogle Scholar
Buckland, W. (1822). Account of an assemblage of fossil teeth and bones of elephant, rhinoceros, hippopotamus, bear, tiger, and hyaena and sixteen other animals: discovered in a cave at Kirkdale, Yorkshire in the year 1821; with a comparative view of 5 similar caverns in various parts of England and others on the Continent. Philosophical Transactions of the Royal Society of London 112, 171–237.CrossRefGoogle Scholar
Burnell, J., Teubner, E., and Miller, A. (1980). Normal maturational changes in bone matrix, mineral, and crystal size in the rat. Calcified Tissue International 310, 13–19.CrossRefGoogle Scholar
Burroni, D., Donahue, R. E., and Pollard, A. M. (2002). The surface alteration features of flint artefacts as a record of environmental processes. Journal of Archaeological Science 29, 1277–1287.CrossRefGoogle Scholar
Buxeda, J., Garrigos, I., Mommsen, H., and Tsolakidou, A. (2002). Alterations of Na, K and Rb concentrations in Mycenaean pottery and a proposed explanation using X-ray diffraction. Archaeometry 44, 187–198.Google Scholar
Buxeda, J., Garrigos, I., Jones, R. E., Kilikoglou, V., Levi, S. T., Maniatis, Y., Mitchell, J., Vagnetti, L., Wardle, K. A., and Andreou, S. (2003). Technology transfer at the periphery of the Mycenaean world: the cases of Mycenaean pottery found in Central Macedonia (Greece) and the Plain of Sybaris (Italy). Archaeometry 45, 263–284.CrossRefGoogle Scholar
Buxton, D. R., and Russel, J. R. (1988). Lignin constituents and cell-wall digestibility of grass and legume stems. Crop Science 28, 553–558.CrossRefGoogle Scholar
Cain, C. J., and Heyn, A. N. J. (1964). X-ray diffraction studies of the crystalline structure of the avian eggshell. Biophysical Journal 4, 23–39.CrossRefGoogle Scholar
Calvert, S. E. (1974). Deposition and diagenesis of silica in marine sediments. In: Pelagic Sediments: On Land and Under the Sea, vol. 1 (ed. Hsü, K. J. and Jenkyns, H. C.). International Association of Sedimentology Special Publications 1, 273–299.Google Scholar
Canti, M. (1997). An investigation of microscopic calcareous spherulites from herbivore dung. Journal of Archaeological Science 24, 219–231.CrossRefGoogle Scholar
Canti, M. G. (1998). Origin of calcium carbonate granules found in buried soils and Quaternary deposits. Boreas 27, 275–288.CrossRefGoogle Scholar
Canti, M. G. (2007). Deposition and taphonomy of earthworm granules in relation to their interpretative potential in Quaternary stratigraphy. Journal of Quaternary Science 22, 111–118.CrossRefGoogle Scholar
Carcaillet, C. (1998). A spatially precise study of Holocene fire history, climate and human impact within the Maurienne Valley, north French Alps. Journal of Ecology 86, 384–396.CrossRefGoogle Scholar
Carlström, D. (1963). A crystallographic study of vertebrate otoliths. Biological Bulletin 125, 441–463.CrossRefGoogle Scholar
Caró, F., and Di Giulio, A. (2004). Reliability of textural analysis of ancient plasters and mortars through automated image analysis. Materials Characterization 53, 243–257.CrossRefGoogle Scholar
Caró, F., Riccardi, M. P., and Mazzilli Savini, M. T. (2008). Characterization of plasters and mortars as a tool in archaeological studies: the case of Lardirago Castle in Pavia, northern Italy. Archaeometry 50, 85–100.Google Scholar
Carter, J. G., and Clark, G. R. (1985). Classification and phylogenetic significance of molluscan shell structure. In: Mollusks: Notes for a Short Course (ed. Broadhead, T. W.), pp. 50–71. Department of Geological Science Studies, University of Tennessee, Knoxville.Google Scholar
Carty, M. W., and Senapti, U. (1998). Porcelain – raw materials, processing, phase evolution, and mechanical behavior. Journal of the American Ceramic Society 81, 3–20.CrossRefGoogle Scholar
Casadio, F., Chiari, G., and Simon, S. (2005). Evaluation of binder/aggregate ratios in archaeological lime mortars with carbonate aggregate: a comparative assessment of chemical, mechanical and microscopic approaches. Archaeometry 47, 671–689.CrossRefGoogle Scholar
Casas, L., Linford, P., and Shaw, J. (2007). Archaeomagnetic dating of Dogmersfield Park brick kiln (southern England). Journal of Archaeological Science 34, 205–213.CrossRefGoogle Scholar
Cazalla, O., Rodriguez-Navarro, C., Sebastian, E., and Cultrone, G. (2000). Aging of lime putty: effects on traditional lime mortar carbonation. Journal of the American Ceramics Society 83, 1070–1076.CrossRefGoogle Scholar
Charola, A. E., and Henriques, F. M. A. (1999). Hydraulicity in lime mortars revisited. Paper presented at the RILEM TC 167-COM International Workshop, University of Paisley, Scotland.
Chefetz, B., Tarchitzky, J., Deshmukh, A. P., Hatcher, P. G., and Chen, Y. (2002). Structural characterization of soil organic matter and humic acids in particle-size fractions of an agricultural soil. Soil Science Society of America Journal 66, 129–141.CrossRefGoogle Scholar
Chernysh, I. G., Pakhovchisin, S. V., and Goncharik, V. P. (1993). The activity of dispersed oxides on natural and exfoliated graphite surfaces. Reaction Kinetics Catalysis Letters 50, 273–277.CrossRefGoogle Scholar
Cho, G., Wu, Y., and Ackerman, J. L. (2003). Detection of hydroxyl ions in bone mineral by solid state NMR spectroscopy. Science 300, 1123–1127.CrossRefGoogle ScholarPubMed
Chu, V., Regev, L., Weiner, S., and Boaretto, E. (2008). Differentiating between anthropogenic calcite in plaster, ash and natural calcite using infrared spectroscopy: implications in archaeology. Journal of Archaeological Science 35, 905–911.CrossRefGoogle Scholar
Cohen-Ofri, I., Weiner, L., Boaretto, E., Mintz, G., and Weiner, S. (2006). Modern and fossil charcoal: aspects of structure and diagenesis. Journal of Archaeological Science 33, 428–439.CrossRefGoogle Scholar
Cohen-Ofri, I., Popovitz-Biro, R., and Weiner, S. (2007). Structural characterization of modern and fossil natural charcoal using high resolution TEM and electron energy loss spectroscopy (EELS). Chemistry: A European Journal 13, 2306–2310.CrossRefGoogle Scholar
Cole, W. E., and Hoskins, J. S. (1957). Clay mineral mixtures and interstratified minerals. In: The Differential Thermal Investigation of Clays (ed. Mackenzie, R. C.), pp. 248–274. Mineralogical Society, London.Google Scholar
Collins, M. J., Muyzer, G., Westbroek, P., Curry, G. B., Sandberg, P. A., Xu, S. J., Quinn, R., and MacKinnon, D. (1991). Preservation of fossil biopolymeric structures: conclusive immunological evidence. Geochimica et Cosmochimica Acta 55, 2253–2257.CrossRefGoogle Scholar
Collins, M. J., Riley, M., Child, A. M., and Turner, W. G. (1995). A basic mathematical simulation of the chemical degradation of ancient collagen. Journal of Archaeological Science 22, 175–183.CrossRefGoogle Scholar
Collins, M. J., Nielsen-Marsh, C. M., Hiller, J., Smith, C. I., Roberts, J. P., Prigodich, R. V., Wess, T. J., Csapó, J., Millard, A. R., and Turner-Walker, G. (2002). The survival of organic matter in bone: a review. Archaeometry 44, 383–394.CrossRefGoogle Scholar
Colomban, P. (2005). Raman μ-spectrometry, a unique tool for on-site analysis and identification of ancient ceramics and glasses. Materials Research Society Symposium Proceedings 852, 1–15.Google Scholar
Condamin, J., Formenti, F., Metais, M. O., Michel, M., and Blond, P. (1976). The applications of gas chromatography to the tracing of oil in ancient amphorae. Archaeometry 18, 195–201.CrossRefGoogle Scholar
Courty, M. A., Goldberg, P., and MacPhail, R. (1989). Soils and Micromorphology in Archaeology. Cambridge University Press, Cambridge.Google Scholar
Coutts, P. J. F. (1970). Bivalve growth patterning as a method of seasonal dating in archaeology. Nature 226, 874.CrossRefGoogle ScholarPubMed
Crane, N. J., Popescu, V., Morris, M. D., Steenhuis, P., and Ignelzi, M. A. (2006). Raman spectroscopic evidence for octacalcium phosphate and other mineral species deposited during intramembraneous mineralization. Bone 39, 431–433.CrossRefGoogle Scholar
Cuif, J.-P., and Dauphin, Y. (2005). The two-step mode of growth in the scleractinian coral skeletons from the micrometre to the overall scale. Journal of Structural Biology 150, 319–331.CrossRefGoogle ScholarPubMed
Currey, J. D. (1984). The Mechanical Adaptations of Bones. Princeton University Press, Princeton, NJ.CrossRefGoogle Scholar
Currey, J. D. (2002). Bones. Princeton University Press, Princeton, NJ.Google Scholar
Curry, G. B., Cusack, M., Walton, D., Endo, K., Clegg, H., Abbott, G., and Armstrong, H. (1991). Biogeochemistry of brachiopod intracrystalline molecules. Philosophical Transactions of the Royal Society of London, Series B 333, 359–365.CrossRefGoogle Scholar
Daculsi, G., Menanteau, J., Kerebel, L. M., and Mire, D. (1984). Length and shape of enamel crystals. Calcified Tissue International 36, 550–555.CrossRefGoogle ScholarPubMed
Darwin, C. (1859). The Origin of Species. John Murray, London.Google Scholar
da Silveira, P. M., Rosário Veiga, M., and Brito, J. (2007). Gypsum coatings in ancient buildings. Construction and Building Materials 21, 126–131.Google Scholar
Dauphin, Y. (2002). Comparison of the soluble matrices of the calcitic prismatic layer of Pinna nobilis (Mollusca, Bivalvia, Pteriomorpha). Comparative Biochemistry and Physiology, Part A 132, 577–590.CrossRefGoogle Scholar
Dauphin, Y., Pickford, M., and Senut, B. (1998). Diagenetic changes in the mineral and organic phases of fossil avian eggshells from Namibia. Applied Geochemistry 13, 243–256.CrossRefGoogle Scholar
Dauphin, Y., Montuelle, S., Quantin, C., and Massard, P. (2007). Estimating the preservation of tooth structures: towards a new scale of observation. Journal of Taphonomy 5, 43–56.Google Scholar
Davies, G., and Ghabbour, E. A. (1998). Humic Substances: Structures, Properties and Uses. Royal Society of Chemistry, Cambridge.Google Scholar
Dean, M. C., Beynon, A. D., Thackeray, J. F., and Macho, G. A. (1993). Histological reconstruction of dental development and age at death of a juvenile Paranthropus robustus specimen, SK 63, from Swartkrans, South Africa. American Journal of Physical Anthropology 91, 401–419.CrossRefGoogle ScholarPubMed
Benedetto, G. E., Laviano, R., Sabbatini, L., and Zambonin, P. G. (2002). Infrared spectroscopy in the mineralogical characterization of ancient pottery. Journal of Cultural Heritage 3, 177–186.CrossRefGoogle Scholar
Deer, W. A., Howie, R. A., and Zussman, J. (1992). An Introduction to the Rock-Forming Minerals. London Scientific and Technical, London.Google Scholar
Deith, M. R. (1983a). Seasonality of shell collecting determined by oxygen isotope analysis of marine shells from Asturian sites in Cantabria. In: Animals and Archaeology, vol. 2, Shell Middens, Fishes and Birds (ed. Grigson, C. and Clutton-Brock, J.). British Archaeological Reports International Series 183, 67–76.
Deith, M. R. (1983b). Molluscan calendars: the use of growth line analysis to establish seasonality of shellfish collection at the Mesolithic site of Morton, Fife. Journal of Archaeological Science 10, 423–440.CrossRefGoogle Scholar
Della Casa, P. (2005). Lithic resources in the early prehistory of the Alps. Archaeometry 47, 221–234.CrossRefGoogle Scholar
DeNiro, M. J. (1985). Postmortem preservation and alteration of in vivo bone collagen isotope ratio in relation to palaeodietry reconstraction. Nature, 317, 806–809.CrossRefGoogle Scholar
DeNiro, M. J., and Epstein, S. (1978). Influence of diet on the distribution of carbon isotopes in animals. Geochimica et Cosmochimica Acta 42, 495–506.CrossRefGoogle Scholar
DeNiro, M. J., and Weiner, S. (1988a). Chemical, enzymatic and spectroscopic characterization of collagen and other organic fractions from prehistoric bones. Geochimica et Cosmochimica Acta 52, 2197–2206.CrossRefGoogle Scholar
DeNiro, M. J., and Weiner, S. (1988b). Organic matter within crystalline aggregates of hydroxyapatite: a new substrate for stable isotopic and possibly other biochemical analyses of bone. Geochimica et Cosmochimica Acta 52, 2415–2423.CrossRefGoogle Scholar
Dennis, J. E., Xiao, S. Q., Agarval, M., Fink, D. J., Heuer, A. H., and Caplan, A. I. (1996). Microstructure of matrix and mineral components of eggshells from white leghorn chicken (Gallus gallus). Journal of Morphology 228, 287–306.3.0.CO;2-#>CrossRefGoogle Scholar
Derry, L. A., Kurtz, A. C., Ziegler, K., and Chadwick, O. A. (2005). Biological control of terrestrial silica cycling and export fluxes to watersheds. Nature 433, 728–731.CrossRefGoogle ScholarPubMed
Vaux, R. (1973). Archaeology and the Dead Sea Scrolls. Oxford University Press, Oxford.Google Scholar
Dixon, J., and Weed, S. (1977). Minerals in Soil Environments. Soil Science Society of America, Madison, WI.
Dodd, J., and Stanton, R. J. (1981). Paleoecology: Concepts and Applications. John Wiley, New York.Google Scholar
Doner, H. E., and Lynn, W. C. (1977). Carbonate, halide, sulfate, and sulfide minerals. In: Minerals in Soil Environments (ed. Dixon, J. B. and Weed, S. B.), pp. 75–98. Soil Science Society of America, Madison, WI.Google Scholar
Douglass, A. E. (1921). Dating our prehistoric ruins. Natural History 21, 27–30.Google Scholar
Douglass, A. E. (1941). Crossdating in dendrochronology. Journal of Forestry 39, 825–831.Google Scholar
Dove, P. M., Weiner, S., and DeYoreo, J. J. (2003). Biomineralization. Reviews in Mineralogy and Geochemistry vol. S54.Google Scholar
Dressler, V. L., Pozebon, D., Flores, E. L. M., Paniz, J. N. B., and Flores, E. M. M. (2002). Potentiometric determination of fluoride in geological and biological samples following pyrohydrolytic decomposition. Analytica Chimica Acta 466, 117–123.CrossRefGoogle Scholar
Dudd, S. N., and Evershed, R. P. (1998). Direct demonstration of milk as an element of archaeological economies. Science 282, 1478–1481.CrossRefGoogle ScholarPubMed
Dufour, E., Cappetta, H., Denis, A., Dauphin, Y., and Mariotti, A. (2000). La diagenese des otolithes par la comparaison des donnees microstructurales, mineralogiques et geochimiques; application aux fossiles du Pliocene du Sud-Est de la France. Bulletin de la Societe Geologique de France 171, 521–532.CrossRefGoogle Scholar
Eckert, K. A., and Kunkel, T. A. (1991). DNA polymerase fidelity and the polymerase chain reaction. In: PCR Methods and Application (eds. Bentley, D., Gibbs, R., Green, E. and Myers, R. pp. 17–24. Cold Spring Harbour Laboratory Press, New York.Google Scholar
Eckmeier, E., Gerlach, R., Gehrt, E., and Schmidt, M. W. I. (2007). Pedogenesis of Chernozems in central Europe – a review. Geoderma 139, 288–299.CrossRefGoogle Scholar
Edwards, K. J., Whittington, G., Robinson, M., and Richter, D. (2005). Paleoenvironments, the archaeological record and cereal pollen detection at Clickimin, Shetland, Scotland. Journal of Archaeological Science 32, 1741–1756.CrossRefGoogle Scholar
Edwards, N. T. (1975). Effects of temperature and moisture on carbon dioxide evolution in a mixed deciduous forest floor. Soil Science Society of America Proceedings 39, 361–365.CrossRefGoogle Scholar
Eframov, I. A. (1940). Taphonomy: a new branch of paleontology. Pan-American Geology 74, 81–93.Google Scholar
Efstratiou, N. (1984). Ethnoarchaeological research in Thrace. Archaeology 13, 20–26.Google Scholar
Efstratiou, N. (1990). Prehistoric habitations and structures in northern Greece: an ethnoarchaeological case study. Bulletin de Correspondance Hellénique 19, 33–41.Google Scholar
Eiland, M. L., and Williams, Q. (2000). Infrared spectroscopy of ceramics from Tell Brak, Syria. Journal of Archaeological Science 27, 993–1006.CrossRefGoogle Scholar
Elbaum, R., Albert, R. M., Elbaum, M., and Weiner, S. (2003). Detection of burning of plant materials in the archaeological record by changes in the refractive indices of siliceous phytoliths. Journal Archaeological Science 30, 217–226.CrossRefGoogle Scholar
Elbaum, R., Melamed-Bessudo, C., Boaretto, E., Galili, E., Lev-Yadun, S., Levy, A. A., and Weiner, S. (2006). Ancient olive DNA in pits: preservation, amplification and sequence analysis. Journal Archaeological Science 33, 77–88.CrossRefGoogle Scholar
Elbaum, R., Zaltzman, L., Burgert, I., and Fratzl, P. (2007). The role of wheat awns in the seed dispersal unit. Science 316, 884–886.CrossRefGoogle ScholarPubMed
Elbaum, R., Melamed-Bessudo, C., Tuross, N., Levy, A. A., and Weiner, S. (2009). New methods to isolate organic materials from silicified phytoliths reveal fragmented glycoproteins but no DNA. Quaternary International 193, 11–19.CrossRefGoogle Scholar
El Hiveris, S. O. (1978). Nature of resistance to Striga hermonithica (Del) Benth. parasitism in some Sorghum vulgare (Pers.) cultivars. Weed Research 27, 305–311.Google Scholar
Elias, R. W., Hirao, Y., and Patterson, C. C. (1982). The circumvention of the natural biopurification of calcium along nutrient pathways by atmospheric inputs of industrial lead. Geochimica et Cosmochimica Acta 46, 2561–2580.CrossRefGoogle Scholar
Eliyahu-Behar, A., Regev, L., Shalev, Y., Shilstein, S., Sharon, G., Gilboa, A., and Weiner, S. (2009). Functional analysis of a pyrotechnological installation from the Roman period at Tel-Dor, Israel: casting pit for bronze objects. Journal of Field Archaeology, 34, 135–151.CrossRefGoogle Scholar
El Mansouri, M., El Fouikar, A., and Saint Martin, B. (1996). Correlation between 14C ages and aspartic acid racemization at the Upper Paleolithic site of the Abri Pataud (Dordogne, France). Journal of Archaeological Science 23, 803–809.CrossRefGoogle Scholar
Elsdon, T. S., and Gillanders, B. M. (2003). Reconstructing patterns of fish based on environmental influences on otolith chemistry. Reviews in Fish Biology and Fisheries 13, 219–235.CrossRefGoogle Scholar
Epstein, S., Buchsbaum, R., Lowenstam, H. A., and Urey, H. C. (1953). Revised carbonate-water isotopic temperature scale. Bulletin of the Geological Society of America 64, 1315–1326.CrossRefGoogle Scholar
Ericson, J. E. (1985). Strontium isotope characterization in the study of prehistoric human ecology. Journal of Human Evolution 14, 503–514.CrossRefGoogle Scholar
Esslemont, G., Maher, W., Ford, P., and Krikowa, F. (2000). The determination of phosphorus and other elements in plant leaves by ICP-MS after low-volume microwave digestion with nitric acid. Atomic Spectroscopy 21, 42–45.Google Scholar
Everett, D. H. (1959). An Introduction to the Study of Chemical Thermodynamics. Longmans, Greem, London.Google Scholar
Evershed, R. P., Dudd, S. N., Lockheart, M. J., and Jim, S. (2001). Lipids in archaeology. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 331–349. John Wiley, Chichester, UK.Google Scholar
Faix, O., Bremer, J., Schmidt, O., and Stevanovic, T. (1991). Monitoring of chemical changes in white-rot degraded beech wood by pyrolysis – gas chromatography and Fourier-transform infrared spectroscopy. Journal of Analytical and Applied Pyrolysis 21, 147–162.CrossRefGoogle Scholar
Farmer, V. C. (1974). The Infrared Spectra of Minerals. Mineralogical Society, London.CrossRefGoogle Scholar
Farmer, V. C., Delbos, E., and Miller, J. D. (2005). The role of phytolith formation and dissolution in controlling concentrations of silica in soil solutions and streams. Geoderma 127, 71–79.CrossRefGoogle Scholar
Featherstone, J. D. B., Pearson, S., and LeGeros, R. Z. (1984). An infrared method for the quantification of carbonate in carbonated apatites. Caries Research 18, 63–66.CrossRefGoogle ScholarPubMed
Ferraiolo, J. A. (1982). A systematic classification of nonsilicate minerals. Bulletin of the American Museum of Natural History 172, 1–237.Google Scholar
Ferraiolo, J. A. (2008). A Systematic Classification of Nonsilicate Minerals. Ferraiola, J. A. jfer@erols.com, Bowie, MD, 1–457.Google Scholar
Feynman, R. P. (1998). The Meaning of It All: Thoughts of a Citizen Scientist. Perseus Books, Reading, MA.Google Scholar
Fitton Jackson, S. (1956). The fine structure of developing bone in the embryonic fowl. Proceedings of the Royal Society of London, Series B 146, 270–280.CrossRefGoogle Scholar
Fladmark, K. R. (1982). Microdebitage analysis: initial considerations. Journal of Archaeological Science 9, 205–220.CrossRefGoogle Scholar
Fleming, S. (1976). Dating in Archaeology. St. Martin's Press, New York.Google Scholar
Fogel, M. L., Tuross, N., and Owsley, D. W. (1989). Nitrogen isotope tracers. Carnegie Institution of Washington Yearbook 88, 133–134.Google Scholar
Forbes, R. J. (1957). Studies in Ancient Technology. Vol. 5. Brill, Leiden, Netherlands.Google Scholar
Franceschi, V. R., and Horner, H. T. (1980). Calcium oxalate crystals in plants. Botanical Review 46, 361–427.CrossRefGoogle Scholar
Franklin, R. E. (1951). Crystallite growth in graphitizing and non-graphitizing carbons. Proceedings of the Royal Society of London, Series A 209, 196–218.CrossRefGoogle Scholar
Fraysse, F., Pokrovsky, O. S., Schott, J., and Meunier, J.-D. (2006). Surface properties, solubility and dissolution kinetics of bamboo phytoliths. Geochimica et Cosmochimica Acta 70, 1939–1951.CrossRefGoogle Scholar
Freestone, I. C., and Tite, M. S. (1986). Refractories in the ancient and preindustrial world. In: High-Technology Ceramics: Past, Present and Future (ed. Kingery, W. D.), pp. 35–63. American Ceramic Society, Westerville, OH.Google Scholar
Freund, F., and Knobel, R. M. (1977). Distribution of fluorine in hydroxyapatite studied by infrared spectroscopy. Journal of the Chemical Society Dalton 12, 1136–1140.CrossRefGoogle Scholar
Friedman, I., and Smith, R. L. (1960). A new dating method using obsidian: Part 1, the development of the method. American Antiquity 25, 476–522.CrossRefGoogle Scholar
Friedman, I., and Trembour, F. W. (1983). Obsidian hydration dating update. American Antiquity 48, 544–547.CrossRefGoogle Scholar
Furlan, V., and Bissegger, P. (1975). Les mortiers anciens: histoire et essais d'analyse scientifique. Revue suisse d'Art et d'Archéologie 32, 166–178.Google Scholar
Gago-Duport, L., Briones, M. J. I., Rodríguez, J. B., and Covelo, B. (2008). Amorphous calcium carbonate biomineralization in the earthworm's calciferous gland: pathways to the formation of crystalline phases. Journal of Structural Biology 162, 422–435.CrossRefGoogle ScholarPubMed
Garrison, E. G. (2003). Techniques in Archaeological Geology. Springer, Berlin.CrossRefGoogle Scholar
Garrod, D. A. E. (1957). The Natufian culture: the life and economy of a Mesolithic people in the Near East. Proceedings of the British Academy 43, 211–227.Google Scholar
Garti, J., Kunin, P., Delarea, J., and Weiner, S. (2002). Calcium oxalate and sulphate-containing structures on the thallial surface of the lichen Ramalina lacera: response to polluted air and simulated acid rain. Plant, Cell and Environment 25, 1591–1604.CrossRefGoogle Scholar
Gauldie, R. W. (1986). Vaterite otoliths from chinook salmon (Oncorhynchus tshawytscha). New Zealand Journal of Marine and Freshwater Research 20, 209–217.CrossRefGoogle Scholar
Gautron, J., Rodriguez-Navarro, A. B., Gómez-Morales, J., Hernández-Hernández, M. A., Dunn, I. C., Bain, M. M., García-Ruiz, M., and Nys, Y. (2007). Evidence for the implication of chicken eggshell matrix proteins in the presence of shell mineralization. In: Biomineralization: From Paleontology to Materials Science (ed. Arias, J. L. and Fernández, M. S.), pp. 145–154. Editorial Universitaria, Santiago, Chile.Google Scholar
Geiger, S. B., and Weiner, S. (1993). Fluoridated carbonatoapatite in the intermediate layer between glass ionomer and dentin. Dental Materials 9, 33–36.CrossRefGoogle ScholarPubMed
Genestar, C., and Palou, J. (2006). SEM-FTIR spectroscopic evaluation of deterioration in an historic coffered ceiling. Analytic Bioanalytic Chemistry 384, 987–993.CrossRefGoogle Scholar
Gernaey, A. M., Waite, E. R., Collins, M. J., and Craig, O. E. (2001). Survival and interpretation of archaeological proteins. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 323–329. John Wiley, Chichester, UK.Google Scholar
Gheradi, S. (1862). Sul magnetismo polare de palazzi ed altri edifizi in Torino. Il Nuovo Cimento 16, 384–404.Google Scholar
Gifford, D. P. (1978). Ethnoarchaeological observations of natural processes affecting cultural materials. In: Explorations in Ethnoarchaeology (ed. Gould, R. A.), pp. 77–101. University of New Mexico Press, Albuquerque.Google Scholar
Gifford-Gonzalez, D. P. (1998). Gender and early pastoralists in East Africa. In: Gender in African Prehistory (ed. Kent, S.), pp. 115–137. Altamira Press, Walnut Creek, CA.Google Scholar
Gifford-Gonzalez, D. P., Damrosch, D. B., Damrosch, D. R., Pryor, J., and Thunen, R. L. (1985). The third dimension in site structure: an experiment in trampling and vertical dispersal. American Antiquity 50, 803–818.CrossRefGoogle Scholar
Gilbert, M., Hansen, A., Willerslev, E., Rudbeck, L., Barnes, I., Lynnerup, N., and Cooper, A. (2003a). Characterization of genetic miscoding lesions caused by postmortem damage. American Journal of Human Genetics 72, 48–61.CrossRefGoogle ScholarPubMed
Gilbert, M. T. P., Willerslev, E., Hansen, A. J., Barnes, I., Rudbeck, L., Lynnerup, N., and Cooper, A. (2003b). Distribution patterns of postmortem damage in human mitochondrial DNA. American Journal of Human Genetics 72, 32–47.CrossRefGoogle ScholarPubMed
Giraud-Guille, M. M. (1988). Twisted plywood architecture of collagen fibrils in human compact bone osteons. Calcified Tissue International 42, 167–180.CrossRefGoogle ScholarPubMed
Glaser, B., Haumaier, L., Guggenberger, G., and Zech, W. (1998). Black carbon in soils: the use of benzenecarboxylic acids as specific markers. Organic Geochemistry 29, 811–819.CrossRefGoogle Scholar
Glaser, B., Balashov, E., Haumaier, L., Guggenberger, G., and Zech, W. (2000). Black carbon in density fractions of anthropogenic soils of the Brazilian Amazon region. Organic Geochemistry 31, 669–678.CrossRefGoogle Scholar
Glob, P. V. (1965). The Bog People: Iron-Age Man Preserved. New York Review of Books, New York.Google Scholar
Goffer, Z. (2007). Archaeological Chemistry. 2nd edition. John Wiley, Hoboken, NJ.CrossRefGoogle Scholar
Goldberg, P. (2001). Some micromorphological aspects of prehistoric cave deposits. Cahiers d'archéologie du CELAT 10, 161–175.Google Scholar
Goldberg, P., and MacPhail, R. I. (2006). Practical and Theoretical Geoarchaeology. Blackwell, Malden, MA.Google Scholar
Goldberg, P., Weiner, S., Bar-Yosef, O., Xu, Q., and Liu, J. (2001). Site formation processes at Zhoukoudian. Journal Human Evolution 41, 483–530.CrossRefGoogle ScholarPubMed
Goldberg, P., Laville, H., Meignen, L., and Bar-Yosef, O. (2007). Stratigraphy and geoarchaeological history of Kebara Cave, Mount Carmel. In: Kebara Cave, Mt. Carmel, Israel: The Middle and Upper Paleolithic Archaeology, Part 1 (ed. Bar-Yosef, O. and Meignen, L.), pp. 49–89. Peabody Museum of Archaeology and Ethnology, Harvard University, Cambridge, MA.Google Scholar
Goodfriend, G. A. (1987). Evaluation of amino acid racemization/epimerization dating using radiocarbon-dated fossil land snails. Geology 15, 698–700.2.0.CO;2>CrossRefGoogle Scholar
Goodfriend, G. A. (1992). Rapid racemization of aspartic acid in mollusc shells and potential for dating over recent centuries. Nature 357, 399–401.CrossRefGoogle Scholar
Goodfriend, G. A., and Ellis, G. L. (2000). Stable carbon isotopic record of middle to late Holocene climate changes from land snail shells at Hinds Cave, Texas. Quaternary International 67, 47–60.CrossRefGoogle Scholar
Goodfriend, G. A., Collins, M. J., Fogel, M. L., Macko, S. A., and Wehmiller, J. F. (2000). Perspectives in Amino Acid and Protein Geochemistry. Oxford University Press, New York.Google Scholar
Gorecki, P. (1985). Ethnoarchaeology: the need for a post-mortem enquiry. World Archaeology 17, 175–191.CrossRefGoogle Scholar
Goren, Y., and Goldberg, P. (1991). Petrographic thin sections and the development of Neolithic plaster production in northern Israel. Journal of Field Archaeology 18, 131–138.Google Scholar
Goren, Y., Finkelstein, I., and Na'aman, N. (2004). Inscribed in Clay: Provenance Study of the Amarna Tablets and Other Ancient Near Eastern Texts. Emery and Claire Yass Publications in Archaeology, Tel Aviv.
Goren-Inbar, N., Werker, E., and Feibel, C. (2002). The Acheulian Site of Gesher Benot Yaakov, Israel: The Wood Assemblage. Oxbow Books, Oxford.Google Scholar
Goren-Inbar, N., Alperson, N., Kislev, M. E., Simchoni, O., Melamed, Y., Ben-Nun, A., and Werker, E. (2004). Evidence of hominin control of fire at Gesher Benot Ya'aqov, Israel. Science 304, 725–727.CrossRefGoogle ScholarPubMed
Gosselain, O. P. (1992). Bonfire of the enquiries – pottery firing temperatures in archaeology: what for?Journal of Archaeological Science 19, 243–259.CrossRefGoogle Scholar
Gotliv, B., Robach, J. A., and Veis, A. (2006). The composition and structure of bovine peritubular dentin: mapping by time of flight secondary ion mass spectrometry. Journal of Structural Biology 156, 320–333.CrossRefGoogle Scholar
Gould, R. A. (1980). Living Archaeology. Cambridge University Press, Cambridge.Google Scholar
Gourdin, W., and Kingery, W. (1975). The beginnings of pyrotechnology: Neolithic and Egyptian lime plaster. Journal of Field Archaeology 2, 133–150.Google Scholar
Green, R. E., Krause, J., Ptak, S. E., Briggs, A. W., Ronan, M. T., Simons, J. F., Du, L., et al. (2006). Analysis of one million base pairs of Neanderthal DNA. Nature 444, 330–336.CrossRefGoogle ScholarPubMed
Greenfield, H. J., Fowler, K. D., and van Schalkwyk, L. O. (2005). Where are the gardens? Early Iron Age horticulture in the Thukela River Basin of South Africa. World Archaeology 37, 307–328.CrossRefGoogle Scholar
Greensmith, J. T. (1963). Clastic quartz, provenance and sedimentation. Nature 197, 345–347.CrossRefGoogle Scholar
Grögler, N., Geiss, J., Grünenfelder, M., and Houtermans, F. G. (1966). Isotopenuntersuchungen zur Bestimmung der Herkunft römischer Bleirohre und Bleibarren. Zeitschrift für Naturforschung 21a, 1167–1172.Google Scholar
Grün, R. (1989). Electron spin resonance (ESR) dating. Quaternary International 1, 65–109.CrossRefGoogle Scholar
Grün, R. (2001). Trapped charge dating (ESR, TL, OSL). In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 47–62. John Wiley, Chichester, UK.Google Scholar
Grün, R., Stringer, C., McDermott, F., Nathane, R., Porat, N., Robertson, S., Taylor, L., Mortimer, G., Eggins, S., and McCulloch, M. (2005). U-series and ESR analyses of bones and teeth relating to the human burials from Skhul. Journal of Human Evolution 49, 316–334.CrossRefGoogle ScholarPubMed
Grupe, G., and Hummel, S. (1991). Trace element studies on experimentally cremated bone. I. Alteration of the chemical composition at high temperatures. Journal of Archaeological Science 18, 177–186.CrossRefGoogle Scholar
Gueta, R., Natan, A., Addadi, L., Weiner, S., Refson, K., and Kronik, L. (2007). Local atomic order and infrared spectra of biogenic calcite. Angewandte Chemie International Edition 46, 291–294.CrossRefGoogle ScholarPubMed
Gujer, W., and Zehnder, A. J. B. (1983). Conversion processes in anaerobic digestion. Water Science and Technology 15, 127–167.CrossRefGoogle Scholar
Haberle, S. G., and Ledru, M.-P. (2001). Correlations among charcoal records of fires from the past 16,000 years in Indonesia, Papua New Guinea and central South America. Quaternary Research 55, 97–104.CrossRefGoogle Scholar
Hackett, C. J. (1981). Microscopical focal destruction (tunnels) in exhumed human bones. Medical Science and Law 21, 243–265.CrossRefGoogle ScholarPubMed
Hagelberg, E., Sykes, B. C., and Hedges, R. E. M. (1989). Ancient bone DNA amplified. Nature342, 485.Google ScholarPubMed
Hardy, K., Blakeney, T., Copeland, L., Kirkham, J., Wrangham, R., and Collins, M. J. (2009). Starch granules, dental calculus and new perspectives on ancient diet. Journal of Archaeological Science 36, 248–255.CrossRefGoogle Scholar
Hare, P. E. (1980). Organic geochemistry of bone and its relation to the survival of bone in the natural environment. In: Fossils in the Making: Vertebrate Taphonomy and Paleoecology (ed. Behrensmeyer, A. K. and Hill, A. P.), pp. 208–219. University of Chicago Press, Chicago.Google Scholar
Hare, P. E., and Abelson, P. H. (1965). Amino acid composition of some calcified proteins. Carnegie Institution of Washington Yearbook 64, 223–234.Google Scholar
Hare, P. E., and Abelson, P. H. (1967). Racemization of amino acids in fossil shells. Carnegie Institution of Washington Yearbook 66, 526–528.Google Scholar
Hare, P. E., and Estep, M. L. F. (1983). Carbon and nitrogen isotopic composition of amino acids in modern and fossil collagens. Carnegie Institution of Washington Yearbook 82, 410–414.Google Scholar
Hare, P. E., Hoering, T. C., and King, K. (1980). Biogeochemistry of Amino Acids. John Wiley, New York.Google Scholar
Hare, P. E., Fogel, M. L., Stafford, T. W., Mitchell, A. D., and Hoering, T. C. (1991). The isotopic composition of carbon and nitrogen in individual amino acids isolated from modern and fossil proteins. Journal of Archaeological Science 18, 277–292.CrossRefGoogle Scholar
Harrison, C. C. (1996). Evidence for intramineral macromolecules containing protein from plant silicas. Phytochemistry 41, 37–42.CrossRefGoogle ScholarPubMed
Harrison, C. C., and Lu, Y. (1994). In vivo and in vitro studies of polymer controlled silicification. Bulletin de l'Institut océanographique, Monaco 14, 151–158.Google Scholar
Harry, K. G., and Johnson, A. (2004). A non-destructive technique for measuring ceramic porosity using liquid nitrogen. Journal of Archaeological Science 31, 1567–1575.CrossRefGoogle Scholar
Hart, J. P., and Matson, R. G. (2008). The use of multiple discriminant analysis in classifying prehistoric phytolith assemblages recovered from cooking residues. Journal of Archaeological Science 36, 430–433.Google Scholar
Haslam, M. (2004). The decomposition of starch grains in soils: implications for archaeological residue analyses. Journal of Archaeological Science 31, 1715–1734.CrossRefGoogle Scholar
Hata, T. Y. O., Kobayashi, E., Yamane, K., and Kikuchi, K. (2000). Onion-like graphitic particles observed in wood charcoal. Journal of Wood Science 46, 89–92.CrossRefGoogle Scholar
Hauschka, P. V., Lian, J. B., and Gallop, P. M. (1975). Direct identification of the calcium-binding amino acid, γ-carboxyglutamate in mineralized tissue. Proceedings of the National Academy of Sciences of the United States of America 72, 3925–3929.CrossRefGoogle ScholarPubMed
Hayes, M. H. B. (1998). Humic substances: progress towards more realistic concepts of structures. In: Humic Substances: Structures, Properties and Uses (ed. Davies, G., Ghabbour, E. A., and Khairy, K. A.), pp. 1–27. Royal Society of Chemistry, Cambridge.Google Scholar
Hedges, R. E. M. (2001). Overview – dating in archaeology; past, present and future. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 3–8. John Wiley, Chichester, UK.Google Scholar
Hedges, R. E. M. (2002). Bone diagenesis: an overview of processes. Archaeometry 44, 319–328.CrossRefGoogle Scholar
Hedges, R. E. M., Thorp, J. A., and Tuross, N. C. (1995). Is tooth enamel carbonate a suitable material for radiocarbon dating?Radiocarbon 37, 285–290.CrossRefGoogle Scholar
Heimann, R. B., and Maggetti, M. (1981). Experiments on simulated burial of calcareous terra sigillata (mineralogical change): preliminary results. British Museum Occasional Paper 19, 163–177.Google Scholar
Hein, A., Kilikoglou, V., and Kassianidou, V. (2007). Chemical and mineralogical examination of metallurgical ceramics from a Late Bronze Age copper smelting site in Cyprus. Journal of Archaeological Science 34, 141–154.CrossRefGoogle Scholar
Heinemeier, J., Jungner, H., Lindroos, A., Ringbom, A., Konow, T., and Rud, N. (1997). AMS C-14 dating of lime mortar. Nuclear Instruments and Methods in Physics Research, Section B 123, 487–495.CrossRefGoogle Scholar
Heinrich, E. W. (1965). Microscopic Identification of Minerals. McGraw-Hill, New York.Google Scholar
Henderson, J. (2000). The Science and Archaeology of Materials: An Investigation of Inorganic Materials. Routledge, London.Google Scholar
Henderson, J. (2001). Glass and glaze. In: Handbook of Archaeological Science (ed. Brothwell, D. R. and Pollard, A. M.), pp. 471–482. John Wiley, Chichester, UK.Google Scholar
Henry, A. G., and Piperno, D. R. (2008). Using plant microfossils from dental calculus to recover human diet: a case study from Tell al-Raqa'i, Syria. Journal of Archaeological Science 35, 1943–1950.CrossRefGoogle Scholar
Heron, C., and Evershed, R. P. (1993). The analysis of organic residues and the study of pottery use. In: Archaeological Method and Theory, vol. 5 (ed. Schiffer, M. B.), pp. 247–284. University of Arizona Press, Tucson.Google Scholar
Heron, C., and Pollard, A. M. (1988). The analysis of natural resinous materials from Roman amphoras. In: Science and Archaeology, Glasgow 1987: Proceedings of a Conference on the Application of Scientific Methods to Archaeology, vol. 196 (ed. Slater, E. A. and Tate, J. O.), pp. 429–447. British Archaeological Reports, British Series, Oxford.Google Scholar
Herron, C. (2001). Geochemical prospecting. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 565–573. John Wiley, Chichester, UK.Google Scholar
Herz, N. (1992). Provenance determination of Neolithic to classical Mediterranean marbles by stable isotopes. Archaeometry 34, 185–194.CrossRefGoogle Scholar
Hess, D., Coker, D. J., Loutsch, J. M., and Russ, J. (2008). Production of oxalates in vitro by microbes isolated from rock surfaces with prehistoric paints in the Lower Pecos region, Texas. Geoarchaeology 23, 3–11.CrossRefGoogle Scholar
Higham, T. F. G., and Horn, P. L. (2000). Seasonal dating using fish otoliths: results from the Shag River Mouth site, New Zealand. Journal of Archaeological Science 27, 439–448.CrossRefGoogle Scholar
Higuchi, R., Bowman, B., Freiberger, M., Ryder, O. A., and Wilson, A. C. (1984). DNA sequences from the quagga, an extinct member of the horse family. Nature 312, 282–284.CrossRefGoogle ScholarPubMed
Hill, C., and Forti, P. (1997). Cave Minerals of the World. 2nd edition. National Speleological Society, Huntsville, AL.
Hiller, C. R., Robinson, C., and Weatherell, J. A. (1975). Variation in the composition of developing rat incisor enamel. Calcified Tissue Research 18, 1–12.CrossRefGoogle Scholar
Hillman, G. (1996). Late Pleistocene changes in wild plant-foods available to hunter-gatherers of the northern Fertile Crescent: possible preludes to cereal cultivation. In: The Origins and Spread of Agriculture and Pastoralism in Eurasia (ed. Harris, D. R.), pp. 159–203. UCL Press, London.Google Scholar
Hillson, S. (1986). Teeth. Cambridge University Press, Cambridge.Google ScholarPubMed
Hodder, I. (1999). The Archaeological Process. Blackwell, Oxford.Google Scholar
Hodge, A. J., and Petruska, J. A. (1963). Recent studies with the electron microscope on ordered aggregates of the tropocollagen molecule. In: Aspects of Protein Structure (ed. Ramachandran, G. N.), pp. 289–300. Academic Press, New York.Google Scholar
Hodson, M. J., White, P. J., Mead, A., and Broadley, M. R. (2005). Phylogenetic variation in the silicon composition of plants. Annals of Botany 96, 1027–1046.CrossRefGoogle Scholar
Holden, J. L., Phakey, P. P., and Clement, J. G. (1995). Scanning electron microscope observations of incinerated human femoral bone: a case study. Forensic Science International 74, 29–45.CrossRefGoogle ScholarPubMed
Holliday, V. T. (2004). Soils in Archaeological Research. Oxford University Press, New York.Google Scholar
Hood, M. A., and Meyers, S. P. (1977). Rates of chitin degradation in an estuarine environment. Journal of Oceanography 33, 328–334.CrossRefGoogle Scholar
Hughen, K., Lehman, S., Southon, J., Overpeck, J., Marchal, O., Herring, C., and Turnbull, J. (2004). 14C activity and global carbon cycle changes over the past 50,000 years. Science 303, 202–207.CrossRefGoogle ScholarPubMed
Hull, K. L. (1987). Identification of cultural site formation processes through microdebitage analysis. Antiquity 52, 772–783.CrossRefGoogle Scholar
Hull, K. L. (2001). Reasserting the utility of obsidian hydration dating: a temperature-dependent empirical approach to practical temporal resolution with archaeological obsidians. Journal of Archaeological Science 28, 1025–1040.CrossRefGoogle Scholar
Humphrey, L. T., Dirks, W., Dean, M. C., and Jeffries, T. E. (2008). Tracking dietary transitions in weanling baboons (Papio hamadryas anubis) using strontium/calcium ratios in enamel. Folia Primatalogica 79, 197–212.CrossRefGoogle ScholarPubMed
Humphreys, G. S., Hunt, P. A., and Buchanan, R. (1987). Wood-ash stone near Sydney, N.S.W.: a carbonate pedological feature in an acidic soil. Australian Journal of Soil Research 25, 115–124.CrossRefGoogle Scholar
Huntley, D. J., Godfrey-Smith, D. I., and Thewalt, M. L. W. (1985). Optical dating of sediments. Nature 313, 105–107.CrossRefGoogle Scholar
Hustedt, F. (1942). Aerophile Diatomeen in der nordwestdeutschen Flora. Berigte der Deutschen Botanischen Gesellschaft 60, 55–73.Google Scholar
Hutchinson, G. (1950). Survey of contemporary knowledge of biogeochemistry. 3. The biogeochemistry of vertebrate excretion. Bulletin of the American Museum of Natural History 96, 1–554.Google Scholar
Hutson, S. R., Stanton, T. W., Magnoni, A., Terry, R., and Craner, J. (2007). Beyond the buildings: formation processes of ancient Maya houselots and methods for the study of non-architectural space. Journal of Anthropological Archaeology 26, 442–473.CrossRefGoogle Scholar
Huxley, T. H. (1863). Essays: On Our Knowledge of the Causes of the Phenomena of Organic Nature. J. A. Mays, London.
Ikeya, M. (1978). Electron spin resonance as a method of dating. Archaeometry 33, 153–199.Google Scholar
Inbar, Y., and Chen, Y. (1990). Humic substances formed during the compasting of organic matter. Soil Science Society of America Journal 54, 1316–1323.CrossRefGoogle Scholar
Isaac, G. (1968). Traces of Pleistocene hunters: an East African example. In: Man the Hunter (ed. Lee, R. and DeVore, I.), pp. 253–261. Aldine, Chicago.Google Scholar
Jacobs, Z., Wintle, A. G., Roberts, R. G., and Duller, G. A. T. (2008). Equivalent dose distributions of single grains of quartz at Sibudu, South Africa: context, causes and consequences for optical dating of archaeological deposits. Journal of Archaeological Science 35, 1808–1820.CrossRefGoogle Scholar
James, S. R. (1989). Hominid use of fire in the Lower and Middle Pleistocene. Current Anthropology 30, 1–11.CrossRefGoogle Scholar
Jans, M. M. E., Nielsen-Marsh, C. M., Smith, C. I., Collins, M. J., and Kars, H. (2004). Characterisation of microbial attack on archaeological bone. Journal of Archaeological Science 31, 87–95.CrossRefGoogle Scholar
Jansma, M. J. (1982). Diatom analysis from some prehistoric sites in the coastal area of the Netherlands. Acta Geologica Academiae Scientiarum Hungaricae 25, 229–236.Google Scholar
Jansonius, J., and McGregor, D. C. (1996). Palynology: Principles and Applications. American Association for Palynology Foundation, College Station, TX.Google Scholar
Jodaikin, A., Weiner, S., and Traub, W. (1984). Enamel rod relations in the developing rat incisor. Journal of Ultrastructure Research 89, 324–332.CrossRefGoogle ScholarPubMed
Johnson, B. J., and Miller, G. H. (1997). Archaeological applications of amino acid racemization. Archaeometry 39, 265–287.CrossRefGoogle Scholar
Johnson, B. J., Miller, G. H., Fogel, M. L., and Beaumont, P. B. (1997). The determination of late Quaternary paleoenvironments at Equus cave, South Africa, using stable isotopes and amino acid racemization in ostrich eggshell. Palaeogeography, Palaeoclimatology, Palaeoecology 136, 121–137.CrossRefGoogle Scholar
Johnson, B. J., Vogel, M. L., and Miller, G. H. (1998). Stable isotopes in modern ostrich eggshell: a calibration for modern paleoenvironmental applications in semi-arid regions of South Africa. Geochimica et Cosmochimica Acta 62, 2451–2461.CrossRefGoogle Scholar
Jones, D., and Wilson, M. J. (1986). Biomineralization in crustose lichens. In: Biomineralization in Lower Plants and Animals (ed. Leadbeater, B. S. C. and Riding, R.), pp. 91–105. Clarendon Press, Oxford.Google Scholar
Jones, J. D., and Vallentyne, J. R. (1960). Biogeochemistry of organic matter – I. Polypeptides and amino acids in fossils and sediments in relation to geothermometry. Geochimica et Cosmochimica Acta 21, 1–34.CrossRefGoogle Scholar
Jones, M. K., and Colledge, S. (2001). Archaeobotany and the transition to agriculture. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 393–401. John Wiley, Chichester, UK.Google Scholar
Jones, R. L., and Beavers, A. H. (1963). Some mineralogical and chemical properties of plant opal. Soil Science 96, 375–379.CrossRefGoogle Scholar
Jowsey, J. (1961). Age changes in human bone. Clinical Orthopaedics 17, 210–218.Google Scholar
Kalish, J. M. (1991). Determinants of otolith chemistry: seasonal variation in the composition of blood plasma, endolymph and otoliths of bearded rock cod Pseudophycis barbatus. Marine Ecology Progress Series 74, 137–159.CrossRefGoogle Scholar
Kandel, A. W., and Conard, N. J. (2005). Production sequences of ostrich eggshell beads and settlement dynamics in the Geelbek Dunes of the Western Cape, South Africa. Journal of Archaeological Science 32, 1711–1721.CrossRefGoogle Scholar
Karasik, A., and Smilansky, U. (2008). 3D scanning technology as a standard archaeological tool for pottery analysis: practice and theory. Journal of Archaeological Science 35, 1148–1168.CrossRefGoogle 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 22, 775–796.CrossRefGoogle Scholar
Karkanas, P., Bar-Yosef, O., Goldberg, P., and Weiner, S. (2000). Diagenesis in prehistoric caves: the use of minerals that form in situ to assess the completeness of the archaeological record. Journal of Archaeological Science 27, 915–929.CrossRefGoogle Scholar
Karkanas, P., 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 29, 721–732.CrossRefGoogle Scholar
Karkanas, P., Koumouzelis, M., Kozolowski, J. K., Sitlivy, V., Sobczyk, K., Berna, F., and Weiner, S. (2004). The earliest evidence for clay hearths: Aurignacian features in Klisoura Cave 1, southern Greece. Antiquity 78, 513–525.CrossRefGoogle Scholar
Karkanas, P., Shahack-Gross, R., Ayalon, A., Bar-Matthews, M., Barkai, R., Frumkin, A., Gopher, A., and Stiner, M. C. (2007). Evidence for habitual use of fire at the end of the Lower Paleolithic: site formation processes at Qesem Cave, Israel. Journal of Human Evolution 53, 197–212.CrossRefGoogle ScholarPubMed
Kastner, M., Keene, J. B., and Gieskes, J. M. (1977). Diagenesis of siliceous oozes – I. Chemical controls on the rate of opal-A to opal-CT transformation – an experimental study. Geochimica et Cosmochimica Acta 41, 1041–1951.CrossRefGoogle Scholar
Katz, E. P., and Li, S. (1973). Structure and function of bone collagen fibrils. Journal of Molecular Biology 80, 1–15.CrossRefGoogle ScholarPubMed
Katzenberg, M. A. (1993). Age differences and population variation in stable isotope variations from Ontario, Canada. In: Prehistoric Human Bone: Archaeology at the Molecular Level (ed. Lambert, J. B. and Grupe, G.), pp. 39–62. Springer, Berlin.CrossRefGoogle Scholar
Katzenburg, M. A., and Harrison, R. G. (1997). What's in a bone? Recent advances in archaeological bone chemistry. Journal of Archaeological Science 5, 265–293.CrossRefGoogle Scholar
Kaufman, A., Broecker, W. S., Ku, T. L., and Thurber, D. L. (1971). The status of U-series methods of mollusc dating. Geochimica et Cosmochimica Acta 35, 1155–1183.CrossRefGoogle Scholar
Kelly, E. F., Amundson, R. G., Marino, B. D., and DeNiro, M. J. (1991). Stable isotope ratios in carbon in phytoliths as a quantitative method of monitoring vegetation and climate change. Quaternary Research 35, 222–233.CrossRefGoogle Scholar
Kennish, M. J., and Olsson, R. K. (1975). Effects of thermal discharges on the microstructural growth of Mercenaria mercenaria. Environmental Geology 1, 41–64.CrossRefGoogle Scholar
Kenyon, K. M. (1960). Archaeology in the Holy Land. Ernest Benn, London.Google Scholar
Kidwell, S. M., and Holland, S. M. (2002). The quality of the fossil record: implications for evolutionary analyses. Annual Review of Ecological Systematics 33, 561–588.CrossRefGoogle Scholar
Kingery, W. D., Bowen, H. K., and Uhlmann, D. R. (1960). Introduction to Ceramics. John Wiley, New York.Google Scholar
Kingery, W. D., and Francl, J. (1954). Fundamental study of clay: XIII, drying behavior and plastic properties. Journal of the American Ceramic Society 37, 596–602.CrossRefGoogle Scholar
Kingery, W. D., Vandiver, P. B., and Prickett, M. (1988). The beginnings of pyrotechnology, Part II: production and use of lime and gypsum plaster in the Pre-Pottery Neolithic Near East. Journal of Field Archaeology 15, 219–244.CrossRefGoogle Scholar
Kirchmann, H., and Witter, E. (1992). Composition of fresh, aerobic and anaerobic farm animal dungs. Bioresource Technology 40, 137–142.CrossRefGoogle Scholar
Kitano, Y., and Hood, D. W. (1962). Calcium carbonate crystal forms formed from sea water by inorganic processes. Journal of the Oceanographic Society of Japan 18, 141–145.CrossRefGoogle Scholar
Kittrick, J. (1977). Mineral equlibria in the soil system. In: Minerals in Soil Environments (ed. Dixon, J. and Weed, S.), pp. 1–25. Soil Science Society of America, Madison, WI.Google Scholar
Klein, R. G., Kruz-Uribe, K., Halkett, D., Hart, T., and Parkington, J. E. (1999). Paleoenvironmental and human behavioral implications of the Boegoeberg 1 Late Pleistocene hyena den, northern Cape Province, South Africa. Quaternary Research 52, 393–403.CrossRefGoogle Scholar
Knauth, L. P. (1994). Silica: physical behavior, geochemistry and materials applications. Reviews in Mineralogy 29, 233–258.Google Scholar
Knudson, K. J., Frink, L., Hoffman, B., and Price, T. D. (2004). Chemical characterization of Arctic soils: activity area analysis in contemporary Yup'ik fish camps using ICP-AES. Journal of Archaeological Science 31, 443–456.CrossRefGoogle Scholar
Koehl, M. A. R. (1982). Mechanical design of spicule-reinforced connective tissue: stiffness. Journal of Experimental Biology 98, 239–267.Google Scholar
Koon, H. E. C., Nicholson, R. A., and Collins, M. J. (2003). A practical approach to the identification of low temperature heated bone using TEM. Journal of Archaeological Science 30, 1393–1399.CrossRefGoogle Scholar
Koren, Z. C. (2005). The first optimal all-Murex all-natural purple dyeing in the eastern Mediterranean in a millennium and a half. In: Dyes in History and Archaeology, 20, 136–149. Archetype, London.
Kramer, C. (1982). Village Ethnoarchaeology. Academic Press, London.Google Scholar
Kramer, I. R. H. (1951). The distribution of collagen fibrils in the dentine matrix. British Dental Journal 91, 1–7.Google ScholarPubMed
Kriasakul, M., and Mitterer, R. M. (1978). Isoleucine epimerization in peptides and proteins: kinetic factors and applications to fossil proteins. Science 201, 1011–1014.CrossRefGoogle Scholar
Kronick, P. L., and Cooke, P. (1996). Thermal stabilization of collagen fibers by calcification. Connective Tissue Research 33, 275–282.CrossRefGoogle ScholarPubMed
Kühl, N., and Litt, T. (2003). Quantitative time series reconstruction of Eemian temperature at three European sites using pollen data. Vegetation History and Archaeobotany 12, 205–214.CrossRefGoogle Scholar
Labeyrie, J., and Delibrias, G. (1964). Dating of old mortars by the carbon-14 method. Nature, 742.Google Scholar
Lalueza Fox, C., Juan, J., and Albert, R. M. (1996). Phytolith analysis on dental calculus, enamel surface, and burial soil: information about diet and paleoenvironment. American Journal of Physical Anthropology 101, 101–113.Google Scholar
Lambert, J. B. (1997). Traces of the Past: Unraveling the Secrets of Archaeology through Chemistry. Perseus Books, Reading, UK.Google Scholar
Lambert, J. B., Johnson, S. C., and Poinar, G. O. (1995). Resin from Africa and South America: criteria for distinguishing between fossilized and recent resin based on NMR spectroscopy. American Chemical Society Symposium Series 617, 193–202.CrossRefGoogle Scholar
Lammie, D., Bain, M. M., and Wess, T. J. (2005). Microfocus X-ray scattering investigations of eggshell nanotexture. Journal of Synchrotron Radiation 12, 721–726.CrossRefGoogle ScholarPubMed
Lanning, F. C., and Eleuterius, L. N. (1992). Silica and ash in seeds of cultivated grains and native plants. Annals of Botany 69, 151–160.CrossRefGoogle Scholar
Lanting, J. N., and Brindley, A. L. (1998). Dating cremated bone: the dawn of a new era. Journal of Irish Archaeology 9, 1–8.Google Scholar
Lanting, J. N., and Plicht, J. (1998). Reservoir effects and apparent 14C ages. Journal of Irish Archaeology 9, 151–165.Google Scholar
Lanting, J. N., Aerts-Bijma, A., and Plicht, H. (2001). Dating of cremated bones. Radiocarbon 43, 249–254.CrossRefGoogle Scholar
Law, I. A., and Hedges, R. E. M. (1989). A semi-automated bone pretreatment of older and contaminated samples. Bone 31, 247–253.Google Scholar
Laws, R. M. (1952). A new method of age determination for mammals. Nature 169, 972.CrossRefGoogle ScholarPubMed
Lee-Thorp, J. (2002). Two decades of progress towards understanding fossilization processes and isotopic signals in calcified minerals. Archaeometry 44, 435–446.CrossRefGoogle Scholar
Lees, S. (1987). Considerations regarding the structure of the mammalian mineralized osteoid from viewpoint of the generalized packing model. Connective Tissue Research 16, 281–303.CrossRefGoogle ScholarPubMed
Legeros, R. Z. (1991). Calcium Phosphates in Oral Biology and Medicine. Karger, Basel, Switzerland.Google ScholarPubMed
Legeros, R. Z., and Legeros, J. P. (1984). Phosphate minerals in human tissue. In: Phosphate Minerals (ed. Nriagu, J. O. and Moore, P. B.), pp. 351–385. Springer, Berlin.CrossRefGoogle Scholar
Legeros, R. Z., Balmain, N., and Bonel, G. (1987). Age-related changes in mineral of rat and bovine cortical bone. Calcified Tissue International 41, 137–144.CrossRefGoogle Scholar
Lehninger, A. L. (1982). Principles of Biochemistry. Worth, New York.Google Scholar
Lehr, J. R., McClellan, G. H., Smith, J. P., and Frasier, A. W. (1968). Characterization of apatites in commercial phosphate rocks. Coll. Int. Phosphates Mineraux Solides, Toulouse 2, 29–44.Google Scholar
Leny, L., and Casteel, R. W. (1975). Simplified procedure for examining charcoal specimens for identification. Journal of Archaeological Science 2, 153–159.CrossRefGoogle Scholar
Leute, U. (1987). An Introduction to Physical Methods in Archaeology and History of Art. Wiley-VCH, Weinheim, Germany.Google Scholar
Levi-Kalisman, Y., Raz, S., Weiner, S., Addadi, L., and Sagi, I. (2001). XAS study of the structure of a biogenic “amorphous” calcium carbonate phase. Dalton Transactions 21, 3977–3982.Google Scholar
Lev-Yadun, S. (2007). Wood remains from archaeological excavations: a review with a Near Eastern perspective. Israel Journal of Earth Sciences 56, 139–162.CrossRefGoogle Scholar
Lewis, I. C. (1982). Chemistry of carbonization. Carbon 20, 519–529.CrossRefGoogle Scholar
Libby, W. F., Anderson, E. C., and Arnold, J. R. (1949). Age determination by radiocarbon content: world-wide assay of natural radiocarbon. Science 109, 227–228.CrossRefGoogle ScholarPubMed
Lieberman, D. E. (1993). Life history variables preserved in dental cementum microstructure. Science 261, 1162–1164.CrossRefGoogle ScholarPubMed
Lieberman, D. E., and Meadow, R. H. (1992). The biology of cementum increments (with an archaeological application). Mammal Review 22, 57–77.CrossRefGoogle Scholar
Lindroos, A., Heinemeier, J., Ringbom, A., Brasken, M., and Sveinbjornsdottir, A. (2007). Mortar dating using AMS C-14 and sequential dissolution: Examples from medieval, non-hydraulic lime mortars from the Aland Islands, SW Finland. Radiocarbon 49, 47–67.CrossRefGoogle Scholar
Lippmann, F. (1973). Sedimentary Carbonate Minerals. Springer, Berlin.CrossRefGoogle Scholar
Liritzis, I. (1994). A new dating method by thermoluminescence of carved megalithic stone building. Comptes Rendus de l'Académie des sciences 319, 603–610.Google Scholar
Littmann, E. W. (1957). Ancient Mesoamerican mortars, plasters and stuccos: Comalcalco, Part I. American Antiquity 23, 135–140.CrossRefGoogle Scholar
Liu, D., Wagner, H. D., and Weiner, S. (2000). Bending and fracture of compact circumferential and osteonal lamellar bone of the baboon tibia. Journal Material Sciences Material and Medicine 11, 49–60.CrossRefGoogle ScholarPubMed
Liu, D., Weiner, S., and Wagner, H. D. (1999). Anisotropic mechanical properties of lamellar bone using miniature cantilever bending specimens. Journal of Biomechanics 32, 647–654.CrossRefGoogle ScholarPubMed
Livingstone Smith, A. (2001). Bonefire II: the return of pottery firing temperatures. Journal of Archaeological Science 28, 991–1003.CrossRefGoogle Scholar
Longinelli, A. (1965). Oxygen isotopic composition of orthophosphate from shells of living marine organisms. Nature 207, 716.CrossRefGoogle Scholar
Longinelli, A. (1984). Oxygen isotopes in mammal bone phosphate: a new tool for paleohydrological and paleoclimatalogical research?Geochimica et Cosmochimica Acta 48, 385–390.CrossRefGoogle Scholar
Loucaidies, S., Cappellen, P., and Behrends, T. (2008). Dissolution of biogenic silica from land to ocean: role of salinity and pH. Limnology and Oceanography 53, 1614–1621.Google Scholar
Lowenstam, H. A., and Fitch, J. E. (1981). Vaterite formation of higher teleost fishes. Paper presented at the 62nd annual meeting of the Pacific Division of the American Association for the Advancement of Science.
Lowenstam, H. A., and Weiner, S. (1989). On Biomineralization. Oxford University Press, New York.Google Scholar
Lozinski, J. (1973). Rare earth elements in fossil bones. Mineralogia Polonica 2, 29–35.Google Scholar
Lucas, P. W. (2004). Dental Functional Morphology. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Luckenback, A. H., Holland, C. G., and Allen, R. O. (1975). Soapstone artifacts: tracing prehistoric trade patterns in Virginia. Science 187, 57–58.CrossRefGoogle Scholar
Luke, C., Tycott, R. H., and Scott, R. W. (2006). Petrographic and stable isotope analyses of late classic Ulúa marble vases and potential sources. Archaeometry 48, 13–29.CrossRefGoogle Scholar
Lux, A., Luxová, M., Hattori, T., Inanaga, S., and Sugimoto, Y. (2002). Silicification in sorghum (Sorghum bicolor) cultivars with different drought resistance. Physiologia Plantarum 115, 87–92.CrossRefGoogle Scholar
Luxán, M. P., Dorrego, F., and Laborde, A. (1995). Ancient gypsum mortars from St Engracia (Zaragoza, Spain): characterization, identification of additives and treatments. Cement and Concrete Research 25, 1755–1765.CrossRefGoogle Scholar
Luz, B., and Kolodny, Y. (1989). Oxygen isotope variations in bone phosphate. Applied Geochemistry 4, 317–323.CrossRefGoogle Scholar
Luz, B., Kolodny, Y., and Horowitz, M. (1984). Fractionation of oxygen isotopes between mammalian bone-phosphate and environmental drinking water. Geochimica et Cosmochimica Acta 48, 1689–1693.CrossRefGoogle Scholar
Maby, J. C. (1932). The identification of wood and wood charcoal fragments. Analyst 57, 2–8.CrossRefGoogle 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 31, 175–191.CrossRefGoogle Scholar
Madejova, J. (2003). FTIR techniques in clay mineral studies. Vibrational Spectroscopy 31, 1–10.CrossRefGoogle Scholar
Madella, M., Jones, M. K., Goldberg, P., Goren, Y., and Hovers, E. (2002). The exploitation of plant resources by Neanderthals in Amud Cave (Israel): the evidence from phytolith studies. Journal of Archaeological Science 29, 703–719.CrossRefGoogle Scholar
Madella, M., Alexandre, A., and Ball, T. (2005). International code for phytolith nomenclature. Annals of Botany 96, 253–260.CrossRefGoogle ScholarPubMed
Maggetti, M. (1982). Phase analysis and its significance for technology and origin. In: Archaeological Ceramics (ed. Olin, J. S. and Franklin, A. D.), pp. 121–133. Smithsonian Institution Press, Washington, DC.Google Scholar
Mahamid, J., Sharir, A., Addadi, L., and Weiner, S. (2008). Amorphous calcium phosphate is a major component of the forming fin bones of zebrafish: indications for an amorphous precursor phase. Proceedings of the National Academy of Sciences of the United States of America 105, 12748–12753.CrossRefGoogle ScholarPubMed
Makarewicz, C., and Tuross, N. (2006). Foddering by Mongolian pastoralists is recorded in the stable carbon (d13C) and nitrogen (d13N) isotopes of caprine dentinal collagen. Journal of Archaeological Science 33, 862–870.CrossRefGoogle Scholar
Mangerud, J. (1972). Radiocarbon dating of marine shells, including a discussion of apparent age of recent shells from Norway. Boreas 1, 143–172.CrossRefGoogle Scholar
Maniatis, Y., and Tite, M. S. (1981). Technological examination of Neolithic–Bronze Age pottery from central and southeast Europe and from the Near East. Journal of Archaeological Science 8, 59–76.CrossRefGoogle Scholar
Mannion, A. M. (2007). Fossil diatoms and their significance in archaeological research. Oxford Journal of Archaeology 6, 131–147.CrossRefGoogle Scholar
Marín-Arroyo, A. B., Fosse, P., and Vigne, J.-D. (2009). Probable evidences of bone accumulation by Pleistocene bearded vulture at the archaeological site of El Mirón Cave (Spain). Journal of Archaeological Science 36, 284–296.CrossRefGoogle Scholar
Maritan, L., Mazzoli, C., and Freestone, I. (2007). Modeling changes in mollusc shell internal microstructure during firing: implications for temperature estimation in shell-bearing property. Archaeometry 49, 529–541.CrossRefGoogle Scholar
Marshall, G. W., Habelitz, S., Gallagher, R., Balooch, M., Balooch, G., and Marshall, S. J. (2001). Nanomechanical properties of hydrated carious human dentin. Journal of Dental Research 80, 1768–1771.CrossRefGoogle ScholarPubMed
Marxen, J. C., Becker, W., Finke, D., Hasse, B., and Epple, M. (2003). Early mineralization in Biomphalaria glabrata: microscopic and structural results. Journal of Molluscan Studies 69, 113–121.CrossRefGoogle Scholar
Mason, B. (1958). Principles of Geochemistry. 3rd edition. John Wiley, New York.Google Scholar
Mata, M. P., Peacor, D. L., and Gallart-Marti, M. D. (2002). Transmission electron microscopy (TEM) applied to ancient pottery. Archaeometry 44, 155–176.CrossRefGoogle Scholar
Matthews, W., French, C., Lawrence, T., and Cutler, D. (1996). Multiple surfaces: the micromorphology. In: On the Surface: Catalhoyuk 1993–95 (ed. Hodder, I.), pp. 301–342. MacDonald Institute for Research and British Institute for Archaeology, Ankara, Cambridge.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 activities. World Archaeology 29, 281–308.CrossRefGoogle Scholar
Mazar, A., Namdar, D., Panitz-Cohen, N., Neumann, R., and Weiner, S. (2008). Iron Age beehives in Tel Rehov in the Jordan Valley. Antiquity 82, 629–639.CrossRefGoogle Scholar
Mbida, C. M., van Neer, W., Doutrelepont, H., and Vrydaghs, L. (2000). Evidence for banana cultivation and animal husbandry during the first millennium b.c. in the forest of southern Cameroon. Journal of Archaeological Science 27, 151–162.CrossRefGoogle Scholar
McConnell, D. (1952). The crystal chemistry of carbonate apatites and their relation to calcified tissues. Journal of Dental Research 31, 53–63.CrossRefGoogle Scholar
McConnell, D. (1969). The mineralogy of the apatites and their relation to biologic precipitation. Proceedings of the North American Paleontological Convention, 1525–1535.Google Scholar
McDonnell, J. G. (2001). Pyrotechnology. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 493–505. John Wiley, Chichester, UK.Google Scholar
McGeehin, J., Burr, G. S., Jull, A. J. T., Reines, D., Gosse, J., Davis, P. T., Muhs, D., and Southon, J. R. (2001). Stepped combustion 14C dating of sediment: a comparison with established techniques. Radiocarbon 43, 255–261.CrossRefGoogle Scholar
McKinley, J. I., and Bond, J. M. (2001). Cremated bone. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 281–292. John Wiley, Chichester, UK.Google Scholar
McNaughton, S. J., and Tarrants, J. L. (1983). Grass leaf silicification: natural selection for an inducible defense against herbivores. Proceedings of the National Academy of Sciences of the United States of America 80, 790–791.CrossRefGoogle ScholarPubMed
Meldrum, F. C., and Cölfen, H. (2008). Controlling mineral morphologies in biological and synthetic systems. Chemical Reviews 108, 4332–4432.CrossRefGoogle ScholarPubMed
Mendelovici, E. (1997). Comparative study of the effects of thermal and mechanical treatments on the structures of clay minerals. Journal of Thermal Analysis 49, 1385–1397.CrossRefGoogle Scholar
Mercier, N., Valladas, H., Joron, J., Schiegl, S., Bar-Yosef, O., and Weiner, S. (1995). Thermoluminescence dating and the problem of geochemical evolution of sediments. Israel Journal of Chemistry 35, 137–141.CrossRefGoogle Scholar
Mercier, N., Valladas, H., Froget, L., Joron, J.-L., Reyss, J.-L., Weiner, S., Goldberg, P., et al. (2007). Hayonim cave: a TL-based chronology for this Levantine Mousterian sequence. Journal Archaeological Science 24, 1064–1077.CrossRefGoogle Scholar
Meredith, N., Sheriff, M., Setchell, D., and Sivanson, S. (1996). Measurements of the microhardness and Young's modulus of human enamel and dentin using an indentation technique. Archives of Oral Biology 41, 539–545.CrossRefGoogle ScholarPubMed
Michel, V., Ildefonse, P., and Morin, G. (1996). Assessment of archaeological bone and dentine preservation from Lazaret Cave (Middle Pleistocene) in France. Palaeogeography, Palaeoclimatology, Palaeoecology 126, 109–119.CrossRefGoogle Scholar
Middleton, J. (1844). On fluorine in bones, its source and its application to the determination of bone age. Proceedings of the Geological Society of London 4, 431–433.Google Scholar
Millard, A. R. (2000). A model for the effect of weaning on nitrogen isotope ratios in humans. In: Perspectives in Amino Acid and Protein Geochemistry (ed. Goodfriend, G. A., Collins, M. J., Fogel, M. L., Macko, S. A., and Wehmiller, J. F.), pp. 51–59. Oxford University Press, New York.Google Scholar
Miller, G. H., Beaumont, P. B., Jull, A. J. T., and Johnson, B. (1992). Pleistocene geochronology and paleothermometry from protein diagenesis in ostrich eggshells: implications for the evolution of modern humans. Philosophical Transactions of the Royal Society of London, Series B 337, 149–157.CrossRefGoogle Scholar
Miller, J. (1954). The microradiographic appearance of dentin. British Dental Journal 97, 7–9.Google Scholar
Miller, N. F. (1996). Seed eaters of the ancient Near East: human or herbivore?Current Anthropology 37, 521–528.CrossRefGoogle Scholar
Miller Rosen, A., and Weiner, S. (1994). Identifying ancient irrigation: a new method using opaline phytoliths from emmer wheat. Journal of Archaeological Science 21, 125–132.CrossRefGoogle Scholar
Milliman, J. D. (1974). Marine Carbonates. Springer, New York.Google Scholar
Mills, M. G. L., and Mills, M. E. J. (1977). An analysis of bones collected at hyaena breeding dens in the Gemsbok National Parks (Mammalia: Carnivora). Annals of the Transvaal Museum 30, 145–155.Google Scholar
Minnis, P. E. (1981). Seeds in archaeological sites: sources and some interpretive problems. American Antiquity 46, 143–152.CrossRefGoogle Scholar
Mithen, S., Jenkins, E., Jamjoum, K., Niumat, S., Nortcliff, S., and Finlayson, B. (2008). Experimental crop growing in Jordan to develop methodology for the identification of ancient crop irrigation. World Archaeology 40, 7–25.CrossRefGoogle Scholar
Moffat, D., and Butler, R. J. (1986). Rare earth element distribution patterns in Shetland steatite – consequences for artifact provenancing studies. Archaeometry 28, 101–115.CrossRefGoogle Scholar
Moore, P. B. (1984). Crystallochemical aspects of the phosphate minerals. In: Phosphate Minerals (ed. Nriagu, J. O. and Moore, P. B.), pp. 155–170. Springer, Berlin.CrossRefGoogle Scholar
Morariu, V. V., Bogdan, M., and Ardelean, I. (1977). Ancient pottery: its pore structure. Archaeometry 19, 187–192.CrossRefGoogle Scholar
Morgenstein, M., and Redmount, C. A. (2005). Using portable energy dispersive X-ray fluorescence (EDXRF) for on-site analysis of ceramic sherds at El Hibeh, Egypt. Journal of Archaeological Science 32, 1613–1623.CrossRefGoogle Scholar
Moropoulou, A., Bakolas, A., and Bisbikou, K. (2000). Investigation of the technology of historic mortars. Journal of Cultural Heritage 1, 45–58.CrossRefGoogle Scholar
Morris, R. W., and Kittleman, L. R. (1967). Piezoelectric properties of otoliths. Science 158, 368–370.CrossRefGoogle Scholar
Müller, W., Fricke, H., Halliday, A. N., McCulloch, M. T., and Wartho, J. (2003). Origin and migration of the Alpine Iceman. Science 302, 862–866.CrossRefGoogle ScholarPubMed
Murayama, E., Herbomel, P., Kayakami, A., Takeda, H., and Nagasawa, H. (2005). Otolith matrix proteins OMP-1 and Otolin-1 are necessary for normal otolith growth and their anchoring onto the sensory maculae. Mechanisms of Development 122, 791–803.CrossRefGoogle ScholarPubMed
Namdar, D., Neumann, R., Sladezki, Y., Haddad, N., and Weiner, S. (2007). Alkane composition variations between darker and lighter colored comb beeswax. Apidologie 38, 453–461.CrossRefGoogle Scholar
Namdar, D., Neumann, R., Goren, Y., and Weiner, S. (2009). The contents of unusual cone-shaped vessels (cornets) from the Chalcolithic of the Southern Levant. Journal of Archaeological Science 36, 629–636.CrossRefGoogle Scholar
Nanci, A. (2003). Ten Cate's Oral Histology. 6th edition. Mosby, St. Louis, MO.Google Scholar
Nash, S. E. (1999). Time, Trees, and Prehistory. University of Utah Press, Salt Lake City.Google Scholar
Navazo, M., Colina, A., Domínguez-Bella, S., and Benito-Calvo, A. (2008). Raw stone material supply for Upper Pleistocene settlements in Sierra de Atapuerca (Buergos, Spain): flint characterization using petrographic and geochemical techniques. Journal of Archaeological Science 35, 1961–1973.CrossRefGoogle Scholar
Nesse, W. D. (1991). Introduction to Optical Mineralogy. Oxford University Press, New York.Google Scholar
Newesely, H. (1989). Fossil bone apatite. Applied Geochemistry 4, 233–245.CrossRefGoogle Scholar
Nichols, G. J., Cripps, J. A., Collinson, M. E., and Scott, A. C. (2000). Experiments in waterlogging and sedimentology of charcoal: results and implications. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 43–56.CrossRefGoogle Scholar
Nicholson, R. A. (1993). A morphological investigation of burnt animal bone and an evaluation of its utility in archaeology. Journal of Archaeological Science 20, 411–428.CrossRefGoogle Scholar
Nielsen-Marsh, C. M., and Hedges, R. E. M. (1997). Dissolution experiments on modern and diagenetically altered bone and their effects of the infrared splitting factor. Bulletin de Societe Geologique Francaise 168, 485–490.Google Scholar
Nielsen-Marsh, C. M., and Hedges, R. E. M. (1999). Bone porosity and the use of mercury intrusion porosimetry in bone diagenesis studies. Archaeometry 41, 165–174.CrossRefGoogle Scholar
Nielsen-Marsh, C. M., Hedges, R. E. M., Mann, T., and Collins, M. J. (2000). A preliminary investigation of the application of differential scanning calorimetry to the study of collagen degradation in archaeological bone. Thermochimica Acta 365, 129–139.CrossRefGoogle Scholar
Nielsen-Marsh, C. M., Gandhi, H., Shapiro, B., Cooper, A., Hauschka, P. V., and Collins, M. J. (2002). Sequence preservation of osteocalcin protein and mitochondrial DNA in bison bones older than 55 ka. Geology 30, 1099–1102.2.0.CO;2>CrossRefGoogle Scholar
Nocete, F., Álex, E., Nieto, J. M., Sáez, R., and Bayona, M. R. (2005). An archaeological approach to regional environmental pollution in the south-western Iberian Peninsula related to third millenium BC mining and metallurgy. Journal of Archaeological Science 32, 1566–1576.CrossRefGoogle Scholar
Noonan, J. P., Coop, G., Kudaravalli, S., Smith, D., Krause, J., Alessi, J., Chen, F., et al. (2006). Sequencing and analysis of Neanderthal genomic DNA. Science 314, 1113–1118.CrossRefGoogle ScholarPubMed
Norris, E., Norris, C., and Steen, J. B. (1975). Regulation and grinding ability of grit in the gizzard of Norwegian Willow Ptarmigan (Lagopus lagopus). Poultry Science 54, 1839–1843.CrossRefGoogle Scholar
Norton, D. A., and Ogden, J. (1987). Dendrochronology: a review with emphasis on New Zealand applications. New Zealand Journal of Ecology 10, 77–95.Google Scholar
Nriagu, O. J. (1976). Phosphate-clay mineral relations in soils and sediments. Canadian Journal Earth Sciences 13, 717–736.CrossRefGoogle Scholar
Nriagu, O. J. (1984). Phosphate minerals: their properties and general modes of occurrence. In: Phosphate Minerals (ed. Nriagu, J. O. and Moore, P. B.), pp. 1–136. Springer, Berlin.CrossRefGoogle Scholar
Ntinou, M. (2002). Charcoal analysis at Sarakini: An ethnoarchaeological case study. In: El paisaje en el norte de Grecia desde el Tardiglaciar al Antlántico: Formaciones vegetales, Recursos y Usos. British Archaeological Reports International Series 1038, 115–129.Google Scholar
Nudelman, F., Chen, H. H., Goldberg, H. A., Weiner, S., and Addadi, L. (2007). Lessons from biomineralization: comparing the growth strategies of mollusk shell prismatic and nacreous layers in Atrina rigida. Faraday Discussions 136, 9–25.CrossRefGoogle ScholarPubMed
Oades, J. M. (1989). An introduction to organic matter in mineral soils. In: Minerals in Soil Environments, vol. 1 (ed. Dixon, J. B. and Weed, S. B.), pp. 89–159. Soil Science Society of America, Madison, WI.Google Scholar
Oakley, K. P. (1948). Fluorine and relative dating of bones. Advancement of Science 16, 336–337.Google Scholar
Oakley, K. P. (1970). On man's use of fire, with comments on tool-making and hunting. In: Social Life of Early Man (ed. Washburn, S. L.), pp. 176–193. Aldine, Chicago.Google Scholar
Oberlin, A. (1984). Carbonization and graphitization. Carbon 22, 521–541.CrossRefGoogle Scholar
O'Connell, J. F. (1987). Alyawara site structure and its archaeological implications. American Antiquity 52, 74–108.CrossRefGoogle Scholar
Odriozola, C. P., and Pérez, M. H. (2006). The manufacturing process of 3rd millennium BC bone based incrusted pottery decoration from the Middle Guadiana river basin (Badajoz, Spain). Journal of Archaeological Science 34, 1794–1803.CrossRefGoogle Scholar
Olson, E. A., and Broecker, W. S. (1958). Sample contamination and reliability of radiocarbon dates. Transactions of the New York Academy of Sciences, Series II 20, 593–604.CrossRefGoogle Scholar
Olszta, M. J., Cheng, X., Jee, S. S., Kumar, R., Kim, Y.-Y., Kaufman, M. J., Douglas, E. P., and Gower, L. B. (2007). Bone structure and formation: a new perspective. Material Science and Engineering R 58, 77–116.CrossRefGoogle Scholar
Orme, B. (1974). Twentieth-century prehistorians and the idea of ethnographic parallels. Man 9, 199–212.CrossRefGoogle Scholar
Osawa, E., Hirose, Y., Kimura, A., and Shibuya, M. (1997). Fullerenes in Chinese ink: a correction. Fullerene Science and Technology 5, 177–194.CrossRefGoogle Scholar
Ostwald, W. Z. (1879). Studies on the formation and transformation of solid phases. Physical Chemistry 22, 289–330.Google Scholar
Ostwald, W. Z. (1896). Lehrbruck der Allgemeinen Chemie. Leipzig, Germany.Google Scholar
Ozawa, T. (1965). A new method of analyzing thermogravimetric data. Bulletin of the Chemical Society of Japan 38, 1881–1886.CrossRefGoogle Scholar
Pääbo, S. (1985). Molecular cloning of ancient Egyptian mummy DNA. Nature 314, 644–645.CrossRefGoogle ScholarPubMed
Pääbo, S., Higuchi, R. G., and Wilson, A. C. (1989). Ancient DNA and the polymerase chain reaction: The emerging field of molecular archaeology. Journal of Biological Chemistry 264, 9709–9712.Google ScholarPubMed
Panella, G. (1971). Fish otoliths: daily growth layers and periodical patterns. Science 173, 1124–1126.CrossRefGoogle ScholarPubMed
Panshin, A. J., and DeZeeuw, C. (1970). Textbook of Wood Technology. McGraw-Hill, New York.Google Scholar
Paris, O., Zollfrank, C., and Zickler, G. A. (2005). Decomposition and carbonisation of wood biopolymers – a microstructural study of soft wood pyrolysis. Carbon 43, 53–66.CrossRefGoogle Scholar
Park, R., and Epstein, S. (1961). Metabolic fractionation of C12 & C13 in plants. Plant Physiology 36, 133–138.CrossRefGoogle Scholar
Parker, F. S. (1971). Infrared Spectroscopy in Biochemistry, Biology, and Medicine. Plenum Press, New York.CrossRefGoogle Scholar
Parker, R. B., and Toots, H. (1970). Minor elements in fossil bone. Bulletin of the Geological Society of America 81, 925–932.CrossRefGoogle Scholar
Peacock, D. P. S. (1967). The heavy mineral analysis of pottery: a preliminary report. Archaeometry 10, 97–100.CrossRefGoogle Scholar
Pearsall, D. M. (1978). Phytolith analysis of archaeological soils: evidence for maize cultivation in formative Ecuador. Science 199, 177–178.CrossRefGoogle Scholar
Peregrine, P. N. (2004). Cross-cultural approaches in archaeology: comparative ethnology, comparative archaeology, and archaeoethnology. Journal of Archaeological Research 12, 281–309.CrossRefGoogle Scholar
Perlman, I., and Asaro, F. (1969). Pottery analysis by neutron activation. Archaeometry 11, 21–52.CrossRefGoogle Scholar
Pernicka, E., Begemann, F., Schmitt-Strecker, S., Todorova, H., and Kuleff, I. (1997). Prehistoric copper in Bulgaria: its composition and provenance. Eurasia Antiqua 3, 41–180.Google Scholar
Perry, C. C. (2003). Silicification: the process by which organisms capture and mineralize silica. In: Biomineralization: Reviews in Mineralogy and Geochemistry, vol. 54 (ed. Dove, P. M., DeYoreo, J. J., and Weiner, S.), pp. 291–327. Mineralogical Society of America / Geochemical Society, Washington, DC.Google Scholar
Person, A., Bocherens, H., Saliege, J. F., Paris, F., Zeitoun, V., and Gerard, M. (1995). Early diagenetic evolution of bone phosphate – an X-ray-diffractometry analysis. Journal of Archaeological Science 22, 211–221.CrossRefGoogle Scholar
Picon, M. (1976). Remarques préliminares sur deux types d'altération de la composition chimique des céramiques an cours du temps. Figlina 1, 159–166.Google Scholar
Pike, A. W. G., Hedges, R. E. M., and Calsteren, P. (2002). U-series dating of bone using the diffusion-adsorption model. Geochimica et Cosmochimica Acta 66, 4273–4286.CrossRefGoogle Scholar
Piperno, D. R. (1988). Phytolith Analysis: An Archaeological and Geological Perspective. Academic Press, San Diego, CA.Google Scholar
Piperno, D. R. (2006). Phytoliths: A Comprehensive Guide for Archaeologists and Paleoecologists. Altamira Press, Lanham, MD.Google Scholar
Piperno, D. R., Weiss, E., Holst, I., and Nadel, D. (2004). Processing of wild cereal grains in the Upper Paleolithic revealed by starch grain analysis. Nature 430, 670–673.CrossRefGoogle Scholar
Plummer, T. W., Kinyua, A. M., and Potts, R. (1994). Provenancing of hominid and mammalian fossils from Kanjera, Kenya, using EDXRF. Journal of Archaeological Science 21, 553–563.CrossRefGoogle Scholar
Pobeguin, T. (1943). Les oxalates de calcium chez quelques angiospermes. Annales des Sciences Naturelles Botanique 4, 1–95.Google Scholar
Pokroy, B., Qunitana, J. P., Caspi, E. N., Berner, A., and Zolotoyabko, E. (2004). Anisotropic lattice distortions in biogenic aragonite. Nature Materials 3, 900–902.CrossRefGoogle ScholarPubMed
Politi, Y., Metzler, R. A., Abrecht, M., Gilbert, B., Wilt, F. H., Sagi, I., Addadi, L., Weiner, S., and Gilbert, P. U. P. A. (2008). Transformation mechanism of amorphous calcium carbonate into calcite in the sea urchin larval spicule. Proceedings of the National Academy of Sciences of the United States of America 105, 17362–17366.CrossRefGoogle ScholarPubMed
Pollard, A. M., and Heron, C. (2008). Archaeological Chemistry. Royal Society of Chemistry, Cambridge.Google Scholar
Posner, A. S., Harper, R. A., and Muller, S. A. (1965). Age changes in the crystal chemistry of bone apatite. Annals of the New York Academy of Sciences 131, 737–742.CrossRefGoogle ScholarPubMed
Post, P. W., and Donner, D. D. (2005). Frostbite in a pre-Columbian mummy. American Journal of Physical Anthropology 37, 187–191.CrossRefGoogle Scholar
Prescott, J. R., and Robertson, G. B. (1997). Sediment dating by luminescence: a review. Radiation Measurements 27, 893–922.CrossRefGoogle Scholar
Price, T. D., Grupe, G., and Schröter, P. (1994). Reconstruction of migration patterns in the Bell Beaker period by stable strontium isotope analysis. Applied Geochemistry 9, 413–417.CrossRefGoogle Scholar
Price, T. D., Burton, J. H., and Bentley, R. A. (2002). The characterization of biologically available strontium isotope ratios for the study of prehistoric migration. Archaeometry 44, 117–135.CrossRefGoogle Scholar
Priestley, J. (1770). Experiments and observations on charcoal. Philosophical Transactions 14, 211–227.CrossRefGoogle Scholar
Qian, Y., Engel, M. H., Goodfriend, G. A., and Macko, S. A. (1995). Abundance and stable isotopic composition of amino acids in molecular weight fractions of fossil and artificially aged mollusk shells. Geochimica et Cosmochimica Acta 59, 1113–1124.CrossRefGoogle Scholar
Quitmyer, I. R., and Jones, D. S. (1997). The sclerochronology of hard clams, Mercenaria spp., from the south eastern U.S.A.: a method for elucidating the zooarchaeological records of seasonal resource procurement and seasonality in prehistoric shell middens. Journal of Archaeological Science 24, 825–840.CrossRefGoogle Scholar
Ramaswamy, K., and Kamalakkannan, M. (1995). Infrared study of the influence of temperature on clay minerals. Journal of Thermal Analysis 44, 629–638.CrossRefGoogle Scholar
Rapp, G., and Hill, C. L. (1998). Geoarchaeology. Yale University Press, New Haven, CT.Google Scholar
Rebollo, N. R., Cohen-Ofri, I., Popovitz-Biro, R., Bar-Yosef, O., Meignen, L., Goldberg, P., Weiner, S., and Boaretto, E. (2008). Structural characterization of charcoal exposed to high and low pH: implications for 14C sample preparation and charcoal preservation. Radiocarbon 50, 289–307.CrossRefGoogle Scholar
Reiche, I., Vignaud, C., and Menu, M. (2002). The crystallinity of ancient bone and dentine: new insights by transmission electron microscopy. Archaeometry 44, 447–459.CrossRefGoogle Scholar
Reiche, I., Favre-Quattropani, L., Vignaud, C., Bocherens, H., Charlet, L., and Menu, M. (2003). A multi-analytical study of bone diagenesis: the Neolithic site of Bercy (Paris, France). Measurement Science Technology 14, 1608–1619.CrossRefGoogle Scholar
Reilly, P. (1989). Data visualization in archaeology. IBM Systems Journal 28, 569–570.CrossRefGoogle Scholar
Reimer, P. J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Bertrand, P. G., Blackwell, C. E., et al. (2004a). Radiocarbon calibration from 0–26 cal kyr BP. Radiocarbon 46, 1029–1058.CrossRefGoogle Scholar
Reimer, P. J., Baillie, M. G. L., Bard, E., Bayliss, A., Beck, J. W., Bertrand, C., Blackwell, P. G., et al. (2004b). IntCal04:Calibration issue. Radiocarbon 46, 1029–1058.CrossRefGoogle Scholar
Reinhard, K. J., and Danielson, D. R. (2005). Pervasiveness of phytoliths in prehistoric southwestern diet and implications for regional and temporal trends for dental microwear. Journal of Archaeological Science 32, 981–988.CrossRefGoogle Scholar
Renfrew, C., and Bahn, P. (1991). Archaeology: Theories, Methods and Practice. Thames and Hudson, London.Google Scholar
Renfrew, C., Dixon, J., and Cann, J. (1966). Obsidian and early cultural contact in the Near East. Proceedings of the Prehistoric Society 32, 30–72.CrossRefGoogle Scholar
Rey, C., Collins, B., Goehl, T., Dickson, I. R., and Glimcher, M. J. (1989). The carbonate environment in bone mineral: a resolution-enhanced Fourier transform infrared spectroscopy study. Calcified Tissue International 45, 157–164.CrossRefGoogle ScholarPubMed
Rey, C., Renugopalakrishnan, V., Collins, B., and Glimcher, M. J. (1991). Fourier transform infrared spectroscopic study of the carbonate ions in bone mineral during aging. Calcified Tissue International 49, 251–258.CrossRefGoogle ScholarPubMed
Rey, C., Miquel, J. L., Facchini, L., Legrand, A. P., and Glimcher, M. J. (1995). Hydroxyl groups in bone mineral. Bone 16, 583–586.CrossRefGoogle ScholarPubMed
Rheren, T., and Pernicka, E. (2007). Coins, artefacts and isotopes – archaeometallurgy and archaeometry. Archaeometry 50, 232–248.CrossRefGoogle Scholar
Rhoads, D. C., and Lutz, R. A. (1980). Skeletal Growth of Aquatic Organisms. Plenum Press, New York.CrossRefGoogle Scholar
Rhode, D. (2003). Coprolites from Hidden Cave revisited: evidence for site occupation history, diet and sex of occupants. Journal of Archaeological Science 30, 909–922.CrossRefGoogle Scholar
Rice, P. (1987). Pottery Analysis: A Source Book. University of Chicago Press, Chicago.Google Scholar
Righi, D., and Meunier, A. (1995). Origin of clays by rock weathering and soil formation. In: Origin and Mineralogy of Clays and the Environment (ed. Velde, B.), pp. 43–161. Springer, Berlin.CrossRefGoogle Scholar
Riley, T. (2008). Diet and seasonality in the Lower Pecos: evaluating coprolite data sets with cluster analysis. Journal of Archaeological Science 35, 2726–2741.CrossRefGoogle Scholar
Rink, W. J., Schwarcz, H. P., Weiner, S., Goldberg, P., Stiner, M. C., Meignen, L., and Bar-Yosef, O. (2004). Age of the Mousterian industry at Hayonim Cave, northern Israel, using electron spin resonance and 230Th/234U methods. Journal of Archaeological Science 31, 953–964.CrossRefGoogle Scholar
Robbins, L. H. (1973). Turkana material culture viewed from an archaeological perspective. World Archaeology 5, 209–214.CrossRefGoogle Scholar
Robinson, R. A. (1952). An electron microscopy study of the crystalline inorganic components of bone and its relationship to the organic matrix. Journal of Bone and Joint Surgery 34, 389–434.CrossRefGoogle Scholar
Robinson, R. A. (1979). Bone tissue: composition and function. Johns Hopkins Medical Journal 145, 10–24.Google ScholarPubMed
Rogers, A. K. (2008). Obsidian hydration dating: accuracy and resolution limitations imposed by intrinsic water variability. Journal of Archaeological Science 35, 2009–2016.CrossRefGoogle Scholar
Rollo, F., Ermini, L., Luciani, S., Marota, I., Olivieri, C., and Luiselli, D. (2006). Fine characterization of the Iceman's mtDNA haplogroup. American Journal of Physical Anthropology 130, 557–564.CrossRefGoogle ScholarPubMed
Rosen, A. (1986). Cities of Clay – The Geoarchaeology of Tells. University of Chicago Press, Chicago.Google Scholar
Rosen, A. M. (1987). Phytolith studies at Shiqmim. In: Shiqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982–1984) (ed. T. E. Levy). British Archaeological Reports International Series 356, 243–249.Google Scholar
Rosen, A. M. (1992). Preliminary identification of silica skeletons from Near Eastern archaeological sites. In: Phytolith Systematics: Emerging Issues (ed. Rapp, G. and Mulholland, S. C.), pp. 129–147. Plenum Press, New York.CrossRefGoogle Scholar
Rosen, A. M., and Weiner, S. (1994). Identifying ancient irrigation: a new method using opaline phytoliths from emmer wheat. Journal of Archaeological Science 21, 125–132.CrossRefGoogle Scholar
Russ, J., Hyman, M., Shafer, H. J., and Rowe, M. W. (1990). Radiocarbon dating of prehistoric rock paintings by selective oxidation of organic carbon. Nature 348, 710–711.CrossRefGoogle Scholar
Rypkema, H. A., Lee, W. E., Galaty, M. L., and Haws, J. (2007). Rapid, in-stride soil phosphate measurement in archaeological survey: a new method tested in Loudoun County, Virginia. Journal of Archaeological Science 34, 1859–1867.CrossRefGoogle Scholar
Sahin, N. (2004). Isolation and characterization of mesophilic, oxalate-degrading Streptomyces from plany rhizosphere and forest soils. Naturwissenschaften 91, 498–502.CrossRefGoogle Scholar
Salamon, M., Tuross, N., Arensburg, B., and Weiner, S. (2005). Relatively well preserved DNA is present in the crystal aggregates of fossil bones. Proceedings of the National Academy of Sciences of the United States of America 102, 13783–13788.CrossRefGoogle ScholarPubMed
Saleh, N., Deutsch, D., and Gil-Av, E. (1993). Racemization of aspartic acid in the extracellular matrix proteins of primary and secondary dentin. Calcified Tissue International 53, 103–110.CrossRefGoogle ScholarPubMed
Sangster, A. G., and Parry, D. W. (1969). Some factors in relation to bulliform cell silicification in the grass leaf. Annals of Botany 33, 315–323.CrossRefGoogle Scholar
Sangster, A. G., and Parry, D. W. (1981). Ultrastructure of silica deposits in higher plants. In: Silicon and Siliceous Structures in Biological Systems (ed. Simpson, T. L. and Volcani, B. E.), pp. 383–407. Springer, New York.CrossRefGoogle Scholar
Sanson, G. D., Kerr, S. A., and Gross, K. A. (2006). Do silica phytoliths really wear mammalian teeth?Journal of Archaeological Science 34, 526–531.CrossRefGoogle Scholar
Saxon, A., and Higham, C. (1969). A new research method for economic prehistorians. American Antiquity 34, 303–311.CrossRefGoogle Scholar
Sayre, E. V., and Dodson, R. W. (1957). Neutron activation study of Mediterranean potsherds. American Journal of Archaeology 61, 35–41.CrossRefGoogle Scholar
Scheffer, T. C., and Cowling, E. B. (1966). Natural resistance of wood to microbial deterioration. Annual Review of Phytopathology 4, 147–170.CrossRefGoogle Scholar
Schieber, J., Krinsley, D., and Riciputi, L. (2000). Diagenetic origin of quartz silt in mudstones and implications for silica cycling. Nature 406, 981–985.CrossRefGoogle ScholarPubMed
Schiegl, S., Lev-Yadun, S., Bar-Yosef, S., El Goresy, A., and Weiner, S. (1994). Siliceous aggregates from prehistoric wood ash: a major component of sediments in Kebara and Hayonim caves (Israel). Israel Journal Earth Sciences 43, 267–278.Google Scholar
Schiegl, S., 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 Archaeological Science 23, 763–781.CrossRefGoogle Scholar
Schiegl, S., Goldberg, P., Pfretzschner, H. U., and Conard, N. J. (2003). Paleolithic burnt bone horizons from the Swabian Jura: distinguishing between in situ fire places and dumping areas. Geoarchaeology 18, 541–565.CrossRefGoogle Scholar
Schiffer, M. B. (1983). Toward the identification of formation processes. American Antiquity 48, 675–706.CrossRefGoogle Scholar
Schiffer, M. B. (1986). Radiocarbon dating and the “old wood” problem: the case of the Hohokam chronology. Journal of Archaeological Science 13, 13–30.CrossRefGoogle Scholar
Schiffer, M. B. (1987). Formation Processes of the Archaeological Record. University of New Mexico Press, Albuquerque.Google Scholar
Schmidt, M., Botz, R., Rickert, D., Bohrmann, G., Hall, S. R., and Mann, S. (2001). Oxygen isotopes of marine diatoms and relations to opal-A maturation. Geochimica et Cosmochimica Acta 65, 201–211.CrossRefGoogle Scholar
Schmidt, M. W. I., Skjemstad, J. O., and Jäger, C. (2002). Carbon isotope geochemistry and nanomorphology of soil black carbon: black chernozemic soils in central Europe originate from ancient biomass burning. Global Biogeochemical Cycles 16, 70–1–70-8.CrossRefGoogle Scholar
Schmidt, W. J. (1936). Uber die Kristallorientierung im Zahnschmelz. Naturwissenschaften 24, 361.CrossRefGoogle Scholar
Schoeninger, M. J. (1979). Diet and status at Chalcatzingo: some empirical and technical aspects of strontium analysis. American Journal of Physical Anthropology 51, 295–310.CrossRefGoogle ScholarPubMed
Schoeninger, M. J., and DeNiro, M. J. (1984). Nitrogen and carbon isotopic composition of bone collagen from marine and terrestrial animals. Geochimica et Cosmochimica Acta 48, 625–639.CrossRefGoogle Scholar
Schubert, P. (1986). Petrographic modal analysis – a necessary complement to chemical analysis of ceramic course ware. Archaeometry 28, 163–178.CrossRefGoogle Scholar
Schultz, T. P., Curry Templeton, M., and McGinnis, G. D. (1985). Rapid determination of lignocellulose by diffuse reflectance Fourier transform infrared spectrometry. Analytical Chemistry 57, 2867–2869.CrossRefGoogle Scholar
Schurr, M. R., and Gregory, D. A. (2002). Fluoride dating of faunal materials by ion-selective electrode: high resolution relative dating at an early agricultural period site in the Tucson basin. Antiquity 67, 281–299.CrossRefGoogle Scholar
Schwarcz, H. P. (1980). Absolute age determination of archaeological sites by uranium series dating of travertines. Archaeometry 22, 3–24.CrossRefGoogle Scholar
Schwarcz, H. P., and Grün, R. (1992). Electron spin resonance (ESR) dating of the origin of modern man. Philosophical Transactions of the Royal Society of London, Series B 337, 145–148.CrossRefGoogle ScholarPubMed
Schwarcz, H. P., and Rink, W. J. (2001). Dating methods for sediments of caves and rockshelters with examples from the Mediterranean region. Geoarchaeology 16, 355–371.CrossRefGoogle Scholar
Schwedt, A., and Mommsen, H. (2007). On the influence of drying and firing of clay on the formation of trace element profiles within pottery. Archaeometry 49, 495–509.CrossRefGoogle Scholar
Schwedt, A., Mommsen, H., and Zacharias, N. (2004). Post-depositional elemental alterations in pottery: Neutron activation analyses of core and surface samples. Archaeometry 46, 85–101.CrossRefGoogle Scholar
Scott, A. C. (2000). The pre-Quaternary history of fire. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 281–329.CrossRefGoogle Scholar
Scott, E. M., Bryant, C., Cook, G. T., and Naysmith, P. (2003). Is there a fifth radiocarbon international intercomparison (VIRI)?Radiocarbon 45, 493–495.CrossRefGoogle Scholar
Scurfield, G., and Michell, A. J. (1973). Crystals in woody stems. Botanical Journal of the Linnaen Society 66, 277–289.CrossRefGoogle Scholar
Scurfield, G., Anderson, C. A., and Segnit, E. R. (1974). Silica in woody stems. Australian Journal of Botany 22, 211–229.CrossRefGoogle Scholar
Sealy, J. (2001). Body tissue chemistry and paleodiet. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 269–279. John Wiley, Chichester, UK.Google Scholar
Segal, I., Nathan, Y., Zbenovich, B., and Barzilay, E. (2005). Geochemical characterization of flint and chert artifacts from the Modi'in area. Israel Journal of Earth Sciences 54, 229–236.CrossRefGoogle 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 35, 965–982.CrossRefGoogle Scholar
Shahack-Gross, R., Shemesh, A., Yakir, D., and Weiner, S. (1996). Oxygen isotopic composition of opaline phytoliths: potential for terrestrial climatic reconstruction. Geochimica et Cosmochimica Acta 60, 3949–3953.CrossRefGoogle Scholar
Shahack-Gross, R., Bar-Yosef, O., and Weiner, S. (1997). Black-colored bones in Hayonim Cave, Israel: differentiating between burning and oxide staining. Journal of Archaeological Science 24, 439–446.CrossRefGoogle 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 30, 439–459.CrossRefGoogle Scholar
Shahack-Gross, R., Berna, F., Karkanas, P., and Weiner, S. (2004). Bat guano and preservation of archaelogical remains in cave sites. Journal of Archaeological Science 31, 1259–1272.CrossRefGoogle Scholar
Shahack-Gross, R., Albert, R. M., Gilboa, A., Nagar-Hilman, 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 32, 1417–1431.CrossRefGoogle Scholar
Shahack-Gross, R., Ayalon, A., Goldberg, P., Goren, Y., Ofek, B., Rabinovich, R., and Hovers, E. (2008). Formation processes of cemented features in karstic cave sites revealed using stable oxygen and carbon isotopic analyses: a case study at Middle Paleolithic Amud Cave, Israel. Geoarchaeology 23, 43–62.CrossRefGoogle Scholar
Shemesh, A. (1990). Crystallinity and diagenesis of sedimentary apatites. Geochimica et Cosmochimica Acta 54, 2433–2438.CrossRefGoogle Scholar
Shimoda, S., Aoba, T., Moreno, E. C., and Miake, Y. (1990). Effect of solution composition on morphological and structural features of carbonated calcium apatites. Journal of Dental Research 69, 1731–1740.CrossRefGoogle ScholarPubMed
Shipman, P., Foster, G. F., and Schoeninger, M. (1984). Burnt bones and teeth: an experimental study of colour, morphology, crystal structure and shrinkage. Journal of Archaeological Science 11, 307–325.CrossRefGoogle Scholar
Shoval, S. (1993). The burning temperature of a Persian-period pottery kiln at Tel Michal, Israel, estimated from the composition of slag-like material formed in its wall. Journal of Thermal Analysis 39, 1157–1168.Google Scholar
Shoval, S., Ginott, Y., and Nathan, Y. (1991). A new method for measuring the crystallinity index of quartz by infrared spectroscopy. Mineralogical Magazine 55, 579–582.CrossRefGoogle Scholar
Shoval, S., Graft, M., Beck, P., and Kirsh, Y. (1993). Thermal behavior of limestone and monocrystalline calcite tempers during firing and their use in ancient vessels. Journal of Thermal Analysis 40, 263–273.CrossRefGoogle Scholar
Shoval, S., Yofe, O., and Nathan, Y. (2003). Distinguishing between natural and recarbonated calcite in oil shales. Journal of Thermal Analysis and Calorimetry 71, 883–892.CrossRefGoogle Scholar
Siani, G., Paterne, M., Michel, E., Sulpizio, R., Sbrana, A., Arnold, M., and Haddad, G. (2001). Mediterranean sea surface radiocarbon reservoir age changes since the Last Glacial Maximum. Science 294, 1917–1920.CrossRefGoogle ScholarPubMed
Sillen, A. (1991). Solubility profiles of synthetic apatites and of modern and fossil bones. Journal of Archaeological Science 18, 385–397.CrossRefGoogle Scholar
Sillen, A., and Smith, P. (1984). Weaning patterns are reflected in strontium-calcium ratios of juvenile skeletons. Journal of Archaeological Science 11, 237–245.CrossRefGoogle Scholar
Sillen, A., Hall, G., and Armstrong, R. (1998). 87Sr/86Sr ratios in modern and fossil food-webs of the Sterkfontian Valley: implications for early hominid habitat preference. Geochimica et Cosmochimica Acta 62, 2463–2478.CrossRefGoogle Scholar
Simkiss, K. (1968). The structure and formation of the shell and shell membranes. In: Egg Quality: A Study of the Hen's Egg (ed. Carter, T. C.), pp. 3–25. Oliver and Boyd, Edinburgh.Google Scholar
Simkiss, K., and Wilbur, K. (1989). Biomineralization: Cell Biology and Mineral Deposition. Academic Press, San Diego, CA.Google Scholar
Simms, S. R. (1988). The archaeological structure of a Bedouin camp. Journal of Archaeological Science 15, 197–211.CrossRefGoogle Scholar
Simpson, I. A., Dockrill, S. J., Bull, I. D., and Evershed, R. P. (1998). Early anthropogenic soil formation at Toft Ness, Sanday, Orkney. Journal of Archaeological Science 25, 729–746.CrossRefGoogle Scholar
Simpson, I. A., Vésteinsson, O., Adderly, W. P., and McGovern, T. H. (2003). Fuel resource utilization in landscapes of settlement. Journal of Archaeological Science 30, 1401–1420.CrossRefGoogle Scholar
Smith, C. I., Craig, O. C., Prigodich, R. V., Nielsen-Marsh, C. M., Jans, M. M. E., Vermeer, C., and Collins, M. J. (2005). Diagenesis and survival of osteocalcin in archaeological bone. Journal of Archaeological Science 32, 105–113.CrossRefGoogle Scholar
Smith, H. I. (1899). The ethnological arrangement of archaeological material. Report of the Museums Association of the United Kingdom 1898.
Smith, T. M., Reid, D. J., and Sirianni, J. E. (2006). The accuracy of histological assessments of dental development and age at depth. Journal of Anatomy 208, 125–138.CrossRefGoogle Scholar
Solano, M. L., Iriarte, F., Ciria, P., and Negro, M. J. (2001). Performance characteristics of three aeration systems in the composting of sheep manure and straw. Journal of Agricultural Engineering Research 79, 317–329.CrossRefGoogle Scholar
Soltes, E. J., and Elder, T. J. (1981). Pyrolysis of Organic Chemicals from Biomass. CRC Press, Boca Raton, FL.Google Scholar
Sondi, I., and Slovenic, D. (2003). The mineralogical characteristics of the Lamboglia 2 Roman-age amphorae from the central Adriatic (Croatia). Archaeometry 45, 251–262.CrossRefGoogle Scholar
Soudry, D., and Nathan, Y. (2001). Diagenetic trends of fluorine concentration in Negev phosphorites, Israel: implications for carbonate fluorapatite composition during phophogenesis. Sedimentology 48, 723–743.CrossRefGoogle Scholar
Sponheimer, M., and Lee-Thorp, J. A. (1999). Alteration of enamel carbonate environments during fossilization. Journal of Archaeological Science 26, 143–150.CrossRefGoogle Scholar
Sponheimer, M., and Lee-Thorp, J. A. (2006). Enamel diagenesis at South African Australopith sites: implications for paleoecological reconstruction with trace elements. Geochimica et Cosmochimica Acta 70, 1644–1654.CrossRefGoogle Scholar
Stafford, T. W, Hare, P. E., Currie, L., Jull, A. J. T., and Donahue, D. J. (1991). Accelerator radiocarbon dating at the molecular level. Journal of Archaeological Science, 18, 35–72CrossRefGoogle Scholar
Stankiewicz, B. A., Hutchins, J. C., Thomson, R., Briggs, D. E. G., and Evershed, R. P. (1998). Assessment of bog-body tissue preservation by pyrolysis-gas chromatography/mass spectrometry. Rapid Communications in Mass Spectrometry 11, 1884–1890.3.0.CO;2-5>CrossRefGoogle Scholar
Stanley, R. G., and Linskens, H. F. (1974). Pollen: Biology, Biochemistry, Management. Springer, Berlin.CrossRefGoogle Scholar
Stein, J. K., and Farrand, W. R. (2001). Sediments in Archaeological Context. University of Utah Press, Salt Lake City.Google Scholar
Stenzel, H. B. (1963). Aragonite and calcite as constituents of adult oyster shells. Science 142, 232–233.CrossRefGoogle ScholarPubMed
Stern, L. A., Johnson, G. D., and Chamberlain, C. P. (1994). Carbon isotope signature of environmental change found in fossil ratite eggshells from a South Asian Neogene sequence. Geology 22, 419–422.2.3.CO;2>CrossRefGoogle Scholar
Stevenson, M. G. (1985). The formation of artifact assemblages at workshops/habitation sites: models from Peace Point in northern Alberta. American Antiquity 50, 63–81.CrossRefGoogle Scholar
Stiner, M., Kuhn, S., Weiner, S., and Bar-Yosef, O. (1995). Differential burning, recrystallization, and fragmentation of archaeological bone. Journal of Archaeological Science 22, 223–237.CrossRefGoogle Scholar
Stos-Gale, Z. A., Maliotis, G., Gale, N. H., and Annetts, N. (1997). Lead isotope characteristics of the Cyprus copper ore deposits applied to provenance studies of copper oxhide ingots. Archaeometry 39, 83–124.CrossRefGoogle Scholar
Stott, A. W., Evershed, R. P., Jim, S., Jones, V., Rogers, J. M., Tuross, N., and Ambrose, S. (1999). Cholesterol as a new source of paleodietary information: experimental approaches and archaeological applications. Journal of Archaeological Science 26, 705–716.CrossRefGoogle Scholar
Stott, A. W., Berstan, R., Evershed, P., Hedges, R. E. M., Bronk-Ramsey, C., and Humm, M. J. (2001). Radiocarbon dating of single compounds isolated from pottery cooking vessel residue. Radiocarbon 43, 191–197.CrossRefGoogle Scholar
Stuiver, M., Kromer, B., Becker, B., and Ferguson, C. W. (1986). Radiocarbon age calibration back to 13,300 years BP and the 14C age matching of the German oak and US Bristlecone pine chronologies. Radiocarbon 28, 969–979.CrossRefGoogle Scholar
Stumm, W., and Morgan, J. J. (1970). Aquatic Chemistry. John Wiley, New York.Google Scholar
Stutz, A. J. (2002). Polarizing microscopy identification of chemical diagenesis in archaeological cementum. Journal of Archaeological Science 29, 1327–1347.CrossRefGoogle Scholar
Suetsugu, Y., and Tanaka, J. (1999). Crystal growth of carbonate apatite using a CaCO3 flux. Journal of Materials Science: Materials in Medicine 10, 561–566.Google ScholarPubMed
Suga, S. (1984). The role of fluoride and iron in mineralization of fish enameloid. In: Tooth Enamel, vol. 4 (ed. Fearnhead, R. W. and Suga, S.), pp. 472–477. Elsevier, Amsterdam.Google Scholar
Suga, S., Taki, Y., Wada, K., and Ogawa, M. (1991). Evolution of fluoride and iron concentrations in the enameloid of fish teeth. In: Mechanisms and Phylogeny of Mineralization in Biological Systems (ed. Suga, S. and Nakahara, H.), pp. 439–446. Springer, Tokyo.CrossRefGoogle Scholar
Sumper, M., and Kröger, N. (2004). Silica formation in diatoms: the function of long-chain polyamines and silaffins. Journal of Materials Chemistry 14, 2059–2065.CrossRefGoogle Scholar
Surge, D., and Walker, K. J. (2005). Oxygen isotopic composition of modern and archaeological otoliths from the estuarine hardhead catfish (Ariopsis felis) and their potential to record low latitude climate change. Palaeogeography, Palaeoclimatology, Palaeoecology 228, 179–191.CrossRefGoogle Scholar
Surovell, T. A., and Stiner, M. C. (2001). Standardizing infrared measures of bone mineral crystallinity: an experimental approach. Journal of Archaeological Science 28, 633–642.CrossRefGoogle Scholar
Sutcliffe, A. (1970). Spotted hyaena: crusher, gnawer, digester and collector of bones. Nature 227, 1110–1113.CrossRefGoogle ScholarPubMed
Swider, J. R., Hackley, V. A., and Winter, J. (2003). Characterization of Chinese ink in size and surface. Journal of Cultural Heritage 4, 175–186.CrossRefGoogle Scholar
Sykes, G. A., Collins, M. J., and Walton, D. I. (1995). The significance of a geochemically isolated intracrystalline organic fraction within biominerals. Organic Geochemistry 23, 1059–1065.CrossRefGoogle Scholar
Sykes, N. J., White, J., Hayes, T. E., and Palmer, M. R. (2006). Tracking animals using strontium isotopes in teeth: the role of fallow deer (Dama dama) in Roman Britain. Antiquity 80, 948–959.CrossRefGoogle Scholar
Tan, K. H. (2003). Humic Matter in Soil and the Environment. Marcel Dekker, New York.CrossRefGoogle Scholar
Taylor, A., and Gurney, E. (1961). Solubilities of potassium and ammonium taranakites. Journal of Physical Chemistry 65, 1613–1616.CrossRefGoogle Scholar
Taylor, H. P., O'Neil, J. R., and Kaplan, I. R. (1991). Stable Isotope Geochemistry: A Tribute to Samuel Epstein. Special Publication 3. Geochemical Society, San Antonio, TX.Google Scholar
Taylor, R. E. (2001). Radiocarbon dating. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 23–34, John Wiley, Chichester, UK.Google Scholar
Tchernov, E., and Valla, F. F. (1997). Two new dogs, and other Natufian dogs, from the Southern Levant. Journal of Archaeological Science 24, 65–95.CrossRefGoogle Scholar
Teilhard de Chardin, P., and Young, C. C. (1929). Preliminary report on the Choukoutien fossiliferous deposit. Bulletin of the Geological Society of China 8, 173–202.CrossRefGoogle Scholar
Ten Cate, A. R. (1989). Oral Histology: Development, Structure and Function. Mosby, St. Louis, MO.Google Scholar
Termine, J. D. (1984). The tissue-specific proteins of the bone matrix. In: The Chemistry and Biology of Mineralized Tissues (ed. Butler, W. T.), pp. 94–97. Ebsco Media, Birmingham, AL.Google Scholar
Termine, J. D., and Posner, A. S. (1966). Infra-red determination of percentage of crystallinity in apatitic calcium phosphates. Nature 211, 268–270.CrossRefGoogle ScholarPubMed
Termine, J. D., Belcourt, A. B., Conn, K. M., and Kleinman, H. K. (1981). Mineral and collagen-binding proteins of fetal calf bone. Journal of Biological Chemistry 256, 10403–10408.Google ScholarPubMed
Terri, J. A., and Stowe, L. G. (1976). Climatic patterns and the distribution of C4 grasses in North America. Oecologia 23, 1–12.CrossRefGoogle Scholar
Terry, R. E., Fernández, F. G., Parnelli, J. J., and Inomata, T. (2004). The story in the floors: chemical signatures of ancient and modern Maya activities at Aguateca, Guatemala. Journal of Archaeological Science 31, 1237–1250.CrossRefGoogle Scholar
Thellier, E. (1938). Sur l'aimantation des terres cuities et ses applications geophysiques. Annal Institute de Physique du Globe of Paris 16, 157–302.Google Scholar
Théry, I., Gril, J., Vernet, J. L., Meignen, L., and Maury, J. (1996). Coal used for fuel at two prehistoric sites in southern France: Les Canalettes (Mousterian) and Les Usclades (Mesolithic). Journal of Archaeological Science 23, 509–512.CrossRefGoogle Scholar
Thiebault, S. (1988). Paleoenvironment and ancient vegetation of Baluchistan based on charcoal analysis of archaeological sites. Proceedings of the Indian National Science Academy 54, 501–509.Google Scholar
Thieme, H. (1997). Lower Paleolithic hunting spears from Germany. Nature 385, 807–810.CrossRefGoogle Scholar
Thompson, D. (1942). On Growth and Form. 2nd edition. Cambridge University Press, Cambridge.Google Scholar
Tindall, J. A., and Kunkel, J. R. (1999). Unsaturated Zone Hydrology. Prentice Hall, Upper Saddle River, NJ.Google Scholar
Tite, M. S., Kilikoglou, V., and Vekinis, G. (2001). Strength, toughness and thermal shock resistance of ancient ceramics, and their influence on technological choice. Archaeometry 43, 301–324.CrossRefGoogle Scholar
Toivanen, T.-J., and Alen, R. (2006). Variations in the chemical composition within pine (Pinus sylvestris) trunks determined by diffuse reflectance infrared spectroscopy and chemometrics. Cellulose 13, 53–61.CrossRefGoogle Scholar
Towe, K. M., and Thompson, G. R. (1972). The structure of some bivalve shell carbonates prepared by ion-beam thinning. Calcified Tissue Research 10, 38–48.CrossRefGoogle ScholarPubMed
Traub, W., Arad, T., and Weiner, S. (1992). Growth of mineral crystals in turkey tendon collagen fibrils. Connective Tissue Research 28, 99–111.CrossRefGoogle Scholar
Trigger, B. G. (1989). A History of Archaeological Thought. Cambridge University Press, Cambridge.Google Scholar
Trombold, C. D., and Israde-Alcantara, I. (2005). Paleoenvironment and plant cultivation on terraces at La Quemada, Zacatecas, Mexico: the pollen, phytolith and diatom evidence. Journal of Archaeological Science 32, 341–353.CrossRefGoogle Scholar
Trueman, C. N. G., Behrensmeyer, A. K., Tuross, N., and Weiner, S. (2003). Mineralogical and compositional changes in bones exposed on soil surfaces in Amboseli National Park, Kenya: diagenetic mechanisms and the role of sediment pore fluids. Journal of Archaeological Science 31, 21–39.Google Scholar
Trueman, C. N., Field, J. H., Dortch, J., Charles, B., and Wroe, S. (2005). Prolonged coexistence of humans and megafauna in Pleistocene Australia. Proceedings of the National Academy of Sciences of the United States of America 102, 8381–8385.CrossRefGoogle ScholarPubMed
Trueman, C. N., Behrensmeyer, A. K., Potts, R., and Tuross, N. (2006). High-resolution records of location and stratigraphic provenance from the rare earth element composition of fossil bones. Geochimica et Cosmochimica Acta 70, 4343–4355.CrossRefGoogle Scholar
Trueman, C. N., Privat, K., and Field, J. (2008). Why do crystallinity values fail to predict the extent of diagenetic alteration of bone mineral?Palaeogeography, Palaeoclimatology, Palaeoecology 266, 160–167.CrossRefGoogle Scholar
Tsibiridou, F. (2000). Les Pomak dans la Thrace greque: Discours ethnique et practiques Socioculturelles. L'Harmattan, Paris.Google Scholar
Tsartsidou, G., Lev-Yadun, S., Albert, R. M., Miller-Rosen, A., Efstratiou, N., and Weiner, S. (2007). The phytolith archaeological record: strengths and weaknesses evaluated based on a quantitative modern reference collection from Greece. Journal of Archaeological Science 34, 1262–1275.CrossRefGoogle Scholar
Tsartsidou, G., Lev-Yadun, S., Efstratiou, N., and Weiner, S. (2008). Ethnoarchaeological study of phytolith assemblages from an agro-pastoral village in northern Greece (Sarakini): development and application of a Phytolith Difference Index. Journal of Archaeological Science 35, 600–613.CrossRefGoogle Scholar
Tuross, N., Behrensmeyer, A. K., and Eanes, E. D. (1989a). Sr increases and crystallinity changes in taphonomic and archaeological bone. Journal of Archaeological Science 16, 661–672.CrossRefGoogle Scholar
Tuross, N., Behrensmeyer, A. K., Eanes, E. D., and Fisher, D. L. (1989b). Molecular preservation and crystallographic alterations in a weathering sequence of wildebeest bones. Applied Geochemistry 4, 261–270.CrossRefGoogle Scholar
Urey, H. C. (1947). The thermodynamic properties of isotopic substances. Journal of the Chemical Society, 562–581.CrossRefGoogle ScholarPubMed
Vafiadou, A., Murray, A. S., and Liritzis, I. (2007). Optically stimulated luminescence (OSL) dating investigations of rock and underlying soil from three case studies. Journal of Archaeological Science 34, 1659–1669.CrossRefGoogle Scholar
Valla, F. R., Khalaily, H., Valladas, H., Kaltnecker, E., Bocquentin, F., Cabellos, T., Bar-Yosef Mayer, D. E., et al. (2007). Les Fouilles de Ain Mallaha (Eynan) de 2003 à 2005: Quatrième Rapport Préliminaire. Journal of the Israel Prehistoric Society 37, 135–383.Google Scholar
Valladas, H., Reyss, J. L., Joron, J. L., Valladas, G., Bar-Yosef, O., and Vandermeersch, B. (1988). Thermoluminescence dating of Mousterian Troto-Cro-Magnon remains from Israel and the origin of modern man. Nature 331, 614–616.CrossRefGoogle Scholar
Kooij, G. (2002). Ethnoarchaeology in the Near East. In: Moving Matters: Ethnoarchaeology in the Near East (ed. Wendrich, W. and Kooij, G.), pp. 29–44. CNWS, Leiden, Netherlands.Google Scholar
Merwe, N. J., and Vogel, J. C. (1978). 13C content of human collagen as a measure of prehistoric diet in woodland North America. Nature 276, 815–816.CrossRefGoogle ScholarPubMed
Plicht, J., Beck, J. W., Bard, E., Baillie, M. G. L., Blackwell, P. G., Buck, C. E., Friedrich, M., et al. (2004). NOTCAL04 – comparison/calibration 14C records 26–50Cal KYR BP. Radiocarbon 46, 1225–1238.CrossRefGoogle Scholar
Klinken, G. J. (1999). Bone collagen quality indicators for paleodietry and radiocarbon measurements. Journal of Archaeological Science 26, 687–695.CrossRefGoogle Scholar
Vavilin, V. A., Rytov, S. V., and Lokshina, L. Y. (1996). A description of hydrolysis kinetics in anaerobic degradation of particulate organic matter. Bioresource Technology 56, 229–237.CrossRefGoogle Scholar
Veis, A. (2003). Mineralization in organic matrix frameworks. In: Reviews in Mineralogy and Chemistry, vol. 54 (ed. Dove, P. M., DeYoreo, J. D., and Weiner, S.), pp. 249–289. Mineralogical Society of America/Geochemical Society, Washington, DC.Google Scholar
Veis, A., and Perry, A. (1967). The phosphoprotein of the dentin matrix. Biochemistry 6, 2409–2416.CrossRefGoogle ScholarPubMed
Velde, B., and Druc, I. C. (1999). Archaeological Ceramic Materials. Springer, Berlin.CrossRefGoogle Scholar
Verri, G., Barkai, R., Bordeanu, C., Gopher, A., Hass, M., Kubik, P., Montanari, E., et al. (2004). Flint mining in prehistory recorded by in situ produced cosmogenic 10Be. Proceedings of the National Academy of Sciences of the United States of America 101, 7880–7884.CrossRefGoogle ScholarPubMed
Verri, G., Barkai, R., Gopher, A., Hass, M., Kubik, P., Paul, M., Ronen, A., Weiner, S., and Boaretto, E. (2005). Flint procurement strategies in the Late Lower Palaeolithic recorded by in situ produced cosmogenic 10Be in Tabun and Qesem Caves (Israel). Journal of Archaeological Science 32, 207–213.CrossRefGoogle Scholar
Very, J.-M., and Baud, C.-A. (1984). X-ray diffraction of calcified tissues. In: Methods of Calcified Tissue Preparation (ed. Dickson, G. R.), pp. 369–390. Elsevier, Amsterdam.Google Scholar
Villaseñor, I., and Price, C. A. (2008). Technology and decay of magnesian lime plasters: the sculptures of the funerary crypt of Palque, Mexico. Journal of Archaeological Science 35, 1030–1039.CrossRefGoogle Scholar
Vogel, J. C., and Merwe, N. J. (1977). Isotopic evidence for early maize cultivation in New York State. American Antiquity 42, 238–242.CrossRefGoogle Scholar
Vogel, J. C., Visser, E., and Fuls, A. (2001). Suitability of ostrich eggshell for radiocarbon dating. Radiocarbon 43, 133–137.CrossRefGoogle Scholar
Post, L. (1944). The prospect for pollen analysis in the study of the earth's climatic history. New Phytologist 45, 193–217.CrossRefGoogle Scholar
Vote, E. (2002). Discovering Petra: archaeological analysis in VR. Computer Graphics in Art History and Archaeology, September–October, 38–50.Google Scholar
Wada, K. (1977). Allophane and imogolite. In: Minerals in Soil Environments (ed. Dixon, J. B. and Weed, S. B.), pp. 603–638. Soil Science Society of America, Madison, WI.Google Scholar
Wada, K. (1989). Allophane and imogolite. In: Minerals in Soil Environments, vol. 1 (ed. Dixon, J. B. and Weed, S. B.), pp. 1051–1088. Soil Science Society of America, Madison, WI.Google Scholar
Wainwright, S. A., Biggs, W. D., Currey, J. D., and Gosline, J. M. (1976). Mechanical Design in Organisms. Princeton University Press, Princeton, NJ.Google Scholar
Waksman, S. (1952). Soil Microbiology. John Wiley, New York.CrossRefGoogle Scholar
Wang, L., Nancollas, G. H., Henneman, Z. J., Klein, E., and Weiner, S. (2006). Nanosized particles in bone and dissolution insensitivity of bone mineral. Biointerphases, 1, 106–111.CrossRefGoogle ScholarPubMed
Wang, R. Z., and Weiner, S. (1998). Human root dentin: structural anisotropy and Vickers microhardness isotropy. Connective Tissue Research 39, 269–279.CrossRefGoogle ScholarPubMed
Wang, R. Z., Addadi, L., and Weiner, S. (1997). Design strategies of sea urchin teeth: structure, composition and micromechanical relations to function. Philosophical Transactions of the Royal Society of London, Series B 352, 469–480.CrossRefGoogle Scholar
Watabe, N. (1963). Decalcification of thin sections for electron microscope studies of crystal-matrix relationship in mollusk shells. Journal of Cell Biology 18, 701–703.CrossRefGoogle Scholar
Watchman, A. (1990). A summary of occurrences of oxalate-rich crusts in Australia. Rock Art Research 7, 44–50.Google Scholar
Watkinson, D. (2001). Maximizing the lifespans of archaeological objects. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 649–659. John Wiley, Chichester, UK.Google Scholar
Watson, P. J. (1979). Archaeological Ethnography in Western Iran. Viking Fund Publications in Anthropology. University of Arizona Press, Tucson.Google Scholar
Watson, P. J., LeBlanc, S. A., and Redman, C. L. (1984). Archaeological Explanation. Columbia University Press, New York.Google Scholar
Wattez, J., Courty, M. A., and Macphail, R. I. (1990). Burnt organo-mineral deposits related to animal and human activities in caves. In: Soil Micromorphology: A Basic and Applied Science (ed. Douglas, L. A.), pp. 431–439. Elsevier, Amsterdam.Google Scholar
Weaver, A. R., Kissel, D. E., Chen, F., West, L. T., Adkins, W., Rickman, D., and Luvali, J. C. (2004). Mapping soil pH buffering capacity of selected fields in the coastal plain. Soil Science Society of America Journal 68, 662–668.CrossRefGoogle Scholar
Webb, E. A., and Longstaffe, F. J. (2000). The oxygen isotopic composition of silica phytoliths and plant water in grasses: implications for the study of paleoclimate. Geochimica et Cosmochimica Acta 64, 767–780.CrossRefGoogle Scholar
Webb, M. A. (1999). Cell-mediated crystallization of calcium oxalates in plants. Plant Cell 11, 751–761.CrossRefGoogle ScholarPubMed
Weiner, J. S. (1955). The Piltdown Forgery. Oxford University Press, Oxford.Google Scholar
Weiner, S. (1985). Organic matrix-like macromolecules associated with the mineral phase of sea urchin skeletal plates and teeth. Journal of Experimental Zoology 234, 7–15.CrossRefGoogle Scholar
Weiner, S. (2008). Archaeology, archaeological science and integrative archaeology. Israel Journal of Earth Sciences 56, 57–61.CrossRefGoogle Scholar
Weiner, S., and Addadi, L. (1997). Design strategies in mineralized biological materials. Journal of Materials Chemistry 7, 689–702.CrossRefGoogle Scholar
Weiner, S., and Bar-Yosef, O. (1990). States of preservation of bones from prehistoric sites in the Near East: a survey. Journal of Archaeological Science 17, 187–196.CrossRefGoogle Scholar
Weiner, S., and Dove, P. M. (2003). An overview of biomineralization processes and the problem of the vital effect. In: Biomineralization: Reviews in Mineralogy and Geochemistry, vol. 54 (ed. Dove, P. M., DeYoreo, J. J., and Weiner, S.), pp. 1–29. Mineralogical Society of America/Geochemical Society, Washington, DC.Google Scholar
Weiner, S., and Goldberg, P. (1990). On-site Fourier transform infrared spectrometry at an archeological excavation. Spectroscopy 5, 46–50.Google Scholar
Weiner, S., and Price, P. A. (1986). Disaggregation of bone into crystals. Calcified Tissue International 39, 365–375.CrossRefGoogle ScholarPubMed
Weiner, S., and Traub, W. (1986). Organization of hydroxyapatite crystals within collagen fibrils. FEBS Letters 206, 262–266.CrossRefGoogle ScholarPubMed
Weiner, S., and Wagner, H. D. (1998). The material bone: structure–mechanical function relations. Annual Reviews of Material Sciences 28, 271–298.CrossRefGoogle Scholar
Weiner, S., Lowenstam, H. A., and Hood, L. (1976). Characterization of 80-million year old mollusk shell proteins. Proceedings of the National Academy of Sciences of the United States of America 73, 2541–2545.CrossRefGoogle ScholarPubMed
Weiner, S., Kustanovich, Z., Gil-Av, E., and Traub, W. (1980). Dead Sea Scroll parchments: unfolding of the collagen molecules and racemization of aspartic acid. Nature 287, 820–823.CrossRefGoogle Scholar
Weiner, S., Goldberg, P., and Bar-Yosef, O. (1993). Bone preservation in Kebara Cave, Israel using on-site Fourier transform infrared spectroscopy. Journal of Archaeological Science 20, 613–627.CrossRefGoogle Scholar
Weiner, S., Xu, Q., Goldberg, P., Liu, J., and Bar-Yosef, O. (1998). Evidence for the use of fire at Zhoukoudian, China. Science 281, 251–253.CrossRefGoogle ScholarPubMed
Weiner, S., Goldberg, P., and Bar-Yosef, O. (1999a). Overview of ash studies in two prehistoric caves in Israel: implications to field archaeology. In: The Practical Impact of Science on Near Eastern and Aegean Archaeology, Wiener Laboratory Monograph 3 (ed. Pike, S. and Gitin, S.), pp. 85–90. Archetype, London.Google Scholar
Weiner, S., Traub, W., and Wagner, H. D. (1999b). Lamellar bone: structure-function relations. Journal of Structural Biology 126, 241–255.CrossRefGoogle ScholarPubMed
Weiner, S., Veis, A., Beniash, E., Arad, T., Dillon, J. W., Sabsay, B., and Siddiqui, F. (1999c). Peritubular dentin formation: crystal organization and the macromolecular constituents in human teeth. Journal of Structural Biology 126, 27–41.CrossRefGoogle ScholarPubMed
Weiner, S., Addadi, L., and Wagner, H. (2000). Materials design in biology. Materials Science and Engineering C 11, 1–8.CrossRefGoogle Scholar
Weiner, S., Goldberg, P., and Bar-Yosef, O. (2002). Three dimensional distribution of minerals in the sediments of Hayonim Cave, Israel: diagenetic processes and archaeological implications. Journal of Archaeological Science 29, 1289–1308.CrossRefGoogle Scholar
Weiner, S., Sagi, I., and Addadi, L. (2005). Choosing the path less travelled. Science 309, 1027–1028.CrossRefGoogle Scholar
Weiner, S., Berna, F., Cohen, I., Shahack-Gross, R., Albert, R. M., Karkanas, P., Meignen, L., and Bar-Yosef, O. (2007). Mineral distributions in Kebara Cave: diagenesis and its effect on the archaeological record. In: Kebara Cave, Mt. Carmel, Israel: The Middle and Upper Paleolithic Archaeology, Part 1 (ed. Bar-Yosef, O. and Meignen, L.), pp. 131–146. Peabody Museum of Archaeology and Ethnology, Harvard University, Cambridge, MA.Google Scholar
Weiss, E., and Kislev, M. E. (2004). Plant remains as indicators for economic activity: a case study from Iron Age Ashkelon. Journal of Archaeological Science 31, 1–13.CrossRefGoogle Scholar
Weiss, E., and Kislev, M. E. (2007). Plant remains as a tool for reconstruction of the past environment, economy, and society: archaeobotany in Israel. Israel Journal of Earth Sciences 56, 163–173.CrossRefGoogle Scholar
Weiss, I. M., Tuross, N., Addadi, L., and Weiner, S. (2002). Mollusk larval shell formation: amorphous calcium carbonate is a precursor for aragonite. Journal of Experimental Zoology 293, 478–491.CrossRefGoogle ScholarPubMed
Werb, Z. (1992). The role of metalloproteinases and their inhibitors in matrix remodeling in mineralized tissue. In: Chemistry and Biology of Mineralized Tissues (ed. Slavkin, H. and Price, P. A.), pp. 321–327. Elsevier Science, Amsterdam.Google Scholar
Whitbread, I. K. (2001). Ceramic petrology, clay geochemistry and ceramic production – from technology to the mind of the potter. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 449–459. John Wiley, Chichester, UK.Google Scholar
White, D. J. (1997). Dental calculus: recent insights into occurrence, formation, prevention, removal and oral health effects of supragingival and subgingival deposits. European Journal of Oral Science 105, 508–522.CrossRefGoogle ScholarPubMed
White, J. R. (1978). Archaeological and chemical evidence for the earliest American use of raw coal as a fuel in ironmaking. Journal of Archaeological Science 5, 391–393.CrossRefGoogle Scholar
Whittemore, O., and Halsey, G. (1983). Pore structure characterized by mercury porosimetry. In: Advances in Materials Characterization: Materials Science Research, vol. 15 (ed. Rossington, D., Condrate, R., and Snyder, R.), pp. 147–158. Plenum Press, New York.CrossRefGoogle Scholar
Wilding, L. P. (1967). Radiocarbon dating of biogenetic opal. Science 156, 166–167.CrossRefGoogle ScholarPubMed
Williams-Thorpe, O., Potts, P. J., and Webb, P. C. (1999). Field portable non-destructive analysis of lithic archaeological samples by X-ray fluorescence using a mercury iodide detector: comparison with wavelength-dispersive XRF and a case study in British stone axe provenancing. Journal of Archaeological Science 26, 215–237.CrossRefGoogle Scholar
Wilson, E. O. (1998). Consilience. Alfred A. Knopf, New York.Google Scholar
Wilson, L., and Pollard, A. M. (2001). The provenance hypothesis. In: Handbook of Archaeological Sciences (ed. Brothwell, D. R. and Pollard, A. M.), pp. 507–517. John Wiley, Chichester, UK.Google Scholar
Wilson, R. W., Millero, F. J., Taylor, J. R., Walsh, P. J., Christensen, V., Jennings, S., and Grosell, M. (2009). Contributions of fish to the marine inorganic carbon cycle. Science 323, 359–362.CrossRefGoogle Scholar
Woods, W. I. (1977). The quantitative analysis of soil phosphate. American Antiquity 42, 248–252.CrossRefGoogle Scholar
Woot-Tsuen, W. L. (1968). Food composition table for use in Africa. Food and Agriculture Organization of the United Nations. http://www.fao.org/DOCREP/003/X6877E/X6877E02.htm.
Wright, L. E., and Schwarcz, H. P. (1996). Infrared and isotopic evidence for diagenesis of bone apatite at Dos Pilas, Guatemala: paleodietry implications. Journal of Archaeological Science 23, 933–944.CrossRefGoogle Scholar
Wyckoff, R. W. G. (1972). The Biochemistry of Animal Fossils. Scientechnica, Bristol, UK.Google Scholar
Xu, S., Sun, K., Cui, F. Z., and Landis, W. J. (2003). Organization of apatite crystals in human woven bone. Bone 32, 150–162.Google Scholar
Yasuda, Y. (2002). Origins of pottery and agriculture in East Asia. In: The Origins of Pottery and Agriculture (ed. Yasuda, Y.), pp. 119–142. Lustre Press, New Delhi.Google Scholar
Yellin, J. E. (1977). Archaeological Approaches to the Present. Academic Press, New York.Google Scholar
Yenner, A. K., Ösban, H., Kaptan, E., Pehlivan, A. N., and Goodway, M. (1989). Kestel: an Early Bronze age source of tin ore in the Taurus Mountains, Turkey. Science 244, 200–203.CrossRefGoogle Scholar
Yizhaq, M. (2004). Characterizing and dating of the Early PPNB layer at the site of Motza. M.S. thesis, Weizmann Institute of Science, Rehovot, Israel.Google Scholar
Yizhaq, M., Mintz, G., Cohen, I., Khalally, H., Weiner, S., and Boaretto, E. (2005). Quality controlled radiocarbon dating of bones and charcoal from the early Pre-Pottery Neolithic B (PPNB) of Motza (Israel). Radiocarbon 47, 193–206.CrossRefGoogle Scholar
Yoshida, S., Ohnishi, Y., and Kitagishi, K. (1959). The chemical nature of silicon in the rice plant. Soil and Plant Food 5, 23–27.CrossRefGoogle Scholar
Zaslansky, P., Friesem, A. A., and Weiner, S. (2006a). Structure and mechanical properties of the soft zone separating bulk dentin and enamel in crowns of human teeth: insight into tooth function. Journal of Structural Biology 153, 188–199.CrossRefGoogle ScholarPubMed
Zaslansky, P., Shahar, R., Friesem, A. A., and Weiner, S. (2006b). Relations between shape, materials properties and function in biological materials using laser speckle interferometry: in-situ tooth deformation. Advanced Functional Materials 16, 1925–1936.CrossRefGoogle Scholar
Zeller, E. J. (1968). Use of electron spin resonance for measurement of natural radiation damage. In: Thermoluminescence of Geological Materials (ed. McDougall, D. J.), pp. 271–279. Academic Press, London.Google Scholar
Zevenhoven-Onderwater, M., Blomquist, J. P., Skrifvars, B. J., Backman, R., and Hupa, M. (2000). The prediction of behaviour of ashes from five different solid fuels in fluidised bed combustion. Fuel 79, 1353–1361.CrossRefGoogle Scholar
Zhang, L., and Gellerstedt, G. (2001). NMR observation of a new lignin structure, a spiro-dienone. Chemical Communications 24, 2744–2745.CrossRefGoogle Scholar
Zhao, Z. (1998). The middle Yangtze region in China is one place where rice was domesticated: phytolith evidence from the Diaotonghuan Cave, northern Jiangxi. Antiquity 72, 885–897.Google Scholar
Ziv, V., and Weiner, S. (1994). Bone crystal sizes: a comparison of transmission electron microscopic and X-ray diffraction line width broadening techniques. Connective Tissue Research 30, 165–175.CrossRefGoogle ScholarPubMed
Zohary, D., and Hopf, M. (2001). Domestication of Plants in the Old World. Oxford University Press, Oxford.Google Scholar
zur Nedden, D., Wicke, K., Knapp, R., Deidler, H., Wilfing, H., and Weber, G. (1994). New findings of the Tyrolean “Ice Man”: archaeological and CT-body analysis suggest personal disaster before death. Journal of Archaeological Science 21, 809–818.CrossRefGoogle 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
  • Stephen Weiner, Weizmann Institute of Science, Israel
  • Book: Microarchaeology
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811210.016
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
  • Stephen Weiner, Weizmann Institute of Science, Israel
  • Book: Microarchaeology
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811210.016
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
  • Stephen Weiner, Weizmann Institute of Science, Israel
  • Book: Microarchaeology
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511811210.016
Available formats
×